0% found this document useful (0 votes)
412 views527 pages

Handbook of II-VI Semiconductor-Based Sensors and Radiation Detectors

This document provides the preface to a three-volume handbook on II-VI semiconductor-based sensors and radiation detectors. It introduces the editor and provides an overview of the scope and purpose of the handbook series. The preface describes how II-VI semiconductors have unique properties making them suitable for a wide range of applications, including photodetectors, solar cells, lasers, and sensors. It notes that while previous books have covered aspects of II-VI semiconductors, this handbook aims to provide a comprehensive and up-to-date resource covering the materials, technologies, and applications of II-VI semiconductor sensors and detectors.

Uploaded by

prvthesap
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
412 views527 pages

Handbook of II-VI Semiconductor-Based Sensors and Radiation Detectors

This document provides the preface to a three-volume handbook on II-VI semiconductor-based sensors and radiation detectors. It introduces the editor and provides an overview of the scope and purpose of the handbook series. The preface describes how II-VI semiconductors have unique properties making them suitable for a wide range of applications, including photodetectors, solar cells, lasers, and sensors. It notes that while previous books have covered aspects of II-VI semiconductors, this handbook aims to provide a comprehensive and up-to-date resource covering the materials, technologies, and applications of II-VI semiconductor sensors and detectors.

Uploaded by

prvthesap
Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 527

Ghenadii Korotcenkov Editor

Handbook of II-VI
Semiconductor-Based
Sensors and Radiation
Detectors
Volume 2, Photodetectors
Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors
Ghenadii Korotcenkov
Editor

Handbook of II-VI
Semiconductor-Based
Sensors and Radiation
Detectors
Volume 2, Photodetectors
Editor
Ghenadii Korotcenkov
Department of Physics and Engineering
Moldova State University
Chisinau, Moldova

ISBN 978-3-031-20509-5    ISBN 978-3-031-20510-1 (eBook)


https://doi.org/10.1007/978-3-031-20510-1

© The Editor(s) (if applicable) and The Author(s), under exclusive license to Springer Nature
Switzerland AG 2023
This work is subject to copyright. All rights are solely and exclusively licensed by the Publisher, whether
the whole or part of the material is concerned, specifically the rights of translation, reprinting, reuse of
illustrations, recitation, broadcasting, reproduction on microfilms or in any other physical way, and
transmission or information storage and retrieval, electronic adaptation, computer software, or by similar
or dissimilar methodology now known or hereafter developed.
The use of general descriptive names, registered names, trademarks, service marks, etc. in this publication
does not imply, even in the absence of a specific statement, that such names are exempt from the relevant
protective laws and regulations and therefore free for general use.
The publisher, the authors, and the editors are safe to assume that the advice and information in this book
are believed to be true and accurate at the date of publication. Neither the publisher nor the authors or the
editors give a warranty, expressed or implied, with respect to the material contained herein or for any
errors or omissions that may have been made. The publisher remains neutral with regard to jurisdictional
claims in published maps and institutional affiliations.

This Springer imprint is published by the registered company Springer Nature Switzerland AG
The registered company address is: Gewerbestrasse 11, 6330 Cham, Switzerland
Preface

Binary and ternary semiconductors of II-VI group (ZnS, ZnSe, ZnTe, CdS, CdSe,
CdTe, HgTe, HgS, HgSe, HgCdTe, CdZnTe, CdSSe, and HgZnTe) are very popular
among researchers because of their remarkable physical and chemical properties,
which, as a group, are unique. II-VI compounds possess a very wide spectrum of
electronic and optical properties. Most materials of group II-VI are semiconductors
with a direct band gap and high optical absorption and emission coefficients. In
addition, binary II-VI compounds are easily miscible, providing a continuous range
of properties. As results, the II-VI semiconductors possess band gap, varying over a
wide range. Therefore, II-VI compounds can serve as efficient light emitters, such
as light diodes and lasers, solar cells, and radiation detectors operating in the range
from IR to UV and X-ray. II-VI compound-based devices can also cover terahertz
range. Besides common photovoltaic applications, II-VI semiconductors are also
potential candidates for a variety of electronic, electro-optical, sensing, and piezo-
electric devices. In particular, nanoparticles of II-VI semiconductors, such as quan-
tum dots, one-dimensional structures, and core-shells structures, can be used for
development of gas sensors, electrochemical sensors, and biosensors. These semi-
conductors, when downsized to nanometer, have become the focus of attention
because of their tunable band structure, high extinction coefficient, possible multi-
ple exciton generation, and unique electronic and transport properties. It is impor-
tant that II-VI semiconductors can be easily prepared in high quality epitaxial,
polycrystalline, and nanocrystalline films. The concentration of charge carriers can
also vary in II-VI semiconductors in wide range due to doping. Thus, the use of
II-VI films represents an economical approach to the synthesis of semiconductors
for various applications. It should be noted that the range of technical applications
for II-VI compounds goes beyond the better-known semiconductors such as Si, Ge,
and some of III-V compounds.
Formally, metal oxides such as CdO and ZnO also belong to II-VI compounds.
However, we will not cover them in this book. In recent years, these compounds
have been allocated to a separate group, “metal oxides,” and many books have been
devoted to their discussion, in contrast to other II-VI compounds. In particular,

v
vi Preface

those who are interested in exactly these compounds, we can recommend the Metal
Oxides series which is published by Elsevier.
The aim of this three-volume book is to provide an updated account of the state
of the art of multifunctional II-VI semiconductors, from fundamental sciences and
material sciences to their applications as various sensors and radiation detectors,
and, based on this knowledge, formulate new goals for further research. This book
provides interdisciplinary discussion of a wide range of topics, such as synthesis of
II-VI compounds, their deposition, processing, characterization, device fabrication,
and testing. Topics of the recent remarkable progresses in application of nanoparti-
cles, nanocomposites, and nanostructures consisting of II-VI semiconductors in
various devices are also covered. Both experimental and theoretical approaches
were used for this analysis.
Currently, there exist books on II-VI semiconductors. However, some of them
were published too long ago and cannot reflect the current state of research in this
area. Other published books focus on a limited number of topics, from which topics
related to various sensor applications such as gas sensors, humidity sensors, and
biosensors are almost completely excluded. When considering photodetectors, the
focus is also only on the analysis of IR photodetectors. Although sensors operating
in the visible, ultraviolet, terahertz, and X-ray ranges also hold great promise for
applications. With these books, we will try to close this gap.
Our three-volume book Handbook of II-VI Semiconductor-Based Sensors and
Radiation Detectors is the first to cover both chemical sensors and biosensors and
all types of photodetectors and radiation detectors based on II-VI semiconductors.
It contains a comprehensive and detailed analysis of all aspects of the application of
II-VI semiconductors in these devices. This makes these books very useful and
comfortable to use. Combining this information in three volumes, united by com-
mon topics, should help readers in finding the necessary information on required
subject.
Chapters in Handbook of II-VI Semiconductor-Based Sensors and Radiation
Detectors. Vol. 1: Materials and Technologies describe the physical, chemical, and
electronic properties of II-VI compounds, which give rise to an increased interest in
these semiconductors. Technologies that are used in the development of various
devices based on II-VI connections are also discussed in detail in this volume.
Handbook of II-VI Semiconductor-Based Sensors and Radiation Detectors. Vol.
2: Photodetectors focuses on the consideration of all types of optical detectors,
including IR detectors, visible detectors, and UV detectors. This consideration
includes both the fundamentals of the operation of detectors and the peculiarities of
their manufacture and use. An analysis of new trends in development of II-VI
semiconductors-­based photodetectors is also given.
Handbook of II-VI Semiconductor-Based Sensors and Radiation Detectors. Vol.
3: Sensors, Biosensors and Radiation Detector describes the use of II-VI com-
pounds in other fields such as radiation detectors, gas sensors, humidity sensors,
optical sensors, and biosensors. The chapters in this volume provide a comprehen-
sive overview of the manufacture, parameters, and applications of these devices.
Preface vii

We believe that these books will enable the reader to understand the present sta-
tus of II-VI semiconductors and their role in the development of new generation of
photodetectors, sensors, and radiation detectors. I am very pleased that many well-­
known experts with extensive experience in the development and research of II-VI
semiconductor sensors and radiation detectors were involved in the preparation of
the chapters of these books.
The target audience for this series of books are scientists and researchers work-
ing or planning to work in the field of materials related to II-VI semiconductors, i.e.,
scientists and researchers whose activities are related to electronics, optoelectron-
ics, chemical and bio sensors, electrical engineering, and biomedical applications. I
believe this three-volume book may also be of interest to practicing engineers and
project managers in industries and national laboratories who would like to develop
II-VI semiconductor-based radiation sensors and detectors but do not know how to
do it, and how to select the optimal II-VI semiconductor for specific applications.
With numerous references to an extensive resource of recently published literature
on the subject, these books can serve as an important and insightful source of valu-
able information, providing scientists and engineers with new ideas for understand-
ing and improving existing II-VI semiconductor devices.
I believe that these books will be very useful for university students, doctoral
students, and professors. The structure of these books offers the basis for courses in
materials science, chemical engineering, electronics, optoelectronics, environmen-
tal control, chemical sensors, photodetectors, radiation detectors, biomedical appli-
cations, and many others. Graduate students may also find the book very useful in
their research and understanding of the synthesis of II-VI semiconductors, study,
and application of this multifunctional material in various devices. We are confident
that all of them will find the information useful for their activities.
Finally, I thank all the authors who contributed to these books. I am grateful that
they agreed to participate in this project and for their efforts to prepare these chap-
ters. This project would not have been possible without their participation. I am also
very grateful to Springer for the opportunity to publish this book with their help. I
would like also to inform that my activity related to editing this book was funded by
the State Program of the Republic of Moldova project 20.80009.5007.02.
I am also grateful to my family and wife, who always support me in all my
endeavors.

Chisinau, Moldova Ghenadii Korotcenkov


Contents

Part I IR Detectors Based on II–VI Semiconductors


1 
Introduction in IR Detectors������������������������������������������������������������������    3
Ghenadii Korotcenkov
2 
Photoconductive and Photovoltaic IR Detectors����������������������������������   23
Rada Savkina and Oleksii Smirnov
3 II–VI Compound Semiconductor Avalanche Photodiodes
for the Infrared Spectral Region: Opportunities and Challenges������   53
K. -W. A. Chee
4 IR Detectors Array����������������������������������������������������������������������������������   79
Ghenadii Korotcenkov
5 New Trends and Approaches in the Development of Photonic
IR Detector Technology �������������������������������������������������������������������������� 107
Ghenadii Korotcenkov and Igor Pronin
6 II-VI Semiconductor-Based Unipolar Barrier Structures
for Infrared Photodetector Arrays �������������������������������������������������������� 135
A. V. Voitsekhovskii, S. N. Nesmelov, S. M. Dzyadukh, D. I. Gorn,
S. A. Dvoretsky, N. N. Mikhailov, G. Y. Sidorov, and M. V. Yakushev
7 
Infrared Sensing Using Mercury Chalcogenide Nanocrystals������������ 155
Emmanuel Lhuillier, Tung Huu Dang, Mariarosa Cavallo,
Claire Abadie, Adrien Khalili, John C. Peterson,
and Charlie Gréboval
8 
Graphene/HgCdTe Heterojunction-Based IR Detectors���������������������� 183
Shonak Bansal, M. Muthukumar, and Sandeep Kumar

ix
x Contents

Part II II–VI Semiconductors–Based Detectors for Visible


and UV Spectral Regions
9 
CdTe-Based Photodetectors and Solar Cells ���������������������������������������� 205
Alessio Bosio
10 
CdSe – Based Photodetectors for Visible-­NIR Spectral Region���������� 231
Hemant Kumar and Satyabrata Jit
11 
CdS-Based Photodetectors for Visible-UV Spectral Region���������������� 251
Nupur Saxena, Tania Kalsi, and Pragati Kumar
12 
ZnTe-Based Photodetectors for Visible-UV Spectral Region�������������� 281
Jiajia Ning
13 ZnSe-Based Photodetectors�������������������������������������������������������������������� 301
Ghenadii Korotcenkov
14 ZnS-Based UV Detectors������������������������������������������������������������������������ 333
Sema Ebrahimi, Benyamin Yarmand, and Nima Naderi
15 
Photodetectors Based on II-VI Multicomponent Alloys���������������������� 349
Ghenadii Korotcenkov and Tetyana Semikina

Part III New Trends in Development of II–VI Semiconductors–Based


Photodetectors
16 
Nanowire-Based Photodetectors for Visible-UV Spectral Region ������ 371
Ghenadii Korotcenkov and Victor V. Sysoev
17 QDs of Wide Band Gap II–VI Semiconductors Luminescent
Properties and Photodetector Applications ������������������������������������������ 399
M. Abdullah, Baqer O. Al-Nashy, Ghenadii Korotcenkov,
and Amin H. Al-Khursan
18 Solution-Processed Photodetectors�������������������������������������������������������� 427
Shaikh Khaled Mostaque, Abdul Kuddus, Md. Ferdous Rahman,
Ghenadii Korotcenkov, and Jaker Hossain
19 Multicolor Photodetectors���������������������������������������������������������������������� 453
Paweł Madejczyk
20 
Flexible Photodetectors Based on II-VI Semiconductors�������������������� 469
Mingfa Peng and Xuhui Sun
21 Self-Powered Photodetector�������������������������������������������������������������������� 495
Hemant Kumar and Satyabrata Jit

Index������������������������������������������������������������������������������������������������������������������ 517
About the Editor

Ghenadii Korotcenkov received his PhD in physics


and technology of semiconductor materials and devices
in 1976 and his Doctor of Science degree (doctor habili-
tate) in physics of semiconductors and dielectrics in
1990. He has more than 50-year experience as a teacher
and scientific researcher. For a long time he was a leader
of gas sensor group and manager of various national
and international scientific and engineering projects
carried out in the Laboratory of Micro- and
Optoelectronics, Technical University of Moldova,
Chisinau, Moldova. International foundations and pro-
grams such as the CRDF, the MRDA, the ICTP, the
INTAS, the INCO-COPERNICUS, the COST, and NATO have supported his
research. From 2007 to 2008, he carried out his research as an invited scientist at
Korea Institute of Energy Research (Daejeon). Then, from 2008 to 2018, Dr.
G. Korotcenkov was a research professor in the School of Materials Science and
Engineering at Gwangju Institute of Science and Technology (GIST) in Korea.
Currently, G. Korotcenkov is a chief scientific researcher at Moldova State
University, Chisinau, Moldova.Scientists from the former Soviet Union know the
results of G. Korotcenkov's research in the study of Schottky barriers, MOS struc-
tures, native oxides, and photoreceivers based on III–Vs compounds such as InP,
GaP, AlGaAs, and InGaAs. His current research interests since 1995 include mate-
rial sciences, focusing on metal oxide film deposition and characterization (In2O3,
SnO2, ZnO, TiO2), surface science, thermoelectric conversion, and design of physi-
cal and chemical sensors, including thin film gas sensors.G. Korotcenkov is the
author or editor of 45 books and special issues, including the 11-volume Chemical
Sensors series published by Momentum Press; 15-volume Chemical Sensors series
published by Harbin Institute of Technology Press, China; 3-volume “Porous
Silicon: From Formation to Application” issue published by CRC Press; 2-volume
Handbook of Gas Sensor Materials published by Springer; 3-volume Handbook of
Humidity Measurements published by CRC Press; and 6 proceedings of the

xi
xii About the Editor

international conferences published by Trans Tech Publ., Elsevier, and EDP


Sciences. In addition, currently he is a series editor of Metal Oxides book series
published by Elsevier. Since 2017, more than 35 volumes have been published
within this series.G. Korotcenkov is the author and coauthor of more than 650 sci-
entific publications, including 31 review papers, 38 book chapters, and more than
200 peer-­reviewed articles published in scientific journals (h-factor=42 (Web of
Science), h=44 (Scopus) and h=59 (Google scholar citation), 2022). He is the holder
of 17 patents. He presented more than 250 reports at national and international con-
ferences, including 17 invited talks. G. Korotcenkov, as a cochairman or member of
program, scientific, and steering committees, has participated in the organization of
more than 40 international scientific conferences. Dr. G. Korotcenkov is a member
of editorial boards of five scientific international journals. His name and activities
have been listed by many biographical publications including Who’s Who. His
research activities have been honored by the National Prize of the Republic of
Moldova (2022), the Honorary Diploma of the Government of the Republic of
Moldova (2020), an award of the Academy of Sciences of Moldova (2019), an
award of the Supreme Council of Science and Advanced Technology of the Republic
of Moldova (2004), the Prize of the Presidents of the Ukrainian, Belarus, and
Moldovan Academies of Sciences (2003), Senior Research Excellence Award of the
Technical University of Moldova (2001, 2003, 2005), the National Youth Prize of
the Republic of Moldova in the field of science and technology (1980), among oth-
ers. Some of his research results and published books have won awards at interna-
tional exhibitions. G. Korotcenkov also received a fellowship from the International
Research Exchange Board (IREX, United States, 1998), Brain Korea 21 Program
(2008–2012), and BrainPool Program (Korea, 2007–2008 and 2015–2017). https://
www.scopus.com/authid/detail.uri?authorId=6701490962 https://publons.com/
researcher/1490013/ghenadii-­korotcenkov/ https://scholar.google.com/citations?us
er=XR3RNhAAAAAJ&hl https://www.researchgate.net/profile/G_Korotcenkov
Contributors

Claire Abadie Sorbonne Université, CNRS, Institut des NanoSciences de Paris,


INSP, Paris, France
M. Abdullah Nasiriya Nanotechnology Research Laboratory (NNRL), College of
Science, University of Thi-Qar, Nasiriya, Iraq
Department of Physics, College of Science, University of Thi-Qar, Nasiriya, Iraq
Amin H. Al-Khursan Nasiriya Nanotechnology Research Laboratory (NNRL),
College of Science, University of Thi-Qar, Nasiriya, Iraq
Baqer O. Al-Nashy Department of Physics, College of Science, University of
Misan, Omarah, Iraq
Shonak Bansal Electronics and Communication Engineering Department,
Chandigarh University, Gharuan, India
Alessio Bosio Department of Mathematical, Physical and Computer Sciences,
University of Parma v.le delle Scienze, Parma, Italy
Mariarosa Cavallo Sorbonne Université, CNRS, Institut des NanoSciences de
Paris, INSP, Paris, France
K. -W. A. Chee School of Electronic and Electrical Engineering, Kyungpook
National University, Daegu, Republic of Korea
School of Electronics Engineering, College of IT Engineering, Kyungpook National
University, Daegu, Republic of Korea
Tung Huu Dang Sorbonne Université, CNRS, Institut des NanoSciences de Paris,
INSP, Paris, France
S. A. Dvoretsky Rzhanov Institute of Semiconductor Physics of the Siberian
Branch of the RAS, Novosibirsk, Russia
S. M. Dzyadukh National Research Tomsk State University, Tomsk, Russia
Sema Ebrahimi Materials and Energy Research Center (MERC), Karaj, Iran

xiii
xiv Contributors

Light, Nanomaterials, Nanotechnologies (L2n) Laboratory, CNRS EMR7004, The


University of Technology of Troyes, Troyes Cedex, France
Department of Physics and Mathematics, University of Hull, Cottingham Road, UK
G.W.Gray Centre for Advanced Materials, University of Hull, Cottingham Road, UK
D. I. Gorn National Research Tomsk State University, Tomsk, Russia
Charlie Gréboval Sorbonne Université, CNRS, Institut des NanoSciences de
Paris, INSP, Paris, France
Jaker Hossain Solar Energy Laboratory, Department of Electrical and Electronic
Engineering, University of Rajshahi, Rajshahi, Bangladesh
Satyabrata Jit Department of Electronics Engineering, Indian Institute of
Technology (Banaras Hindu University), Varanasi, India
Tania Kalsi Nano-Materials and Device Lab, Department of Nanoscience and
Materials, Central University of Jammu, Rahya-Suchani, Jammu & Kashmir, India
Adrien Khalili Sorbonne Université, CNRS, Institut des NanoSciences de Paris,
INSP, Paris, France
Ghenadii Korotcenkov Department of Physics and Engineering, Moldova State
University, Chisinau, Moldova
Abdul Kuddus Solar Energy Laboratory, Department of Electrical and Electronic
Engineering, University of Rajshahi, Rajshahi, Bangladesh
Graduate School of Science and Engineering, Saitama University, Saitama, Japan
Hemant Kumar Department of Electronics and Communication Engineering,
Jaypee Institute of Information Technology, Noida, India
Pragati Kumar Nano-Materials and Device Lab, Department of Nanoscience and
Materials, Central University of Jammu, Rahya-Suchani, Jammu & Kashmir, India
Sandeep Kumar ICAR-Central Institute for Subtropical Horticulture, Lucknow,
Uttar Pradesh, India
Emmanuel Lhuillier Sorbonne Université, CNRS, Institut des NanoSciences de
Paris, INSP, Paris, France
Paweł Madejczyk Military University of Technology, Warsaw, Poland
N. N. Mikhailov Rzhanov Institute of Semiconductor Physics of the Siberian
Branch of the RAS, Novosibirsk, Russia
Shaikh Khaled Mostaque Solar Energy Laboratory, Department of Electrical and
Electronic Engineering, University of Rajshahi, Rajshahi, Bangladesh
M. Muthukumar ICAR-Central Institute for Subtropical Horticulture, Lucknow,
Uttar Pradesh, India
Contributors xv

Nima Naderi Materials and Energy Research Center (MERC), Karaj, Iran
Photonics Research Centre, University of Malaya, Kuala Lumpur, Malaysia
S. N. Nesmelov National Research Tomsk State University, Tomsk, Russia
Jiajia Ning Key Laboratory of Physics and Technology for Advanced Batteries,
Ministry of Education, College of Physics, Jilin University, Changchun, China
Mingfa Peng School of Electronic and Information Engineering, Jiangsu Province
Key Laboratory of Advanced Functional Materials, Changshu Institute of
Technology, Changshu, Jiangsu, People’s Republic of China
John C. Peterson The James Franck Institute, The University of Chicago,
Chicago, IL, USA
Igor Pronin Department of Nano- and Microelectronics, Penza State University,
Penza, Russia
Md. Ferdous Rahman Department of Electrical and Electronic Engineering,
Begum Rokeya University, Rangpur, Bangladesh
Rada Savkina National University “Kyiv-Mohyla Academy”, Kyiv, Ukraine
V. Lashkaryov Institute of Semiconductor Physics at NAS of Ukraine, Kyiv, Ukraine
Nupur Saxena Organisation for Science Innovations and Research, Garha
Pachauri, India
Tetyana Semikina Lashkarev Institute of Semiconductor Physics, National
Academy of Science of Ukraine, Kiev, Ukraine
G. Y. Sidorov Rzhanov Institute of Semiconductor Physics of the Siberian Branch
of the RAS, Novosibirsk, Russia
Oleksii Smirnov V. Lashkaryov Institute of Semiconductor Physics at NAS of
Ukraine, Kyiv, Ukraine
Xuhui Sun Institute of Functional Nano & Soft Materials (FUNSOM), and Jiangsu
Key Laboratory for Carbon-based Functional Materials and Devices, Soochow
University, Suzhou, Jiangsu, People’s Republic of China
Victor V. Sysoev Yuri Gagarin State Technical University of Saratov, Saratov, Russia
A. V. Voitsekhovskii National Research Tomsk State University, Tomsk, Russia
M. V. Yakushev Rzhanov Institute of Semiconductor Physics of the Siberian
Branch of the RAS, Novosibirsk, Russia
Benyamin Yarmand Materials and Energy Research Center (MERC), Karaj, Iran
Part I
IR Detectors Based on II–VI
Semiconductors
Chapter 1
Introduction in IR Detectors

Ghenadii Korotcenkov

1.1 Introduction

Infrared (IR) range is the range of electromagnetic radiation from 0.78 to 1000 μm,
which is divided into sub-ranges:
• near IR, NIR: 0.78–1 μm;
• short wavelength IR, SWIR: 1–3 μm;
• medium wavelength IR, MWIR: 3–6 μm;
• long wavelength IR, LWIR: 6–12 μm;
• very long wavelength IR, VLWIR: 12–20 μm;
• far infrared region (FIR): 20–1000 μm
In comparison with visible and ultraviolet rays, infrared radiation has small energy,
for example 1.24 eV at λ = 1 μm, 0.12 eV at 10 μm, and ~0.01 eV at 100 μm.
For the first time the presence of infrared radiation was found in 1800 in the
process of experiments conducted by the English astronomer William Herschel. A
clearer understanding of IR radiation was obtained in 1900 through Plank’s law
(Eq. 1.1). According to Plank’s law, every physical object spontaneously emits radi-
ation in a wide range of wavelengths. The peak wavelength of the radiation corre-
sponds to the equilibrium temperature of the object. Spectral radiation emittance,
calculated according to Plank’s law, is shown in Fig. 1.1. As is seen, the peak radia-
tion of objects at room temperature (~300 К), is ~10 μm. The surface of the Sun,
which has a temperature of ~6000 K, has a maximum radiation in the visible range,
although radiation in the IR region is also present.
1
M     C1 5 exp  C2 / ·T   1 (1.1)

G. Korotcenkov (*)
Department of Physics and Engineering, Moldova State University, Chisinau, Moldova
e-mail: ghkoro@yahoo.com

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 3


G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_1
4 G. Korotcenkov

Fig. 1.1 Spectral emittance of objects at given equilibrium temperatures. (Reprinted from Karim
and Andersson [10]. Published 2013 by IOP as open access)

Table 1.1 Main applications of infrared photodetectors


Community Applications
Military Reconnaissance, navigation, night vision, guided missiles.
Commercial
Civil Police, firemen, border post.
Environment Pollution control, natural resources, energy savings, meteorology.
Industry Maintenance, fabrication processes control, nondestructive tests, optical
communications.
Medical Thermography.
Science
Astronomy Observation of the universe in the infrared region.
Physics, IR spectroscopy.
chemistry

Where T – absolute temperature (K), C1 – first radiation constant = 3.74·104


Wμm4/cm2, C2 – second radiation constant = 1.44· 104 μmK, λ – wavelength (μm).
It should be noted that the development of the IR technologies has intensified
significantly only in the last 40–50 years. The main applications of the IR devices
are shown in Table 1.1. During this period, significant advances have been made in
the development of various IR photoreceivers (see Fig. 1.2) and as a result, it is cur-
rently impossible to imagine many extremely important applications without the
use of IR detectors. IR detectors have become the basis of space surveillance sys-
tems, ballistic missile launch detection systems, non-contact temperature measure-
ment, motion sensors, IR spectroscopy, night vision devices, warhead homing
1 Introduction in IR Detectors 5

Fig. 1.2 History of the development of infrared detectors and systems. (Adapted from Rogalski
et al. [21]. Published 2020 by PAS as open access)

systems, and holographic information recording and processing systems. IR sensors


are widely used in astronomy, medicine, ultra-long range optical communication
systems, rangefinders, meteorological exploration, climatology, laser search and
visualization, etc. [4, 8, 12, 30]. It is the photodetectors, which in the overwhelming
majority of cases determines such basic parameters of IR systems as range, sensitiv-
ity, spectral scope, noise immunity, resolution, dynamic range and other character-
istics of the equipment.

1.2 IR Photodetectors

Two types of detectors are currently used to detect infrared radiation – photonic
(cooled and uncooled) and thermal.

1.2.1 Thermal (Non-selective) IR Detectors

Thermal radiation detectors are non-selective devices, i.e. they have the same spec-
tral characteristic over a wide range of the electromagnetic spectrum (up to hun-
dreds of micrometers) [4, 23, 34]. The operation of thermal radiation detectors is
based on the conversion of radiation energy into heat, and then into electrical energy.
In bolometers and thermistors, an increase in the temperature of the receiver changes
its electrical conductivity, in thermopiles, a thermo-EMF appears, in a pyroelectric
receiver, the value of the surface charge changes, and in dielectric bolometers using
6 G. Korotcenkov

ferroelectric capacitors (barium strontium titan) as sensing elements, its dielectric


changes with a change in temperature constant and, therefore, the capacitance of the
capacitor. Thermal detector manufacturing technologies have reached a certain
degree of perfection and predetermined a number of advantages, due to which sen-
sors of this type occupy a quantitatively dominant position in the market of the IR
detectors. Their advantages are well known. In its most general form, this is the
simplicity of the design and the absence of the need for cryogenic temperatures,
which leads to significantly lower power consumption for power supply. In addi-
tion, there is practically no need for service. For pyroelectrics and thermopiles, this
is the absence of both a power supply and the need for temperature compensation.
As a result, the noise (1/f) of such devices is sharply reduced. Competitive price also
plays a significant role when choosing a thermal sensor [23].
In the development of thermal radiation detectors, the maximum effect was
achieved in recent years, when silicon micro-electromechanical systems (MEMS)
technology began to be used in their manufacture [2, 3, 13, 34]. This made it pos-
sible to manufacture micro-sized elements of thermal sensors and provide them
with good thermal isolation. In particular, the state-of-the-art technology has made
it possible to fabricate monolithic silicon-based focal plane arrays (FPAs) with more
than 105 pixels and a pixel size of 50 μm. When using vanadium oxide (VOx) as a
sensitive material for MEMS microbolometers, fabricated using Si-based MEMS
technology, a temperature noise equivalent (NETD) of 40–50 mK for an aperture
number of 1 has been achieved, which is an undeniable progress [16, 29]. The
achievement of such results has led to the fact that of all types of IR thermal detec-
tors, uncooled MEMS-based detectors are of the greatest interest. Their appearance
and structure are shown in Fig. 1.3. The interest in such devices is due to the fact
that MEMS-based detector arrays combine the ability to form images of very good
quality with the low cost and ease of use inherent in uncooled detectors. While their
sensitivity does not come close to that of the best HgCdTe detector arrays, uncooled
MEMS-based detector arrays provide a level of performance that most users are
satisfied with. Currently MEMS-based detector arrays, including microbolometer

Fig. 1.3 (a) Modern 25 μm Microbolometer pixel (https://movitherm.com/), (b) Pyroelectric sen-
sor. Sketch of the pyroelectric sensor with an integrated resonant absorber. (c) The CMOS-MEMS
thermopile after drop-coating of graphene. ((b) Reprinted from Kuznetsov et al. [13]. Published
2018 by Nature as open access; (c) Reprinted from Chen and Chen [3]. Published 2020 by MDPI
as open access)
1 Introduction in IR Detectors 7

Table 1.2 Thermal detectors


Mechanism Comment Advantages Disadvantages
nn
Mercury in glass Thermometer Simple Low R
Very slow
Gas in chamber Golay cell Available Very slow
Simple
Thermal coefficient of resistance
Carbon resistor First bolometer Easy to model Need TLHe (4 K)
Easy to make
Vanadium oxide (VOx) Microbolometer FPAs Available Low R
Silicon processing
Good stability
Amorphous Si Microbolometer FPAs Available Less sensitive than VOx
Silicon processing
Thermal EMF (Seebeck effect)
Thermocouple One junction pair Slow, low R
Thermopile Many junction pairs Available Slow, low R
MEMS thermopile Simple
Inexpensive
Ferroelectric and pyroelectric
Barium strontium titanate First high volume Available Needs chopper
AC coupled High R
Inexpensive
Source: Reprinted with permission from Vincent et al. [33]. Copyright 2016: Wiley

arrays and MEMS-based thermoelectric (thermopile) arrays with acceptable param-


eters, are commercially available worldwide [6, 34]. Comparative characteristics of
thermal detectors are given in Table 1.2.
However, thermal detectors also have disadvantages that turn out to be quite
significant when choosing an IR sensor for work in applications, requiring high
parameters, such as military equipment [29]. Such disadvantages are the structural
and technological complexity of thermal insulation of many detector pixels from
each other and from the substrate, their sensitivity to temperature fluctuations and
vibrations with relatively small relative changes in the electrical characteristics of
the material per one degree of change in the object temperature, the inertia of the
spectral response and low uniformity of the image. For thermal detectors, there is
also the problem of decreasing pixel size. For example, the pixel sensitivity of a
microbolometer is highly dependent on the pixel area; therefore, there is a need to
maximize both the optical absorption area and thermal insulation. Unfortunately,
when the pixel size is reduced, this condition cannot be fulfilled, and therefore,
when using conventional single-level micromachining processes, performance deg-
radation is observed as the unit cell size is reduced below 40 μm. That is why, in
most MEMS bolometers, the pixel size does not decrease below 50 μm. Due to the
problems mentioned earlier, when using thermal detectors, there are often problems
with obtaining a clear image at long distances [29].
8 G. Korotcenkov

1.2.2 Photonic Radiation Detectors

The main group of IR detectors in terms of various applications is the so-called


photon or quantum radiation detectors [23]. Photonic radiation detectors ensure the
conversion of the incident photon flux into an electrical signal due to the direct
interaction of photons with the electronic subsystem of the receiver material (see
Fig. 1.4). The sensitivity of the photon detector is proportional to the number of
absorbed photons. Such a receiver is selective, i.e. it only reacts to photon of radia-
tion with a certain frequency (wavelength). In other words, photonic detectors
respond to photons whose energy exceeds certain threshold values, for example, the
semiconductor band gap (“intrinsic” detectors).
In turn, photonic radiation detectors are divided into receivers with an external
photoelectric effect (photomultipliers) and receivers with an internal photoelectric
effect (photoresistors, photodiodes, phototransistors, etc.). In modern infrared sys-
tems, receivers with an internal photoelectric effect are most widely used. In receiv-
ers of this type, three main physical phenomena are used, caused by the effect of
radiation on a semiconductor: the phenomenon of photoconductivity, photovoltaic
and photoelectric effects. Since IR radiation has small energy (energy is inversely
proportional to wavelength), these detectors are cooled down to cryogenic tempera-
tures in order to increase infrared detection efficiency/sensitivity. Quantum detec-
tors react very quickly to IR radiation (response time is order of μs), but they have
response curves with a detectivity that varies greatly with wavelength.
Photoconductors The operation of a photoresistor (PR) or photoconductor (PC) is
based on the change in the electrical conductivity of the sensitive layer during irra-
diation (read Chap. 2, Vol. 2). It is important to note that the first IR photonic receiv-
ers were photoconductive because of the simplicity of the technology, and the
relative ease of achieving near-ideal infrared performance and excellent reliability.
At the same time, it was found that the most sensitive PRs are also the most inertial.
For a number of them, a direct relationship has been established between the PR

Fig. 1.4 Photonic mechanisms of excitation of the electron subsystem in photonic detectors: (1)
intrinsic excitation, (2) impurity excitation, (3) absorption by free carriers, (4) excitation in
Schottky diodes. (Idea from Kulchitsky et al. [12])
1 Introduction in IR Detectors 9

sensitivity threshold and the time constant τ. Often, the mobility of electrons and
holes in a semiconductor is very different, as a result of which faster carriers can
pass through the detector several times before the carriers recombine. This provides
a gain mechanism. Operating temperature is another important factor. Decreasing
the temperature of the sensitive layer expands the spectral range of its operation in
the IR spectral region and increases its integral sensitivity. With cooling, the noise
decreases, and, consequently, the detecting ability of the PR increases. In addition,
upon cooling, the time constant and its dark resistance increase.
The advantages of PR are small size and weight, lower supply voltage compared
to photoemission detectors and the ability to work in a wide spectral range. The PRs
usually have a very high integral sensitivity and low power consumption not exceed-
ing several watts. The disadvantages of photodetectors of this class include increased
inertia, a significant dependence of characteristics and parameters on temperature, a
small linear zone of the energy characteristic, and the dependence of the output
signal on the illumination area of the sensitive layer. Currently, the main materials
used for the manufacture of IR photoresistors are CdS, PbS, PbSe, InSb, Ge:Au and
HgCdTe. Photo detectors on CdS, PbS, PbSe, InSb, and HgCdTe are intrinsic, and
Ge:Au-based devices are extrinsic IR photoconductive detectors.
Photodiodes A photodiode (PD) is usually called a semiconductor radiation detec-
tor based on the use of one-way conductivity of a p-n junction, upon illumination of
which an electromotive force (EMF) is formed (photovoltaic mode), or in the pres-
ence of a voltage source in the photodiode circuit, its reverse current changes (pho-
todiode mode). Currently, materials for the manufacture of IR photodiodes are
mainly Ge, Si, InSb, as well as ternary compounds such as InGaAs and HgCdTe.
The features of the operation of HgCdTe-based photodiodes are described in [1] and
the Chaps. 2, 3, 4, 5, 6, 7, and 8. There are many methods for manufacturing IR PDs,
each of which has its own strengths and weaknesses, depending on the specific
design of the photodetector and its purpose. For HgCdTe, this information can be
found in [17, 19] and Chap. 15, Vol. 1.
The spectral sensitivity of PDs changes when switching from photovoltaic mode
to photodiode mode. This sensitivity also depends on the temperature of the semi-
conductor material used for PDs fabrication. With decreasing temperature, the spec-
tral characteristic and its maximum shift to the short-wavelength region, and the
dark current as well as the noise level also decreases. Therefore, IR detectors, espe-
cially those designed to work in the LWIR spectral range, are forced to operate at
low temperatures, down to cryogenic temperatures. In many cases, the temperature
of operation is a critical parameter, because the cooling system often dominates the
size, weight and reliability of the detector system.
The time constant of the photodiode τPD largely depends on the method of its
manufacture and the size of the photosensitive area. For PDs, the τPD value is usu-
ally close to 10−5 s; for diffusion p-n junction with small areas, τPD can reach 10−6 s.
In special PDs with a small thickness, τPD ~ 10−10 s can be achieved. Comparatively
large dark currents when conventional photodiodes are switched on in the
10 G. Korotcenkov

photodiode mode make it impossible to use them for measuring low light flows. In
this case, it is necessary to work in the photovoltaic mode, in which the detecting
ability of the system is determined not by the low noise of the receiver, but by noises
of its electronic circuit or subsequent electronic links.
To improve the sensitivity for detecting infrared radiation, avalanche photodi-
odes (APDs) are often used [16]. Avalanche photodiodes include areas of high elec-
trical field. Carrier multiplication is carried out by transferring sufficient kinetic
energy to the carrier to create an additional electron-hole pair by impact ionization.
There is always some excess noise associated with multiplication, but this can be
minimized with designs that allow one carrier to be multiplied. The ideal APD is an
inexpensive device with low dark noise, wide spectral and frequency response, and
a gain from 1 to 106 or more. The characteristics of avalanche diodes based on
HgCdTe are discussed in the Chap. 3, Vol. 2.
In addition to the above devices, it is possible to note the Schottky barrier photo-
diodes belonging to the group of photoemissive detectors (see Fig. 1.4). They are
characterized by a relatively simple manufacturing technology and their parameters
are close to those of p-i-n PDs. The most popular Schottky-barrier IR detector is the
PtSi-p-Si detector, which can be used for IR detection in the 3–5 μm spectral range
[11]. Radiation is transmitted through the p-type silicon and is absorbed in the metal
PtSi (Platinum silicide) layer, producing hot holes which are then emitted over the
potential barrier into the silicon, leaving the silicide negatively charged. This funda-
mental difference in the detection mechanism underlies the unique properties of
Schottky IR detectors, including their exceptional spatial uniformity and their modi-
fied Fowler spectral response. The Schottky barrier height of the PtSi detector is
~0.22 eV, corresponding to a cutoff wavelength of ~5.6 μm. Due to the Fowler
dependence, the quantum efficiency (QE) of the PtSi detector in the 3–5 μm MWIR
regime is relatively low [14]. The main advantage of the Schottky-barrier detectors
is that they can be fabricated as monolithic arrays in a standard silicon process [23].
Photodiodes formed on the basis of heterostructures are also present in a large
number on the market of IR detectors [5]. Some options for using heterostructures
in the development of IR detectors are described in Chaps. 5 and 6, Vol. 2.
Phototransistors A phototransistor (PT) is a semiconductor device with photocur-
rent amplification properties with two p-n junctions, in which there is a directional
movement of current carriers. Phototransistors have a high quantum efficiency
(about 100). However, the presence of the second p-n junction leads to a significant
increase in noise. Therefore, it is often preferable to use photodiodes with an addi-
tional stage of the signal amplifier, the noise of which affects the detection ability of
the device less than the noise arising in the phototransistor. The disadvantages of
phototransistors also include: significant instability of parameters and characteris-
tics over time when the ambient temperature changes; and lower detection ability
than photodetectors. It should be noted that some PTs have a “blind spot” in the
center of the sensitive layer due to the shading of a part of the base by the emitter.
Therefore, when using them, it is necessary to distribute the flow over the entire
photosensitive surface of the PT.
1 Introduction in IR Detectors 11

Table 1.3 Infrared detectors used in single detector assemblies


Detector (μm)
Common applications, and comments
Si (PV) 0.1–
Optical communication, fire sensing, light and laser power
1.1
measurement, photon counting
InGaAs (PV) 0.7–
Optical communications, FTIR, gas detection, light and laser power
1.8
measurement, tunable diode laser spectroscopy (TDLS), moisture
analyzers. Replaces Ge (faster)
InAs (PV) 0.9– FTIR, non-contact temperature measurement, laser monitoring, gas
3.5 analyzers, spectrophotometry.
InSb (PC, PV) 1.0– FTIR, spectrophotometry, thermometry, remote sensing, gas
5.5 analysis
InAsSb (PV) 3.3– Gas measurement (CH4, CO2, CO, NH3, O3, etc.), flame monitoring
11 (CO2 resonance radiation), radiation thermometry
PbS (PC) 1–3.6 Non-contact temperature measurement, spark detecting, flame
control, moisture measurement, spectrophotometry. These detectors
are used especially when large-area detectors are required because
they are significantly less expensive than comparable III-V
(InGaAs) detectors.
PbSe (PC) 1–5.8 Gas analysis, laser power measurement, medical CO2 detection,
non-contact temperature measurement, flame detection, fire
detection, moisture monitoring.
HgCdTe (PC, PV) 1.0– FTIR, industrial process control, heat-seeking guidance, laser
16 warning receiver, laser monitoring, temperature monitoring, gas
detection, remote sensing. Используется вместо PbSe and PbS
where high detectivity is needed, low bias voltage, selective peak
wavelength response, and fast response times.
Thermopile 0.1– Fire sensing, intrusion, laser power, temperature measurement, gas
100 detection.
Pyroelectric 0.1– Fire sensing, motion detection, laser power, temperature
(piezoelectric) 100 measurement. Most sensitive of thermal detectors, sensitive to
vibration
Bolometer and 0.1– FTIR, astronomy
microbolometer 100
Golay cell and 0.1– FTIR
microgolay cell 100
FTIR Fourier-transform infrared spectroscopy, NDIR nondispersive infrared spectroscopy, PC
photoconductive, PV photovoltaic

As for the most common commercial and science applications of IR detectors,


they are listed in Table 1.3. A description of some of these applications can be found
in [33].

1.2.3 IR Photodetectors Array

A certain share in the market of modern IR photodetectors is occupied by photode-


tector (PD) arrays developed for thermal imaging equipment [12] and for astro-
nomical application [8]. The main distinctions of photodetector arrays from single
photodetectors are listed in the Table 1.4.
12 G. Korotcenkov

Table 1.4 Comparison of two general IR detector configurations


eature Single detector assembly PDs array
Detectors (pixels) A few: 1, 2, 4, perhaps 16 Thousands, millions
Equipment and software Can be simple Specialized
Output and responsivity measurement AC (usually) DC (usually)
Record data for every element Yes Yes
Report data for every element Yes – tabular No, or graphical only
Report array statistics Optional Essential
Figures of merit Basic Basic + some unique
Relative cost Low Expensive
Quantity per year Many Few
Electronics Discrete or small ASICs ROIC
Primary applications Various nonimaging Imaging
Source: Data extracted from Vincent et al. [33]
ASICs application-specific ICs, ROIC a readout- integrated circuit

A focal plane array (FPA) of photodetectors is created by combining separate,


usually identical, photosensitive elements on a single chip. The individual detectors
in an array are often referred to as pixels, short for picture elements; pixel is gener-
ally thought of as the smallest single component of a digital image. However, the
process of developing an integrated detector array is much more complicated than
manufacturing a separate detector element (read the Chap. 4, Vol. 2). A fundamental
limitation in the design of detector arrays is that light easily hits adjacent pixels in
the array, resulting in false counts or crosstalk. There are approaches that can be
taken to mitigate this limitation, but they add additional complexity to production.
In addition, the manufacturing process of the array becomes even more complex
due to the requirement to maintain low leakage currents in individual pixels, which
makes the manufacturing process even more cumbersome.
It is important to note that the first HgCdTe-based photodetector arrays were
made on the basis of photoconductive detectors. HgCdTe photoconductive detectors
have been in routine production since the early 1980s and are often called first-­
generation detectors. They have been designed for spectral ranges of 8–12 μm and
3–5 μm. Detectors operated at 80 K for LWIR and at 80–200 K for MWIR spectral
range. The reliable performance, low levels of defects and easily understood physics
has led to a long product life for photoconductive arrays. However, the size of the
matrix in such detectors was limited, and the first-generation thermal imaging sys-
tems had to use sophisticated optics to scan the infrared image over the matrix to
construct the scene. Therefore, photoconductive HgCdTe detectors have been very
successful in producing arrays of up to a few hundred elements for use in first-­
generation thermal imaging systems. The limitations of photoconductive detectors
emerged when the need arose for very large focal plane arrays. In addition, the low
impedance of the photoconductor made it unsuitable for injecting into silicon charge
transfer devices or field effect transistors; hence, each element required a lead-out
1 Introduction in IR Detectors 13

Fig. 1.5 Materials used for development of large format FPAs. (https://www.raytheon.com)

through the vacuum encapsulation to an off-focal plane amplifier. All this signifi-
cantly limited the capabilities of matrices based on photoconductive detectors. In
this regard, the use of photoconductive detectors in the development of photodetec-
tors arrays has now been limited. Currently, photodiodes and heterostructures of
various types are the basis of such photodetectors arrays. For a more detailed dis-
cussion of photodetectors arrays, see Chap. 4, Vol. 2.
The market for high performance thermal imaging cameras based on FPAs is
currently split between two material systems: InSb and HgCdTe [12, 31]. FPAs
present on the market are shown on Fig. 1.5. InSb has a large share of the market of
cooled thermal imaging cameras because the first FPAs were originally obtained
from this material. At the same time, HgCdTe has several key advantages that could
make it the material of choice for the development of third generation thermal imag-
ing cameras. HgCdTe can operate at much higher temperatures ( Toper max
< 200 K ) ,
than InSb ( Toper <100–120 К), and therefore MCT-based cameras can be smaller
max

and consume less power. For FPAs designed to operate in the SWIR range, other
materials can also be used, mainly InGaAs (https://www.flir.com). However,
InGaAs / InP-based FPAs can have high noise in the spectral range above 1.7 μm
due to defects caused by the lattice mismatch between InGaAs and InP, which is
used as substrate for growing InGaAs epitaxial layers.
As for astronomical applications, in addition to InSb and HgCdTe, they also use
Si:As-based FPAs [8, 31]. The fabrication of FPAs on the base of Si:As does not
present any particular difficulties, as standard technologies are used and large, high
quality silicon wafers are available. In addition, there are no problems with thermal
expansion mismatch between the FPA and the multiplexer. As a result, the best
14 G. Korotcenkov

samples of matrices based on Si:As with a sensitivity up to 28–40 μm have more


than 2 k × 2 k pixels with a pixel size of 18 μm [32]. For Si:As devices, the more
challenging task is to develop a multiplexer capable of operating at temperatures
around 10 K required to achieve low detector dark currents. However, this tempera-
ture is outside the range in which standard commercial silicon CMOS devices oper-
ate. The development of special multiplexers requires additional costs, which
sometimes compensate for the advantages of Si:As. This is due to the fact that the
market for astronomical Si:As-based devices is very small [8].

1.2.4 Photosensitive Materials for IR Technology

If we turn to the history of the development of IR technology [4, 22, 23], then in the
early stages it was based on the use of complex semiconductors. The first practical
IR detector was manufactured in 1933 on the basis of lead sulfide (PbS) with an
operating sensitivity range of up to 3 μm. In the late 1940s, this detection wave-
length was increased to 5 μm through the use of lead selenide (PbSe) and lead tel-
luride (PbTe). In the 1950s, research began on semiconductor compounds from
III-V, IV-IV and II-VI groups, which led to the emergence of new photosensitive
materials suitable for the development of various IR photodetectors. InSb,
InAs1 − xSbx, Pb1 − xSnxTe (LTT), InGaAs and HgCdTe (MCT) were one of them. A
large number of other combinations of compounds were also tested, but in most
cases, there were problems with crystal growth or there were limitations on doping,
which made them unsuitable for the manufacture of IR devices.
It should be noted that in the late 1960s – early 1970s, PbSnTe was actively
developed in parallel with HgCdTe [7]. Unlike HgCdTe, PbSnTe of the required
quality was easy to grow, and therefore high-quality IR photodiodes on their basis
for the LWIR spectral range were demonstrated relatively quickly. However, in the
late 1970s, work on PbSnTe-based detectors was discontinued [1]. There were two
reasons for this: a high dielectric constant and a large mismatch in the temperature
coefficient of expansion (TCE) of PbSnTe with TCE of the Si. The scanning-based
IR imaging systems required relatively fast response times so that the scanned
image was not blurred. The high dielectric constant did not allow the development
of devices with the low capacitance required for fast operation. Large TCE, led to
the destruction of indium bonds in the hybrid structure between the silicon-based
reader and the PbSnTe-based detector array during repeated thermal cycling from
room temperature to cryogenic operating temperature.
Concerning current IR technology (www.hamamatsu.com; www.teledynejud-
son.com; www.photonicsolutions.co.uk; https://vigo.com.pl), it is mainly based on
InSb, InGaAs and HgCdTe [17]. One should note that the use of HgCdTe allowed
to cover a wider range of wavelengths of infrared radiation compared to other mate-
rials. Depending on the composition of Hg1-хCdхTe, photodetectors based on this
material can cover the entire spectral IR range from SWIR to VLWIR (see Fig. 1.6).
In addition, HgCdTe has nearly ideal properties for the development of electronic
1 Introduction in IR Detectors 15

Fig. 1.6 The cut-off wavelength for Hg1-хCdхTe as a function of composition, x

avalanche photodiodes (e-APDs) for the infrared region (read the Chap. 3, Vol. 2).
The high mobility ratio of HgCdTe contributes to the creation of conditions under
which single-carrier multiplication with low noise figures take place. Such IR detec-
tors are effective when used in photon-starved applications, such as long-range
imaging and astronomy [1]. Other advantages of HgCdTe can be found in the Chap.
4, Vol. 1.
Currently, HgCdTe-based detectors are mainly used for spectral ranges of
1–2.5 μm, 3–5 μm and 8–14 μm. InSb is more suitable for the spectral range of
3–5 μm, while InGaAs is more suitable for the spectral range of 0.4–2.3 μm [9, 17].
The spectral areas of their application are shown in Figs. 1.5 and 1.7. However, even
until now, inexpensive polycrystalline thin-film photoconductive detectors based on
PbS and PbSe remain the preferred choice for many applications in the spectral
range of 1–3 μm and 3–5 μm. It should be noted that in recent decades there has also
been interest in such compounds as Hg1-xMnxTe, Hg1-xZnxTe and Hg1-xCdxSe as
potential alternatives to HgCdTe for infrared detectors. Photoconductive and photo-
voltaic detectors have been reported using these materials, but the devices have not
yet reached technology perfection like Hg1-xCdxTe [26, 27].
We must also not forget about extrinsic (photon energies smaller than the band-
gap) photoconductive detectors, which were developed in the early 1950s. Research
has shown that using extrinsic photoconductive response from germanium, doped
by copper, mercury, zinc, and gold allows developing IR receivers capable to work
in the 8–14 μm (LWIR) and 14–30 μm (VLWIR) spectral range. However, they
must operate at very low temperatures (30–50 К) to achieve similar performance to
intrinsic detectors, and they sacrifice quantum efficiency to avoid the need for thick
detectors.
16 G. Korotcenkov

Fig. 1.7 Basic semiconductor materials for cooled thermal imaging arrays and operating spectral
ranges. This Figure also shows the specific absorption bands of the atmosphere. (Data extracted
from Rogalski [22, 23] and Piotrowski and Rogalski [17])

There are also Si-based intrinsic and extrinsic IR photoconductive detectors.


Silicon intrinsic IR detectors are usually designed and fabricated as Schottky-barrier
detectors (PtSi detector). Such detectors can be used for detection in the range of
3–5 μm. As for longer wavelengths, in this area can be used extrinsic silicon IR
photodetectors, such as Si:As or Si:Sb [20]. The developed devices are designed to
operate in the range of 5–40 μm. However, the sensitivity range can be extended up
to approximately 300 μm. As well as Ge-based IR PDs, operating temperatures of
Si-based extrinsic detectors lies in the cryogenic temperature range (8–30 K). A
detailed review of a bulk Si and Ge IR detectors can be found in [28].
It is important to note that in recent years there has been a significant increase in
interest in extrinsic IR photoconductive detectors based on germanium and espe-
cially silicon [20]. This interest is due to the creation of multi-element FPAs for use
in space and land infrared astronomy, as well as on spacecraft for various purposes.
Advances in Si-based FPAs technology, the creation of low-noise semiconductor
preamplifiers and deep-cooled multiplexers, as well as unique designs of extrinsic
silicon IR PDs and deep-cooling equipment have achieved record-breaking detec-
tivity, close to the emission limit even under exceedingly low backgrounds in
space [15].

1.2.5 Comparison of Thermal and Photonic Infrared Detectors

Comparing the relative response of photon and thermal detectors as a function of


wavelength with either a vertical scale of W−1 or photon−1, then as seen in Fig. 1.8,
photon detectors show a linear increase in response on the dependences, recalculated
1 Introduction in IR Detectors 17

Fig. 1.8 Relative spectral response for a photon and thermal detector for (a) constant incident
radiant power and (b) photon flux, respectively. (Reprinted with permission from Rogalski et al.
[24]. Copyright 2000: SPIE)

per unit power of the incident radiation, until the cutoff wavelength is reached. An
increase in response is associated with an increase in the number of photons to
achieve the same power with increasing wavelength. The cutoff wavelength is deter-
mined by the detector material used. At the same time, thermal detectors tend to be
spectrally flat in this case. This is due to the fact that their response is proportional to
the energy absorbed, which, when recalculated per unit of incident radiation power,
does not depend on the wavelength. If the sensor response is recalculated per photon,
then the photon detectors in this case are usually flat, and the thermal sensors have a
linearly decreasing response.
Compared to thermal detectors, cooled photonic sensors have a NETD of
10–20 mK, and this indicator practically does not change over a wide range of inte-
gration times (5–7 ms and more). The detectivity of photonic detectors is about two
orders of magnitude higher than that of thermal detectors. This allows large aperture
ratios to be used when designing thermal imaging cameras [29]. In addition, photon-­
type detectors have better signal-to-noise ratio and faster response time. For many
applications, thermal detectors may not provide the required response time.
Consequently, heat detectors are not suitable for infrared thermal imaging cameras
that use higher frame rates and multispectral performance. At the same time, photon
detectors require deep cryogenic cooling, and this leads to a complication of the
device, an increase in its geometric dimensions, weight and high power consump-
tion. Only in the last decade, due to progress in the development of photodetectors,
it was possible to increase the operating temperatures from 77 to 150 K. In addition,
if the service life of the photodetector array of the photon detector itself can be
determined for decades, then the service life of the cooler does not exceed 30 thou-
sand operating hours (the best models), after which the cooler must be replaced. All
this leads to an increase in the cost of sensors of this type in comparison with ther-
mal detectors, as well as to an increase in the cost of their operation. Nevertheless,
the price criterion, as a rule, is not decisive in applications such as military and
18 G. Korotcenkov

astronomy, where the technical characteristics of photodetectors come first, and not
their price. Therefore, often in such applications, photonic detectors dominate, as
devices with the best parameters.

1.2.6 Parameters Characterizing IR Photodetectors

The main parameters characterizing IR detectors are the followings:


• Dark current – current measured in the absence of radiation under the operating
conditions.
• Photo sensitivity (Responsibility) – is the output voltage (or output current) per
watt of incident energy when noise is ignored.
• Quantum efficiency – Quantum efficiency (QE) represents the percentage of
photo-generated electron-hole pairs to the number of incoming photons.
• Integral sensitivity – sensitivity to non-monochromatic radiation of a given
spectral range.
• Spectral sensitivity – dependence of the sensitivity on the radiation wavelength.
• Current sensitivity – the sensitivity of the photodetector, in which the measured
electrical value is the photocurrent, and the voltage sensitivity is the sensitivity
when the measuring value is the voltage at the output of the radiation detector.
• Noise of the radiation detector – the output chaotic signal at the output of the
photodetector measured in the absence of radiation. Noise does not allow regis-
tering arbitrarily small signals that become imperceptible against its background,
i.e. restrict the limiting capabilities of the device.
• Noise Equivalent Power (NEP) – NEP is a measure of flux in order to generate
a signal that is equal to the noise level.
• Noise Equivalent Temperature Difference (NETD) – The minimum tempera-
ture difference, producing a signal level equal to the noise of the detector, is
defined as NETD. NETD is an important figure of merit, which can be consid-
ered in order to estimate thermal detection performance.
• Detectivity (D*) – the value of the minimum radiation flux, which produces a
signal at the output equal to the intrinsic noise. It is inversely proportional to the
square root of the area of the radiation receiver. It is measured in 1/W.
• Specific Detectivity – The detectivity multiplied by the square root of the product
of a frequency bandwidth of 1 Hz and an area of 1 cm2. Measured in cm2Hz1/2/W.
• Response time, or time constant τ – the time required to establish a signal at the
output corresponding to the input action. The time constant τ determines the
cutoff frequency of modulation of the signal at the input of the photodetector.
• Working temperature – the maximum temperature of the sensor and the environ-
ment at which the sensor is able to perform its functions correctly.
• Photodetector resistance – for photoresistors, the dark resistance measured in
the absence of radiation is considered as a photodetector resistance. For photodi-
odes, the value of the differential resistance RD is usually given, which is equal
1 Introduction in IR Detectors 19

to the ratio of small increments of the signal voltage to the photocurrent under
the given operating conditions.
Most of the important parameters of photonic photodetectors are controlled by the
noise level. The nature of this noise and approaches to reduce it are given in
Table 1.5.
As for the spectral detectivity of various photodetectors, this data for a number
of commercially available IR detectors are shown in Fig. 1.9. The temperatures at
which these parameters were determined are also indicated there. It is seen that in
many cases spectral detectivity approaches the theoretical limit values.

1.2.7 The Role of the Atmosphere in IR Technology

All IR detectors, except for sensors used in space work in the Earth’s atmosphere,
which has its own specific absorption bands (see Fig. 1.7). The absorption bands of
water vapor with a center of 6.3 μm and of carbon dioxide with centers of 2.7 and
15 μm limit the transmission of radiation by the atmosphere in the wavelength range
of 2–20 μm, determining the position of the so-called atmospheric transparency
windows: 3–5 and 8–12 μm (Fig. 1.2). According to the established classification,
these transparency windows correspond to the mid- (MWIR) and long-wave (LWIR)
ranges of the IR spectrum. In the international photometric system, the position of
the transparency windows is standardized corresponding to the wavelength (λ ± Δλ):
• in the visible and early near infrared range: B, V, R, J – up to 1.2 μm;
• H- range – (1.6 ± 0.1) μm;
• К- range – (2.2 ± 0.3) μm;

Table 1.5 Limits to photonic detector performance


Limits Noise origin How to reduce?
Fundamental Background radiation noise Spatial and spectral filtering
Signal photon noise Cannot be reduced
Heterodyne photon noise Cannot be reduced
Less Auger thermal generation Selection of semiconductors, non-­
fundamental equilibrium depletion
Internal radiative generation Design of the detector
Radiative generation from adjacent Design of the detector
elements
Technological Shockley-read generation Elimination of Shockley-read centers
Thermal generation at surfaces, Improved surface treatment and
interfaces and contacts technology
Low frequency noise Zero bias operation, improved
technology
Amplifier noise Improved electronic interface
Source: Reprinted from Piotrowski [18]. Published 2004 by PAS as open access
20 G. Korotcenkov

Fig. 1.9 Comparison of the D* of various commercially available infrared detectors when oper-
ated at the indicated temperature. Chopping frequency is 1000 Hz for all detectors except the
thermopile, thermocouple, thermistor bolometer, Golay cell and pyroelectric detector. For these
detectors a chopping frequency is 10 Hz. Each detector is assumed to view a hemispherical sur-
round (2π field of view) at a temperature of 300 K. Theoretical curves for the background-limited
D* (dashed lines) for ideal photovoltaic and photoconductive detectors and thermal detectors are
also shown. PC photoconductive detector, PV photovoltaic detector. (Reprinted from Rogalski
[25]. Published 2000 by MYU K.K. as open access)

• L- range – (3.6 ± 0.45) μm;


• М- range – (4.6 ± 0.5) μm;
• N- range – (10 ± 2) μm;
• Q- range – (20 ± 0.4) μm.
Between the transparency bands there are bands of complete absorption of IR radia-
tion by the atmosphere, mainly by carbon dioxide СО2 (2.6–2.9; 4.2–4.4 μm) and
water vapor Н2О (5–8 μm). This means that when measured in these spectral ranges,
the Earth’s atmosphere absorbs most of the radiation coming from the object. Thus,
in order to detect the IR signal, one must use the so-called atmospheric transparency
windows. These spectral ranges have their advantages in different applications.
Infrared spectral range 0.78–3 μm (NIR and SWIR) is used in fiber-optic communi-
cation lines, devices for external observation of objects and equipment for chemical
analysis. SWIR ends at 2.5 μm and differs from medium and long wavelengths
(MWIR and LWIR) that objects in this spectral range generate extremely little
intrinsic thermal radiation. Observation in the SWIR range is carried out due to
reflected light, that is, illumination from some powerful source is required, so this
range is in many ways similar to the visible range. All wavelengths from 2 to 5 μm
1 Introduction in IR Detectors 21

(SWIR and MWIR) are used in pyrometers and gas analyzers that monitor the level
of contamination in a various gas environment. It is also believed that the spectral
range of 3–5 μm (MWIR) is more suitable for systems that capture images of objects
with a high intrinsic temperature or in applications where contrast is required higher
than sensitivity. Compared to MWIR, the LWIR range is more suitable for operation
in smoky and dusty conditions. It was also found that measurements in the LWIR
range are not affected by solar reflections and sea surface gloss. It has been experi-
mentally established that in the LWIR range, the interference from the radiation of
the inhomogeneity of the sky, re-reflected from the rough sea surface, is about 10
times less than in MWIR range. This range of 8–15 μm (LWIR), popular for special
applications, is also used wherever it is necessary to see and recognize any objects
in the fog.

Acknowledgments This research was funded by the State Program of the Republic of Moldova,
project 20.80009.5007.02.

References

1. Baker IM (2017) II-VI narrow bandgap semiconductors: optoelectronics. In: Kasap S, Capper
P (eds) Springer handbook of electronic and photonic materials. Springer, pp 867–896
2. Bao A, Lei C, Mao H, Li R, Guan Y (2019) Study on a high performance MEMS infrared
thermopile detector. Micromachines 10:877
3. Chen S-J, Chen B (2020) Research on a CMOS-MEMS infrared sensor with reduced graphene
oxide. Sensors 20:4007
4. Corsi C (2010) History highlights and future trends of infrared sensors. J Mod Opt
57(18):1663–1686
5. Ghimire H, Jayaweera PVV, Somvanshi D, Lao Y, Unil Perera AG (2020) Recent prog-
ress on extended wavelength and split-off band heterostructure infrared detectors.
Micromachines 11:547
6. Graf A, Arndt M, Sauer M, Gerlach G (2007) Review of micromachined thermopiles for infra-
red detection. Meas Sci Technol 18(7):R59–R75
7. Harman TC, Melngailis J (1974) Narrow gap semiconductors. In: Wolfe R (ed) Applied solid
state science. Academic, New York
8. Hodapp KW, Hall DNB (2006) Introduction to detectors: possible status in 2010–2020.
In: Whitelock P, Dennefeld M, Leibundgut B (eds) Proceedings IAU symposium, No. 232.
International Astronomical Union, pp 40–51
9. Joshi AM, Ban VS, Mason S, Lange MJ, Kosonocky WF (1992) 512 and 1024 element lin-
ear InGaAs detector arrays for near-infrared (1–3 μm) environmental sensing. Proc SPIE
1735:287–295
10. Karim A, Andersson JY (2013) Infrared detectors: advances, challenges and new technologies.
IOP Conf Ser: Mater Sci Eng 51:012001
11. Kimata M (2001) Metal silicide Schottky infrared detector arrays. In: Capper P, Elliott CT
(eds) Infrared detectors and emitters: materials and devices. Kluwer Academic Publishers,
Boston, pp 77–98
12. Kulchitsky NA, Naumov AV, Startsev VV (2020) Infrared focal plane array detectors: “post
pandemic” development trends. Part 1. Photonics Russia 14(3):234–244. (In Russian)
13. Kuznetsov SA, Paulish AG, Navarro-Cía M, Arzhannikov AV (2018) Selective pyroelectric
detection of millimetre waves using ultra-thin metasurface absorbers. Sci Rep 6:21079
22 G. Korotcenkov

14. Lin TL, Park JS, George T, Jones EW, Fathauer RW, Maserjian J (1993) Long-wavelength PtSi
infrared detectors fabricated by incorporating a p+ doping spike grown by molecular beam
epitaxy. Appl Phys Lett 62(25):3318–3320
15. McCreight CR, McKelvey ME, Goebel JH, Anderson GM, Lee JH (1986) Detector arrays for
low-background space infrared astronomy. Laser Focus/Electro-Optics 22:128–133
16. Norton PR (2006) Third-generation sensors for night vision. Opto-Electron Rev 14:283–296
17. Piotrowski J, Rogalski A (2007) High-operating-temperature infrared photodetectors. SPIE,
Bellingham
18. Piotrowski J (2004) Uncooled operation of IR photodetectors. Opto-Electron Rev
12(1):111–122
19. Reine MB (2001) HgCdTe photodiodes for IR detection: a review. Proc SPIE 4288:266–277
20. Rieke GH (2007) Infrared detector arrays for astronomy. Annu Rev Astron Astrophys
45:77–115
21. Rogalski A, Kopytko M, Martyniuk P (2020) 2D material infrared and terahertz detectors:
status and outlook. Opto-Electron Rev 28:107–154
22. Rogalski A (2012) History of infrared detectors. Opto-Electron Rev 20:279–308
23. Rogalski A (2003) Infrared detectors: status and trends. Prog Quant Electron 27:59–210
24. Rogalski A, Adamiec K, Rutkowski J (2000) Narrow-gap semiconductor photodiodes. SPIE-­
The International Society for Optical Engineering, Bellingham
25. Rogalski A (2000) Infrared detectors at the beginning of the next millennium. Sens Mater
12(5):233–288
26. Rogalski A (1991) Hg1-xMnxTe as a new infrared detector material. Infrared Phys 31:117–166
27. Rogalski A (1989) Hg1-xZnxTe as a potential infrared detector material. Prog Quant Electron
13:299–253
28. Sclar N (1984) Properties of doped silicon and germanium infrared detectors. Prog Quant
Electron 9:149–257
29. Smuk S, Kochanov Y, Petroshenko MP, Solomitskii D (2014) IRnova long-wavelength infra-
red sensors based on quantum wells. Komponenti Tehnologia 1:20–25. (In Russian)
30. Sobrino JA, Del Frate F, Drusch M, Jiménez-Muñoz JC, Manunta P, Regan A (2016) Review
of thermal infrared applications and requirements for future high-resolution sensors. IEEE
Trans Geosci Remote Sens 54:2963–2972
31. Starr B, Mears L, Fulk C, Getty J, Beuville E, Boe R et al (2016) RVS large format arrays for
astronomy. Proc SPIE 9915:99152X
32. Tan CL, Mohseni H (2018) Emerging technologies for high performance infrared detectors.
Nano 7(1):169–197
33. Vincent JD, Hodges SE, Vampola J, Stegall M, Pierce G (2016) Fundamentals of infrared and
visible detector operation and testing, 2nd edn. Wiley, Hoboken
34. Xu D, Wang Y, Xiong B, Li T (2017) MEMS-based thermoelectric infrared sensors: a review.
Front Mech Eng 12(4):557–566
Chapter 2
Photoconductive and Photovoltaic IR
Detectors

Rada Savkina and Oleksii Smirnov

2.1 Introduction to Photoconductive and Photovoltaic IR


Detectors on II-VI Semiconductors

The photoconductivity and photovoltaic effect-based devices are the most widely
exploited photon detectors of the infrared (IR) radiation. As we already know from
the previous chapters, photon detectors have significant advantages over other tech-
nologies in the field of detecting IR radiation such as fast response, high sensitivity,
and wavelength selectivity. A few excellent scientific works have been published on
IR detectors and most of them have been devoted to HgCdTe. HgCdTe is a variable-­
gap II-VI semiconductor, which is used most often in the production of IR photon
detectors since it able to cover a wide wavelength range from 1 to 25 μm by control-
ling the Hg content. Prof. Antoni Rogalski, one of the world’s leading researches in
the field of IR optoelectronics, on the pages of his one of the last reviews concluded
that: “HgCdTe … has triggered the rapid development of the three “detector genera-
tions” considered for military and civilian applications …” [45]. In the opinion of
Antoni Rogalski, unique position of this material is conditioned by “…composition-­
dependent tailorable energy band gap over the entire 1–30 μm range, large optical
coefficients that enable high quantum efficiency, and favorable inherent recombina-
tion mechanisms that lead to long carrier lifetime and high operating tempera-
ture” [42].

R. Savkina (*)
National University “Kyiv-Mohyla Academy”, Kyiv, Ukraine
V. Lashkaryov Institute of Semiconductor Physics at NAS of Ukraine, Kyiv, Ukraine
e-mail: rk.savkina@ukma.edu.ua; savkina@nas.gov.ua; rada.k.savkina@gmail.com
O. Smirnov
V. Lashkaryov Institute of Semiconductor Physics at NAS of Ukraine, Kyiv, Ukraine
e-mail: alex_tenet@isp.kiev.ua

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 23


G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_2
24 R. Savkina and O. Smirnov

This chapter provides data about photoconductive and photovoltaic infrared


detectors manufactured from HgCdTe, as well as from the alternative ternary alloy
systems, such as HgZnTe and HgMnTe. Their design, performance, advantage, and
disadvantages are evaluated and compared. Infrared photon detectors operating in
the middle (3–5 μm) and long wavelength (8–14 μm) infrared spectral range require
cryogenic cooling to achieve useful performance. Background-limited performance
is typically not achieved without significant cooling of the IR photon detectors. At
the same time, there are known some concepts of the high operating temperature
photon detection proposed and implemented to improve the performance of IR pho-
todetectors near room temperature, which will also be described in this chapter.

2.2 Hg-Based Materials for IR Photon Detectors

Narrow-gap mercury-cadmium-telluride technologies are well developed now, and


today this material is one of the basic semiconductors for photon detectors from
near IR (wavelength λ ∼ 1.5 μm) to long IR (λ ∼ 20 μm) and is used in large-scale
arrays with silicon CMOS readouts. We will not focus here on the properties of
HgCdTe. The advantages of these compounds are discussed in the Chap. 4, Vol. 1.
We only point out that there are several important features denoted in [37] that make
HgCdTe an excellent material for infrared detectors:
–– HgCdTe can be made n-type or p-type by several relatively convenient methods,
at carrier concentrations required for most high-performance photodiode archi-
tectures. There exists a clear trend the use of extrinsic acceptor doping at photo-
diods manufacturing to avoid the strong Shockley-Read recombination associated
with the Hg vacancy.
–– Device-quality material can be grown by a variety of methods – both liquid phase
epitaxy and vapor phase epitaxy methods such as MBE and MOVPE.
–– The lattice mismatch between HgTe and CdTe is small, approximately 0.3%.
This allows epitaxial growth of high-quality HgCdTe films on IR-transparent
CdTe or nearly-lattice-matched IR-transparent CdZnTe substrates, with disloca-
tion densities in the low-105 cm−2 range. IR-transparent substrates that are less
costly, are available in much larger areas, and are more rugged than CdTe and
CdZnTe, such as sapphire and silicon, can be used for epitaxial growth of
HgCdTe films with dislocation densities that are acceptably low (mid-106 cm−2)
for many important photodiode applications.
–– The thermal expansion coefficient of HgCdTe sufficiently closed to that of sili-
con as well as the relatively low dielectric constant (εs = 18 × ε0) of HgCdTe have
opened important technological possibilities. This is important for fast response
in laser pulse detectors where small RC time constants, for suppressing pream-
plifier noise below the detector noise and for technologically viable hybrid
arrangements of HgCdTe detector arrays.
2 Photoconductive and Photovoltaic IR Detectors 25

Table 2.1 The main physical properties for HgZnTe and HgMnTe solid solutions
Lattice constant a(x), (nm) at Density ρ(x) (g/ The composition of the
300 K cm3) at 300 K solid solution
Hg1-xZnxTe 0.6461–0.0361x 8.05–2.41x 0.10 ≤ x ≤ 0.40
Hg1-xMnxTe 0.6461–0.0121x 8.12–3.37x 0.08 ≤ x ≤ 0.30
Energy gap Eg (eV)
Hg1-xZnxTe −0.3 + 0.0324x0.5 + 2.731x-0.629x2 + 0.533x3 + 5.3 × 10−4 T(1–0.76x0.5–1.29x)
Hg1-xMnxTe −0.253 + 3.466x + 4.9 × 10−4Tx-2.55 × 10−3 T
Intrinsic carrier concentration, ni (cm−3)
Hg1-xZnxTe (3.067 + 11.37x + 6.548·10−3 T-3.633·10−2Tx) × 1014·Eg3/4·T3/2exp(−5802Eg/T)
Hg1-xMnxTe (4.615–1.59x + 2.64 × 10−3 T-1.7 × 10−2Tx + 34.15x2) × 1014·Eg3/4·T3/2exp(−5802
Eg/T)
Momentum matrix element P
(eV cm) Spin-orbit splitting energy Δ (eV)
Hg1-xZnxTe 8.5 × 10−8 1.0
Hg1-xMnxTe (8.35–7.94x) × 10−8 1.08
Effective masses Mobilities (cm2/Vs)
Hg1-xZnxTe me*/m = 5.7 × 10−16·Eg/P2 μe = 9 × 108b/T2a a = (0.14/x)0.6 b = (0.14/x)7.5
mh*/m = 0.6 μh = μe/100
Hg1-xMnxTe me*/m = 5.7 × 10−16·Eg/P2 μe = 9 × 108b/T2a a = (0.095/x)0.6
mh*/m = 0.5 b = (0.095/x)7.5
μh = μe/100
Static dielectric constant ε High frequency dielectric constant ε∞
Hg1-xZnxTe 20.206–15.153x + 6.5909x2– 13.2–19.1916x + 19.496x2-6.458x3
0.951826x3
Hg1-xMnxTe 20.5–32.6x + 25.1x2 15.2–28.8x + 28.2x2
Source: Data extracted from Ref. [43]

Among the narrow-gap II-VI semiconductors, HgZnTe and HgMnTe have been
studied as potential alternatives to HgCdTe solid solution. The main physical prop-
erties for both ternary alloys are presented in the Table 2.1. Data was obtained from
the Ref. [43]. Both HgZnTe and HgMnTe exhibit compositional-dependent optical
and transport properties like HgCdTe material with the same energy gap. Some
physical properties of alternative alloys indicate on structural advantage in compari-
son with HgCdTe [39, 44]. For example, introducing ZnTe in HgTe decreases statis-
tically the ionicity of the bond improving the stability of the HgZnTe alloy.
Interdiffusion coefficient is about 10 times lower in HgZnTe than in HgCdTe. In the
range of temperatures typical for IR detectors operation (77 K), the spin-­independent
properties of HgMnTe are practically identical to the properties of HgCdTe, dis-
cussed exhaustively in the literature. But the question of HgMnTe lattice stability is
rather ambiguous.
The first HgZnTe photoconductive detectors were fabricated by Z. Nowak and
M.E. Ejsmont in the early 1970s (see Ref. in Rogalski [39]). Then, it was shown that
26 R. Savkina and O. Smirnov

Hg0.885Zn0.15Te can be used as a material for high-quality ambient-temperature


10.6 μm photoconductors with detectivity around 108 cm Hz1/2 W−1 [35]. The
research group at Santa Barbara Research Center has developed very long wave-
length HgZnTe photoconductive detectors with λ = l7 μm at temperature ≥65 K
with peak detectivity D* = 8 × 1010 cm Hz1/2 W−1 obtained for the best wafers [32].
The best quality n+-p HgZnTe photodiodes were manufactured by ion implantation
technique. The measurements reveal comparable value of R0A for both HgZnTe and
HgCdTe photodiodes at 77 K in 1980s [2].
In comparison with HgCdTe, HgZnTe detectors are easier to prepare due to their
relatively high hardness. The maximum of microhardness for HgZnTe is more than
twice that one for HgCdTe (see Fig. 2.1). Moreover, HgZnTe is the material that is
more resistant to dislocation formation and plastic deformation than HgCdTe. It was
found that fixed charges at the anodic oxide-HgZnTe interface are lower than that of
HgCdTe and is around 2 × 1010 cm−2 at 90 K.
Based on HgMnTe, mainly p-n junction photodiodes are created. Good quality
p-n HgMnTe and HgCdMnTe junctions manufactured by annealing in Hg-saturated
atmospheres of as-grown, p-type samples were made with the detectivities in the
3- to 5- and 8- to 12-μm spectral ranges closed to the background limit [6]. Other
type of detector – avalanche photodiodes, was developed on the base HgMnTe and
HgCdMnTe compounds. The R0A product for such detector was 2.62 × 102 Ω cm2,
which is equivalent to a detectivity value of 1.9 × 1011 cm Hz1/2 W−1 at 300 K [46].

Fig. 2.1 Microhardness H versus composition x of HgZnTe and HgCdTe. (Reprinted with permis-
sion from Ref. [48]. Copyright 1986: Elsevier)
2 Photoconductive and Photovoltaic IR Detectors 27

At the same time, neither HgZnTe nor HgMnTe have ever been systematically
explored in the device context [43]. The prime reason for this is a serious problems
encountered in these crystals’ growth including:
–– the essential separations between the liquids and solidus curves resulting in the
high segregation coefficients;
–– a weak variation of the growth temperature causes a large composition variation;
–– high mercury pressure over the melt both in the case of HgZnTe and HgMnTe is
an unfavorable condition for the growth of homogeneous bulk crystals.

2.3 Photoconductive and Photovoltaic IR Detectors: Design,


Performance, Advantage, and Disadvantages

2.3.1 Photonic Mechanism of Detection

Photonic mechanism of detection (see Fig. 2.2) consists in direct conversion of


incident photons into conducting electrons either bound to lattice atoms (intrinsic
absorption) or to impurity atoms (extrinsic, impurity absorption) or with free elec-
trons within a material. A key difference between intrinsic and extrinsic detectors is
that extrinsic detectors require much cooling to achieve high sensitivity at a given
spectral response cutoff in comparison with intrinsic detectors. According to the
classification presented in [21], there are two main types of photon detectors that
work with the majority and the minority charge carriers. If the dominant carrier is
the majority carrier, then the sensing is photoconductive (PC) in nature. For minor-
ity carrier devices, both photoconductive and photovoltaic (PV) modes of detection
can be utilized. Historical perspective on PC and PV detectors in HgCdTe was
reviewed in [19], and the theory of photoconductors and a conventional photodiode
can be found in monographs [41, 43].

Fig. 2.2 Fundamental A B C


optical excitation processes
in semiconductors: A
intrinsic absorption, B
extrinsic absorption, C free Cb
carriers’ absorption, Cb
conduction band, Vb
valence band

Vb
28 R. Savkina and O. Smirnov

2.3.2 Photonic Detector Characterization

The following parameters are typically used to characterize and compare the perfor-
mance of different materials systems and device architectures: responsivity (Rv),
noise equivalent power (NEP), and detectivity (D) [41]. The responsivity is a func-
tion of the wavelength of the incident radiation. The responsivity of IR detector is
defined as the ratio of the electrical output signal (voltage Vs, or current Is) to an
input signal power in the form of a known photon flux (Pλ):

Vs I
Rv  or Ri  s
P P (2.1)

The unit of responsivity is volts per watt (V/W) or amperes per watt (A/W).
The noise equivalent power is the signal level that produce a signal-to-noise ratio
of 1. It can be written in terms of responsivity:

Vn I n
=
NEP = (2.2)
Rv Ri

where Vn and In are a noise voltage and noise current respectively.


The unit of NEP is watt. If NEP is quoted for a fixed reference bandwidth, which
is often assumed to be 1 Hz, this “NEP per unit bandwidth” has a unit off watts per
square root hertz (W/Hz1/2).
The detectivity D is the reciprocal of NEP:

1
D= (2.3)
NEP

Both NEP and detectivity are functions of electrical bandwidth (∆f) and detector
area (Ad), so a normalized detectivity D* suggested in [15, 16] is defined as

 Ad  f 
1/ 2

D  (2.4)
NEP

and is expressed in unit cm Hz1/2 W−1, which recently is called “Jones.” The
importance of D* is that this figure of merit permits comparison of detectors of the
same type but having different areas.
The detectivity D* of IR photodetector is limited by generation and recombination
rates G and R in the active region of the device Ad. For a given wavelength and operat-
ing temperature, the highest device performance can be obtained by maximizing the
ratio η/[(G + R)t]1/2. This means that high quantum efficiency η can be obtained with
a thin device. A quantum efficiency is the number of photocarriers generated per num-
ber of incident photons for a specific semiconductor with reflectivity r and is given by
η = (1−r)·(1−exp(−αx)), where 0 ≤ η ≤ 1. If both the generation and recombination
rates and electrical and optical area of IR detector are equal, the detectivity of an opti-
mized photodetector is limited by thermal processes and it can be expressed as
2 Photoconductive and Photovoltaic IR Detectors 29

1/ 2
  
D  0.31 k (2.5)
hc  G 

where 1 ≤ k ≤ 2 and is dependent on the contribution of recombination and back-


side reflection. The ratio of the absorption coefficient to the thermal generation rate,
α/G, is the fundamental figure of merit of any material intended for infrared
photodetection.
An important factor influencing the performance of an IR detector is unwanted
fluctuations in the measured signal defined as noise. All detectors are limited in the
minimum radiation power that they can detect, by noise that is determined by the
fluctuation of the charge carriers generated both thermally within the detector (the
internal noise) and by the incident background flux of IR radiation that is absorbed
by the detector material (the radiation noise). The ultimate performance of infrared
detectors is reached when the internal noise is low compared to the radiation noise.
The practical operating limit for most infrared detectors is not the signal fluctuation
limit but the background fluctuation limit, also known as the background-limited
infrared photodetector (BLIP) limit. The expression for BLIP detectivity can be
written as

1/ 2
    
DBLIP    (2.6)
hc  2 B 

where ΦB is the total background photon flux density reaching the detector:

 2  2 c d 
0
 B  sin 2  ,

c  4 exp  hc   1 (2.7)
   kT  
  B  

 B  
 2  
sin 2 
B  2  (2.8)

where the integrand is Planck’s photon emittance at temperature TB, θ is a field of


view angle. The background-limited detectivity D* relative to 2π FOV becomes

1

DBLIP  

sin  2 
DBLIP  2  (2.9)

Equation (2.9) holds for photovoltaic detectors. Photoconductive detectors have


a lower D∗ by a factor of 2 :
BLIP
30 R. Savkina and O. Smirnov

1/ 2
    
DBLIP    (2.10)
2hc   B 

The following paragraphs will describe the main properties of the PС and PV IR
detectors. The theory of photoconductive and photovoltaic IR detectors can be
found in monographs [41, 43].

2.3.3 Photoconductive IR Detectors

Photoconductive IR detector is a type of photodetector that are based on semicon-


ductor materials whose conductivity increases under the absorption of incident pho-
ton flux density Φs(λ) resulted in non-equilibrium charge carriers’ generation.
Excitation of electrons into the conduction band or holes into the valence band
occurs from another band or impurity states within the band (impurity-bound states
in energy gap, quantum wells, or quantum dots). Alternative term for photoconduc-
tive detector is photoresistor, light-dependent resistor and photocell [31]. The pho-
toconductive effect involves applying a bias voltage across a uniform piece of
detector material to generate a photocurrent proportional to the photoexcited elec-
tron concentration (see Fig. 2.3). For low-resistance material, PC detector is usually
operated in a constant current circuit. For high-resistance photoconductors, a con-
stant voltage circuit is preferred, and the signal is detected as a change in current in
the bias circuit. Herewith, the basic value describing intrinsic or extrinsic photocon-
ductivity is a short-circuit photocurrent:

Fig. 2.3 Schematic view of unit cells for PC detector


2 Photoconductive and Photovoltaic IR Detectors 31

I ph  Aq s g, (2.11)

where A = wl is a detector area, g is a PC gain represented the number of elec-


trons flowing through the electrical circuit per photon absorbed [26], η is a quantum
efficiency, Φs – is the incident photon flux density. The value of PC gain depends
upon relation between the drift length Ld = vdτ and interelectrode spacing, l.
The basic requirements for high photoconductive responsivity at a given wave-
length λ follow from the expression for the voltage responsivity of PC detector:

Vs   Vb
Rv  
P At hc n0 (2.12)

where the absorbed monochromatic power Pλ = ΦsAhv. In other word, one must
have high quantum efficiency η, long excess carrier lifetime τ, the smallest possible
piece of crystal, low thermal equilibrium carrier concentration n0, and the highest
possible bias voltage Vb. Current gain of photoconductors is their advantage, which
leads to higher responsivity than is possible, for example, with non-avalanching
photovoltaic detectors. However, serious problem of photoconductors operated at
low temperature is nonuniformity of detector element due to recombination mecha-
nisms at the electrical contacts and its dependence on electrical bias.
The frequency dependent responsivity can be determined by the equation

  ef Vb 1
Rv 
At hc n0 1   2 2 1/ 2 (2.13)
ef

where τef is the effective carrier lifetime.


However, there are two limits on applied bias voltage and, respectively, voltage
responsivity increase – Joule heating of the detector element and sweep-out of
minority carriers. It should be noted that for typical detector sizes of 50 × 50 μm2
and long lifetimes of excess carriers (1μs in 8–14 μm devices at 77 K and 10 μs in
3–5 μm devices at higher temperatures), such parameters as contacts, drift, and dif-
fusion have a significant effect on the PC detector performance (read Section
9.1.1.1 in Rogalski [41]).
The total noise voltage of a photoconductor is
2
Vnoise  V j2  VGR
2
 V12/ f (2.14)

where terms in expression are the fundamental types of internal noise sources
usually operative in photoconductive detectors: Johnson–Nyquist (sometime called
thermal) noise (Vj), generation–recombination (G–R) noise (VGR) and the third form
of noise, not amenable to exact analysis, is 1/f noise (V1/f).
Johnson–Nyquist noise is associated with the finite resistance R of the device and
is caused by random movements of charge carriers whose temperature is bigger
32 R. Savkina and O. Smirnov

than 0 K. Their thermal energy increases as temperature increases. Thermal noise


occurs in the absence of external bias as a fluctuating voltage or current depending
upon the method of measurement. The root mean square of Johnson–Nyquist noise
voltage in the bandwidth Δf is given as

VJ2  4 kTRf (2.15)

where k is the Boltzmann constant. This noise has “white” character.


At finite bias currents, the carrier density fluctuations cause resistance variations,
which are observed as noise exceeding Johnson–Nyquist noise. This type of excess
noise in photoconductive detectors is referred to as G–R noise. The G–R noise is due
to the random generation of free charge carriers by the crystal vibrations and their
subsequent random recombination. It leads to conductivity changes that will be
reflected as fluctuations in current flow through the crystal. The (G–R) noise voltage
for equilibrium conditions is equal

Vgr2  2  G  R  lwt  Rqg  f


2
(2.16)

where G and R are the volume generation and recombination rates, g = τ/tt is a
photoconductive gain which can be defined as a ratio of free carrier lifetime, τ, to
transit time, tt, between the ohmic contacts, and l, w, t are the geometric dimensions
of the detector.
Finally, 1/f noise is associated with the presence of potential barriers at the con-
tacts, interior, or surface of the detector. Its reduction to an acceptable level depends
greatly on the processes of preparing the contacts and surfaces. The two most cur-
rent models for the explanation of 1/f noise are assuming the fluctuations in the
mobility of free charge carriers (Hooge’s model) [14] and in the free carrier density
(McWhorter’s model) [49].

2.3.4 Photovoltaic IR Detector

A photovoltaic effect occurs in structures with built-in potential barriers. The most
widely used PV detector is the p-n junction photodiode (see Fig. 2.4a), where a
strong internal electric field exists across the junction even in the absence of radia-
tion. When a photoexcited electron-hole pair are injected optically into the vicinity
of such barriers, the electron and hole are separated by the space-charge field caus-
ing a change in voltage across the open-circuit cell or a current to flow in the short-­
circuited case. The role of the built-in electric field is to cause the charge carriers of
opposite sign to move in opposite directions depending upon the external circuit.
Minority carriers are readily accelerated to become majority carriers on the other
side. This way a photocurrent is generated that shifts the current-voltage character-
istic in the direction of negative or reverse current, as shown in Fig. 2.4b.
2 Photoconductive and Photovoltaic IR Detectors 33

Fig. 2.4 (a) Schematic of typical p-n junction photodiode and (b) its current-voltage characteristic
for the illuminated and nonilluminated diode

The current-voltage (I-V) relationship for photodiode in which the dark current
is only due to diffusion current, is given by the well-known Shockley diode equation:

 qV 
J dark V   J s  exp  1 (2.17)
 kT 

With illumination, the total current density in the p-n junction becomes:

J V ,   J dark V   J ph    (2.18)

where the dark current density, Jdark, depends on V and the photocurrent density,
Jph = ηqAΦ, depends on the photon flux density Φ. Here Js is the saturation current.
As it was noted above, the theory of a conventional photodiode can be found in
monographs [41, 43]. Here we will discuss only the main points. For a photodiode,
the important parameters are open-circuit voltage and short-circuit current. The
open-circuit voltage produced by the accumulation of electrons and holes on the
two sides of the junction can be obtained by multiplying the short-circuit current by
1
 I  , where Vb is the bias voltage. In many applications the photodiode is
R 
 V V Vb
operated at zero-bias voltage Vb = 0 which corresponds to the resistance R0. The R0A
(Ω·cm2) product is a parameter that makes it possible to inform the order of magni-
tude of the dark current and makes it possible to evaluate the quality of the photode-
tector i.e., this is a figure of merit of a PV detector. This product for ideal
diffusion-limited diode is given by [41]:

kT
 R0 A D  (2.19)
qJ s

 1   1/ 2 1   1/ 2 
J s   kT 
1/ 2 2 1/ 2
nq
i
  e   h  (2.20)
 p   e  n  h  

34 R. Savkina and O. Smirnov

where p and n are the hole and electron majority carrier concentrations, τe and τh the
electron and hole lifetimes in the p- and n-type regions, respectively.
Long [26] and Tredwell and Long [47] showed that the fundamental noise mech-
anisms in HgCdTe photodiodes and HgCdTe photoconductors were the same. In the
case of the photodiode at zero-bias voltage Vb = 0 in thermal equilibrium, the result-
ing noise is identical to Johnson –Nyquist noise (see Eq. 2.15), while for a diode
exposed to background flux density Φb this expression becomes

 
VJ2  4 kT  2q 2 AR0  b R0 f
(2.21)
In the absence of a background-generated current, the current noise is equal to
the Johnson noise (4kTR0Δf) at zero bias, and it tends to the usual expression for
shot noise (2qIDΔf) for voltages greater than a few kT in either direction [41]. Due
to the absence of recombination noise, the limiting p–n junction’s noise level can
ideally be √2 times lower than that of the photoconductor.
Detectivity in the case of the photodiode can be determined as

1/ 2
 q  4 kT 
D    2q 2 B 
hc  R0 A  (2.22)
4 kT
Here in the case of background-limited performance and  2q  B , we
2

R0 A
obtain Eq. (2.10), but in the case of t thermal noise limited performance and
4 kT
 2q 2 B , Eq. (2.22) converts to
R0 A
1/ 2
 q  R0 A 
D 
2hc  kT 
(2.23)
It should be noted that for real diodes, the process of carrier transfer is much more
complicated, and to determine the quality of photodiodes, it is necessary to consider
such phenomena as generation–recombination within the depletion region, tunnel-
ing through the depletion region, surface effects, impact ionization, and space-­
charge limited current. For photodiode device structures, the critical region is the
p-n junction, which can be formed using different approaches, including, ion
implantation, Hg diffusion, impurity diffusion, and type conversion. The various
HgCdTe-based p–n junction photodiode architectures are collected and described in
review [37] and are presented in Table 2.2.
Over the years, several kinds of PV IR devices have been developed, including
homojunction and heterojunction photodiodes and metal-insulator-semiconductor
(MIS) photo-capacitors. Among them, p-n junction PV diodes have become the
mainstream technology. In general, there are two groups: one is an n-on-p junction
diode, and the other is a p-on-n junction diode, where the latter device has been
demonstrated to have lower dark current and thus higher operating temperature.
Regarding the fabrication technology, there are generally two classes of p-n
Table 2.2 Cross section views of various photodiode architectures
The front-illuminated planar Hg-diffused n-on-p
homojunction – one of the first HgCdTe photodiode
structures, pioneered by Societe Anonyme de
Telecommunications [27] and used successfully to
achieve electrical bandwidths of >2 GHz for CO2 laser
detection at 9.6–10.6 μm. It is formed in p-type bulk-­
grown material with Hg vacancies as the acceptors. It is
(a) the first reported use of an interdiffused CdTe layer for a
HgCdTe device passivation.
The back-illuminated planar ion implanted n+-n-p
homojunction was the first architecture to be used for
hybrid HgCdTe FPAs. The p-type absorber layer was
grown by LPE onto an IR-transparent substrate such as
CdTe, CdZnTe or sapphire. Factor-of-ten higher R0A
values for this structure were attributed to higher lifetime
in the p-type absorber layer [7].
(b)
The back-illuminated mesa P-on-n heterojunction has
become the most widely applicable junction architecture
for bump-mounted hybrid HgCdTe FPAs. The absorber
layer is doped with In at 1 × 1015 cm−3 or less. The p-type
layer is doped with As at (1÷4) × 1017 cm−3. There is a
thin (3÷5 μm) interdiffused layer between the n-type
absorber and the substrate, with compositional grading
for preventing recombination at interface. The first
(c) back-illuminated double-layer heterojunction photodiodes
in HgCdTe with N-on-p polarity was reported by [38].
The back-illuminated As-implanted p-n-N planar
buried-junction heterostructure is the first photodiode
structure to exploit the low-temperature (~175 °C) growth
and bandgap engineering provided by MBE [5]. The
unique feature of this structure is that the junction is
buried below the top wide-bandgap n-type layer. Another
feature is the wide-gap n-type buffer layer between the
absorber layer and the substrate for prevents carrier
(d) recombination at interface.
P-π-N non-equilibrium photodiode introduced by Elliott
and Ashley and known as HOT (High Operating
Temperature) detector [3]. Ashley and Elliott proposed
the device configuration for suppression of the Auger
process by reducing the electron and hole concentrations
below their equilibrium values. Two configurations – pνn+
and nπp+, are possible to achieve such nonequilibrium
devices. More information about HOT detectors – see
Sect. 2.5.
(e)
This is the lateral collection photodiode as implemented by
Rockwell in their back-illuminated As-implanted planar
buried-junction heterostructure. The lateral collection device
uses several very-small-area photodiodes to cover a larger
optical area through lateral collection of photocurrent. There
is also lateral collection of diffusion current. The key benefit
that the lateral collection architecture has for HgCdTe is to
reduce the number of tunneling-related defects that are
intercepted by the junction depletion region, which is
(f) particularly important for lower temperature operation, e.g.,
40 K for LWIR, and for very large optical areas.
Source: Data extracted from Ref. [37]
36 R. Savkina and O. Smirnov

Fig. 2.5 Schematic diagram of (a) mesa structure and (b) planar structure of p-on-n HgCdTe
photodiodes. (Reprinted with permission from Ref. [25]. Copyright 2015: AIP Publishing)

junction photodiodes processing structure used: mesa structures and planar struc-
tures, which are illustrated in Fig. 2.5. The processing procedure of a planar device
structure usually involves surface passivation, window etching, p-n junction forma-
tion, followed by metal contact deposition, whereas that of mesa device structures
usually involves p-n junction formation, mesa isolation etching, followed by surface
passivation and metal contact formation [25].
A popular alternative to the simple p-n photodiodes especially for ultrafast pho-
todetection in optical communication, measurement, and sampling systems is the
p-i-n photodiode. A detailed discussion of such photodetectors can be found in
Donati [8].
Schottky barrier photodiodes have also found application as IR detectors. These
devices reveal some advantages over p–n junction photodiodes: fabrication simplic-
ity, absence of high temperature diffusion processes, and high speed of response
[41]. This is a majority carrier device with the thermionic emission process which
much more efficient than the diffusion process. Therefore, for a given built-in volt-
age, the saturation current in a Schottky diode is several orders of magnitude higher
than in the p–n junction. Recent studies have demonstrated that the combination of
n-type HgCdTe and p-type graphene can construct a Schottky junction to effectively
dissociate photogenerated electron-hole pairs, resulting in a highest external quan-
tum efficiency of 69.06% at 3 μm and maximum high responsivity of 2.6 A/W in
visible to mid-infrared wavelength region (0.5÷5.2) μm [53]. A heterostructure of
Bi-Layer Graphene-CdTe-HgCdTe for MWIR photodetector that has an Ohmic
contact for the electrons but a Schottky barrier for the holes was proposed by [10].
The properties of various insulator/HgCdTe interfaces have attracted numerous
experimental investigations due to the demand for the high-performance HgCdTe
IR detectors, such as PV detectors or metal-insulator-semiconductor (MIS) photo-
diodes. MIS photodiode consists of a metal gate separated from a semiconductor
surface by an insulator. By applying a negative voltage to the metal electrode, elec-
trons are repelled from the insulator – semiconductor interface, creating a depletion
region. When incident photons create hole–electron pairs, the minority carriers drift
a way to the depletion region and the volume of the depletion region shrinks. The
total amount of charge that a photogate can collect is defined as its well capacity.
2 Photoconductive and Photovoltaic IR Detectors 37

The general theory of MIS devices as applied to HgCdTe is reviewed in [20]. The
passivation effect of ZnS [30], Al2O3 [54], CdTe [51] and other materials deposition
on a HgCdTe MIS detector were investigated.

2.4 High Operation Temperature IR Detector

Usually, IR detectors are used mainly in special areas, so the main task is to opti-
mize their sensitivity, spatial and time resolution. To achieve good signal-to-noise
performance and a very fast response, the photon IR detectors require cryogenic
cooling. To obtain high performance HgCdTe-based detectors, temperatures typi-
cally 80÷200 K are required. Significant cooling is needed to reduce noise and leak-
age currents resulting from the thermal generation and recombination (G-R)
processes near room temperature. The cooling adds substantial cost and bulk to the
IR sensor, thus, in general, the photon IR sensors are relatively complex, costly, and
not highly portable. At the same time affordability is ignored. Today, commercial
and government industries (medicine, coercion and rescue services, transportation,
etc.) are showing increasing interest in “affordable” IR receivers. The civil market
forms requirements to the price, the sizes, convenience in use and accordingly
roughens such parameters, as sensitivity, equivalent noise of a difference of tem-
peratures, and inertia of receivers.
The detectivity of an optimized IR detector (see Eq. 2.5) is depended on the ratio
of the absorption coefficient to the thermal generation α , i.e., is limited by the
G
processes in the device workspace. As we know (see, for example, [19]), the most
important generation and recombination mechanisms in HgCdTe: Shockley-Read-­
Hall (SRH), radiative and Auger. Auger processes become more important and
degrade the performances of HgCdTe-based IR devices as the energy band gap is
decreased and/or the temperature is increased. The Auger mechanism imposes fun-
damental limitations to the LWIR HgCdTe detector performance. The Shockley-­
Read-­ Hall (SRH), recombination mechanism involves the recombination of
electron-hole pairs via defect levels within the energy bandgap of the material.
Therefore, it is not an intrinsic process and its effect on detector noise and external
quantum efficiency can be suppressed as HgCdTe processing improves. To reduce
the dark current and increase the detector operating temperature, it is essential to
suppress these defect-related dark current mechanisms, in which passivation plays
a critical role. Successes of IR devices performance associated with passivation are
reviewed in [25].
There are several strategies of improving IR photon detectors high-temperature
performance such as the suppression of radiative recombination using photon recy-
cling [22], the suppression of both radiative and Auger recombination with carrier
depletion [3, 4], the suppression of Auger recombination using band structure engi-
neering in strained layer superlattices – both for III-V [13] and for II-VI [17] com-
pounds. All three effects provide significant enhancements in detectivities.
Development of the high operating temperatures (HOT) photodetectors based on
38 R. Savkina and O. Smirnov

II-VI compounds were specifically focused on HgCdTe solid solutions [36, 42].
Many new concepts have been implemented and tested to improve their perfor-
mance [18, 29, 45].

2.4.1 Ways to Improve Detector’s Performance


Without Cooling

Key improvements of IR technology through innovations in material growth and


device design have made it possible to create IR devices that operate at room tem-
perature. Next, we will talk about barrier structures, Auger suppressed structures
and photon trapping structures.
A new concept of infrared detector named the nBn detector has been proposed by
Maimon and Wicks [28]. Figure 2.6 shows examples of the barrier IR HOT detec-
tors and their bandgap diagrams, that eliminate the cooling requirements of photo-
detectors operating in the MWIR range. nBn detector is designed to reduce the dark
current (G-R current originating within the depletion layer associated with
Shockley–Read–Hall process) and noise without impeding the photocurrent (sig-
nal). The barrier serves also to reduce the surface leakage current. The n-type semi-
conductor on one side of the barrier constitutes a contact layer for biasing the device,
while the n-type narrow-bandgap semiconductor on the other side of the barrier is a
photon-absorbing layer whose thickness should be comparable to the absorption
length of light in the device, typically several microns. The barrier should be located
near the minority-carrier collector and away from the region of optical absorption.
nBn – detector somewhat resembles a typical p–n photodiode, except that the junc-
tion (space-charge region) is replaced by an electron-blocking unipolar barrier (B),
and that the p-contact is replaced by an n-contact. The structure can filter out major-
ity carriers while collecting minority carriers, similarly to a photodiode. It can be
stated that the nBn design is a hybrid between a PC and PV photodetector.

Fig. 2.6 Schematic energy band diagram of an ideal nBn detector under (a) zero bias and (b)
illumination and low reverse bias V. (Reprinted with permission from Ref. [25]. Copyright 2015:
AIP Publishing)
2 Photoconductive and Photovoltaic IR Detectors 39

The results obtained by [1] indicate that the composition, doping, and thickness
of the barrier layer in MWIR HgCdTe nBn detectors can be optimized to yield per-
formance levels comparable with those of ideal HgCdTe p-n photodiodes. It is also
shown that introduction of an additional barrier at the back contact layer of the
detector structure (nBnn+) leads to substantial suppression of the Auger G-R mech-
anism. This results in an order-of-magnitude reduction in the dark current level
compared with conventional nBn or p-n junction-based detectors, thus enabling
background-limited detector operation above 200 K. The idea to optimize the nBn
structure by an asymmetric barrier - abrupt on the contact side to efficiently block
the majority carriers, and gradual on the absorption layer side to plane down the
remaining potential barrier for the collected photocarriers, is discribed in [12].
The Auger suppressed structure also provides a significant reduction in dark cur-
rent at high operating temperatures. This device structure first proposed by Ashley
and Elliott [3] is formed as the lightly doped absorber located between two layers of
high doped and high bandgap. The absorber region of the device is near intrinsic and
is π−type (lightly doped p-type) or ν−type (lightly doped n-type). In the case of
high reverse bias, the device design (P+/v/N+) reduces the carrier concentration in
the central absorber layers well below thermal equilibrium and, thus, suppresses
Auger processes and effectively reduces the dark current. It has been shown that the
reduction in the dark current is around 12 times at temperatures above
200 K. Numerical simulations also shows that the dark current is significantly
smaller in Auger suppressed structure compared with standard double layer planar
heterostructure (DLPH) ones, and this leads to improved detectivities for a wide
range of temperatures [50]. A MWIR (6 μm cutoff) DLPH detector and a HOT
detector with the same cutoff will operate with D* = 5 × 1010 cm Hz1/2 W−1 detectiv-
ity at ~120 K and ~170 K respectively. By implementing an Auger suppressed
architecture, an improvement of 50 K in the operating temperature opens the pos-
sibility of thermoelectrically (TE) cooled high performance MWIR IR photon
detectors.
There are other ways to improve the performance of IR detectors associated with
size effect of reducing the amount of thermal generation as well as optical optimiza-
tion [25, 33]. Thermally generated dark current is directly proportional to the mate-
rial volume of the detector. Photon trapping technology is an important approach to
reducing dark current and enhancing the detector operating temperature. Photonic
crystal structures can be used to effectively enhance the light absorption in detectors
with small material volume. To reduce dark current without dropping quantum effi-
ciency, it was proposed the photonic crystal structure which is a HgCdTe-based
MWIR [52] and LWIR IR detector [25].
One approach to reducing the material volume of the detector is to reduce the
thickness of the absorption layer. For example, it was found that reducing the thick-
ness of the absorption region of HgCdTe photodiodes operating at 200 K [23] is
good solution to improve the response speed of a detector. On the other hand, for
standard HgCdTe detectors, most of the light is absorbed in the absorption layer
40 R. Savkina and O. Smirnov

Fig. 2.7 Schematic structure of interference enhanced photodetectors: (a) the simplest structure,
interference occurs between the waves reflected at the rear and the front surface of semiconductor
(b) structure immersed between two dielectric layers supplied with backside reflector, and (c)
structure immersed between two photonic crystals. (Reprinted with permission from Ref. [33].
Copyright 2004: Elsevier)

around 10 μm, leading to a high quantum efficiency. However, devices with too thin
absorption layer (to less than 1 μm) would suffer from poor quantum efficiency and
reduced responsivity. Therefore, maintaining high light absorption in a thin HgCdTe
layer becomes the critical point.
A possible way to improve the performance of IR detector is to reduce its thick-
ness that can be achieved by using multiple pass of radiation with a backside reflec-
tor. Various proposed optical resonator structures are shown in Fig. 2.7. More
efficient light absorption can be achieved by utilizing interference phenomena to
setup a resonant cavity within the detector by using two dielectric layers on the front
and rear surface of the detector, the schematic structure of which is shown in
Fig. 2.7b. An important limitation of the optical cavity application is that the gain in
quantum efficiency can be achieved only in narrow spectral regions. Another limita-
tion comes from the fact that efficient optical resonance occurs only for near-­
perpendicular incidence and is less effective for oblique incidence.
An increase in apparent “optical” size of the detector using a various type of opti-
cal concentrators such as optical cones, conical fibers, and other types of optical
concentrators also leads to an increase in the performance of IR photodetector.
More efficient enhancement can be achieved by placing the absorber in optical reso-
nant cavities (see Ref. [18]).
In addition to classical photoresistors and photodiodes, three other types of IR
detectors can operate at near 300 K: magnetic concentration detectors, photoelec-
tromagnetic (or PEM) detectors and Dember effect detectors [33].
2 Photoconductive and Photovoltaic IR Detectors 41

2.4.2 Photoelectromagnetic Effect IR Detectors

The photoelectromagnetic effect is caused by an in-depth diffusion of photogene-


rated carriers whose trajectories are deflected in a magnetic field. The driving force
for diffusion is the gradient of carrier concentration that is caused by the non-­
uniform absorption of radiation in the different layers of bulk semiconductor.
Theoretical analysis indicates that the maximum voltage responsivities of PEM
detectors can be reached in relatively strong magnetic fields B ~μe−1 for samples
with high resistance. At room temperature the radiation is almost uniformly absorbed
within the diffusion length in HgCdTe material since the ambipolar diffusion length
in narrow gap semiconductors is several μm while the absorption of radiation is
relatively weak (1/α ≈ 10 μm). In such cases, a low recombination velocity at the
front surface and a high recombination velocity at the back surface is necessary for
a good PEM detector response.
The best at the time of 2004, uncooled HgCdTe 10.6 μm PEM detectors exhib-
ited experimental voltage responsivity of about 0.1 V/W (width of 1 mm) and detec-
tivities of about 1 × 107 cm Hz1/2 W−1, which is by a factor of ~3 below the predicted
ultimate value 3.4 × 107 cmHz1/2 W−1, [33]. Gaziyev et al. [11] have reported that the
specific detectivity D* of manufactured on the basis of HgCdTe (x = 0,2) monocrys-
tals PEM detector for MWIR region with a maximum responsivity near 6 μm was
about (0.8÷0.9) 108 cmHz1/2 W−1 on frequency of 1200 Hz.
As of today, company VIGO System S.A. has developed uncooled HgCdTe-­
based photovoltaic IR detectors based on PEM effect in the semiconductor with
peak detectivity on the level 1.6 × 108 cmHz1/2 W−1 (see Table 2.3). The devices are
designed for the maximum performance at 10.6 μm and especially useful as a large
active area device to detect CW (Continuous Wave) and low frequency modulated
radiation. These devices are mounted in specialized packages with incorporated
magnetic circuit inside. Zinc selenide anti-reflection coated (wZnSeAR) window
prevents unwanted interference effects and protects against pollution. Spectral
detectivity of HgCdTe-based PEM detectors presented in Table 2.3 is shown in
Fig. 2.8.
Today the measured performance of PEM detectors is the same as that of the PC
detectors for the same spectral range (read Sect. 2.5). In contrast to photoconduc-
tors, PEM detectors do not require electrical biasing. The frequency characteristics

Table 2.3 2.0–12.0 μm HgCdTe ambient temperature PEM detectors with ZnSe window.
Optimum wavelength λopt = 10.6 μm
HgCdTe-­ Current
based PEM Optical area, responsivity-optical Peak Time
detector A0 Resistance, area length product, Detectivity, Da constant, t
20 °C (mm × mm) R (Ω) Ri (A·mm/W) (cm Hz1/2 W−1) (nsec)
PEM series 1×1 ≥40 ≥0.002 ≥2.0 × 107 ≤1.2
PEMIa series 2×2 40–100 ≥0.01 ≥1.6 × 108
Source: Data extracted from https://vigo.com.pl/
a
Optically immersed
42 R. Savkina and O. Smirnov

Fig. 2.8 Spectral detectivity of HgCdTe-based PEM detector (a) with optical immersion and (b)
without optical immersion; (c) detector image and (d) cross section of a back side illuminated ele-
ment of PEM detector. (Reprinted from https://vigo.com.pl/)

of a PEM detector is flat over a wide frequency range, starting from constant cur-
rent. This is due to lack of low frequency noise and a very short response time [33].
PEM detectors have been also fabricated with other HgTe-based ternary compound
such as HgZnTe and HgMnTe [34].

2.4.3 Magnetoconcentration IR Detectors

Unlike the PEM detector, the device based on the magnetoconcentration effect is
biased with an electric field perpendicular to the magnetic field. As a result, the
Lorentz force directs current carriers to the surface of high surface recombination
velocity, the concentration of electrons and holes in the bulk of the material is
diminished and thus the regions become depleted. This leads to reduction of ther-
mally generated noise. At the same time, such devices exhibited a large low-­
frequency noise when biased to achieve a sufficient depletion of semiconductor that
have been severe obstacles to their widespread applications. Improvement of perfor-
mance was observed at frequency >100 kHz.
2 Photoconductive and Photovoltaic IR Detectors 43

2.4.4 Dember Effect IR Detectors

Detectors based on the Dember effect have also been proposed [9]. Dember effect
is known as an effect of photodiffusion – the appearance of the potential difference
in the direction of radiation absorption and bulk photodiffusion. Two conditions are
required for the photovoltage generation: the distribution of photogenerated carriers
should be nonuniform and the diffusion coefficients of electrons and holes must be
different. The best performance is achievable for a device with thickness of the
order of a diffusion length, with a low surface recombination velocity and a low
reflection coefficient at the illuminated front surface and with a large recombination
velocity and a large reflection coefficient at the nonilluminated back surface
(Fig. 2.9b). Since the device is not biased and the noise voltage is determined by the
Johnson–Nyquist thermal noise. Detectivities as high as ≈ 2.4 × 108 cmHz1/2 W−1
and of ≈ 2.2 × 109 cmHz1/2 W−1 are predicted for optimized 10.6 μm devices at 300
and 200 K, respectively (see Fig. 2.9a).

2.5 PC and PV IR Detectors Manufacturing

The first report about photoconductive effect in HgCdTe was published by Lawson
et al. in 1959 [24]. Historical review of the further development of HgCdTe-based
photoconductors is summarized in [40] and in monograph Infrared Detectors [41].
Infrared photon detector technology has reached a high level of sophistication. The
results of study in this area made it possible to move to the stage of device
manufacturing.

Fig. 2.9 The calculated normalized responsivity (RvA), detectivity (D*) and bulk recombination
time (τ) of uncooled 10.6-μm HgCdTe Dember detector as a function of acceptor concentration.
(Reprinted with permission from Ref. [9]. Copyright 1991: Elsevier)
44 R. Savkina and O. Smirnov

Table 2.4 HgCdTe photoconductive detectors produced by Teledyne Judson Technologies


HgCdTe-based PC Active
detector series Size Cutoff Peak Time
J15n* (square), Wavelength, Wavelength, Peak D*@ constant,
mm μm μm 10KHz, cm t, μsec
Hz1/2 W−1
n = 12 0.25…4 >12 11 ± 1 5 × 1010…1.5 × 1010 0.15÷0.5
n = 14 0.1…2 >13.5 13 5 × 1010…2.5 × 1010 0.5
n = 16 0.1…2 >16.6 14 4 × 1010…2 × 1010 0.3
n = 22 0.25…2 >22 16 1 × 1010…6 × 109 0.1
n = 24 1 >24 20 5 × 109 0.1
*
J15 Series detectors are designed for operation in the 2–26 μm wavelength region and for
cryogenic operation at 77 °K.
Thermoelectrically Active
cooled HgCdTe-­ size Cutoff Peak Peak D*@ 10KHz, Time
based PC detector (square), wavelength, wavelength, cm Hz1/2 W−1 constant,
mm μm μm t, μsec
Series J15TE2 0.25…1 4.0 ± 0.25 ~4 4 × 1010 5
−40 °C 0.25…1 4.5 ± 0.25 ~4.4 2.5 × 1010 3
0.25…1 5.0 ± 0.25 ~4.8 1.5 × 1010 2
Series J15TE3 0.1…1 >5 ~4.8 3 × 1010 2
−65 °C 0.25…1 – 10.6 2 × 108 1…5
Series J15TE4 0.1…1 >5 ~4.8 6 × 1010 2
−80 °C 0.25…1 – 10.6 6 × 108 1…5
Source: Data extracted from www.teledynejudson.com

Fig. 2.10 Spectral detectivity of HgCdTe-based PC detector series J15n (a) and thermoelectri-
cally cooled HgCdTe-based PC detector series J15TE (b) produced by Teledyne Judson
Technologies. (Reprinted from www.teledynejudson.com)
2 Photoconductive and Photovoltaic IR Detectors 45

Teledyne Judson Technologies (TJT) is a global leader for IR sensors from the
Visible to VLWIR spectrum from 1969 year. This company produces the back-
ground limited PC detectors with state-of-the-art performance (see Table 2.4) for
thermal imaging, CO2 laser detection, industrial process control, FTIR spectros-
copy, missile guidance, and night vision. J15 Series PC HgCdTe detectors is com-
posed of a thin layer (10–20 μm) of HgCdTe with metalized contact pads defining
the active area. Spectral detectivity of HgCdTe-based PC detector series J15n is
shown in Fig. 2.10a. Its value reaches 1011 cm Hz1/2 W−1 for MWIR spectral region
and is at the level of 1010 cm Hz1/2 W−1 and higher for the LWIR spectral region, i.e.,
these detectors work at the BLIP level. These detectors are low impedance devices,

Fig. 2.11 Spectral detectivity of HgCdTe-based PV detector series J19 with 2.8 μm cutoff (a) and
5 μm cutoff (b) produced by Teledyne Judson Technologies. (Reprinted from www.teledyne-
judson.com)
46 R. Savkina and O. Smirnov

typically 10–150 ohms, and require a low voltage noise preamplifier. It should be
noted that the series of thermoelectrically cooled HgCdTe PC detectors designed for
industrial and military applications demonstrates high detectivity at the level
(1÷5) × 1010 cm Hz1/2 W−1 in the 2–5 μm wavelength region without liquid nitrogen
cooling (see Fig. 2.10b).
J19 Series PV HgCdTe detectors are high-quality photodiodes for use in the
500 nm to 2.8 μm and 500 nm to 5.0 μm spectral ranges (see Fig. 2.11a, b). Unlike
the photoconductors commonly used in the 500 nm to 5.0 μm region, HgCdTe pho-
todiodes operate in the photovoltaic mode and do not require a bias current for
operation. According to the manufacturer this makes J19 detectors the better choice

Table 2.5 HgCdTe photovoltaic detectors produced by Teledyne Judson Technologies


Active size Shunt Dark Peak D*@ 1
diameter, impedance, current@ KHz, cm
HgCdTe-based PV detector series J19 mm Ohm −0.1 V, A Hz1/2 W−1
J19:2.8-18C (*↓) 0.25 1.5 × 104 2.0 × 10−6 2.8 × 1010
22 °C 1.00 1.5 × 103 2.0 × 10−5 3.5 × 1010
J19TE1:2.8-66C 0.25 2.0 × 105 1.0 × 10−7 1.1 × 1011
−20 °C 1.00 2.0 × 104 1.0 × 10−6 1.4 × 1011
J19TE2:2.8-66C 0.25 1.5 × 106 2.0 × 10−8 2.9 × 1011
−40 °C 1.00 1.5 × 10 5
2.0 × 10−7 3.7 × 1011
J19TE3:2.8-66C 0.25 8.0 × 10 6
5.0 × 10−9 5.9 × 1011
−65 °C 1.00 8.0 × 10 5
5.0 × 10−8 7.4 × 1011
J19TE4:2.8-3CN 0.25 3.2 × 10 7
3.0 × 10−9 8.0 × 1011
−85 °C 1.00 3.2 × 10 6
3.0 × 10−8 1.0 × 1012
J19TE4:2.8-3VN 0.25 6.4 × 10 7
2.0 × 10−9 8.6 × 1011
−90 °C 1.00 6.4 × 10 6
2.0 × 10−8 1.1 × 1012
*
Series detectors are high-quality HgCdTe photodiodes for use in the 500 nm to 2.8 μm spectral
range. Data for detectors with active size 1 mm; 50% Cutoff Wavelength is 2.8 μm; Peak
Wavelength is 2.6 μm; Peak Responsivity is 1.3 A/W. FOV is 180°, 60° FOV at
−90 °C. Maximum reverse bias voltage for all detectors is 0.2 V.
J19TE1:5-66C (**↓) 0.25 4.0 × 102 5.0 × 10−5 5.6 × 109
−20 °C 1.00 4.0 × 101 5.0 × 10−4 4.7 × 109
J19TE2:5-66C 0.25 1.0 × 103 2.0 × 10−5 1.1 × 1010
−40 °C 1.00 1.0 × 102 2.0 × 10−4 1.0 × 1010
J19TE3:5-66C 0.25 3.2 × 10 3
6.0 × 10−5 2.2 × 1010
−65 °C 1.00 3.2 × 10 2
6.0 × 10−4 2.4 × 1010
J19TE4:5-3CN 0.25 7.2 × 10 3
3.0 × 10−6 3.8 × 1010
−80 °C 1.00 7.2 × 10 2
3.0 × 10−5 4.6 × 1010
J19TE4:5-3VN 0.25 1.2 × 10 4
2.0 × 10−6 5.4 × 1010
−90 °C 1.00 1.2 × 10 3
2.0 × 10−5 6.8 × 1010
**
Series detectors are high-quality HgCdTe photodiodes for use in the 500 nm to 5.0 μm
spectral range. Data for detectors with active size 1 mm; 50% Cutoff Wavelength is 5.0 μm;
Peak Wavelength is 4.5 μm; Peak Responsivity is 1.0–2.2 A/W. FOV is 180°, 45° FOV at
−90 °C. Maximum reverse bias voltage for all detectors is 0.5 V.
Source: Data extracted from www.teledynejudson.com
2 Photoconductive and Photovoltaic IR Detectors 47

for DC and low-frequency applications, as it does not exhibit the low frequency or
1/f noise characteristic of the PbS, PbSe and HgCdTe photoconductors. The J19TE
detectors are mounted on thermoelectric coolers (TEC) where one-stage (TE1) is
used for −20 °C operation and subsequent stages (TE2, TE3, TE4) are used to
achieve lower temperatures (see Table 2.5). Cooling an HgCdTe photodiode reduces
noise and improves detectivity. Cooling also increases shunt resistance HgCdTe
photodiodes and improve their response at longer wavelengths with a reduction in
temperature. J19TE detectors offer superior pulse response for applications in mon-
itoring and detecting high-speed pulsed lasers.
Another company, VIGO System S.A., has developed a unique technology for
manufacturing instruments for quick and convenient detection of 1–16 μm infrared
radiation. IR detectors operate in ambient temperature or are cooled with simple and
inexpensive thermoelectric coolers (see Table 2.6). PC series features uncooled IR

Table 2.6 HgCdTe photoconductive detectors produced by VIGO System S.A.


HgCdTe-based PC Active size Peak Time
series*/PCI series** (square), Resistance wavelength, Peak D*, cm constant,
20 °C mm2 R, Ω μm Hz1/2 W−1 t, nsec
PC-5/PCI-5 From ≤1200 5.0 ≥1.5 × 109/6.0 × 109 ≤5000
PC-6/PCI-6 0.05 × 0.05 ≤600 6.0 ≥7.0 × 108/2.5 × 109 ≤500
PC-9/PCI-9 to 4 × 4 ≤300 9.0 ≥1.0 × 108/5 × 108 ≤10
PC-10.6/PCI-10.6 ≤120 10.6 ≥1.9 × 107/1.0 × 108 ≤3
*
1.0–12.0 μm HgCdTe ambient temperature photoconductive detectors.
**
1.0–12.0 μm HgCdTe ambient temperature, optically immersed photoconductive detectors
Thermoelectrically Active size Peak Time
cooled HgCdTe-­ (square), Resistance wavelength, Peak D*, cm constant,
based PC detector mm R, Ω μm Hz1/2 W−1 t, nsec
PC-2TE series* From ≤1200 5.0 ≥2.0 × 1010 ≤20,000
230 °K 0.05 × 0.05 ≤800 6.0 ≥6.0 × 109 ≤4000
to 2 × 2 ≤400 9.0 ≥9.0 × 108 ≤40
≤300 10.6 ≥4.0 × 108 ≤10
≤200 12.0 ≥1.0 × 108 ≤3
≤150 13.0 ≥4.0 × 107 ≤2
PC-3TE series ** From ≤400 9.0 ≥1.5 × 109 ≤60
210 °K 0.05 × 0.05 ≤300 10.6 ≥4.5 × 108 ≤20
to 2 × 2 ≤300 12.0 ≥1.8 × 108 ≤5
≤300 13.0 ≥1.2 × 108 ≤4
PC-4TE series *** From ≤500 9.0 ≥1.9 × 109 ≤80
195 °K 0.05 × 0.05 ≤400 10.6 ≥5.0 × 108 ≤30
to 2 × 2 ≤400 12.0 ≥4.0 × 108 ≤7
≤400 13.0 ≥2.0 × 108 ≤6
≤300 14.0 ≥1.0 × 108 ≤5
*
1.0–14.0 μm HgCdTe two-stage thermoelectrically cooled PC detectors
**
1.0–15.0 μm HgCdTe three-stage thermoelectrically cooled PC detectors
***
1–16 μm HgCdTe four-stage thermoelectrically cooled PC detectors
Source: Data extracted from https://vigo.com.pl/
48 R. Savkina and O. Smirnov

PC detectors based on sophisticated HgCdTe heterostructures for the best perfor-


mance and stability. The devices should operate in optimum bias voltage and cur-
rent readout mode. Performance at low frequencies is reduced due to 1/f noise. The
1/f noise corner frequency increases with the cut-off wavelength.
By 2003, VIGO had been manufacturing detectors with the use of Isothermal
Vapour Phase Epitaxy (ISOVPE) of HgCdTe. It did not require costly devices, but
on the other hand it did not enable fabrication of complicated semiconductor struc-
tures, like ones in which the band gap could be increased or decreased in consecu-
tive layers of a heterostructure. The possibility of free spatial shaping of the band
gap and of doping level adjustment was ensured by a Metal Organic Chemical
Vapour Deposition (MOCVD) reactor commissioned in 2003.
Optical immersion technique was used to improve parameters of VIGO System
detectors (PCI series, see Table 2.6). Optical immersion means the use of a certain
type of lens, which is an integral part of an IR detector. The lens enables collecting
an amount of optical radiation falling on the device larger than the one that could be
collected only due to the physical area of the device. As a result, the detector picks
up more usable signal – as much as a larger device – while retaining a smaller area.
A detector with an immersion lens is best suited for operation with low power of
optical signal, which means the applications where the highest detectivity of the
detector is required. The use of optical immersion enables improving the signal-to-­
noise ratio almost 11 times without any need for additional customer’s interference
with the measurement system. The degree of detector parameters improvement
depends on the material from which the immersion lens is made. In the VIGO
System detectors comprising a GaAs substrate and an integrated immersion lens
made of the same material, the refractive index of the lens is equal to 3.3. That
means the detectivity is improved 3.3 times in a detector with a hemispherical lens,
and nearly 11 times in a detector with a hyperhemispherical lens.
Figure 2.12 shows the performance of the PC HgCdTe devices produced by
VIGO System S.A. Without optical immersion MWIR photovoltaic detectors are
sub−BLIP devices with performance close to the generation−recombination limit,
but well−designed optically immersed devices approach BLIP limit when thermo-
electrically cooled with 2−stage Peltier coolers. Situation is less favorable for >8−
μm LWIR photovoltaic detectors; they show detectivities below the BLIP limit by
an order of magnitude. Typically, the devices are used at zero bias.
HgCdTe ambient temperature PV detectors with anti-fringing technology
(PV-5-AF series) applied for gas detection, monitoring and analysis (CO, HF, NH3,
C2H2, CH4, C2H6, HCl, H2CO, SO2, CO2, N2O) are presented in the Table 2.7. To
make these detectors immune to unwanted optical fringing effects, VIGO developed
anti-fringing technology (internal modification of substrate’s surface) and success-
fully applied it. Characteristics of the uncooled IR PV multiple junction detectors
(PVM series) designed for the maximum performance at 10.6 μm and applied for
the laser power monitoring are also shown in the Table 2.7.
Figure 2.13 shows peak detectivity D* of IR PV HgCdTe-based detectors pro-
duced by VIGO System S.A. without (1) and with (2) optical immersion. Reverse
bias may significantly increase response speed and dynamic range. It also results in
2 Photoconductive and Photovoltaic IR Detectors 49

Fig. 2.12 Typical spectral detectivity of HgCdTe detectors with 2−stage TE coolers (solid lines).
The best experimental data (white dots) are measured for detectors with FOV equal 36°. BLIP
detectivity is calculated for FOV = 2π. Black dots are measured for detectors with 4−stage TE
cooler. (Reprinted from Ref. [29]. Published 2013 by Polish Academy of Sciences (PAN) as
open access)

Table 2.7 HgCdTe ambient temperature and minor cooling photovoltaic detectors produced by
VIGO System S.A.
Active
size Peak Peak D*, Time
HgCdTe-based PV detectors (square), Resistance wavelength, cm constant,
20 °C mm2 R, Ω μm Hz1/2 W−1 t, nsec
PV-5-AF1 × 1-TO39-NW-90 1×1 ~8 4.4 ± 0.2 ~1.45 × 109 ≤570
PV-5-AF0.1 × 0.1-TO39-NW-90 0.1 × 0.1 ~265 ~3.55 × 109 ≤177
PVM-10.6-1 × 1-TO39-NW-90 1×1 ≥30 8.5 ± 1.5 ≥2.0 × 107 ≤1.5
PVM-2TE-10.6-1 × 1-TO8-­ ≥90 8.5 ± 2.0 ≥2.0 × 108 ≤4
wZnSeAR-70 (230 °K)
Source: Data extracted from https://vigo.com.pl/

improved performance at high frequencies, but 1/f noise that appears in biased
devices may reduce performance at low frequencies. The value of the detectivity
increases both with low cooling and in optically immersed PV devices. Moreover,
the combination of these methods allowed VIGO System S.A to achieve the peak
parameters which are comparable with detectivity of the PV devices obtained by
Teledyne Judson Technologies (see Table 2.5 and Fig. 2.11).
50 R. Savkina and O. Smirnov

Fig. 2.13 Peak detectivity of IR photovoltaic detectors based on sophisticated HgCdTe hetero-
structures for the best performance and stability without (1) and with (2) optical immersion. The
arrow indicates the direction of the operating temperature decreasing of the detector – from 20 °C
(lower D* value) to 195 °K (four-stage thermoelectric cooler). (Data extracted from https://vigo.
com.pl/)

References

1. Akhavan ND, Jolley G, Umana-Membreno GA, Antoszewski J, Faraone L (2015) Theoretical


study of midwave infrared HgCdTe nBn detectors operating at elevated temperatures. J
Electron Mater 44:3044–3055
2. Ameurlaine J, Rousseau A, Nguyen-Duy T, Triboulet R (1988) (HgZn)Te infrared photovol-
taic detectors. Proc SPIE 0929:14–20
3. Ashley T, Elliott CT (1985) Non-equilibrium devices for infrared detection. Electron Lett
21(10):451–452
4. Ashley T, Elliott CT, Harker AT (1986) Non-equilibrium modes of operation for infrared
detectors. Infrared Phys 26(5):303–315
5. Bajaj J (2000) State-of-the-art HgCdTe infrared devices. Proc SPIE 3948:42–54
6. Becla P (1986) Infrared photovoltaic detectors utilizing Hg1− xMnxTe and Hg1− x− yCdxMnyTe
alloys. J Vac Sci Technol A 4(4):2014–2018
7. Destefanis G, Chamonal JP (1993) Large improvement in HgCdTe photovoltaic detector per-
formance at LETI. J Electron Mater 22:1027
8. Donati S (2021) Photodetectors: devices, circuits and applications, 2nd edn. Wiley-IEEE
Press, New York
9. Djuric Z, Piotrowski J (1991) Dember IR photodetectors. Solid State Electron 34:265–269
10. Ganguly S, Tonni FF, Ahmed SZ, Ghuman P, Babu S, Dhar NK et al (2021) Dissipative quan-
tum transport study of a bi-layer graphene-CdTe-HgCdTe heterostructure for MWIR photode-
tector. In: 2021 IEEE research and applications of photonics in defense conference (RAPID).
IEEE, pp 1–2
11. Gaziyev FN, Nasibov IA, Ibragimov TI, Huseynov EK (2005) HgCdTe based PEM detector
for middle range of IR spectrum. Proc SPIE 5834:123–132
12. Gravrand O, Boulard F, Ferron A et al (2015) A new nBn IR detection concept using HgCdTe
material. J Electron Mater 44:3069–3075
2 Photoconductive and Photovoltaic IR Detectors 51

13. Grein CH, Young PM, Flatté ME, Ehrenreich H (1995) Long wavelength InAs/InGaSb infra-
red detectors: optimization of carrier lifetimes. J Appl Phys 78(12):7143–7152
14. Hooge FN (1969) 1/f noise is no surface effect. Phys Lett 29A:123–140
15. Jones RC (1952) Performance of detectors for visible and infrared radiation. In: Morton L (ed)
Advances in electronics, vol 5. Academic, New York, pp 27–30
16. Jones RC (1959) Phenomenological description of the response and detecting ability of radia-
tion detectors. Proc IRE 47:1495–1502
17. Jung HS, Grein CH, Becker CR (2003) Superlattices for very long wavelength infrared detec-
tors. In: Materials for infrared detectors III, vol 5209. International Society for Optics and
Photonics, pp 90–98
18. Kalinowski P, Mikołajczyk J, Piotrowski A, Piotrowski J (2019) Recent advances in manu-
facturing of miniaturized uncooled IR detection modules. Semicond Sci Technol 34:033002
19. Kasap S, Willoughby A, Capper P, Garland J (2011) Mercury cadmium telluride: growth, prop-
erties and applications. Wiley
20. Kinch MA (1981) Metal-insulator-semiconductor infrared detectors. Semicond Semimet
18:313–378
21. Kinch MA (2007) Fundamentals of infrared detector materials, vol 76. SPIE Press, Bellingham/
Washington, DC
22. Kopytko M, Jóźwikowski K, Martyniuk P, Rogalski A (2019) Photon recycling effect in small
pixel pin HgCdTe long wavelength infrared photodiodes. Infrared Phys Technol 97:38–42
23. Kopytko M, Martyniuk P, Madejczyk P, Jóźwikowski K, Rutkowski J (2018) High frequency
response of LWIR HgCdTe photodiodes operated under zero-bias mode. Opt Quant Electron
50(2):1–12
24. Lawson WD, Nielson S, Putley EH, Young AS (1959) Preparation and properties of HgTe and
mixed crystals of HgTe-CdTe. J Phys Chem Solids 9:325–329
25. Lei W, Antoszewski J, Faraone L (2015) Progress, challenges, and opportunities for HgCdTe
infrared materials and detectors. Appl Phys Rev 2(4):041303
26. Long D (1980) Photovoltaic and photoconductive infrared detectors. In: Keys RJ (ed) Optical
and infrared detectors, Part of the book series “topics in applied physics”, vol 19. Springer,
Berlin/New York, pp 101–147
27. Maille JHP, Salaville A (1979) Semiconductor Devices. U.S. Patent 4,132,999, January 2
28. Maimon S, Wicks GW (2006) nBn detector, an infrared detector with reduced dark current and
higher operating temperature. Appl Phys Lett 89:151109
29. Martyniuk P, Rogalski A (2013) HOT infrared photodetectors. Opto-Electron Rev
21(2):240–258
30. Meena VS, Mehata MS (2021) Investigation of grown ZnS film on HgCdTe substrate for pas-
sivation of infrared photodetector. Thin Solid Films 731:138751
31. Paschotta R (2008) Photoconductive detectors. In: Encyclopedia of laser physics and technol-
ogy, 1st edn. Wiley-VCH. https://www.rp-­photonics.com/photoconductive_detectors.html
32. Patten EA, Kalisher MH, Chapman GR, Fulton JM, Huang CY, Norton PR et al (1991)
HgZnTe for very long wavelength infrared applications. J Vac Sci Technol B 9(3):1746–1751
33. Piotrowski J, Rogalski A (2004) Uncooled long wavelength infrared photon detectors. Infrared
Phys Technol 46:115–131
34. Piotrowski J, Rogalski A (2007) High-operating temperature infrared photodetectors. SPIE
Press, Bellingham
35. Piotrowski J, Adamiec K, Maciak A, Nowak Z (1989). ZnHgTe as a material for ambient tem-
perature 10.6 μm photodetectors. Appl Phys Lett 54(2):143–144
36. Piotrowski J, Pawluczyk J, Piotrowski A, Gawron W, Romanis M, Kłos K (2010) Uncooled
MWIR and LWIR photodetectors in Poland. Opto-Electron Rev 18:318–327
37. Reine MB (2001) HgCdTe photodiodes for IR detection: a review. Proc SPIE 4288:266–277
38. Riley KJ, Lockwood AH (1980) HgCdTe hybrid focal-plane arrays. Proc SPIE 217:206
39. Rogalski A (1992) New ternaiy alloy systems for infrared detectors Proc SPIE 1845:52–60
52 R. Savkina and O. Smirnov

40. Rogalski A (2005) HgCdTe infrared detector material: history, status and outlook. Rep Prog
Phys 68:2267–2336. https://doi.org/10.1088/0034-­4885/68/10/R01
41. Rogalski A (2010) Infrared detectors, 2nd edn. CRC Press, Boca Raton, p 898
42. Rogalski A (2011) Recent progress in infrared detector technologies. Infrared Phys Technol
54(3):136–154
43. Rogalski A (2019) Infrared and terahertz detectors, 3nd edn. CRC Press, Boca Raton, p 1044
44. Rogalski A, Adamiec K, Rutkowski J (2000) Narrow-gap semiconductor photodiodes. Vol. 77.
SPIE Press: Chap. 8 HgZnTe and HgMnTe Photodiodes, pp. 337–360
45. Rogalski A, Martyniuk P, Kopytko M, Hu W (2021) Trends in performance limits of the HOT
infrared photodetectors. Appl Sci 11:501
46. Shin SH, Pasko JG, Lo DS, Tennant WE, Anderson JR, Gorsk M et al (1986) Hg1−x−yMnxCdyTe
alloys for 1.3−1.8 μm photodiode applications. MRS Online Proc Library 89:267–274
47. Tredwell TJ, Long D (1977) Detection of long wavelength infrared moderate temperatures.
Final report, NASA Lyndon B. Johnson Space Center Contract NAS9-14180, 5s (NASA
Accession No. N78-13876)
48. Triboulet R, Lasbley A, Toulouse B, Granger R (1986) Growth and characterization of bulk
HgZnTe crystals. J Cryst Growth 79(1–3):695–700
49. Van der Ziel A (1959) Fluctuation phenomena in semiconductors. Butterworths, London
50. Velicu S, Grein CH, Emelie PY, Itsuno A, Phillips JD, Wijewarnasuriya P (2010) Non-­
cryogenic operation of HgCdTe infrared detectors. Quant Sens Nanophotonic Devices VII
7608:760820
51. Wang X, He K, Chen X, Li Y, Lin C, Zhang Q et al (2020) Effect of annealing on the electro-
physical properties of CdTe/HgCdTe passivation interface by the capacitance–voltage charac-
teristics of the metal–insulator–semiconductor structures. AIP Adv 10(10):105102
52. Wehner JGA, Smith EPG, Venzor GM, Smith KD, Ramirez AM, Kolasa BP et al (2011)
HgCdTe photon trapping structure for broadband mid-wavelength infrared absorption.
J Electron Mater 40(8):1840–1846
53. You C, Deng W, Liu M, Zhou P, An B, Wang B et al (2021) Design and performance study of
hybrid graphene/HgCdTe mid-infrared photodetector. IEEE Sensors J 21(23):26708–26715
54. Zhang P, Ye ZH, Sun CH, Chen YY, Zhang TN, Chen X et al (2016) Passivation effect of
atomic layer deposition of Al2O3 film on HgCdTe infrared detectors. J Electron Mater
45(9):4716–4720
Chapter 3
II–VI Compound Semiconductor
Avalanche Photodiodes for the Infrared
Spectral Region: Opportunities
and Challenges

K. -W. A. Chee

3.1 Introduction

Avalanche photodiodes (APDs) are highly sensitive semiconductor devices that


harness the photoelectric effect to convert optical energy into electrical energy.
Inherently, the APDs deliver a high current gain, and hence high sensitivity, upon
the application of a reverse bias voltage, and they enjoy a distinctly higher signal-­
to-­noise ratio (SNR) compared to that of conventional PIN photodiodes, as well as
a short response time, and low dark current. Nevertheless, APDs invariably require
a higher operating voltage. The primordial phenomenon for the internal gain is
called impact ionization, which increases the photocurrent flow in response to an
incident optical power. Under a sufficiently strong electric field, the highly ener-
getic charge carriers generated by impact ionization may trigger cascades of ioniza-
tion events akin to avalanching. Figure 3.1 clarifies such a process of avalanche
multiplication via impact ionization. The electrons and holes will drift in opposite
directions in the electric field. Further, the avalanche gain of the APD will undesir-
ably amplify the dark current and the shot noise due to the electrical junction(s); in
fact, the stochastic nature of the APD gain is liable for the excess noise to the pho-
tocurrent output. Hence, compared to the classical PIN photodiode, the APDs tend
to generate a higher level of noise; the noise current is generally amplified by a
factor that is at least comparable to the avalanche gain, thereby affecting the SNR.
As demonstrated by McIntyre [2], the total spectral shot noise current, <ishot> can
be analytically described depending on the mean avalanche gain as follows:

K. undefined.-W. A. Chee (*)


School of Electronic and Electrical Engineering, Kyungpook National University,
Daegu, Republic of Korea
School of Electronics Engineering, College of IT Engineering, Kyungpook National
University, Daegu, Republic of Korea
e-mail: kwac2@cantab.net

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 53


G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_3
54 K. -W. A. Chee

Fig. 3.1 Symbolic process of three-particle generation via impact ionization and avalanching
through consecutive ionization events. (Reprinted with permission from Ref. [1]. Copyright 2020:
Elsevier B.V)

2
ishot = 2q  I dark , s +  M 2  I photo + I dark ,b  F  M    f (3.1)

 1  M  _ 1  _ 1 
through F  M   kM   2   1  k  or F  M   + 2 1 in the
 M  k  M   k 
pure electron or hole injection limit, respectively; the main noise source is assign-
able to the statistical Poissonian fluctuations of the impact ionization, which is a
random avalanche process that engenders an excess noise factor, F(M), where M is
the current gain and k is the ratio of the hole to electron ionization coefficients
(β/α). Iphoto, Idark,b and Idark,s are the photocurrent, bulk leakage current and surface
leakage current, respectively, and ∆f is the system bandwidth. Since impact ioniza-
tion occurs for hot carriers under the influence of an adequately strong electric field,
the coefficients α and β are governed by the scattering processes due to the electrons
and holes, and the electric field. Under high-field conditions, the mobile charge car-
riers gain kinetic energy at a faster rate. Hence, α and β increase with the electric
field, so that k tends to unity.
Another performance consideration for APDs is their bandwidth. Emmons [3]
showed that the bandwidth is determined by the transit time of the charge carriers
across the avalanche region to attain the high gain. The higher the required gain, the
lower the bandwidth. Hence, a trade-off relationship in conflicting design require-
ments arises although this limitation can be mitigated by a lower k-value. When the
k-value is lowered, both the excess noise and bandwidth performance can be
enhanced. Analogously, the ionization coefficients of the electrons and holes in the
multiplication region are major determinants of the excess noise and gain-­bandwidth
product. Infrared (IR) APDs fabricated in elemental group IV semiconductors such
as Si or Ge, and certain III–V compound semiconductors, such as InGaAs, nP,
3 II–VI Compound Semiconductor Avalanche Photodiodes for the Infrared Spectral… 55

InAlAs, AlGaAs, or AlGaAsP, exhibit high excess noise at high gain because of
significant initiation by both carrier types in the avalanche process proceeding from
the high scattering rates under a high reverse bias. Indeed, Si APDs generally harbor
lower levels of noise than several of the III–V counterparts [4]; for example, F(M)
is between 2 to 3, and 4 to 5, for Si and III–Vs, respectively. In terms of the optimal
SNR, since the optimum gain depends partially on F(M), the former characteristi-
cally ranges from 50 to 1000 for Si APDs but plummets to 10 through 40 for Ge and
InGaAs APDs. Early efforts in HgCdTe APD development, which featured hole-­
dominated avalanching, i.e., hole-injected APDs (or h-APDs), centred on applica-
tions in short-wave IR (SWIR) optical fiber communications [5–9]. Nevertheless,
the linear-mode APD array has its genesis in the elucidation of the physics of ava-
lanche multiplication in HgCdTe toward noiseless, extraordinarily high gain. Since
the seminal report by Kinch et al. in 2004 on HgCdTe electron-initiated APDs
(e-APDs) [10] – avalanching is dominated by one carrier type, i.e., electrons – cou-
pled with step change advances in the control over the heteroepitaxial growth of
multilayer structures, considerable interest has been roused in exploiting virtually
noise-free gain exceeding 1,000 at room-temperature operation not only in the
SWIR band, but in the mid-wave IR (MWIR) and long-wave IR (LWIR) bands.
Only matched by InAs or AlInAsSb, HgCdTe is the only compound semiconduc-
tor so far reported that is being exploited with attractive avalanche gain characteris-
tics that are overwhelmingly dominated by electrons, attainable at the appropriate
alloy composition [11]. Compared with other materials—including Si, Ge, InGaAsP/
InP, and GaSb/AlAsSb [12, 13]—HgCdTe can bestow low excess noise, short
response time, and high intrinsic avalanche gain in SWIR, MWIR, and LWIR appli-
cations [8, 14–17]. Therefore, HgCdTe APDs are suited for IR wave front sensing,
advanced laser rangefinders, light detection and ranging (LiDAR), astronomy, gated
viewing in SWIR, 3D-imaging radars, long-haul optical fiber communications, pho-
ton counting, etc. [18–23]. Although molecular beam epitaxy (MBE)-grown lattice-­
matched InAlAs/InGaAs heterostructures have led to an impressive gain maximum
of 2,000 (at 235 K), their optimal noise performance constrains the gain to sub-20
with an F(M) value of 2.3 at room temperature [24]. Instead, the liquid-phase epi-
taxy (LPE)-grown HgCdTe e-APDs exhibited a virtually noise-free gain of up to
150 (at 196 K), thus fulfilling the noise and gain performance specifications for
thresholded linear-mode photon counting that would otherwise necessitate multi-
stage upscaling of the InAlAs/InGaAs APD to satisfy the equivalent performance
standards [25]. Given that the MBE-grown HgCdTe IR APDs have already estab-
lished a market presence underscored by their assuring reputation for ultralow dark
current levels, there is enormous interest in increasing the sensitivity and readout
rates of focal plane arrays (FPAs) in third-generation (3G) technology whereby the
high avalanche gain means reduction in the exposure times of cameras. The present
chapter surveys the principles, design, fabrication, engineering, and technologies,
as well as the opportunities and challenges related to II–VI compound semiconduc-
tor APDs for the IR spectral region.
56 K. -W. A. Chee

3.2 Background and Roadmap

Several materials and architectures have been explored for adaptation in photode-
tector technologies [26]. Figure 3.2 provides the technology roadmap in the exploi-
tation of IR materials over the previous decades [27]. Modern IR detector technology
progressed rapidly post-World War 2. The first practical IR detectors used PbS,
providing sensitivity up to ~3μm. Meanwhile, high-performance PbSe and PbTe
detectors were also developed. Particularly, since the Bell Labs demonstration of
the first operating transistor in 1947, InSb, HgCdTe, and Si modern IR detectors
were designed. Moreover, PbSe, PbTe, and InSb detectors extended the spectral
sensitivity to 3–5 μm, of the MWIR regime. The direct bandgap ternary compound
semiconductors, HgCdTe and PbSnTe led to the first advanced progress of photode-
tector technologies. Nevertheless, the greater dielectric constant of PbSnTe limited
its high-frequency performance, and the high thermal expansion coefficient con-
strained its integration in CMOS processes.
The first-generation systems included Si and HgCdTe technologies, which were
based on individual linear elements. These early designs were without multiplexing
functions embedded in the focal plane. The combination of MBE and advanced
photolithographic processes became the cornerstone of the semiconductor industry
that enabled rapid FPA development. Charge coupled device (CCD) imagers were
launched in the 1970s by AT&T Bell Labs., which subsequently ushered in a new
era of second-generation FPAs integrated with focal plane readout integrated

Fig. 3.2 Infrared detector and systems roadmap. (Reprinted with permission from Ref. [28].
Copyright 2009: AIP Publishing)
3 II–VI Compound Semiconductor Avalanche Photodiodes for the Infrared Spectral… 57

circuits (ROICs). The ROICs multiplex the signals, enact signal processing, and
read out the array. The second-generation systems adopted monolithic and hybrid
FPA systems incorporating multiplexing ROICs; signal multiplexing could be per-
formed from a very large photodetector array. At present, the 3G FPAs, which com-
prise several orders of magnitude more pixels and superior on-chip properties
compared to that of the second generation, are under active development.
The PIN photodiodes have been very successfully exploited in high-energy par-
ticle physics experiments [29–32]. Nevertheless, despite the use of a state-of-the-art
amplifier necessary because of the lack of an internal gain, the noise is significant,
composed of several hundred electron fluctuations. The APDs retain an exquisite
internal gain that enables a high SNR, but F(M) limits the optimal current gain.
Since the beginning of the present millennium, the Geiger-mode APD was devel-
oped, allowing detection of individual photons. This high-sensitivity detector tech-
nology has evolved since the first Si single-photon detectors, as depicted in Fig. 3.3,
namely, the planar APD (Fig. 3.3a) developed by RCA company [33], and the
reach-through APD (Fig. 3.3b) of Shockley Semiconductor Laboratory [34], in
the 1960s.

3.3 Alloy Composition and Technology

The addition of elemental Cd opens the bandgap of the Hg-based binary semimetal
HgTe to operate in the IR wavelength range. Hence, the II–VI ternary compound,
HgCdTe, is a highly versatile material system that has become an industry standard
in the fabrication of IR detectors for applications encompassing the spectral region
from 2.2 to 9.7 μm. By carefully adjusting the Hg:Cd fraction, the spectral range
connoting wavelengths from 1 to 30μm [27, 36] can be controlled via the tunable
bandgap between 0 and 1.5 eV of the Hg1-xCdxTe alloy to match the atmospheric
windows of the solar and IR spectra for passive thermal imaging and sensing. For

Fig. 3.3 The earliest single-photon APDs. (a) Planar APD from RCA company and (b) Reach-­
through APD from Shockley Semiconductor Laboratory. (Reproduced with permission from Ref.
[35]. Copyright 2006: Elsevier B.V)
58 K. -W. A. Chee

variable alloy compositions from x = 0.1 to 0.7, the hole to electron impact ioniza-
tion ratio is highly favorable for a low F(M) value. Corresponding to the energy
band structure, the considerable asymmetry between the effective masses of con-
duction band electrons and valence band heavy holes leads to a highly unbalanced
ionization coefficient for the holes and electrons. The threshold impact ionization
energy for the electron is marginally above the bandgap, but that for the hole sur-
passes twice this value. Further, the electron mobility is two orders of magnitude
greater than the hole mobility in HgCdTe [37–41], implying wildly variable scatter-
ing rates of holes and electrons. A high avalanche gain can therefore be achieved
alongside a low F(M) value as the multiplication process is initiated by a single
charge carrier type; electron injection for lower x-values (x < 0.6), i.e., k < 1, or hole
injection for 0.6 ≤ x ≤ 0.7, i.e., k ≥ 1 [39, 42, 49], desirable for the SWIR to LWIR
[11, 43–48] or only SWIR [5–7, 9, 49–51] spectral bands, respectively. The domi-
nant carrier multiplication process in the HgCdTe APD can be limited to one type of
charge carrier, such as the focus of the majority of the work involving low Cd con-
tent, i.e., x < 0.6 for e-APDs [10]. Moreover, a nearly constant exponential gain
increase in the range as high as >100 has been characterized, in tandem with an
approximately constant SNR [11, 52–54]. Not only the gain values are gener-
ally well above 1,000 but the bandwidth is not sensitive to the gain [39]. Even at a
low reverse bias, HgCdTe e-APDs exhibit high gain and low excess noise besides a
THz-scale bandwidth [55]. An F(M) close to unity (1.1–1.4) is achievable in this
material [56]. Conversely, h-APDs tolerate a lower gain (< 100) and a higher F(M)
at the equivalent reverse bias condition.
The high internal gain, high quantum efficiency and sensitivity, low excess noise,
and short response time of HgCdTe APDs cater to applications in the SWIR span-
ning between 1 and 3 μm, the MWIR from 3.2 to 5 μm, and the LWIR from 7.5 to
14 μm [8, 14–17, 46, 52, 57]. The SWIR is harnessed in astronomy and in active
imaging when illuminated by IR sources (e.g., 1.55-μm Nd:YAG lasers); and the
MWIR and LWIR regimes are exploited in passive thermal IR regimes- the latter
usually involves imaging in conditions clouded with fog and smoke. The largest
market opportunities for environmental and planetary science applications are
derived from ground- and space-based instrumentation that emphasize on single-­
photon sensitivity. The ability to achieve high sensitivity depends crucially on the
growth processes adopted. At present, several state-of-the-art thermal imaging sen-
sors are based on InSb having 5.5-μm sensitivity. InSb detectors currently dominate
cooled thermal imaging systems since they were first developed based on this mate-
rial. Importantly, HgCdTe has fundamental properties and favorable band structure
features conducive to producing excellent detectors. High linear gain at low bias,
extremely low noise, high gain-bandwidth product, shorter response time, and gain
independent bandwidth achievable in HgCdTe APDs embody a catalog of attributes
suited for integration in next generation (3G) FPAs. Moreover, HgCdTe detectors
can operate at much higher temperatures than InSb detectors. These features have
an edge over that of the III–V counterparts, and HgCdTe-based cameras can be
made more portable with high power efficiencies.
3 II–VI Compound Semiconductor Avalanche Photodiodes for the Infrared Spectral… 59

3.4 General Architecture and Operation

Figure 3.4 reveals the basic architecture of the APD, composed of a moderately
doped p-absorber and a lightly doped n-avalanche region. As the reverse bias
increases, the junction depletion extends further into the lightly doped n-region with
an electric field rising to a critical level that instigates avalanche multiplication. In

Fig. 3.4 (a) Cross-section of the basic APD and the corresponding electric field across the device
structure under a reverse bias. The maximum electric field strength occurs in the multiplication
region, which entails the lightly doped n-type region. (Reprinted with permission from Ref. [19].
Copyright 2011: Elsevier Ltd.). (b) Energy band diagram at a low (left) or high (right) reverse volt-
age. The Shockley-Read-Hall (SRH) and trap-assisted tunneling (TAT) mechanisms at a low
reverse voltage, and the band-to-band tunneling (BBT) and avalanche mechanisms at a high
reverse voltage, are depicted for the dark current generation. (Reprinted from Ref. [58]. Published
2021 by Nature Portfolio as open access under the terms of the Creative Commons Attribution 4.0
International License)
60 K. -W. A. Chee

the abutting p-absorber, the electric field separates the photogenerated electron-hole
pairs, with one charge carrier type drifting towards the avalanche region. The pho-
tocurrent output is amplified via impact ionization by the injected carriers in the
avalanche layer. Typically, the APD is designed to maximize the minority carrier
injection from the absorber layer into the multiplication region to optimize for low
excess noise. The avalanche region is a crucial determinant of the avalanche gain,
excess noise, and gain-bandwidth product.
Practical APD designs need to be optimized for high gain, high gain-bandwidth
product, low excess noise, short response time, and high temperature stability. The
two examples of architectures commonly being fabricated are the planar n-on-p and
the annular n-on-p high density-vertical integrated photodiode (HDVIP). The basic
cross-sectional structures of a back-illuminated planar n-on-p e-APD and the n-on-p
HDVIP are exemplified in Fig. 3.5. In the planar n-on-p technology, the

Fig. 3.5 Cross-section of the (a) back-illuminated planar n-on-p or (b) front-illuminated n-on-p
HDVIP architecture. (Reprinted with permission from Ref. [19]. Copyright 2011: Elsevier Ltd.)
3 II–VI Compound Semiconductor Avalanche Photodiodes for the Infrared Spectral… 61

electron-­hole pairs generated in the p-absorber upon optical irradiation will


diffuse/drift into the depletion region (or avalanche region) within the n−-region. As
the reverse bias increases, the depletion region extends toward the device surface.
Meanwhile, the avalanche gain, and thus, sensitivity, increases with the electric
field. Hence, the avalanche region is governed by the n-layer thickness. Conversely,
in the HDVIP technology, the depletion region can extend beyond the n-region, into
the p-absorber, upon increasing the reverse bias up to the critical field for the onset
of avalanche multiplication. The electric field is uniform across the depletion region,
leading to a homogeneous avalanche gain and dark current distribution, thereby
averting an adverse impact on sensitivity [55].
The APD may be operated in the linear mode or Geiger mode. In the linear mode,
the bias is applied near the breakdown voltage, so that the rate at which the charge
carriers are collected at the device terminals outstrips the rate at which the electron-­
hole pairs are generated, thus resulting in an avalanche photocurrent decay. With a
finite gain, the individual photons can be distinguished. Therefore, linear-mode
APDs can be leveraged in photon counting applications. In the Geiger mode, an
extraordinarily high non-linear gain is used to sense single-photon signals. The bias
is applied above the breakdown voltage (pragmatically known as overbias), so that
the multiplication rate dominates the collection rate. In this on-state, the photocur-
rent initially increases exponentially under the photo-induced breakdown, and the
electrical junction becomes conductive following an optical input, so that even a
single-photon input can result in strong electrical current pulses. In theory, the gain
is infinite if the bias is held above the breakdown voltage. Within a certain finite
period of time, the electrons and holes accumulate, respectively, at the n- and p-type
edges of the depletion region, thus generating a built-in electric field opposing the
externally applied field. The APD remains in the on-state until the applied bias is
reduced to quench the device, turning off the electrical conduction and thwarting
damage to the junction. As the applied bias is held low to quench the device, the
mid-gap states release the trapped carriers. After a certain amount of elapsed time,
the overbias is restored but there is a low probability of after-pulsing (breakdown
due to the charge carriers released from the deep levels). Contrary to that in the
linear mode, the electrical response is governed by the bias circuit instead of the
optical signal power; this mode of operation is unsuitable for photon counting appli-
cations. Especially pertinent to HgCdTe APDs that display highly temperature-­
sensitive performance, cryogenic cooling employing liquid nitrogen and rescaled
junction areas are usually indispensable to minimize the dark current in the case of
the narrow bandgap property [59], and especially if there are less stringent demands
on the purity of the semiconductor alloy. Geiger APDs are deployed to sense faint
optical sources in applications where a short duty cycle and limited portability are
tolerable.
62 K. -W. A. Chee

3.5 Fabrication and Processing

It is well-known that HgCdTe APDs may encounter high fabrication costs and bot-
tlenecks in mass production that stem from the weak Hg-Te bonding. The challenge
in growing HgCdTe layers is primarily attributable to the high vapor pressure of Hg;
the homogenous composition may be non-trivial to achieve in bulk crystals and
epilayers. In industry, there is a systemic transition over the recent three to four
decades away from the use of bulk grown HgCdTe materials toward epitaxially
grown structures based on the metal-organic vapor phase epitaxy (MOVPE),
LPE, and MBE techniques, to allow highly controllable lattice-matched growth of
compound semiconductors [47, 60–66]. Particularly, Leonardo MW Ltd. (formerly
SELEX Galileo Infrared Ltd.) [64, 67] and Vigo Systems SA adopt MOVPE-growth
and CEA-Leti configures an in-house LPE or MBE system [68, 69] for the fabrica-
tion of their 3G IR FPAs. The epitaxial growth techniques shun the high-pressure
processes imperative for bulk growth. Furthermore, in view of the high degree of
control afforded by epitaxial growth, superior APD performance can be achieved
via the enhanced crystal quality, and the uniformity of the alloy composition, dop-
ing, and layer thicknesses. The absorber, junction, and avalanche regions can be
independently optimized in separate optical absorption, charge transport, and mul-
tiplication (SACM) structures. Advanced APD heterostructures can be easily real-
ized through device localization and engineering (e.g., bandgap, quantum well, or
impact ionization engineering, etc.), enabling high-sensitivity e-APDs, multispec-
tral APD functions with high spectral resolution capabilities, and large format arrays
consisting of highly sensitive pixels offering high spatial resolution. The drive
toward lower manufacturing costs relies on the use of large-area growth techniques
and lower-cost substrate materials.
MOVPE is a large-area growth technique that had earlier enjoyed successes in
their application to a range of III–V compound semiconductor technologies [70].
For example, MOVPE-GaN light emitting diodes became commercially available
by the mid-1990s. II–VI semiconductors having wide or narrow bandgaps are also
grown by MOVPE [71]. The growth temperature for MOVPE is generally higher
than that for MBE, but being lower than that for LPE, cannot attain the crystalline
quality of the latter. Nevertheless, bandgap engineering throughout the device can
be performed with much greater ease. Foundational for the MOVPE growth tech-
nique is the interdiffused multilayer process: when depositing alternating thin layers
of CdTe and HgTe grown under their individually optimized conditions (not exceed-
ing a combined thickness of 200–250 nm), the binary compounds interdiffuse to
form uniform HgCdTe crystals within a relatively short amount of time owing to
their high diffusion coefficients. A short annealing procedure at the growth tempera-
ture at the end of the diffusion process fully homogenizes the HgCdTe layer, yield-
ing a composition contingent on the relative thickness of the individual CdTe and
HgTe layers. Central to the flexibility and versatility of the interdiffused multilayer
process is the notion that the flux of each binary compound can be individually
3 II–VI Compound Semiconductor Avalanche Photodiodes for the Infrared Spectral… 63

controlled and optimized for dopant incorporation, and the total thickness of the
crystalline HgCdTe layer depends on the growth rate and growth duration.
Through the interdiffused multilayer process, the heterolayer doping and thick-
ness can be completely optimized. The intrinsic defects commonly considered in
the grown HgCdTe layers include the Hg vacancies, Hg interstitials, Te antisites,
and several other defect complexes. Under normal MOVPE growth conditions, the
amount of metal vacancies depends on the process parameters. The substrate orien-
tation is a dominant factor in the nucleation growth, dopant incorporation efficiency,
and growth rates, because of the germane variety of step energies involved [72]. On
(100) substrates, the growth rates are slowest, whereas (111)B substrates grant
faster growth rates and suppress the macrodefect density. In the former case, the
substrates are slightly misoriented from the (100) plane to mitigate strain-induced
defect formation. In the latter, the twin boundaries present in the grown layers act
like donors [73], and As [74] (I [63]) dopant incorporation is substantially stymied
(enhanced); the most popularly adopted dopants in MOVPE-HgCdTe layers are As
(p-type) and I (n-type). The (211)B [75] and (552)B [76] substrates have also been
investigated as alternatives. For CdZnTe substrates, the (211)B orientation is favored
for device-quality fabrication, by courtesy of the ease in suppressing the formation
of macrodefects; the (211)B orientation also serves as the preferred growth direc-
tion for MBE. CdZnTe substrates typify the traditional option for the lattice-matched
growth of HgCdTe. Even so, there are several drawbacks. The per unit area substrate
cost is up to two orders of magnitude greater compared to the alternative GaAs. In
the preferred <111> orientation for bulk HgCdTe growth, the uniform alloy compo-
sition control crucial in a LWIR architecture cannot be attained; Zn segregation
along the growth axis leads to uneven Zn concentrations within (100) substrates.
Given the higher MOVPE growth temperature (compared to that in MBE or LPE),
outgassing is enhanced of residual Cu impurities from Te precipitates in the sub-
strates that accumulate in the grown HgCdTe layer leading to fluctuations in the
background doping (Cu acts as an active dopant); hence there is a lack of reproduc-
ibility control of the low n-type doping common in several device architectures.
Attractive substrates for MOVPE growth are Si and GaAs, which require the
introduction of buffer layers to alleviate the lattice mismatch-induced defect genera-
tion, and autodoping effects. On account of the low-cost opportunity, large-area
wafer availability, and thermal matching properties to the ROICs, Si represents an
outstanding substrate candidate. However, challenges include the requisite to main-
tain the preferred (100) orientation as well as the necessity of an in-situ process to
rid the thermal oxide. Solutions have included using Ge [77], GaAs [78], or ZnTe/
CdTe [79] interfacial layers, or utilizing a high-temperature oxide desorption
annealing process [80]. Also readily available in large wafer sizes (up to 150 mm
diameter) and nominally more expensive than Si, GaAs has been considered as the
most important substrate for low-density, device-quality MOVPE-HgCdTe [75, 81].
For example, highly affordable high-performance LWIR FPAs have been fabricated
using GaAs substrates for MOVPE-grown HgCdTe materials bump bonded to
ROICs [82]. The lattice mismatch encountered on GaAs substrates (14.6%) is
smaller than that on Si substrates (19.3%), hence the use of GaAs substrates is a
64 K. -W. A. Chee

priori expected to lead to higher FPA performance by virtue of the correlated supe-
rior crystal quality obtainable for the grown HgCdTe material. Typically, the CdTe
seed/buffer layer thickness is optimized for the crystalline quality of the grown
HgCdTe as well as to minimize the rate of Ga out-diffusion from the substrate that
may induce autodoping effects [75, 81]. In-situ annealing techniques have also been
incorporated into the growth reactor procedure to remove Hg vacancies in order that
the intentionally introduced dopants during growth are activated and dominate the
electrical properties of the fabricated devices [83]. Alternatively, an ultrathin layer
of Cd is introduced intervening the interdiffused multilayer process to generate the
excess metal to counter the Hg vacancies [84].
In IR FPAs, the size and density of hillocks (macrodefects) will determine the
cosmetic (defective pixel clusters) quality. To enhance the surface morphology and
crystal quality of the grown HgCdTe layers, Giess et al. [85] successfully reduced
the pyramidal hillock density to <10 cm−2 at the MOVPE-HgCdTe/GaAs (100)
interface by a priori treatment of the substrate using an alkali metal containing solu-
tion, such as aqueous KOH solution [86]. The adsorbed alkali metal atoms (e.g., Na
or K) on the substrate facilitates the homogeneous nucleation of the CdTe layer on
GaAs [75, 86]. Suh et al. [87] reduced the surface hillock density to 90 cm−2 by
simply occluding the Hg bath with a lid in front of the heating susceptor (see
Fig. 3.6) to suppress the pre-reaction kinetics of the Cd precursor with Hg and
facilitate homogeneous nucleation of HgCdTe on the GaAs (100) substrate. The
ever-challenging issue of reducing the hillock density to <103 cm−2 [73, 88] by
means of the foregoing approaches ensure best-in-class crystallinity of the grown
HgCdTe material for 3G FPA applications. Nevertheless, the thermal mismatch
between the GaAs substrate and Si ROICs remains a technological quandary to be
addressed for large-area FPA manufacture. On the contrary, Si substrates are com-
patible with CMOS processes and so they are desirable for integration with
ROICs. Temperature cycling reliability is also expected to be superior as a conse-
quence of the excellent mechanical integrity and thermal capability.
For MBE-HgCdTe, alternative substrates are GaAs [89], Ge [90], and Si [91].

Fig. 3.6 Schematic illustrating the MOVPE reactor within which a Hg bath is occluded with a lid
before the susceptor. (Reprinted with permission from Ref. [87]. Copyright 2002: Elsevier B.V)
3 II–VI Compound Semiconductor Avalanche Photodiodes for the Infrared Spectral… 65

He et al. [92] compared the MBE-HgCdTe on GaAs (211) B and Si (211) sub-
strates. A buffer layer of CdTe was grown on the substrate prior to the HgCdTe
nucleation to alleviate the lattice mismatch. Figure 3.7a uncovers the extent of lat-
tice tilt for epilayers grown on the (211) substrate, thereby indicating that the tilt
stems intrinsically from the misfit strain and is linearly related to the relative differ-
ence between the in-plane lattice constants. To reduce the density of threading dis-
locations, the CdTe buffer layer was grown on the substrates misoriented toward the
[111] direction by 1.0°. Figure 3.7b shows that the X-ray double-crystal rocking
curve full-width at half-maximum (FWHM) value reduces with the tilt angle
between the plane of the CdTe buffer epilayer and Si (211) substrate from 4.2° to
2.75°, and then becomes largely constant for smaller tilt angles; the FWHM value
remains largely unchanged for the CdTe buffer epilayer grown on the GaAs (211)
substrate. These results imply that the misfit strain can be mitigated using misori-
ented Si substrates corresponding to the angle of lattice tilt. Figure 3.8 compares the
crystal quality of the CdTe buffer epilayer deposited on misoriented Si (211) with
that on GaAs (211). Indeed, the FWHM values of the X-ray rocking curves or etch
pit density (EPD) due to CdTe on the misoriented Si substrate became comparable
with that on the GaAs substrate. Hence, the crystal quality in these two disparate
cases is essentially identical. As the CdTe buffer layer thickness increases, the EPD,
and thus, the threading dislocation density, decreases, owing to the strain mitigation
effects.

a 6 b
(522)
lts ) 240 tilting direction [211]
5 e ti (311
fac of [211]
sur ard
4 (733) tow 220 [111] [011]

3 (944) 200 epilayer


lts )
e ti (111 a substrate
2 u r fac
s a rd 180
tow
FWHM (arcsec)

1 FWHM of CdTe/Si
160
FWHM of CdTe/GaAs
0 (211)
140
–1 CdTe on Si
ZnTe on Si 120
–2 CdTe on GaAs
(955) ZnTe on GaAs 100
–3
CdTe on Ge
(744) CdTe on Cd0.95Zn0.04Te
–4 80
–5 (533)
60
–6
0 2 4 6 8 10 12 14 16 18 20 2.0 2.5 3.0 3.5 4.0
Lattice mismatch (%)

Fig. 3.7 (a) Lattice mismatch dependence of the tilt angle between the CdTe/ZnTe (211) and
(211) substrate. (b) FWHM of X-ray rocking curves of CdTe on Si/GaAs as a function of tilt angle
between the CdTe (211) and Si (211) substrate. (Reprinted with permission from Ref. [92].
Copyright 2007: Elsevier B.V)
66 K. -W. A. Chee

250 108
EPD of CdTe/Si
EPD of CdTe/GaAs
107
200

106
150
105
100
104

50
103
FWHM of CdTe/(211)misorientated Si
FWHM of CdTe/(211) standard Si
FWHM of CdTe/(211) standard GaAs
0 102
0 5 10 15 20

Fig. 3.8 CdTe buffer epilayer thickness dependence of etch pit density (EPD) and the FWHM of
X-ray rocking curves for the various substrates. (Reprinted with permission from Ref. [92].
Copyright 2007: Elsevier B.V)

Large-area FPAs need to have excellent lateral homogeneity of the deposited


active layer materials in terms of alloy composition and crystal quality. Control of
the lateral distribution of the substrate growth temperature and flux dispensation
from the sources with rotation mechanisms, coupled with in-line high-precision
mapping and monitoring techniques (Fourier-transform IR mapping is the most
commonly used), allow MBE growth of uniform HgCdTe layers on large-area sub-
strates [93, 94]. The layer thickness can be characterized by analyzing the Fabry-­
Perot interference fringes. Moreover, the detector performance is directly related to
the surface defect states (or hillocks). The defect propagation in the CdTe buffer
layer may affect the surface quality of the deposited HgCdTe. Nevertheless, with
MBE-grown 3” HgCdTe on the CdTe buffer layer, the surface defect (≥2 μm) den-
sity was measured to be less than 300 or 500 cm−2 when involving the GaAs or Si
substrate, respectively [92]. Provided that the growth parameters are optimized for
homogeneity of crystal quality and chemical composition, the fabricated LWIR and
MWIR FPA based on 3″ epilayers displayed performance parameters at 80 K that
are on par with that fabricated using CdZnTe substrates. This is despite the smaller
lattice mismatch with GaAs of ZnTe compared with CdTe (7.8% versus 14.6%).
Moreover, the identical X-ray rocking curve FWHM and etch pit density measure-
ments from the composite substrates of CdTe/GaAs and CdTe/Si indicate their great
potential as highly affordable substitutes for CdZnTe substrates.
3 II–VI Compound Semiconductor Avalanche Photodiodes for the Infrared Spectral… 67

3.6 Device Concept Design and Engineering

There are several strategies to combat the electronic noise that blights the perfor-
mance of optical transceivers incorporating IR APD technology. Low excess noise
and high quantum efficiency are integral conditions for background-limited perfor-
mance determined by photon statistics [95]. The multiplication layer can be fabri-
cated in wide bandgap material to forestall, e.g., trap-assisted tunneling, to reduce
the overall leakage current. Indeed, the surface leakage currents and 1/f noise con-
duce to the dark current [96], which can in turn be detrimental to the SNR perfor-
mance of the APD. Concurrently, the absorption layer thickness can be reduced to
control the thermally induced leakage current. High-carrier concentration layers can
envelope the absorption layer to suppress the 1/f noise. Significantly, the contact
layers can be doped heavily to reduce the sheet resistance undergirding 1/f noise.
Graded bandgap structures can be fabricated in the absorption layer (bandgap engi-
neering) to enhance the field-assisted transport of charge carriers into the multipli-
cation region to elevate the quantum efficiency.
The temperature stability of the APD performance is an important design and
technology consideration for high-reliability systems. New-generation FPAs need
to operate at higher temperatures so as to benefit from (1) reduced cooler power
dissipated from the Stirling cooling engine and (2) an increased operating span as
well as (3) an improved portability design. Pillans et al. [97] demonstrated using
high-operating-temperature (HOT) MWIR HgCdTe APDs, high-quality imaging at
temperatures as high as 210 K. Compared to the other material systems utilized in
the fabrication of HOT LWIR photodetectors, only the use of the ternary HgCdTe
alloy is appropriate, affording a low doping concentration (1013 cm−3) and a high
Shockley-Real-Hall carrier lifetime (>1 ms) [98]. Nevertheless, LWIR HgCdTe
APDs are known to exhibit a temperature-dependent dark current that can degrade
the SNR performance. Consequently, the device concept of unipolar barrier layer
engineering was introduced into the HgCdTe detector [99–103]. The dark current
was shown to be effectively curtailed by the barrier layer engineering, thus forging
new potential in HOT HgCdTe devices [104–110]. Charge carrier diffusion in the
absorption region can be inhibited and the Auger noise contribution can be sup-
pressed so that the overall dark current performance is enhanced at high temperature
[111–113]. To accommodate suitable amplification properties at high reverse bias,
He et al. [114] developed a pBp structure, consisting of a p-type contact layer and a
graded barrier layer between the absorber and the substrate (see Fig. 3.9), which
significantly lowered the dark current without penalty to the avalanche gain at high-­
temperature operation.
Figure 3.10 portrays the bias voltage and temperature dependence of the dark
current characteristics. The dark current increases monotonically with temperature
at low bias voltages. Nevertheless, a high bias voltage is required to raise the intrin-
sic gain. In the pBp structure, the dark current exhibits a negative temperature coef-
ficient at high bias voltages, therefore favoring high-temperature operation without
compromising the high avalanche gain.
68 K. -W. A. Chee

N+ contact layer N+ contact layer

Multiplication layer Multiplication layer

Absorption layer
Absorption layer Graded-Barrier
layer

P-type contact layer

Fig. 3.9 Comparison of the (a) conventional APD and (b) pBp APD structure. (Reproduced from
Ref. [114]. Published 2020 by the Optica Publishing Group as open access)

Fig. 3.10 Variation of dark current as a function of bias voltage and temperature as a parameter in
the (a) conventional HgCdTe APD and (b) HgCdTe pBp APD. The dark current versus temperature
characteristics at – 0.5 and −7 V are shown in the (c) conventional HgCdTe APD and (d) HgCdTe
pBp APD. (Adapted from Ref. [114]. Published 2020 by the Optica Publishing Group as
open access)

Figure 3.11 compares the carrier concentration distributions, and electric field/
potential profiles in the devices at a high operating bias of – 7 V. Within the multi-
plication layer of the pBp structure, the high electric field separates the photogene-
rated electron-hole pairs and promotes(suppresses) electron(hole)-initiated impact
ionization. Electron depletion in the absorption layer significantly enhances the
3 II–VI Compound Semiconductor Avalanche Photodiodes for the Infrared Spectral… 69

Fig. 3.11 Carrier concentration distributions, and electric field and electric potential profiles in the
(a) conventional APD and (b) pBp APD structures. (Adapted from Ref. [114]. Published 2020 by
the Optica Publishing Group as open access)

dark current performance, and the high operating bias has no appreciable effect on
the unipolar blocking capability of the barrier layer. The barrier layer hinders the
diffusion of electrons into the absorption layer yet facilitates charge injection into
the multiplication layer, and therefore, the dark current is suppressed under a high
reverse voltage.
Using additionally complex multi-heterojunction architectures, dual-waveband
detectors (DWBs) have been developed that allow concurrent readout of optical flux
from discordant IR spectral bands, e.g., MWIR and LWIR [115–117]. Both wave-
bands are heterogeneously incorporated into a single detector, thereby providing the
ability to toggle between the two as required. The cost advantage is significant com-
pared to that of two separate image sensors. Fused imaging can be performed to
delineate the features of importance via spectral variances. A dual-band IR metal-­
organic chemical vapor deposition (MOCVD)-structure studied by Kopytko et al.
[115] is shown in Fig. 3.12 depicting that of the N+-n-P+-p-N+ (capitalized letters
refer to wide bandgap and the lower-case letters denote the absorption regions) struc-
ture, whereby the wide bandgap p+-material As-doped to 5 × 1017 cm−3 is sand-
wiched between lightly doped narrow bandgap absorber materials. The abutting
n- and p-layers having the appropriately matched bandgaps are the MWIR (3 to
4.2 μm) and LWIR (4.2 to 5.2 μm) absorption regions. The absorption layers are
stacked vertically in a fashion permitting transmission of the IR flux through the
MWIR layer (MW1) before the LWIR layer (MW2). The MW1 active area has a Cd
molar ratio of 0.32 and is non-intentionally doped, with a background impurity
concentration of 1015 cm−3 featuring n-type conductivity. The MW2 active layer has
a Cd molar ratio of 0.28 and is As-doped to 5 × 1015 cm−3. The electrical contacts are
formed to the wide bandgap layers I-doped to 2 × 1017 cm−3. The selected band is
determined by the applied bias polarity.
70 K. -W. A. Chee

Fig. 3.12 Cross-section of a Hg1-xCdxTe dual-band IR detector. (Reprinted from Ref. [115].
Published 2019 by Springer as open access under the terms of the Creative Commons Attribution
4.0 International (CC BY) License)

The response time of e-APDs has also been carefully studied and engineered. For
excellent response time characteristics, the following conditions must be met: (1)
reduced diffusion time of the minority carrier to the multiplication region, (2) mini-
mal hole and electron transit time across the avalanche layer, and (3) small RC time
constant arising from the junction capacitance and series resistance of the absorber.
Perrais et al. [54, 118] reported a record-level gain of 2,800 and gain-bandwidth
product of 1.1 THz from the planar MWIR HgCdTe p-APD [54, 118], and an ava-
lanche gain of 3,500 and gain-bandwidth product of 2.1 THz from the front-­
illuminated planar MWIR HgCdTe e-APD [119]. Besides eliminating preferential
avalanche breakdown at the junction edge and minimizing the number of bulk
microplasmas [3], Li et al. [120] demonstrated that with an effective guard ring
design, the dark current, F(M), and mean square noise of the MWIR HgCdTe e-APD
can be reduced while maintaining a high bandwidth and a high-temperature perfor-
mance stability.
3 II–VI Compound Semiconductor Avalanche Photodiodes for the Infrared Spectral… 71

3.7 Concluding Remarks and Outlook

We have surveyed the highly versatile HgCdTe material system employed in


industry-­standard IR APDs suited for state-of-the-art photon counting or 3G FPAs.
The ability to fine-tune the alloy composition of the compound semiconductor
through advanced heteroepitaxial techniques possessing a very high degree of con-
trol of the thin epilayers and defect propagation has led to bandgap engineering and
impact ionization localization opportunities that enable a high avalanche gain, a
high temperature stability performance, low dark current, short response time, and
low excess noise. The basic architecture, and fabrication and processing technolo-
gies, have been discussed. The physics and operating principles of the HgCdTe
APD have been addressed, revealing innovative design approaches to attain record-­
level photoelectric performance in SWIR, MWIR, and LWIR applications. Notably,
since Cd destabilises the weak Hg-Te bond, the use of Zn and Mn as candidate
substitutes has been explored in MBE-HgZnTe and MBE-HgMnTe semiconductor
alloys, respectively, with the former exhibiting Hall mobilities comparable with that
of the best of MBE-HgCdTe crystals on GaAs (100) substrates [121]. While noting
certain challenges in fabrication, performance and cost, insights provided in this
work are anticipated to invigorate interest in the adoption and engineering of II–VI
compound semiconductor device concepts for advanced photodetection
capabilities.

Acknowledgements This research was supported by the Kyungpook National University


Research Fund, 2021, the Kyungpook National University Excellent New Research Fund, the
Dongil Culture and Scholarship Foundation for academic research in 2022, and the Brain Korea
(BK) 21 FOUR project funded by the Ministry of Korea (4199990113966) and the National
Research Foundation of Korea.

References

1. Huntington AS (2020) 1 – Types of avalanche photodiode. In: Huntington AS (ed) InGaAs


avalanche photodiodes for ranging and lidar. Woodhead Publishing, Sawston, pp 1–92
2. McIntyre R (1966) Multiplication noise in uniform avalanche diodes. IEEE Trans Electron
Dev ED-13(1):164–168
3. Emmons R (1967) Avalanche-photodiode frequency response. J Appl Phys 38(9):3705–3714
4. Marshall AR, David JP, Tan CH (2010) Impact ionization in InAs electron avalanche photo-
diodes. IEEE Trans Electron Dev 57(10):2631–2638
5. Shin SH, Pasko JG, Law HD, Cheung DT (1982) 1.22-μm HgCdTe/CdTe avalanche photodi-
odes. Appl Phys Lett 40(11):965–967
6. Alabedra R, Orsal B, Lecoy G, Pichard G, Meslage J, Fragnon P (1985) An Hg0.3Cd0.7Te ava-
lanche photodiode for optical-fiber transmission systems at λ = 1.3 μm. IEEE Trans Electron
Dev 32(7):1302–1306
7. Nguyen Duy T, Meslage J, Pichard G (1985) CMT: the material for fiber optical communica-
tion devices. J Cryst Growth 72(1):490–495
72 K. -W. A. Chee

8. de Lyon T, Baumgratz B, Chapman G, Gordon E, Hunter A, Jack M et al (1999) Epitaxial


growth of HgCdTe 1.55-μm avalanche photodiodes by molecular beam epitaxy. In: Proc.
SPIE 3629, 256-267. https://doi.org/10.1117/12.344562
9. Ve’rie´ C, Raymond F, Besson J, Nguyen Duy T (1982) Bandgap spin-orbit splitting reso-
nance effects in Hg1-xCdxTe alloys. J Cryst Growth 59(1):342–346
10. Kinch MA, Beck JD, Wan CF, Ma F, Campbell J (2004) HgCdTe electron avalanche photo-
diodes. J Electron Mater 33(6):630–639
11. Beck JD, Wan C-F, Kinch MA, Robinson J, Mitra P, Scritchfield R et al (2004) The HgCdTe
electron avalanche photodiode. In: Proc. SPIE, 5564, 44-53. https://doi.org/10.1117/12.565142
12. Song H-Z (2018) Chapter 9: Avalanche photodiode focal plane arrays and their application
to laser detection and ranging. In: Chee KWA (ed) Advances in photodetectors-research and
applications. IntechOpen, pp 145–168
13. Dehzangi A, Li J, Razeghi M (2021) Low noise short wavelength infrared avalanche photo-
detector using sb-based strained layer superlattice. Photonics 8(5):148
14. Campbell JC (2021) Evolution of low-noise avalanche photodetectors. IEEE J Sel Top Quant
Electron 28(2):3800911
15. Rothman J (2018) Physics and limitations of HgCdTe APDs: a review. J Electron Mater
47(10):5657–5665
16. Singh A, Pal R (2017) Infrared avalanche photodiode detectors. Defence Sci J 67(2):159
17. Reine M, Marciniec J, Wong K, Parodos T, Mullarkey J, Lamarre P et al (2008)
Characterization of HgCdTe MWIR back-illuminated electron-initiated avalanche photodi-
odes. J Electron Mater 37(9):1376–1386
18. Pal R (2017) Infrared technologies for defence systems. Def Sci J 67(2):133–134
19. Singh A, Srivastav V, Pal R (2011) HgCdTe avalanche photodiodes: a review. Opt Laser
Technol 43(7):1358–1370
20. Rothman J, Foubert K, Mollard L, Péré-Laperne N, Salvetti F, Kerlain A, Reibel Y (2014)
HgCdTe avalanche photodiodes: Application for infra-red detection. In: Proceedings of
11th International Workshop on Low Temperature Electronics (WOLTE). 07-09 July 2014,
Grenoble. https://doi.org/10.1109/WOLTE.2014.6881011
21. Rogalski A (2003) Infrared detectors: status and trends. Prog Quantum Electron 27(2):59–210
22. de Borniol E, Rothman J, Lefoul X (2016) Time resolved infrared detection with HgCdTe ava-
lanche photodiodes. In: OSA Technical Digest of Lasers Congress 2016 (ASSL, LSC, LAC).
30 October–3 November 2016,Boston, LW4B.2. https://doi.org/10.1364/LSC.2016.LW4B.2
23. Rothman J, de Borniol E, Abergel J, Lasfargues G, Delacourt B, Dumas A et al (2017)
HgCdTe APDs for low-photon number IR detection. In: OSA technical Digest (online)
(Optica Publishing Group), paper MM8C.5. https://doi.org/10.1364/MICS.2016.MM8C.5
24. Huntington AS, Compton MA, Williams GM (2007) Linear-mode single-photon APD detec-
tors. In: Proc. SPIE 6771, 67710Q. https://doi.org/10.1117/12.751925
25. Williams GM, Compton M, Ramirez DA, Hayat MM, Huntington AS (2013) Multi-gain-­
stage InGaAs avalanche photodiode with enhanced gain and reduced excess noise. IEEE J
Electron Dev Soc 1(2):54–65
26. Chee KWA (2018) Introductory Chapter: Photodetectors. In: Chee KWA (ed) Advances in
photodetectors – research and applications, vol 3-8. IntechOpen, London
27. Rogalski A (2011) Recent progress in infrared detector technologies. Infrared Phys Technol
54(3):136–154
28. Rogalski A, Antoszewski J, Faraone L (2009) Third-generation infrared photodetector arrays.
J Appl Physics 105(9):091101
29. Simon A, Kalinka G, Jakšić M, Pastuović Ž, Novák M, Kiss ÁZ (2007) Investigation of radia-
tion damage in a Si PIN photodiode for particle detection. Nucl Instrum Methods Phys Res,
Sect B 260(1):304–308
30. Ahmad I, Betts RR, Happ T, Henderson DJ, Wolfs FLH, Wuosmaa AH (1990) Nuclear spec-
troscopy with Si PIN diode detectors at room temperature. Nuclear Instruments and Methods
3 II–VI Compound Semiconductor Avalanche Photodiodes for the Infrared Spectral… 73

in Physics Research Section A: Accelerators, Spectrometers, Detectors and Associated


Equipment 299(1):201–204
31. Andreani L, Bontempi M, Rossi PL, Rignanese LP, Zuffa M, Baldazzi G (2014) Comparison
between a silicon PIN diode and a CsI(Tl) coupled to a silicon PIN diode for dosimetric
purpose in radiology. Nuclear Instruments and Methods in Physics Research Section A:
Accelerators, Spectrometers, Detectors and Associated Equipment 762:11–15
32. Pantazis J, Huber A, Okun P, Squillante MR, Waer P, Entine G (1994) New, high performance
nuclear spectroscopy system using Si-PIN diodes and CdTe detectors. IEEE Trans Nucl Sci
41(4):1004–1008
33. McIntyre RJ (1961) Theory of microplasma instability in silicon. J Appl Phys 32(6):983–995
34. Haitz RH (1964) Model for the electrical behavior of a microplasma. J Appl Phys
35(5):1370–1376
35. Renker D (2006) Geiger-mode avalanche photodiodes, history, properties and prob-
lems. Nuclear Instruments and Methods in Physics Research Section A: Accelerators,
Spectrometers, Detectors and Associated Equipment 567(1):48–56
36. Norton P (2002) HgCdTe infrared detectors. Opto-electron Rev 10:159–174
37. Ralph SH, Neil TG, Jean G, Janet EH, Andrew G, David CH et al (2005) Photomultiplication
with low excess noise factor in MWIR to optical fiber compatible wavelengths in cooled
HgCdTe mesa diodes. In: Proc. SPIE 5783, 92540P. https://doi.org/10.1117/12.603386
38. Ma F, Li X, Campbell JC, Beck JD, Wan C-F, Kinch MA (2003) Monte Carlo simulations of
Hg0.7Cd0.3Te avalanche photodiodes and resonance phenomenon in the multiplication noise.
Appl Physics Lett 83(4):785–787
39. Beck JD, Wan C-F, Kinch MA, Robinson JE (2001) MWIR HgCdTe avalanche photodiodes.
In: Proc. SPIE 4454, 188-196. https://doi.org/10.1117/12.448174
40. Yoo SD, Kwack KD (1997) Theoretical calculation of electron mobility in HgCdTe. J Appl
Phys 81(2):719–725
41. Murthy OVSN, Venkataraman V, Sharma RK, Vurgaftman I, Meyer JR (2009) Multicarrier
conduction and Boltzmann transport analysis of heavy hole mobility in HgCdTe near room
temperature. J Appl Physics 106(11):113708
42. Orsal B, Alabedra R, Valenza M, Pichard G, Meslage J (1985) Impact ionization
rates for electrons and holes in Hg0.3Cd0.7Te in avalanche photodiodes for opti-
cal fiber transmission systems at λ = 1.3 μm. J Crystal Growth 72:496–503. https://doi.
org/10.1016/0022-0248(85)90197-6
43. Han X, Guo H, Yang L, Zhu L, Yang D, Xie H et al (2022) Dark current and noise anal-
ysis for long-wavelength infrared HgCdTe avalanche photodiodes. Infrared Phys Technol
123:104108
44. Rothman J, Foubert K, Lasfargues G, Largeron C, Zayer I, Sodnik Z et al (2014) High
operating temperature SWIR HgCdTe APDs for remote sensing. In: Proc. SPIE 9254,
92540P. https://doi.org/10.1117/12.2069486
45. Zhu L, Guo H, Deng Z, Yang L, Huang J, Yang D et al (2022) Temperature-Dependent
Characteristics of HgCdTe Mid-Wave Infrared E-Avalanche Photodiode. IEEE J Selected
Topics Quantum Electron 28(2: Optical Detectors):1–9
46. Reine MB, Marciniec JW, Wong KK, Parodos T, Mullarkey JD, Lamarre PA et al (2007)
HgCdTe MWIR Back-illuminated electron-initiated avalanche photodiode arrays. J Electron
Mater 36(8):1059–1067
47. Singh A, Shukla AK, Pal R (2015) HgCdTe e-avalanche photodiode detector arrays. AIP Adv
5(8):087172
48. Sieck A, Benecke M, Eich D, Oelmaier R, Wendler J, Figgemeier H (2018) Short-wave
infrared HgCdTe electron avalanche photodiodes for gated viewing. J Electron Mater
47(10):5705–5714
49. Orsal B, Alabedra R, Valenza M, Lecoy GP, Meslage J, Boisrobert CY (1988) Hg0.4Cd0.6Te1.55-
µm avalanche photodiode noise analysis in the vicinity of resonant impact ionization con-
nected with the spin-orbit split-off band. IEEE Trans Electron Dev 35(1):101–107
74 K. -W. A. Chee

50. Orsal B, Alabedra R, Maatougui A, Flachet JC (1991) Hg0.56Cd0.44Te 1.6- to 2.5-μm avalanche
photodiode and noise study far from resonant impact ionization. IEEE Trans Electron Dev
38(8):1748–1756
51. Royer M, Brossat T, Fragnon P, Meslage J, Pichard G, Duy TNG (1983) Détecteurs HgCdTe
pour télécommunication par fibres optiques. Annales des Télécommunications 38(1):62–72
52. Kerlain A, Bonnouvrier G, Rubaldo L, Decaens G, Reibel Y, Abraham P et al (2012)
Performance of mid-wave infrared HgCdTe e-avalanche photodiodes. J Electron Mater
41(10):2943–2948
53. Asbrock J, Bailey S, Baley D, Boisvert J, Chapman G, Crawford G et al (2008) Ultra-High
sensitivity APD based 3D LADAR sensors: linear mode photon counting LADAR cam-
era for the Ultra-Sensitive Detector program. In: Proc. SPIE 6940, 69402O. https://doi.
org/10.1117/12.783940
54. Perrais G, Gravrand O, Baylet J, Destefanis GL, Rothman J (2007) Gain and dark current
characteristics of planar HgCdTe avalanche photo diodes. J Electron Mater 36:963–970
55. Rothman J, Perrais G, Destefanis G, Baylet J, Castelein P, Chamonal JP (2007) High perfor-
mance characteristics in pin MW HgCdTe e-APDs. In: Proc. SPIE 6542, 654219. https://doi.
org/10.1117/12.723465
56. Rothman J, Lasfargues G, Delacourt B, Dumas A, Gibert F, Bardoux A, Boutillier M (2017)
HgCdTe APDS for time resolved space applications. CEAS Space J 9:507–516
57. G. Finger, I. Baker, M. Downing, D. Alvarez, D. Ives, L. Mehrgan, et al. (2017) Development
of HgCdTe large format MBE arrays and noise-free high speed MOVPE EAPD arrays
for ground based NIR astronomy. in Proc. SPIE 10563, 1056311. DOI: https://doi.
org/10.1117/12.2304270
58. Chen J, Chen J, Li X, He J, Yang L, Wang J et al (2021) High-performance HgCdTe avalanche
photodetector enabled with suppression of band-to-band tunneling effect in mid-wavelength
infrared. npj Quantum Mater 6(1):103
59. Zhou T, Chee KWA (2019) Chapter 7: Overcoming the Bandwidth-Quantum Efficiency
Trade-Off in Conventional Photodetectors. In: Chee KWA (ed) Advances in photodetectors –
research and applications. IntechOpen, London, United Kingdom, pp 115–126
60. Michael DJ, James FA, Anderson C, Steven LB, George C, Gordon E et al (2001) Advances
in linear and area HgCdTe APD arrays for eyesafe LADAR sensors. In: Proc. SPIE 4454,
198-211. https://doi.org/10.1117/12.448175
61. Michael J, Jim A, Steven B, Diane B, George C, Gina C et al (2007) MBE based HgCdTe APDs
and 3D LADAR sensors. In: Proc. SPIE 6542, 65421A. https://doi.org/10.1117/12.724347
62. Mallik S, Hultquist K, Ghosh S, Velicu S, Hyeson J (2005) MBE grown mid-infrared HgCdTe
avalanche photodiodes on Si substrates. In: 63rd Device Research Conference Digest, 20-22
June 2005, Santa Barbara, CA, USA, pp 75-76. https://doi.org/10.1109/DRC.2005.1553062
63. Mitra P, Case FC, Reine MB, Starr R, Weiler MH (1997) Doping in MOVPE of HgCdTe:
orientation effects and growth of high performance IR photodiodes. J Cryst Growth
170(1):542–548
64. Hall DNB, Baker IM, Finger G (2016) Towards the next generation of L-APD MOVPE
HgCdTe arrays: beyond the SAPHIRA 320 × 256. In: Proc. SPIE 9915, 99150O. https://doi.
org/10.1117/12.2234370
65. Baker IM, Maxey C, Hipwood LG, Weller HJ, Thorne P (2012) Developments in MOVPE
HgCdTe arrays for passive and active infrared imaging. In: Proc. SPIE 8542, 85421A. https://
doi.org/10.1117/12.981850
66. Maxey CD, Capper P, Baker IM (2019) Chapter 9: MOVPE growth of cadmium mercury
telluride and applications. In: Irvine S, Capper P (eds) Metalorganic Vapor Phase Epitaxy
(MOVPE). Wiley, New Jersey, pp 293–324
67. Ian B, Chris M, Les H, Keith B (2016) Leonardo (formerly Selex ES) infrared sensors for astron-
omy: present and future. In: Proc. SPIE 9915, 991505. https://doi.org/10.1117/12.2231079
3 II–VI Compound Semiconductor Avalanche Photodiodes for the Infrared Spectral… 75

68. Johan R, Pierre B, Julie A, Sylvain G, Gilles L, Lydie M et al (2018) HgCdTe APDs detector
developments at CEA/Leti for atmospheric lidar and free space optical communications. In:
Proc. SPIE 11180, 111803S. https://doi.org/10.1117/12.2536055
69. Ferret P, Zanatta JP, Hamelin R, Cremer S, Million A, Wolny M, Destefanis G (2000) Status
of the MBE technology at leti LIR for the manufacturing of HgCdTe focal plane arrays. J
Electron Mater 29(6):641–647
70. Behet M, Hövel R, Kohl A, Küsters AM, Opitz B, Heime K (1996) MOVPE growth of III–V
compounds for optoelectronic and electronic applications. Microelectron J 27(4):297–334
71. Mullin JB, Cole-Hamilton DJ, Irvine SJC, Hails JE, Giess J, Gough JS (1990) MOVPE of
narrow and wide gap II–VI compounds. J Cryst Growth 101(1):1–13
72. Snyder DW, Mahajan S, Ko EI, Sides PJ (1991) Effect of substrate misorientation on sur-
face morphology of homoepitaxial CdTe films grown by organometallic vapor phase epitaxy.
Appl Phys Lett 58(8):848–850
73. Capper P, Maxey CD, Whiffin PAC, Easton BC (1989) Substrate orientation effects in
CdxHg1−xTe grown by MOVPE. J Cryst Growth 96(3):519–532
74. Švob L, Chèze I, Lusson A, Ballutaud D, Rommeluère JF, Marfaing Y (1998) Crystallographic
orientation dependence of As incorporation in MOVPE-grown CdTe and corresponding
acceptor electrical state activation. J Cryst Growth 184-185:459–464
75. Gawron W, Madejczyk P, Kłos K, Rutkowski J, Piotrowski A, Rogalski A, Mróz W (2009)
Surface smoothness improvement of HgCdTe layers grown by MOCVD. Bullet Polish Acad
Sci Technical Sci 57(2):139–146
76. Mitra P, Case FC, Glass HL, Speziale VM, Flint JP, Tobin SP, Norton PW (2001) HgCdTe
growth on (552) oriented CdZnTe by metalorganic vapor phase epitaxy. J Electron Mater
30(6):779–784
77. Wang W-S, Bhat I (1995) Growth of high quality CdTe and ZnTe on Si substrates using
organometallic vapor phase epitaxy. J Electron Mater 24(5):451–455
78. Jones CL, Hipwood LG, Shaw CJ, Price JP, Catchpole RA, Ordish M et al (2006) High per-
formance MW and LW IRFPAs made from HgCdTe grown by MOVPE. In: Proc. SPIE 6206,
620610. https://doi.org/10.1117/12.667610
79. Hails JE, Keir AM, Graham A, Williams GM, Giess J (2007) Influence of the silicon substrate
on defect formation in MCT grown on II-VI buffered Si using a combined molecular beam
epitaxy/metal organic vapor phase epitaxy technique. J Electron Mater 36(8):864–870
80. Maruyama K, Nishino H, Okamoto T, Murakami S, Saito T, Nishijima Y et al (1996) Growth
of (111) HgCdTe on (100) Si by MOVPE using metalorganic tellurium adsorption and
annealing. J Electron Mater 25(8):1353–1357
81. Nishino H, Murakami S, Saito T, Nishijima Y, Takigawa H (1995) Dislocation profiles in
HgCdTe(100) on GaAs(100) grown by metalorganic chemical vapor deposition. J Electron
Mater 24(5):533–537
82. Hipwood LG, Jones CL, Walker D, Shaw CJ, Abbott P, Catchpole RA et al (2007) Affordable
high-performance LW IRFPAs made from HgCdTe grown by MOVPE. In: Proc. SPIE 6542,
65420I. https://doi.org/10.1117/12.720647
83. Madejczyk P, Piotrowski A, Gawron W, Kłos K, Pawluczyk J, Rutkowski J et al (2005)
Growth and properties of MOCVD HgCdTe epilayers on GaAs substrates. Opto-electron
Rev 13(3):239–251
84. Piotrowski A, Kłos K (2007) Metal-organic chemical vapor deposition of Hg1−xCdxTe fully
doped heterostructures without postgrowth anneal for uncooled MWIR and LWIR detectors.
J Electron Mater 36(8):1052–1058
85. Giess J, Hails JE, Graham AP, Blackmore GW, Houlton MR, Newey J et al (1995) The role of
surface adsorbates in the metalorganic vapor phase epitaxial growth of (Hg,Cd)Te onto (100)
GaAs Substrates. J Electron Mater 24:1149–1153
86. Suh S-H, Song J-H, Moon S-W (1996) Metalorganic vapor phase epitaxial growth of
hillock free (100) sol HgCdTeGaAs with good electrical properties. J Cryst Growth
159(1–4):1132–1135
76 K. -W. A. Chee

87. Suh S-H, Kim J-S, Kim HJ, Song J-H (2002) Control of hillock formation during MOVPE
growth of HgCdTe by suppressing the pre-reaction of the Cd precursor with Hg. J Cryst
Growth 236(1):119–124
88. Irvine SJC, Giess J, Gough JS, Blackmore GW, Royle A, Mullin JB et al (1986) The poten-
tial for abrupt interfaces in CdxHg1−xTe using thermal and photo-MOVPE. J Cryst Growth
77(1–3):437–451
89. He L, Yang J, Wang S, Wu Y, Fang W (1999) Recent progress in molecular beam epitaxy of
HgCdTe. Adv Mater 11(13):1115–1118
90. Zanatta JP, Ferret P, Theret G, Million A, Wolny M, Chamonal JP, Destefanis G (1998)
Heteroepitaxy of HgCdTe (211)B on Ge substrates by molecular beam epitaxy for infrared
detectors. J Electron Mater 27(6):542–545
91. De Lyon TJ, Rajavel RD, Jensen JE, Wu OK, Johnson SM, Cockrum CA, Venzor GM
(1996) Heteroepitaxy of HgCdTe(112) infrared detector structures on Si(112) substrates by
molecular-­beam epitaxy. J Electron Mater 25(8):1341–1346
92. He L, Chen L, Wu Y, Fu XL, Wang YZ, Wu J et al (2007) MBE HgCdTe on Si and GaAs
substrates. J Cryst Growth 301-302:268–272
93. Nosho BZ, Roth JA, Jensen JE, Pham L (2005) Lateral uniformity in HgCdTe layers grown
by molecular beam epitaxy. J Electron Mater 34(6):779–785
94. Faurie JP, Million A, Boch R, Tissot JL (1983) Latest developments in the growth of
CdxHg1−xTe and CdTe–HgTe superlattices by molecular beam epitaxy. J Vac Sci Technol A
1(3):1593–1597
95. Gordon NT, Baker IM (2001) Assessment of infrared materials and devices. In: Capper P,
Elliott CT (eds) Infrared detectors and emitters: materials and devices. Springer US, Boston,
pp 23–42
96. Bae SH, Lee SJ, Kim YH, Lee HC, Kim CK (2000) Analysis of 1/f noise in LWIR HgCdTe
photodiodes. J Electron Mater 29(6):877–882
97. Pillans L, Ash RM, Hipwood L, Knowles P (2012) MWIR mercury cadmium tellu-
ride detectors for high operating temperatures. In: Proc. SPIE 8353, 83532W. https://doi.
org/10.1117/12.919015
98. Kopytko M, Rogalski A (2022) New insights into the ultimate performance of HgCdTe pho-
todiodes. Sensors Actuators A Phys 339:113511
99. Martyniuk P, Rogalski A (2015) MWIR barrier detectors versus HgCdTe photodiodes.
Infrared Phys Technol 70:125–128
100. Maimon S, Wicks GW (2006) nBn detector, an infrared detector with reduced dark current
and higher operating temperature. Appl Phys Letters 89(15):151109
101. Itsuno AM, Phillips JD, Velicu S (2011) Design and modeling of HgCdTe nBn detectors. J
Electron Mater 40(8):1624–1629
102. Uzgur F, Kocaman S (2019) Barrier engineering for HgCdTe unipolar detectors on alternative
substrates. Infrared Phys Technol 97:123–128
103. Itsuno AM, Phillips JD, Velicu S (2012) Design of an Auger-suppressed unipolar HgCdTe
NBνN photodetector. J Electron Mater 41(10):2886–2892
104. Kopytko M, Kębłowski A, Gawron W, Madejczyk P, Kowalewski A, Jóźwikowski K (2013)
High-operating temperature MWIR nBn HgCdTe detector grown by MOCVD 21(4):402–405
105. Kopytko M, Kębłowski A, Gawron W, Pusz W (2016) LWIR HgCdTe barrier photodiode
with Auger-suppression. Semicond Sci Technol 31(3):035025
106. Kopytko M, Jóźwikowski K (2013) Numerical analysis of current–voltage characteristics of
LWIR nBn and p-on-n HgCdTe photodetectors. J Electron Mater 42(11):3211–3216
107. Kopytko M, Wróbel J, Jóźwikowski K, Rogalski A, Antoszewski J, Akhavan ND et al (2015)
Engineering the bandgap of unipolar HgCdTe-based nBn infrared photodetectors. J Electron
Mater 44(1):158–166
108. Kopytko M, Jóźwikowski K (2015) Generation-recombination effect in MWIR HgCdTe bar-
rier detectors for high-temperature operation. IEEE Trans Electron Dev 62(7):2278–2284
3 II–VI Compound Semiconductor Avalanche Photodiodes for the Infrared Spectral… 77

109. Akhavan ND, Umana-Membreno GA, Gu R, Antoszewski J, Faraone L (2018) Delta doping in
HgCdTe-based unipolar barrier photodetectors. IEEE Transa Electron Dev 65(10):4340–4345
110. Madejczyk P, Gawron W, Martyniuk P, Keblowski A, Pusz W, Pawluczyk J et al (2017)
Engineering steps for optimizing high temperature LWIR HgCdTe photodiodes. Infrared
Phys Technol 81:276–281
111. Rogalski A, Kopytko M, Martyniuk P (2018) Performance prediction of p-i-n HgCdTe long-­
wavelength infrared HOT photodiodes. Appl Opt 57(18):D11–D19
112. Vallone M, Goano M, Bertazzi F, Ghione G, Palmieri A, Hanna S et al (2019) Reducing inter-­
pixel crosstalk in HgCdTe detectors. Optical and Quantum Electronics 52(1):25
113. Vallone M, Goano M, Bertazzi F, Ghione G, Hanna S, Eich D et al (2020) Constraints and per-
formance trade-offs in Auger-suppressed HgCdTe focal plane arrays. Appl Opt 59(17):E1–E8
114. He J, Li Q, Wang P, Wang F, Gu Y, Shen C et al (2020) Design of a bandgap-engineered
barrier-blocking HOT HgCdTe long-wavelength infrared avalanche photodiode. Opt Express
28(22):33556–33563
115. Kopytko M, Gawron W, Kębłowski A, Stępień D, Martyniuk P, Jóźwikowski K (2019)
Numerical analysis of HgCdTe dual-band infrared detector. Opt Quantum Electron 51(3):62
116. Fatih U, Serdar K (2019) A dual-band HgCdTe nBn infrared detector design. In: Proc. SPIE
11129, 1112903. https://doi.org/10.1117/12.2529240
117. Philippe T, Gérard D, Philippe B, Jacques B, Olivier G, Johan R (2008) Advanced HgCdTe
technologies and dual-band developments. In: Proc. SPIE 6940, 69402P. https://doi.
org/10.1117/12.779902
118. Perrais G, Rothman J, Destefanis G, Chamonal J-P (2008) Impulse response time measure-
ments in Hg0.7Cd0.3Te MWIR avalanche photodiodes. J Electron Mater 37(9):1261–1273
119. Perrais G, Derelle S, Mollard L, Chamonal J-P, Destefanis G, Vincent G et al (2009) Study of
the transit-time limitations of the impulse response in mid-wave infrared HgCdTe avalanche
photodiodes. J Electron Mater 38(8):1790–1799
120. Li Q, Wang F, Wang P, Zhang L, He J, Chen L et al (2020) Enhanced performance of HgCdTe
midwavelength infrared electron avalanche photodetectors with guard ring designs. IEEE
Trans Electron Dev 67(2):542–546
121. Faurie JP, Reno J, Sivananthan S, Sou IK, Chu X, Boukerche M, Wijewarnasuriya PS (1986)
Molecular beam epitaxial growth and characterization of HgCdTe, HgZnTe, and HgMnTe on
GaAs(100). J Vac Sci Technol A 4(4):2067–2071
Chapter 4
IR Detectors Array

Ghenadii Korotcenkov

4.1 Introduction

In recent years, the pace of development of thermal imaging technology has signifi­
cantly accelerated [20, 52, 60, 64, 65]. Since its inception, the market for infrared
(IR) thermal imaging technology has grown primarily due to its military applica­
tions. Thermal imaging devices using a matrix of thermal and photon detectors allow
combat operations in poor visibility conditions, detect hidden objects and give target
designation. Thermal imaging cameras in the form of binoculars, optical sights,
guidance and homing systems are widely used in the army, navy and aviation, vision
systems for ground combat robots and perimeter security systems. Today, the mili­
tary sector still provides the market with some growth, but the priorities for its devel­
opment have changed (see Fig. 4.1a). At present the main growth in the market is
provided by the sectors of civil and medical thermography, security and fire surveil­
lance, as well as applications related to ensuring the safety of navigation, control and
protection of water and coastal objects, astrophysics, space and science imaging [9,
10, 23, 24, 32, 35, 45–47, 73]. Such devices are used for rescue operations during
extinguishing fires, rescuing people in smoke conditions, identifying the most dan­
gerous areas with high temperatures. Devices using thermal imaging cameras allow
observation in poor visibility conditions and detecting people with high tempera­
tures in a crowd. According to the forecast of Maxtech International (USA), the
market for civil and military infrared systems, amounting to $10.5 billion in 2017,
will exceed $17 billion in 2023 and should exceed $20 billion in 2025.
In the development of thermal imaging cameras, various approaches can be used.
In particular, such systems can be developed on the basis of (a) single photo­
detectors and two-dimensional (line and frame) scanning using a scanning

G. Korotcenkov (*)
Department of Physics and Engineering, Moldova State University, Chisinau, Moldova
e-mail: ghkoro@yahoo.com

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 79


G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_4
80 G. Korotcenkov

Fig. 4.1 (a) Areas of use of thermal imaging (Yole Development 2018, http://www.yole.fr/); (b)
Relative shares of sales of photonic FPAs on different materials (Maxtech International, https://
maxtech-­intl.com/)

optical-­mechanical system, (b) line of photodetectors and a simplified frame scan,


or (c) grouped several lines (with time delay and accumulation) and a low-speed
sweep system. The latter include vacuum devices with electronic scanning of the
receiving target. However, the presence of optical-mechanical scanning systems
leads to a significant complication of the equipment, a decrease in its manufactur­
ability and reliability. When observing fast moving objects, there is also great dif­
ficulty in creating a high-speed mechanical scanner. Electron beam scanning has the
best performance. But the large dimensions of cathode-ray tubes and the need for
high-voltage power supply, an incandescent cathode, deflecting and focusing sys­
tems do not allow the creation of portable transmitting video cameras. Therefore, in
recent years, the main direction in the development of such systems is the use of
“simultaneously looking” – focal-plane, two-dimensional solid-state multi-element
radiation detectors or Focal-Plane Array (FPA), the operation of which does not
require the use of optical-mechanical scanning systems. All other things being
equal, such thermal imaging cameras outperform scanning systems in terms of the
combination of such parameters as reliability, sensitivity, signal-to-noise ratio,
image quality, speed, dynamic range, and spatial resolution. Other advantages of
such devices are low weight, dimensions and power consumption, noiseless opera­
tion, the ability to communicate with modern computers, video and TV equipment,
and digital image processing in real time. At present, developments have reached
such a level that large-format matrices (with the number of pixels 103 × 103 and
more) are actually already technical analogs of the retina in the organs of vision. For
4 IR Detectors Array 81

example, modern technology allows the manufacture of IR-FPAs, which can have
up to 108 IR detectors, which corresponds to the number of sensitive receptors in the
human eye (~2·108). This provides not only a significant expansion of the spectral
range of vision, but also the implementation of optoelectronic systems with the
maximum achievable values ​​of the threshold sensitivity, inertia and information
capacity.

4.1.1 Materials and Types of IR Detectors

IR-FPAs can be made on the basis of various materials that determine such charac­
teristics of thermal imaging cameras as the operating spectral range, temperature
sensitivity, the need for cooling the array and the speed of the system [20]. There are
two main types of infrared detectors used in thermal imaging cameras: photon
detectors and thermal detectors. Thermal detectors include microbolometric and
pyroelectric detectors. Table 4.1 shows the materials on the basis of which various
types of IR photodetectors arrays are developed.
The most important advantage of bolometric infrared detectors is the ability to
operate without cooling (at temperatures of about 300 K), while most photon detec­
tors operate at cryogenic temperatures (usually at least 77 K). In addition, the cost
of FPAs based on uncooled bolometers in industrial production is two orders of
magnitude lower than the cost of photonic arrays [69]. However, they are noticeably
inferior in other parameters to photonic FPAs.

Table 4.1 Materials used in IR-FPAs of various types


Spectral
Types of IR detectors Material of photodetector array range, μm
Photonic Lead chalcogenide (PbS, PbSe); 1.5–6
Compounds mercury-cadmium-tellurium HgCdTe 1–20
(MCT); 3–5
Indium Antimonide (InSb); 3–5
Platinum silicide – silicon Schottky barrier structures 1.5–40
(PtSi-Si); 8–14
Doped silicon (Si:X) and germanium (Ge:X); 0.3–2.3
Multilayer structures with quantum wells based on
GaAs /AlGaAs (QWIP detectors);
Compound indium-gallium-arsenide (InGaAs)
Thermal Vanadium oxides VxO; 8–14
microbolometric Polycrystalline and amorphous silicon 8–14
Thermal pyroelectric Lead zirconate; 8–12
Barium-strontium niobate and titanate;
Triglycine sulfate
82 G. Korotcenkov

4.1.2 Photonic IR FPAs and Basic Materials Used


to Develop Them

4.1.2.1 Materials Used in the Development of Photonic IR FPAs

Currently, the main materials used in the development of photon-type IR FPAs are
HgCdTe (MCT), InSb, InGaAs, InAs/GaSb and GaAs/AlGaAs. InAs/GaSb hetero­
structures are used to fabricate type-II superlattices (T2SLs), while GaAs/AlGaAs
are used to fabricate quantum well heterostructures (type-I superlattices, T1SLs).
We should also not forget the quantum dots of various materials, based on which in
recent years it has also been possible to develop efficient IR photodetectors, the so-­
called quantum dot IR photodetector (QDIP) (Chap. 7, Vol. 2).
Quantum well IR photodetectors (QWIP) on the base of GaAs/AlGaAs hetero­
structures have a narrow spectral region and a relatively low quantum efficiency
(less than 10%), therefore, they require a longer signal accumulation time than
devices based on InSb and HgCdTe (quantum efficiency is about 90%). However,
the spectral characteristics of QWIP detectors can be flexibly adjusted by changing
the width and height of the quantum wells formed by the GaAs and AlGaAs layers.
In addition, the ability to use commercial manufacturing processes developed for
III-V compounds significantly reduces the cost of the ongoing development. One-­
color IR photodetector arrays based on quantum-dimensional structures with for­
mats 320 × 256 and 640 × 480 are already produced by companies such as Sander,
Lookheed Martin and QWIP Technologies, and these FPAs are already being used
in a number of commercial applications. This is proof that the technology of grow­
ing quantum-dimensional structures (QDS) has become effective.
Interest in FPAs development based on InAs/GaSb type-II superlattices (T2SLs)
has emerged in recent years [59]. There are two primary motivations for develop­
ment of InAs/GaSb T2SLs. The first is related to the existing problems of reproduc­
ibly fabricating high-operability HgCdTe focal plane arrays (FPAs) at reasonable
cost. The second motivation is theoretical predictions of lower Auger recombination
for T2SL detectors compared to HgCdTe, which can be translated into a fundamen­
tal advantage of T2SL over HgCdTe in terms of lower dark current, provided that
other parameters such as Shockley-Read-Hall (SRH) lifetime are equal. At present
InAs/GaSb T2SL photodetectors offer similar performance to HgCdTe at an equiva­
lent cut-off wavelength, but with a sizeable penalty in operating temperature, due to
the inherent difference in SRH lifetimes. It is predicted that since the future infrared
(IR) systems will be based on the room temperature operation of depletion-current
limited arrays with pixel densities that are fully consistent with background- and
diffraction-limited performance due to the system optics, the material system with
long SRH lifetime will be required. Since T2SLs are very much resisted in attempts
to improve its SRH lifetime, currently the only material that meets this requirement
is HgCdTe [59]. However, despite the fact that the modern version of the T2SLs
technology is as yet in its infancy, the rapid progress made in T2SLs over the past
few years has shown great promise for this material for a future development of IR
technologies [19, 41].
4 IR Detectors Array 83

As for FPAs based on QDIP, they are still under development and therefore can­
not yet compete with FPAs based on MCT [9, 27]. Achieving acceptable results
requires significant investment and fundamental research. In connection with the
above, FPAs made of mercury-cadmium-tellurium (MCT) and indium antimonide
(InSb) are currently mainly used for high-sensitivity and long-range thermal imag­
ing cameras. In terms of their parameters, devices based on these materials are close
to theoretical characteristics. It is expected that HgCdTe and InSb will remain the
main materials for IR technology for at least the next 10-15 years. At present,
approximately an equal number of devices are produced based on these compounds,
and this ratio will remain the same. However, in the field of military applications
where high sensitivity and speed are required, MCT still dominates [9]. HgCdTe is
currently the most prevalent material system used in high performance infrared
detectors due to the following reasons [60–62, 64]:
• the tailorable energy band gap covers the entire spectral range of 1–30 μm, which
includes important spectral ranges such as 1–2.5 μm, 3–5 μm, and 8–14 μm,
• large optical coefficients that enable high quantum efficiency, and
• favorable inherent recombination mechanisms that lead to high operating
temperature.

4.1.2.2 Photonic IR FPAs

Modern hybrid multi-element IR photodetectors usually consist of two parts: a


multi-element photosensitive structure and a silicon multiplexer, connected to each
other by group cold welding using indium bumps (Fig. 4.2). A linear or matrix mul­
tiplexer or readout integrated circuits (ROIC) is an integrated circuit that provides
the required electrical modes of operation of photosensitive elements, reads electri­
cal signals obtained as a result of photoelectric conversion of incident IR radiation
by photodetector, performs preprocessing, and analog-digital conversion of the
photo signal, and ultimately largely determines the quality of the resulting image
[44]. Examples of ROICs development by Sofradir and LETI one can find in [6, 26,
29]. One of these ROICs is shown in Fig. 4.3.
Usually, the composition of image signal generators also includes shift registers,
clock generators, analog-to-digital converters, blocks for precision correction of
pixel inhomogeneity, blocks for replacing photosignals from non-working ele­
ments, blocks for subtracting the background signal component. They may include
electronic devices that provide high-frequency reading of image fragments, as well
as devices for integrating two or more colors (if photosensitive elements have sen­
sitivity in two or more spectral ranges), devices for encoding images into conven­
tional colors, etc. The combination of the photosensitive element and the
microelectronic path in a single housing led to a decrease in size, to a decrease in
parasitic capacitances and to improved shielding from external interference, that is,
to an improvement in the threshold and operational characteristics of optoelectronic
equipment. In addition, such a combination made it possible to ensure the transport­
ability of the output signal from photodetectors and to significantly simplify its
84 G. Korotcenkov

Fig. 4.2 Hybrid IR FPA with independently optimized signal detection and readout: (a, b) Cross-­
section of the photodiodes during hybridization using (a) indium bump technique and (b) loophole
technique. (Reprinted from Ref. [61]. Published 2008 by De Gruyter as open access)

Fig. 4.3 The photo of 384 × 288 silicon readout integrated circuit with 25 μm pixel pitch for
MWIR and LWIR HgCdTe based FPA: 1 – digital-to-analog converters (DACs) block, 2 – array of
pixel cells, 3 – column followers and decoder, 4 – preamplifiers and output buffers, 5 – digital
block, 6 – row decoder. (Reprinted from Ref. [83]. Published 2015 by IOP as open access)
4 IR Detectors Array 85

processing [32]. Silicon multiplexer chips are subject to more stringent uniformity
requirements than conventional integrated circuit; and meeting these requirements
are often beyond the scope of standard factory manufacturing processes for making
very large scale integrated circuits. By the type of construction, photonic detectors
arrays can be created both on the basis of complementary metal-oxide-­semiconductor
(CMOS) switches [33], and on the basis of a charge-coupled devices (CCD). CCD
technology is used for not very large scale arrays and their technology is more com­
plicated than the CMOS production line. In comparison with CCDs, the MOS mul­
tiplexers exhibit important advantages. CMOS technology provides a higher
realizable capacity of the storage MOS capacitors, which is important at radiation
wavelengths above 5 micrometers, at which the background illumination is high. In
addition, tighter packing, higher and more uniform electrical parameters at 77 K,
lower control voltages, and greater dynamic range are achieved [64]. Leading world
firms such as SOFRADIR (www.sofradir.com), Indigo Systems Corporation (www.
indigosystems.com), etc., have developed and are currently serially producing a
series of linear (4 × 288, 6 × 460, 1 × 512) and matrix (128 × 128, 320 × 240,
320 × 256, 384 × 288, 640 × 512) silicon multiplexers for long-range (LWIR),
medium (MWIR) and near infrared (NIR) spectral ranges. Thus hybridized assem­
bly of the FPA and ROIC is glued to the surface of the carrier sapphire substrate
from the side of the silicon microcircuit. On the contact pads of the substrate, leads
are wired for reading photosignals, supplying power and control signals. Thermal
sensors are installed on the same substrate and a cooled diaphragm is attached. The
appearance of such assemblies is shown in Fig. 4.4a, b. Image shown in Fig. 4.4c
demonstrates the resolution capability of such FPAs with pixel size smaller than
15 μm [49].

Fig. 4.4 (a) A standard SAPHIRA detector (https://www.leonardocompany.com/en/products/


saphira-­1). The dark rectangle in the center is the HgCdTe APD array, and the temperature sensor
is visible in the upper-left corner. In operation it is mounted to a 68-pin leadless chip carrier.
HgCdTe APD array hybridized to a complementary metal-oxide-semiconductor (CMOS) ROIC;
(b) HgCdTe 1280 × 1024 SCORPIO MW detector with pixel pitch of 15 μm (www.Sofradir.com);
(c) Image taken with SELEX SuperHawk 1280 × 1024 MWIR HgCdTe FPA module with pitch of
8 μm (https://www.leonardocompany-­us.com/). This image was obtained from a liquid argon
cooled FPA in the test room at Selex ES. (Reprinted with permission from Ref. [49]. Copyright
2015: SPIE)
86 G. Korotcenkov

4.2 Photovoltaic HgCdTe-Based FPAs

4.2.1 Technological Developments

Photovoltaic HgCdTe FPAs are based on both p-type and n-type materials. The
highest quality structures are grown on Cd1-yZnyTe (CZT) (y = 0.03–0.05) sub­
strates, matched to MCT in the crystal lattice constant, which makes it possible to
obtain structures with a small amount of dislocation (<105 cm−2). The CZT substrate
growth performed by vertical gradient freeze (VGF) bulk method. Current boules
grown at LETI and Sofradir are above 110 mm in diameters, while maintaining the
very high crystalline quality required for MCT growth, low dislocation density (in
the low 104 cm−2) and low precipitate concentration [11]. Usually crystalline orien­
tation is set to (111) in order to perform the liquid phase epitaxy of the MCT layer
onto this substrate. The Zn content in CdZnTe is optimized to guarantee lattice
matching with the targeted MCT composition. After dicing into 7 × 7 cm wafers, the
substrate surface has to be prepared for MCT growth using liquid phase epitaxy
(PLE), molecular beam epitaxy (MBE) [50] or metalorganic vapour phase epitaxy
(MOVPE) methods [10].
The complex and multistage technology for producing MCTs includes deep
purification of the starting Cd, Hg and Te, the synthesis of precursors, and the
growth of HgCdTe epitaxial layers by various methods. One of the important param­
eters of MCT layers designed for FPAs is the uniformity of the composition over the
area. Moreover, for photodetectors operating in the range of 8–12 μm, the change in
the composition of Hg1-xCdxTe films should not exceed ∆x = 0.0002 by 10 mm. At
present, the main industrial method for the manufacture of epitaxial layers in the
world’s leading companies producing IR photodetectors arrays is the method of
liquid-phase epitaxy on a cadmium-zinc-tellurium compound (CZT) substrate. The
advantages of this method are the relatively low cost and high productivity of the
equipment, automatic additional cleaning of the surface at the initial stage of growth,
additional cleaning from impurities in the growth process, and uniformity of the
composition over the area. In addition, photodetectors fabricated using liquid epi­
taxy exhibited higher efficiency and lower dark currents. However, large-area
CdZnTe substrates remain expensive products with poorly reproducible characteris­
tics. Besides that, the use of single-crystal CdZnTe as substrates, due to the limited
diameter of the grown single crystals, significantly limits the possibilities of creat­
ing megapixel photodetectors. As is known, megapixel IR FPAs require large-area
MCT plates with a given high lateral compositional homogeneity. All this together
with the relatively low yield due to cluster defects and blinking pixels, makes the
cost of the technology using CdZnTe substrates prohibitive. In this regard, technolo­
gies are being developed everywhere that make it possible to exclude the use of
CdZnTe substrates when growing large area HgCdTe epitaxial layers [3, 40]. Studies
have shown that this problem can be solved by using low-temperature molecular
beam epitaxy (MBE) of HgCdTe films. The experiment showed that MBE is the
most developed, flexible method for obtaining such a material in the form of
4 IR Detectors Array 87

hetero-epitaxial structures. The built-in ellipsometric equipment allows high-­


precision control of the composition of the epitaxial MCT film and its changes dur­
ing growth. This opens up wide possibilities for optimizing the design of
heterostructures, which makes it possible to simplify the technology of fabricating
IR FPAs with extremely high parameters. Another important advantage of the MBE
method for growing MCTs is the use of cheap silicon wafers. This means that
HgCdTe growth on large-area Si substrates is to enable larger array formats and
potentially reduced FPA cost compared to FPA fabricated using smaller and more
expensive CdZnTe substrates.
For cooled hybrid large format IR FPAs, there is a reliability issue in thermal
cycling from room temperature to liquid nitrogen temperature. This problem is sim­
plified precisely when using silicon substrates for growing MCTs. But it must be
admitted that the large difference in chemical composition and crystal lattice param­
eters mismatch between MCT and Si makes the problem of fabricating FPAs based
on MCT/Si structures with suitable parameters into an extremely difficult task [45,
46]. Only the optimization of the processes of pre-epitaxial preparation of the sili­
con substrate surface and the conditions for the formation of the CdTe/ZnTe buffer
layer [50], as well as the improvement of MCT growth processes and the use of
post-growth processes made it possible to obtain heteroepitaxial MCT structures by
the MBE method with acceptable parameters; without interphase boundaries, with
a uniform distribution over the surface of morphological V-defects and etching pits
with densities less than 103 cm2 and 106 cm2, respectively [12, 17]. This method
made it possible to grow homogeneous MCT heterostructures on silicon substrates
with a diameter of more than 100 mm. Using MBE system, researchers produced
epitaxial HgCdTe layers on (211) Si substrates with very low macro defect density
and uniform Cd composition across the epitaxial wafers. These HgCdTe/Si compos­
ite wafers have shown growth defect densities less than 10 defects /cm2, approxi­
mately 100 times better than can be achieved on CdZnTe substrates, due to the
better crystalline quality of the starting substrate [8]. Comparative characteristics of
the processes of growing epitaxial MCT layers on CdZnTe and Si substrates are
given in the Table 4.2. As you can see, Si-based substrate technology for growing
HgCdTe epitaxial layers has a number of significant advantages that can signifi­
cantly reduce the cost of FPAs based on HgCdTe.
It is important to note that in the fabrication of devices based on epitaxial HgCdTe
films grown on Si substrates, the same etching, passivation, and metallization
schemes can be used as in the fabrication of detectors based on HgCdTe /CdZnTe
structures. Wherein, Bangs et al. [8] have shown that manufacturing processes for
detectors throughout the area 150 mm HgCdTe/Si wafers usually produced high
performing detector pixels from edge to edge of the photolithographic limits across
the wafer, offering 5 times the printable area as compared with 6 × 6 cm CdZnTe
substrates. Large-format (2 K × 2 K) MWIR FPAs fabricated using large area
HgCdTe layers grown on 6-inch diameter (211) silicon substrates demonstrated
NEDT operability better than 99.9%. Moreover, SWIR and MWIR detector perfor­
mance characteristic for devices fabricated on HgCdTe/Si substrates are compara­
ble to those established for devices fabricated on HgCdTe/CdZnTe wafers. HgCdTe
88 G. Korotcenkov

Table 4.2 Advantages of Si-based substrate technology for HgCdTe material development
Parameter Bulk CdZnTe Si
Maximum size 7 × 7 cm2 150 мм in diameter
Maximum area ∼50 cm2 ∼180 cm2
Scalability No Yes
Cost $220/cm2 ∼$1/cm2
Thermal match to Si ROIC No Yes
Robustness Brittle Hard
Lattice match to MCT Yes No
Surface Smooth Smooth
Orientation available (111) (112)
Substrate quality (dislocations) <10,000 cm2 <100 cm2
Impurities Low Extremely low
Source: Data extracted from [12] and [20]

devices fabricated on both types of substrates have demonstrated very low dark
current, high quantum efficiency and full spectral band fill factor characteristic of
HgCdTe [8, 15].
As for the very technology of manufacturing photodetectors that form a matrix
of photodetectors, this technology does not differ from the technology for manufac­
turing discrete photodetectors described earlier in [40] and Chap. 15, Vol. 1. To
create p-n junctions, the method of ion implantation of B+ or As+, depending on the
type of conductivity of the main absorbing layer, with a dose of (2–3)·1013 cm−2 is
usually used. In this case, the main mechanism for the formation of the n-type doped
region in p-HgCdTe is when boron ions knock out mercury atoms, which pass into
interstices, becoming electrically active defects responsible for the formation of the
n-region of the p-n junction.
A HgCdTe avalanche photodiodes (APDs) are also used to make the focal plane
array [68, 75]. It is believed that HgCdTe avalanche photodiodes (APDs) are par­
ticularly suitable for low flux detection. Indeed, the possibility to pre-amplify the
incoming photonic signal into the photodiode itself (with no major degradation of
the signal to noise ratio [53] is very interesting when the detection performance is
limited by the ROIC noise as it is the case in the very low flux detection for astron­
omy [23, 24]. As for the configuration of the photodetectors themselves that form
the array, they can be made both in planar design and in the form of mesa structures.
However, the best results were obtained using mesa diode technology and MCT
films grown by Metal Organic Vapour Phase Epitaxy (MOVPE) [1]. Advantages of
this technology include high optical efficiency, near perfect modulation transfer
function (MTF) and very low dark current. Photodiodes in FPAs usually operate in
the back-side illumination mode.
Depending on the tasks being solved, epitaxial structures can have both single-­
layer and multilayer structures. So, Gu et al. [25] when developing their FPAs used
the structure shown in Fig. 4.5. An x = 0.4 Auger suppressing barrier is grown
between the substrate and absorption layer to reduce Auger injection and thus
4 IR Detectors Array 89

Fig. 4.5 A multilayer structure of HgCdTe material designed and grown for FPAs operated in
3–5 μm spectral range. (Data extracted from Ref. [25])

deduce the dark current. The x = 0.3 Hg1-xCdxTe was grown as the main absorption
layer to ensure that the device is operating with a 5 μm cut off wavelength at 77 K
(mid wavelength infrared, MWIR). The x = 0.4/0.3 double layer heterostructure was
designed to reduce the generation-recombination dark current. Both sides of the
absorption layer are buffered with a graded layer from x = 0.3 to x = 0.4 to improve
the material quality.
When solving some problems, p-n junctions can be formed directly during the
deposition of epitaxial HgCdTe layers. So, when developing an array of dual-band
detectors, the structures shown in Fig. 4.6a were used. Dual-band detectors extend
the single-color process by simply growing a second n-type absorbing layer on top
of the p-type layer to form two back-to-back photodiode junctions, which is referred
to as an n-p-n or triple-layer heterojunction (TLHJ) device structure. A triple-layer
n-p-n heterojunction was grown by molecular-beam epitaxy (MBE) on 100 mm
(211) Si wafers with ZnTe and CdTe buffer layers [50]. The MWIR/LWIR dual
band epitaxial structures had low macro defect densities (<300 cm−2). Inductively
coupled plasma etched detector arrays with 640 × 480 dual band pixels (20 μm)
were mated to dual-band readout integrated circuits (ROICs) to produce FPAs. The
measured 80 K cutoff wavelengths were 5.5 μm for MWIR and 9.4 μm for LWIR,
respectively. The dual-band FPA architecture achieves nearly simultaneous detec­
tion of two spectral bands while still being producible for pixel dimensions between
30 μm and 20 μm, or even smaller. The shorter-wavelength (band 1) radiation is
90 G. Korotcenkov

Fig. 4.6 (a) Cross-section of single-mesa dual-band detector architecture applied to HgCdTe on
Si; (b) SEM image of 20-micron-unit-cell dual-band detectors; (c) A 6 cm × 6 cm molecular-beam
epitaxy (MBE) triple-layer heterojunction (TLHJ) HgCdTe dual-band L/MWIR detector wafer
with a 512 × 512 30-μm-unit-cell detector array product die layout, with test structure detector die
along the upper and lower portions of the wafer. (Reprinted with permission from Ref. [71].
Copyright 2011: Springer)

detected via the p-on-n junction, while the longer-wavelength (band 2) radiation
passes through the shorter-cutoff n-type absorbing layer and can be detected via the
n-on-p junction nearer to the pixel contact. The FPAs exhibited high pixel operabil­
ity in each band, with noise equivalent differential temperature (NEDT) operability
of 99.98% for the MWIR band and 99.6% for the LWIR band at 84 K [50].
To form a multi-element system of contacts, as a rule, multilayer metallization
based on Cr, Au, In and Ni with different layer thicknesses is used. The intermediate
nickel layer gives the contacts the necessary mechanical strength and stability, while
the indium layer provides ductility and elasticity. The last layer of indium with a
thickness of 3–5 micrometers is thermally processed in a vacuum. As a result,
indium contacts are formed in the form of balls 8–10 micrometers in height, pro­
truding above the crystal surface as it is shown in Fig. 4.6b. Indium columnar micro­
contacts up to 10–12 micrometers in height can be formed by other technological
methods. Indium columnar microcontacts are required for cold welding of two crys­
tals, a photodetectors array and a multiplexer. To do this, two crystals are placed in
parallel one above the other, oriented in accordance with the circuit topology, and
with the help of a specially selected value of the mechanical load interconnected via
indium bumps using flip-chip bonding [32]. The combination of many thousands of
indium microcontacts of the photodiode array crystal and the silicon crystal of the
microcircuit can be carried out using an infrared microscope with visualization of
the abutting microcontacts through the silicon substrate. This method is the most
common in the manufacture of hybrid FPAs. However, it is important to note that in
addition to the indium bump technique a hybrid FPAs detectors and multiplexers
can also be fabricated using loophole interconnection [7]. In this case, the detector
and the multiplexer chips are glued together to form a single chip before detector
fabrication. The photovoltaic detector is formed by ion implantation and loopholes
4 IR Detectors Array 91

are drilled by ion milling, and electrical interconnection between each detector and
its corresponding input circuit is made through a small hole formed in each detector.
The loophole interconnection technology offers more stable mechanical and ther­
mal features than flip-chip hybrid architecture. Despite the great laboriousness of
this process, the loophole interconnection technology offers more stable mechanical
and thermal features than flip-chip hybrid architecture. This is the approach used by
Strong et al. [74] when developing FPAs with extremely small pixel sizes.
One of the most important parameters of matrix photodetectors is the magnitude
of the inter-element coupling. Strong coupling can result in significant blurring of
the thermal image. Three types of interconnection are possible: optical, electrical
and photoelectric. The optical coupling is determined by the quality of the optical
path of the thermal imaging device. Photovoltaic is associated with the diffusion of
photogenerated carriers in the common semiconductor layer of the photodiode
array. The voltage drop determines the electrical coupling when the current flows
from the p-n junctions to the inactive contacts, as well as the interconnection
between the input channels of the cooled silicon microcircuit. It is known that the
properties of IR MCT-based photodiodes substantially depend on the chemical
structure and electronic properties of the surface. Therefore, during the formation of
the matrix, as well as during the manufacture of individual photodetectors, the pas­
sivation of the MCT surface is carried out, which helps to reduce surface currents
and prevents the degradation of structures. As a passivating coating, CdTe and ZnS
layers up to 100 nm thick are usually used [4]. These layers are formed using tech­
nological parameters that exclude prolonged temperature treatments. The latter is
necessary to preserve the chemical composition of MCT films, and hence their
properties. For example, these layers can be deposited by thermal evapora­
tion method.
It should be noted that the FPAs manufacturing technology is constantly being
improved. For example, Teledyne has developed a process for removal the CdZnTe
substrate material, and substrate removal is now standard for all NIR, SWIR and
MWIR focal plane arrays made by Teledyne [36]. Substrate removal involves strip­
ping all of the CdZnTe substrate after hybridization to leave a layer of HgCdTe only
7 to 10 μm thick. After removing the substrate, an anti-reflective (AR) coating is
applied to the HgCdTe. Experiment has shown that removing the substrate has a
number of advantages, including the following:
• Increase in sensitivity to visible light, down to 380 nm, with a significant increase
in quantum efficiency below 1.3 μm (see Fig. 4.7a);
• Elimination of fluorescence from cosmic rays absorbed in the CdZnTe substrate,
which is very important for low light level space applications;
• Elimination of Fabry-Perot fringes that can occur in the substrate with narrow
band illumination, such as in spectrometers.
Teledyne has also developed a process to fabricate lightly doped HgCdTe detectors
that can be fully depleted with minimal (1–2 volt) reverse bias [36]. A completely
depleted detector has removed most of the free electrons, which suppresses the dark
Auger current signal to the point where the “dark current” dominates due to the
92 G. Korotcenkov

Fig. 4.7 (a) Quantum Efficiency of a substrate removed HgCdTe detector with 1.7 μm cutoff; (b)
Increase in operating temperature from the use of fully depleted HgCdTe. (Reprinted from Ref.
[36]. Published 2019 by SPIE as open access)

background radiation seen by the detector. Depending on the cut-off wavelength


and operating temperature, the background radiation is 10–100 times lower than the
dark current. This offers a great advantage for space applications, allowing opera­
tion at much higher temperatures with passive cooling or smaller cryo-coolers. This
means that when developing FPAs you can use less expensive cooling options with
longer cooler operating lifetime. Figure 4.7b shows the increase in operating tem­
perature that can be obtained with the use of fully depleted HgCdTe detectors.

4.2.2 Photonic Cooled Detectors

Unfortunately, in order to achieve the desired result in terms of sensitivity and reso­
lution, deep cryogenic cooling is required for the operation of devices with photon
detectors in the IR region, especially in LWIR and VLWIR spectral regions. All
objects in the infrared region of the spectrum are “self-luminous” if their tempera­
ture is above absolute zero, so the IR receivers themselves can “glow” in the range
of their sensitivity (3–5 and 8–14 micrometers), making it difficult to detect weak
radiation coming from outside. Therefore, in order to increase the detecting ability,
it is necessary to extinguish the intrinsic radiation of the sensitive element, adjacent
diaphragms and other elements of the device. This is achieved by cooling the pho­
todetector to temperatures at which the self-radiation noise (dark current) becomes
negligible. In addition, cooling the receiver prevents excessive heating of sensitive
elements with low heat capacity and ensures the stability of the functional proper­
ties of semiconductor elements.
4 IR Detectors Array 93

For deep (cryogenic) cooling of the FPAs (T = 75–80 K), liquid nitrogen or a
cooling machine operating in a closed Split-Stirling cycle are used. For not deep
cooling (T = 150–250 K) or thermal stabilization of the operation of an uncooled
photodetector array, a thermoelectric cooling system based on Peltier elements is
used. A cooling machines operating on a closed Split-Stirling cycle have low power
consumption and dimensions, which allow the cooling element to be placed inside
a thermal imaging camera. Cooling to operating temperature occurs in 5–8 min,
which takes about 3 watts of power.
A typical design of a cooled photon detectors array is shown in Fig. 4.8. A hybrid
photodetector unit, including an array of photosensitive elements and a silicon read­
out integrated circuit, is mounted in a vacuum case. Cooling of FPAs is provided by
a microcryogenic cooling system (MCS) integrated with the FPA housing.

4.2.3 Performances of MWIR and LWIR FPAs

Most thermal imaging cameras operate in the 3–5 μm (MWIR) and 8–14 μm (LWIR)
spectral ranges, which correspond to the atmospheric transparency windows in the
infrared region. Therefore, we will consider their parameters. Among the character­
istics of FPAs, the main ones are:
–– temperature sensitivity NETD (Noise Equivalent Temperature Difference)
–– temperature difference equivalent to noise;
–– the number of pixels that make up the matrix;
–– image acquisition speed;
–– the need for cooling the matrix.

Fig. 4.8 Appearance of FAPs developed by various firms: (a) Scorpio MW (France) (https://www.
lynred.com/products/scorpio-mw-product-range); (b) CD640-12 MW-μ (USA) (https://www.
leonardodrs.com/); (c) ASTROH-640KPT15A810 (Russia) (https://astrohn.ru/product/
astron-­640krt15a810/); (d) PELICAN DLW (Israel) (https://www.scd.co.il/products/pelican-­d-­lw/)
94 G. Korotcenkov

The best QWIPs have NETD below 10 mK, typical −20 mK, medium −35
mK. QWIP-based FPAs are available in 256 × 256, 320 × 240, 320 × 256, 640 × 512
matrix formats [18]. For devices based on MCT detectors, NETD sensitivity: for the
best models −10 mK, typical −15 mK, medium −20 mK; the resolution of the com­
mercially available FPAs is up to 640 × 512 pixels.
Long-wave infrared (LWIR) operation requires cooling to 80 K, while medium-­
wave infrared (MWIR) operation often requires cooling to 120 K. The update rate
for HgCdTe or InSb matrices usually ranges from 100 to 400 Hz; for FPAs based on
QWIPs at full resolution, this frequency is in the range of 50–250 Hz.
For multi-pixel FPAs, aside from the large format, very high level of perfor­
mance is of great importance in quantum efficiency (QE), dark current and noise
[30]. While the noise is mainly addressed by the ROIC input stage, QE and dark
current are highly dependent on the photodiode structure. To obtain a high QE, it is
necessary to optimize the generation and collection of photo carriers: for this pur­
pose, a high pixel fill factor as well as a sufficiently thick absorbing layer are
required. Concerning dark current, at the low operating temperatures considered
here, the dark currents are most likely depletion currents associated with SRH
recombination defects in the space charge region of the photodiode [23, 24]. The
mitigation of such depletion currents implies the use of the best quality materials as
well as the best narrow gap surface passivation.

4.2.3.1 Cooled FPAs for the Spectral Range of 8–12 μm

Cooled FPAs for the spectral range of 8–12 μm are presented in the Table 4.3. They
are serially produced by the world’s leading manufacturers and are widely repre­
sented on the world market of FPAs. The main formats are 320 × 256 and 640 × 512

Table 4.3 FPAs for the spectral range of 8–12 μm of various world manufacturers
Pixel size, Toper,
Country, firm Brand Format μm Technology K
France, Sofradir Mars L 320 × 256 30 HgCdTe/ 80
CdZnTe
Scorpio LW 640 × 512 15 HgCdTe/Ge 80
Sirius LW 640 × 512 20 QWIP 73
England, Harier LW 640 × 512 24 HgCdTe/ 80
Finmeccanica CdZnTe
Hawk LW 640 × 512 16 HgCdTe/ 80
CdZnTe
Germany, AIM – 640 × 512 15 QWIP 70
USA, DRS CD640-12-B 640 × 480 15 HgCdTe/ 80
CdZnTe
China, GST C615M 640 × 512 15 HgCdTe 80
Russia, Astron A-640KRT15A810 640 × 512 15 HgCdTe/Si 78
Source: Data extracted from [73] and [54]
4 IR Detectors Array 95

pixels. Megapixel FPAs in this spectral range are mainly manufactured on the basis
of MCTs and QWIPs. FPAs with a spectral range of 8–12 micrometers provide the
best temperature sensitivity and noise immunity in smoky and dusty conditions. In
this spectral range, there is a maximum of the intrinsic thermal radiation of bodies
at a temperature of 300 K. For example, the human body, by virtue of being at a
temperature of ~300 K, emits radiation that peaks around 10 μm. According to
Wien’s law of displacement, the maximum of the intrinsic thermal radiation of bod­
ies when they are heated shifts to the short-wave region. This is why the spectral
range of 1–5 μm is better suited for detecting more heated bodies.

4.2.3.2 Cooled Photodetectors Array for the Spectral Range of 3–5 μm

Cooled FPAs for the spectral range of 3–5 μm serially produced by leading compa­
nies are presented in the Table 4.4. It can be seen that InSb and MCT are the main
materials used for the development of such devices. It is important to note that for
MCT-based FPAs, the operating temperature can be increased to 110–120 K with­
out degrading performance, ensuring operation in modes limited by background
radiation [75]. For FPAs based on InSb, such an increase in operating temperature
is impossible. This is due both to different coefficients of thermal expansion of the
band gap (Еg) in MCT and InSb (for HgCdTe dEg/dT > 0, and for InSb dEg/dT < 0),
and the initially shorter-wavelength boundary of the photosensitivity of HgCdTe-­
based photodetectors, developed for the region of 3–5 μm. The red border of the
photosensitivity of Hd1-xCdxTe(x = 0.3)-based photodetectors is about 5 μm, and
based on InSb −5.6 μm at T = 80 K. The main matrix format for FPAs developed for

Table 4.4 FPAs for the spectral range of 3–5 μm of various world manufacturers
Country, firm Brand Format Pixel size, μm Technology Toper, K
France, Lynred Jupiter MW 1280 × 1024 15 HgCdTe 80
France, Sofradir Scorpio MW 640 × 512 15 HgCdTe 80
Jupiter MW 1280 × 1024 15 HgCdTe 80
Israel, SCD Pelican MW 640 × 512 15 InSb 80
Black bird 1920 × 1536 10 InSb 80
Hawk MW 640 × 512 16 HgCdTe/CdZnTe 80
Hawk HD 1280 × 1024 8 HgCdTe/CdZnTe 80
Germany, AIM HiPIR 1280 M 1280 × 1024 15 HgCdTe 80
USA, FLIR Neutrino 640 × 512 15 InSb 80
USA, DRS CD640-12-M 640 × 480 12 HgCdTe 80
Korea, i3System 640-15 K-8 640 × 512 15 InSb 80
China, GST C615M 640 × 512 15 T2SL 80
Russia, Orion – 640 × 512 15 InSb 80
Russia, Sapfir – 320 × 256 30 InSb 80
Source: Data extracted from [73] and [54]
96 G. Korotcenkov

3–5 μm range is 640 × 512 pixels. However, a number of companies makes the
transition to 1280 × 1024 and even 1920 × 1080 and 1920 × 1536 formats [79].
The factors that favor the use of IR FPAs developed for the 3–5 μm spectral range
are a greater contrast, more favorable weather conditions, when with an increase in
the water vapor content in the atmosphere, the transmission in the 3–5 μm region
decreases more slowly than in the region of 8–14 μm. As a result, greater transpar­
ency of the atmosphere is achieved in high humidity conditions, and better resolu­
tion due to lower optical diffraction in this spectral range. More information on the
parameters of commercially available FPAs can be found on the websites of the
companies developing these FPAs (http://www.raytheon.com; http://teledynesi.
com/imaging; http://www.sofradir.com; http://www.aim-­ir.com; http://www.scd.
co.il; http://www.flir.com; https://astrohn.ru; https://www.leonardodrs.com; https://
www.lynred.com; https://www.gst-­ir.net; https://orio-­ir.ru/en/; https://www.gst-
­ir.net).

4.3 Trends in FPAs Development

4.3.1 Pixel Size Reduction

The main trend at present is to reduce the weight, size and power consumption of
photoelectronic modules. Decreasing the pixel size and increasing the format is a
general trend for almost all world developers and manufacturers of IC FPAs (see
Figs. 4.9 and 4.10a) [3, 9, 48]. AIM Infrared Modules (Germany), BAE Systems
(USA), Brandywine photonics LLC (USA), CalSensors Inc. (USA), EGIDE USA
(USA), China Germanium Co. Ltd. (China), FLIR Systems (USA), SCD (Israel),
Raytheon Vision Systems (USA), RICOR (Israel), Selex ES (UK), Thales
Cryogenics (France), Lynred (France), Spectrolab Inc. (USA) and others work in
this direction. By decreasing the pixel size, it is possible to increase their overall
number in a given die size, leading to a higher resolution and opening the way for
new applications such as persistent surveillance. In particular, a decrease in the
pitch and an increase in the format leads to a significant increase in the range of
object recognition. On the other hand, maintaining the same number of pixels while
decreasing the pixel size results in a much smaller crystal size, which allows detec­
tors to be manufactured with smaller dimensions, weight, power, and cost (SWaP-C).
Reducing the pixel dimensions can also reduce the size of the optics to support
lower SWaP-C at the system level [14, 49, 57]. Achieving these goals will naturally
require an improvement in the technology for growing high-quality epitaxial MCT
layers with reduced defectiveness.
640 × 512 pixels format of FPAs with a step of 15 μm is currently the main for­
mat of FPAs and, apparently, in terms of price-quality ratio, it will remain so for the
next 5–10 years. As a rule, photodetectors forming the array have a mesa structure
shown in Fig. 4.10b. Such structure helps to reduce the physical area of the p-n
4 IR Detectors Array 97

Fig. 4.9 Focal plane array size progression. Illustrates the timeline and progression of increasing
array format and size with corresponding reduction in pixel size. (https://www.raytheon.com)
(Reprinted with permission from Ref. [72]. Copyright 2016: SPIE)

Fig. 4.10 (a) Dynamics of pixel size reduction for CdHgTe-based APDs. Data extracted from
[45]; (b) Schematic diagram of Selex ES mesa pixel. (Adapted with permission from Ref. [49].
Copyright 2015: SPIE)
98 G. Korotcenkov

junction, and hence to reduce the dark current. As it is seen in Fig. 4.10b, despite its
small size, optical capture remains across the full pixel area and photons, entering
the pixel, are effectively trapped and optically concentrated into the junction and
absorber [49]. However, the leading FPAs developing companies have already
achieved the megapixel format (1280 × 1024 elements) as commercially available
[21, 49]. SOFRADIR (France) already in 2015 offered 10 μm pitch HgCdTe MWIR
diode array [51]. Sofradir is introducing a 10 μm pixel pitch detector in either XGA
(1024 × 768) or HD720 (1280 × 720) formats, named Daphnis. Selex ES Ltd.
(Leonardo, Great Britain) has developed a technology for manufacturing SuperHawk
matrices of megapixel format in the mid-wave infrared range with a step of 8 μm
[49]. Lynred (France) is going to offer matrices with even smaller pitch, 5–7 μm, in
the near future. It is predicted that the pixel size of long-wave HgCdTe infrared
detectors can be reduced to 5 μm, while that of mid-wave HgCdTe infrared detec­
tors can be reduced to 3 μm. Currently, the prices for such photodetectors arrays are
quite high, which does not allow to FPAs developers to make a massive transition to
it. However, many predict that such a transition will take place by 2025 [45, 46].
Although, it must be admitted that this will not be easy to achieve. This task pres­
ents serious challenges associated with detector processing technologies, lower
optical efficiency, larger cross-talk between pixels, interconnect density, and high
charge capacity. According to estimates made by Dhar et al. [20], to achieve high
sensitivity for LWIR FPAs with 5 μm pixels (for example <30 mK) it is required a
large amount of integrated charge to be placed in very small unit cells. For a 5 μm
planar unit cell, the charge capacity in standard ROIC technology is less than one
million electrons, whereas 8 to 12 million electrons are required for good sensitiv­
ity [20].
In spite of these challenges, Teledyne has demonstrated recently small 40 × 40
test HgCdTe LWIR FPAs with pixel size of 5 μm [77]. Interestingly, that these
devices demonstrated significantly better performance than was predicted [78].
Authors attribute this effect to lower doping and thinner active layer in optimized
architecture resulting in reduction of Auger-limited diffusion dark currents. Full
size, 5 μm pitch 1280 × 720 pixel MWIR and LWIR FPAs have also been reported
by researchers from DRS Technologies [5]. Such a small pitch was achieved by
downscaling DRS’ high-density vertically integrated photodiode (HDVIP) architec­
ture (Fig. 4.10b) with interconnects to silicon readout chip obtained by dry etched
and metalized vertical vias. The collection efficiency, cross-talk, and operability are
shown to be similar to larger pitch HDVIP FPAs reported by these authors previ­
ously [74]. They also observed lower dark current levels (factor of 2–4) in compari­
son to those measured in larger, 12 μm and 15 μm pitch FPAs. So there is reason to
believe that these new developments will enable the development of lower cost
FPAs with resolution ranging from high definition (HD) resolution up to many mil­
lions of pixels.
4 IR Detectors Array 99

4.3.2 Quantum-Dimensional Structures

Another important direction in the development of photodetectors arrays is the


development of technologies based on the formation of quantum-dimensional struc­
tures [58] (see also Chaps. 5, 7 and 8, Vol. 2). At present, such photodetectors arrays
are manufactured using wide-gap III-V semiconductor-based heterostructures.
However, due to quantum-size effects, these structures have sensitivity in the mid­
dle or long-wave infrared spectral ranges.
Currently, the most developed quantum- dimensional photodetectors are photo­
resistors based on structures with multiple quantum wells (QWIP), which, depend­
ing on the material used, form type-1 (GaAs /AlGaAs), type-II (InAs / GaSb and
InAs/GaInSb) and type-III (HgTe/CdTe) superlattices. Such quantum-dimensional
structures also include barrier superlattices and structures based on quantum dots. A
description of these structures and photodetectors based on them can be found in the
Chaps. 5 and 6, Vol. 2. The listed quantum-dimension-based photodetectors, includ­
ing those in the form of photodetectors array, are already commercially available.
According to published forecasts [32], in 10–15 years, IR FPAs, including dual-­
range FPAs, with ultimate sensitivity will be available in formats up to 1 K-1 K (or
103 × 103 pixels) and more.

4.3.3 Monolithic and Multispectral FPAs

Development of monolithic IR photodetectors [37, 55, 66, 80, 81] can also be attrib­
uted to the promising direction of FPAs development. Currently, IR photodetectors
arrays are mainly manufactured using hybrid technology, where an MCT-based
photodiode array and a silicon readout integrated circuit (multiplexer) are formed
separately, and then connected element by element using indium bump technique.
However, the hybrid technology has limitations on the number of pixels and the
resistance of the FPA formed to mechanical and thermal influences. Monolithic IR
photodetectors, in which photosensitive elements are formed in MCT layers grown
in the cells of a silicon multiplexer, make it possible to get rid of the above disad­
vantages [38, 63, 82]. Unfortunately, no significant success has yet been achieved in
this direction. However, studies carried out in recent years [13, 16, 28] have shown
that using MCT quantum dots to fabricate photon detectors forming FPAs on the
silicon ROIC surface can facilitate this task. In particular, the QD layers are amor­
phous what permits fabrication of devices directly onto ROIC substrates, as shown
in Fig. 4.11 with no restrictions on pixel or array size and with a short cycle of
production. In addition, the monolithic integration of QD-based detectors into ROIC
does not require any hybridization steps. The individual pixels are determined by
the area of the metal contact pads located on the top of the ROIC surface. Methods
of wet chemistry can be used to synthesize colloidal nanoparticles. This so-called
top-surface photodetector offers a 100% fill factor and is compatible with
100 G. Korotcenkov

Fig. 4.11 IR monolithic array structure based on CQDs. (Reprinted from Ref. [58]. Published
2021 by MDPI as open access)

Fig. 4.12 An example of an image taken at T = 78 K for an army DBFM 640 × 480 M/LWIR FPA
(# 7586704) in MWIR and LWIR spectral ranges with a field of view of f/5 and a frame rate of
30 Hz. (Reprinted with permission from Ref. [56]. Copyright 2005: SPIE)

post-processing at the top of complementary metal-oxide semiconductor (CMOS)


electronics. It is expected that the successful implementation of this new class of
infrared technology may contribute to low-cost CMOS cameras for a wide variety
of applications [58]. It should be noted that these expectations are met. The first
SWIR cameras based on QD thin-film photodiodes, monolithically fabricated on
silicon ROICs, have already been released [34]. The Acuros camera has resolution
1920 × 1080 (2.1 megapixels, 15-μm pixel pitch) and uses 0.4 to 1.7 μm broadband
spectral response.
Development of two-color and multi-spectral focal plane arrays is another prom­
ising direction in the development of thermal imaging technology based on MCT [3,
9]. Almost all leading firms [50, 71] actively develop two-color and multi-spectral
FPAs. Their use in optoelectronic systems increases the likelihood of target detec­
tion and recognition. For example, Fig. 4.12 demonstrates the capabilities of such a
FPAs developed by Raytheon Vision Systems [56]. In this imagery, the subject is
holding a sheet of plastic that transmits in the MWIR band but absorbs in the LWIR
band. The difference in the images demonstrates the multi-spectral capability of
these FPAs. Thus, high information content and compactness of devices are the
4 IR Detectors Array 101

driving forces of the development of this direction. It is believed that in the near
future, dual-spectral IR FPAs will become commercially available.

4.3.4 High Operation Temperature FPAs

Increasing the operating temperatures of FPAs and the development of microcryo­


genic cooling systems for “high-operation temperature” FPAs operated at interme­
diate temperatures (HOT: ~150–300 K), can also be attributed to important
directions in the development of IR technologies in the near future [3, 31, 51, 58,
73]. If the operating temperature of the FPAs increases without degrading image
quality, then smaller coolers can be used and the SWaP and system cost can be
reduced. How does this, along with pixel size reduction, affect the size of FPAs,
developed and fabricated by DRS Network and Imaging Systems [57], shown in
Fig. 4.13.
For example, Shkedy et al. [67] reported that an increase in FPA operating tem­
perature from 77 to 150 K reduced the cooler power consumption from 25 W to
7.5 W. A more significant advantage can be obtained if the operating temperature is
increased to >200 K, where an inexpensive thermoelectric cooler can be imple­
mented. Typically, the increase in operating temperatures of the MWIR and LWIR
infrared detectors try to resolve via reducing dark current, reducing the rate of heat

Fig. 4.13 (a) Progressive decrease in a DRS production 640 × 480 FPA LW package as a function
of pixel pitch, and (b) the reduction in package size for high operating temperatures (HOT).
(Reprinted with permission from Ref. [57]. Copyright 2014: SPIE)
102 G. Korotcenkov

release in the active region and minimizing the active volume of the detector with­
out reducing the quantum efficiency [3, 9, 22, 42, 49, 58, 70]. According to Kinch
[39], in order to achieve this goal, the doping concentration below 5 × 1013 cm−3 is
required. However, despite the sixty-year history of the development of HgCdTe-­
based photodetectors, its ultimate limit of HOT performance, which lies in the range
of 140–300 K, has not been reached. In most publications, FPAs operating tempera­
tures do not exceed 110–120 K [31, 51] and only a few designs allow FPAs to oper­
ate at higher temperatures [67, 79]. For example, Thorne et al. [79] reported that
their MWIR FPAs (640 × 512 array format, 16 μm pixel pitch) can operate at
155 K. To achieve such operating temperature, they used possibilities of the MOVPE
process. Using mesa diode heterostructures (see Fig. 4.10b), the volume of the
absorber was restricted to limit the volume for diffusion current generation. In addi­
tion, the ability to dope the MCT absorber of p-type ensured low Auger recombina­
tion, and the ability to locate the junction in the region with a higher bandgap
reduced thermal generation within the depletion layer. The distance between the p-n
junction and the absorber, as well as the band gap were also monitored. Even more
successes in this direction are expected in the coming years. For example, Kinch
[39] predicted that large area ultra-small pixel diffraction-limited and background-­
limited photon detecting MW and LW HgCdTe FPAs operating at room temperature
will be available within the next ten years. The basis for such predictions is the long
Shockley-Read-Hall (SRР) carrier lifetime in HgCdTe, which makes this material a
great candidate for FPAs operated at 300 K. Simulations carried out by Akhavan
et al. [2] have shown that the new IR detector architecture based on a unipolar nBn
structure can also help reduce the dark current and increase the operating tempera­
ture. The results obtained showed that photodetectors with such a structure could
operate at temperatures above 200 K. Kopytko et al. [43] reported such HgCdTe
nBn detectors with a cut-off wavelength of 3.5 μm.

Acknowledgments This research was funded by the State Program of the Republic of Moldova,
project 20.80009.5007.02.

References

1. Abbott P, Thorne PM, Arthurs CP (2011) Latest detector developments with HgCdTe grown by
MOVPE on GaAs substrates. Proc SPIE 8012:801236
2. Akhavan ND, Jolley G, Umana-Membreno GA, Antoszewski J, Faraone L (2014) Performance
modeling of bandgap engineered HgCdTe-based nBn infrared detectors. IEEE Trans Electron
Dev 61:3691–3698
3. Antoszewski J, Akhavan ND, Umana-Membreno G, Gu R, Lei W, Faraone L (2015) Recent
developments in Mercury Cadmium Telluride IR detector technology. ECS Trans 69(14):61–75
4. Antoszewski J, Musca CA, Dell JM, Faraone L (2003) Small two-dimensional arrays of mid-­
wavelength infrared HgCdTe diodes fabricated by reactive ion-induced p-to-n type conversion.
J Electron Mater 32:627–632
5. Armstrong JM, Skokan MR, Kinch MA, Luttmer JD (2014) HDVIP five-micron pitch HgCdTe
focal plane arrays. Proc SPIE 9070:907033
4 IR Detectors Array 103

6. Baker I, Maxey C, Hipwood L, Weller H, Thorne P (2012) Developments in MOVPE HgCdTe


arrays for passive and active infrared imaging. Proc SPIE 8542:85421A
7. Baker IM, Ballinga RA (1984) Photovoltaic CdHgTe-silicon hybrid focal planes. Proc SPIE
510:121–129
8. Bangs J, Langell M, Reddy M, Melkonian L, Johnson S, Elizondo L et al (2011) Large format
high operability SWIR and MWIR focal plane array performance and capabilities. Proc SPIE
8012:801234
9. Bhan RK, Dhar V (2019) Recent infrared detector technologies, applications, trends and devel­
opment of HgCdTe based cooled infrared focal plane arrays and their characterization. Opto-­
Electron Rev 27(2):174–193
10. Boulade O, Moreau V, Mulet P, Gravrand O, Cervera C, Zanatta J-P et al (2016) Development
activities on NIR large format MCT detectors for astrophysics and space science at CEA and
SOFRADIR. Proc SPIE 9915:99150C
11. Brellier D, Gout E, Gaude G, Pelenc D, Ballet P, Miguet T, Manzato MC (2014) Bulk growth
of CdZnTe: quality improvement and size increase. J Electron Mater 43(8):2901–2907
12. Brill G, Chen Y, Wijewarnasuriya P, Dhar NK (2009) Infrared focal plane array technology uti­
lizing HgCdTe/Si: successes, roadblocks and material improvements. Proc SPIE 7419:74190L
13. Buurma C, Ciani AJ, Pimpinella RE, Feldman JS, Grein CH, Guyot-Sionnes P (2017) Advances
in HgTe colloidal quantum dots for infrared detectors. J Electron Mater 46:6685–6688
14. Caulfield JT, Wilson JA, Dhar NK (2014) Benefits of oversampled small pixel focal plane
arrays. Proc SPIE 9070:907035
15. Cervera C, Boulade O, Gravrand O, Lobre C, Guellec F, Sanson E, Castelein P (2017)
Ultra-low dark current HgCdTe detector in SWIR for space applications. J Electron Mater
46(10):6142–6149
16. Chatterjee A, Babu PN, Jagtap A, Koteswara Rao KSR (2019) Uncooled mid-wave infrared
focal plane array using band gap engineered mercury cadmium telluride quantum dot coated
silicon ROIC. e-J Surf Sci Nanotechnol 17:95–100
17. Chen Y, Farrell S, Brill G, Wijewarnasuriya P, Dhar NK (2008) Dislocation reduction in CdTe/
Si by molecular beam epitaxy through in-situ annealing. J Crystal Growth 310(24):5303–5307
18. Costard E, Bois P, De Rossi A, Nedelcu A, Cocle O, Gauthier F-H, Audier F (2003) QWIP
detectors and thermal imagers. C R Physique 4:1089–1102
19. Delaunay PY, Nosho BZ, Gurga AR, Terterian S, Rajavel RD (2017) Advances in III-V based
dual-band MWIR/LWIR FPAs at HRL. Proc SPIE 10177:101770T
20. Dhar NK, Dat R, Sood AK (2013) Advances in infrared detector array technology. In: Pyshkin
S (ed) Optoelectronics – advanced materials and devices. INTECH, pp 149–190
21. Dorn ML, Piphera JL, McMurtry C, Hartman S, Mainzer A, McKelvey M et al (2016) Proton
irradiation results for long-wave HgCdTe infrared detector arrays for near-earth object camera.
J Astronom Telesc Instrum Syst 2(3):036002
22. D’Souza AI, Robinson E, Ionescu AC, Okerlund D, de Lyon TJ, Rajavel RD et al (2012)
MWIR InAs1-xSbx nCBn detectors data and analysis. Proc SPIE 8353:835333
23. Gravrand O, Rothman J, Cervera C, Baier N, Lobre C, Zanatta JP et al (2016b) HgCdTe detec­
tors for space and science imaging: general issues and latest achievements. J Electron Mater
45(9):4532–4541
24. Gravrand O, Rothman J, Castelein P, Cervera C, Baier N, Lobre C et al (2016a) Latest achieve­
ments on MCT IR detectors for space and science imaging. Proc SPIE 9819:98191W
25. Gu R, Kala H, Antoszewski J, Umana-Membreno G, Dehdashtiakhavan N, Madni I, Faraone
L (2018) Recent advances in IR imaging focal plane arrays technology at UWA. In: Proc. of
the Conference on Optoelectronic and Microelectronic Materials and Devices (COMMAD),
9–13 Dec. 2018. Perth, WA, Australia, 18673702, pp 11–12. https://doi.org/10.1109/
COMMAD.2018.8715245
26. Guellec F, Boulade O, Cervera C, Moreau V, Gravrand O, Rothman J, Zanatta J (2014) ROIC
development at CEA for SWIR detectors: pixel circuit architectures for space applications and
trade-offs, ICSO conference. Proc SPIE 10563:105630K
104 G. Korotcenkov

27. Gunapala SD, Bandara SV, Hill CJ, Ting DZ, Liu JK, Rafol SB et al (2007) Demonstration
of 640 х 512 pixels long-wavelength infrared (LWIR) quantum dot infrared photodetector
(QDIP) imaging focal plane array. Infrared Phys Technol 50:149–155
28. Guyot-Sionnest P, Roberts JA (2015) Background limited mid-infrared photodetection with
photovoltaic HgTe colloidal quantum dots. Appl Phys Lett 107:91115
29. Fièque B, Martineau L, Sanson E, Chorier P, Boulade O, Moreau V, Geoffray H (2011) Infrared
ROIC for very low flux and very low noise applications. Proc SPIE 8176:81761I
30. Fièque B, Lamoure A, Salvetti F, Aufranc S, Gravrand O, Badano G et al (2018) Development
of astronomy large focal plane array “ALFA” at Sofradir and CEA. Proc SPIE 10709:1070905
31. Figgemeier H, Hanna S, Eich D, Mahlein K-M, Fick W, Schirmacher W, Thöt R (2016) State
of the art of AIM LWIR and VLWIR MCT 2D focal plane detector arrays for higher operating
temperatures. Proc SPIE 9819:98191C
32. Filachov AM, Taubkin IL, Trishenkov MA (2015) A review on advances in the solid-state
photoelectronics. Uspehi Prikladnoi Phiziki 3(2):162–168. (in Russian)
33. Finger G, Dorn R, Meyer M, Mehrgan L, Moorwood AFM, Stegmeier J (2006) Interpixel
capacitance in large format CMOS hybrid arrays. Proc SPIE 6276:62760F
34. Hafz SB, Scimeca M, Sahu A, Ko D-K (2019) Colloidal quantum dots for thermal infrared
sensing and imaging. Nano Convergence 6:7
35. Ishimwe R, Abutaleb K, Ahmed F (2014) Applications of thermal imaging in agriculture—a
review. Adv Remote Sens 3:128–140
36. Jerram P, Beletic J (2019) Teledyne’s high performance infrared detectors for space missions.
Proc SPIE 11180:111803D
37. Jiang J, Tsao S, Mi K, Razeghi M, Brown GJ, Jelen C, Tidrow MZ (2005) Advanced mono­
lithic quantum well infrared photodetector focal plane array integrated with silicon readout
integrated circuit. Infr Phys Technol 46:199–207
38. Joshi AM (1998) The next generation of monolithic infrared detector arrays. AIP Conf
Proc 420:67
39. Kinch MA (2014) State-of-the-art infrared detector technology. SPIE Press, Bellingham
40. Kinch MA (2010) HgCdTe: recent trends in the ultimate IR semiconductor. J Electron Mater
39(7):1043–1052
41. Klipstein PC, Avnon E, Azulai D, Benny Y, Fraenkel R, Glozman A et al (2016) Type II super­
lattice technology for LWIR detectors. Proc SPIE 9819:98190T
42. Knowles P, Hipwood L, Pillans L, Ash R, Abbott P (2011) MCT FPAs at high operating tem­
peratures. Proc SPIE 8185:818505
43. Kopytko M, Kębłowski A, Gawron W, Kowalewski A, Rogalski A (2014) MOCVD grown
HgCdTe barrier structures for HOT conditions. IEEE Trans Electron Dev 61(11):3803–3807
44. Kozlov AI (2010) Design features and some implementations of silicon multiplexers for IR
photodetectors. J Opt Technol 77(7):421–428
45. Kulchitsky NA, Naumov AV, Startsev VV (2020a) Infrared focal plane array detectors: “post
pandemic” development trends. Part I. Photonics Russia 14(3):234–244
46. Kulchitsky NA, Naumov AV, Startsev VV (2020b) Infrared focal plane array detectors: “post
pandemic” development trends. Part II. Photonics Russia 14:320–330
47. Lahiri BB, Bagavathiappan S, Jayakumar T, Philip J (2012) Medical applications of infrared
thermography: a review. Infr Phys Technol 55:221–235
48. Liu M, Wang C, Zhou L-Q (2019) Development of small pixel HgCdTe infrared detectors.
Chin Phys B 28(3):037804
49. McEwen RK, Jeckells D, Bains S, Weller H (2015) Developments in reduced pixel geometries
with MOVPE grown MCT arrays. Proc SPIE 9451:94512D
50. Patten EA, Goetz PM, Viela FA, Olsson K, Lofgrren DF, Vodicka JG, Johnson SM (2010)
High-performance MWIR/LWIR dual-band 640 × 480 HgCdTe/Si FPA’s. J Electron Mater
39(10):2215–2219
51. Péré-Laperne N, Rubaldo L, Kerlain A, Carrère E, Dargent L, Taalat R, Berthoz J (2015) 10
μm pitch design of HgCdTe diode array in Sofradir. Proc SPIE 9370:937022
4 IR Detectors Array 105

52. Peric D, Livada B, Peric M, Vujic S (2019) Thermal imager range: predictions, expectations,
and reality. Sensors 19:3313
53. Perrais G, Gravrand O, Baylet J, Destefanis G, Rothman J (2007) Gain and dark current char­
acteristics of planar HgCdTe avalanche photo diodes. J Electron Mater 36(8):963–970
54. Popov V (2020) Modern cooled IR photodetectors. Systemi Bezopasnosti 3:68–70. (in
Russian)
55. Pusino V, Xie C, Khalid A, Steer MJ, Sorel M, Thayne IG, Cumming DRS (2016) InSb photo­
diodes for monolithic active focal plane arrays on GaAs substrates. IEEE Trans Electron Dev
63(8):3135–3142
56. Radford WA, Patten EA, King DF, Pierce GK, Vodicka J, Goetz P, Venzor G et al (2005) Third
generation FPA development status at Raytheon vision systems. Proc SPIE 5783:331–339
57. Robinson J, Kinch M, Marquis M, Littlejohn D, Jeppson K (2014) Case for small pixels: sys­
tem perspectives and FPA challenges. Proc SPIE 9100:91000I
58. Rogalski A, Martyniuk P, Kopytko M, Hu W (2021) Trends in performance limits of the HOT
infrared photodetectors. Appl Sci 11(2):501
59. Rogalski A, Martyniuk P, Kopytko M (2017) InAs/GaSb type-II superlattice infrared detec­
tors: future prospect. Appl Phys Rev 4:031304
60. Rogalski A (2009) Infrared detectors for the future. Acta Phys Polonica A 116(3):389–405
61. Rogalski A (2008) New material systems for third generation infrared photodetectors. Opto-­
Electron Rev 16(4):458–482
62. Rogalski A (2005) HgCdTe infrared detector material: history, status and outlook. Rep Prog
Phys 68:2267–2336
63. Rogalski A (2004) Optical detectors for focal plane arrays. Opto-Electron Rev 12(2):221–245
64. Rogalski A (2000) Infrared detectors at the beginning of the next millennium. Sens Mater
12(5):233–288
65. Szajewska A (2017) Development of the thermal imaging camera (TIC) technology. Procedia
Eng 172:1067–1072
66. Sánchez FJ, Rodrigo MT, Vergara G, Lozano M, Santander J, Torquemada MC et al (2005)
Progress on monolithic integration of cheap IR FPAs of polycrystalline PbSe. Proc SPIE
5783:441–447
67. Shkedy L, Brumer M, Klipstein P, Nitzani M, Avnon E, Kodriano Y et al (2016) Development
of 10μm pitch XBn detector for Low SWaP MWIR applications. Proc SPIE 9819:98191D
68. Singh A, Shukla AK, Pal R (2015) HgCdTe e-avalanche photodiode detector arrays. AIP Adv
5:087172
69. Sizov F (2015) IR-photoelectronics: photon or thermal detectors? Outlooks. Sens Electron
Microelectron Technol 12(1):26–53. (in Russian)
70. Smith KD, Wehner JGA, Graham RW, Randolph JE, Ramirez AM, Venzor GM et al (2012)
High operating temperature mid-wavelength infrared HgCdTe photon trapping focal plane
arrays. Proc SPIE 8353:83532R
71. Smith EPG, Venzor GM, Gallagher AM, Reddy M, Petterson JM, Lofgreen DD, Randolph
JE (2011) Large-format HgCdTe dual-band long-wavelength infrared focal-plane arrays. J
Electron Mater 40(8):1630–1636
72. Starr B, Mears L, Fulk C, Getty J, Beuville E, Boe R et al (2016) RVS large format arrays for
astronomy. Proc SPIE 9915:99152X
73. Startsev V, Naumov A (2018) Modern photodetectors of the infrared spectrum and develop­
ment trends. Tehnologii Zashiti 5:66–70. (in Russian)
74. Strong RL, Kinch MA, Armstrong J (2013) Performance of 12-μm- to 15-μm-pitch MWIR and
LWIR HgCdTe FPAs at elevated temperatures. J Electron Mater 42:3103–3107
75. Sun X, Abshire JB, Krainak MA, Lu W, Beck JD, Sullivan WW III et al (2019) HgCdTe ava­
lanche photodiode array detectors with single photon sensitivity and integrated detector cooler
assemblies for space lidar applications. Opt Eng 58(6):067103
76. Sun X, Abshire JB, Beck JD (2014) HgCdTe e-APD detector arrays with single photon sensi­
tivity for space lidar applications. Proc SPIE 9114:91140K
106 G. Korotcenkov

77. Tennant WE, Gulbransen DJ, Roll A, Carmody M, Edwall D, Julius A et al (2014) Small-pitch
HgCdTe photodetectors. J Electron Mater 43:3041–3046
78. Tennant WE (2012) Interpreting mid-wave infrared MWIR HgCdTe photodetectors. J Prog
Quantum Electron 36:273–292
79. Thorne P, Gordon J, Hipwood LG, Bradford A (2013) 16 Megapixel 12 μm array develop­
ments at Selex ES. Proc SPIE 8704:87042M
80. Xie C, Aziz M, Pusino V, Khalid A, Steer M, Thayne IG, M. Sorel M., Cumming
D.R.S. (2017) Single-chip, mid-infrared array for room temperature video rate imaging.
Optica 4(12):1498–1502
81. Xie C, Pusino V, Khalid A, Aziz M, Steer MJ, Cumming DRS (2016) A new monolithic
approach for mid-IR focal plane arrays. Proc SPIE 9987:99870T
82. Zanio K (1990) HgCdTe on Si for hybrid and monolithic FPAs. Proc SPIE 1308:180–193
83. Zverev AV, Makarov Yu S, Mikhantiev EA, Sabinina IV, Sidorov GY, Dvoretskiy SA (2015)
384 × 288 readout integrated circuit for MWIR and LWIR HgCdTe based FPA. J Phys Conf
Series 643:012055
Chapter 5
New Trends and Approaches
in the Development of Photonic IR
Detector Technology

Ghenadii Korotcenkov and Igor Pronin

5.1 Introduction

As will be shown in subsequent chapters (Chaps. 4 and 15, Vol. 1), HgCdTe is an
expensive material. Therefore, in an attempt to replace this material with cheaper
ones in certain applications, new approaches to the development of new materials
sensitive to infrared radiation have been proposed. At the same time, the task of
increasing operating temperatures and increasing the efficiency of devices was
solved [67, 72]. As a result of these developments, new technologies have appeared
that make it possible to form structures with unique photoelectric properties on the
basis of III-V and II-VI compounds. These include superlattices, barrier photocon-
ductors (so-called barrier structures), and structures with multiple quantum wells
(MQWs) [35, 49]. The technology of using quantum dots (QDs) in the manufacture
of infrared detectors has also received significant development [72, 87].

5.2 High Operating Temperature (HOT) Detectors

One of the main disadvantages of photonic IR photodetectors is the need to cool


them down to low temperatures to achieve the required sensitivity parameters. As a
rule, photodetectors operate at temperatures from 8 to 80 K. Such low temperatures
are necessary to reduce the dark current density, which is high in all detectors with
a narrow band gap. However, this greatly complicates their use and increases the

G. Korotcenkov (*)
Department of Physics and Engineering, Moldova State University, Chisinau, Moldova
e-mail: ghkoro@yahoo.com
I. Pronin
Department of Nano- and Microelectronics, Penza State University, Penza, Russia

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 107
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_5
108 G. Korotcenkov and I. Pronin

cost of the devices being developed. Therefore, the search for solutions that allow
increasing the operating temperatures up to 150–300 K is an important task, the
solution of which gives a greater economic effect. For example, when using a
Stirling cooling motor for cooling, simply raising the operating temperature from
80 K to 150 K results in approximately a half of energy consumption. The cooling
time of the device to the set temperature is also reduced. Higher temperatures also
significantly increase the mean time between failures of the Stirling cooling engine.
A number of concepts to increase the operating temperature of photodetectors
have been proposed [35, 60, 67]. However, all of them, as a rule, are aimed at reduc-
ing the dark current, which can be classified into two groups [39]:
1. inherent mechanisms, which depend only on the intrinsic material properties:
(a) diffusion current due to Auger or radiative recombination in the n-region or
p-region, and
(b) band-to-band tunneling current;
2. defect-related mechanisms, which require surface or bulk defects, located within
the depletion region or within a diffusion length of either side of the deple-
tion region:
(a) diffusion current due to SRH recombination in the n-region or p-region,
(b) generation–recombination within the depletion region,
(c) trap-assisted tunneling, and
(d) surface generation current from surface states.
Therefore, any approaches, which can suppress or eliminate one or more of these
dark current mechanisms, will be helpful in enhancing the operating temperature of
photodetectors and FPAs. In particular, significant improvements have been obtained
by suppression of Auger thermal generation in excluded photoconductors and
extracted photodiodes. As is known, thermal generation and recombination in nar-
row gap semiconductors at near room temperature is determined by the Auger
mechanism [14]. It was established that this problem can be solved using reverse
biased N+-p-P+ photodiodes with lightly doped absorber and non-equilibrium mode
of operation. Under strong depletion, the majority carrier concentration saturates at
the extrinsic level while the concentration of minority carriers is reduced below the
extrinsic level. A device operating in non-equilibrium mode was first proposed by
Ashley and Elliott [4] in the mid-1980s. Ноwever, these non-equilibrium devices
require significant bias currents and exhibit excessive low frequency 1/f noise that
extends up to MHz rang [15, 37].
At low temperatures, the diffusion-limited component of the dark current domi-
nates in the dark current, while at high temperatures; the dark current components
associated with defects dominate in the dark current. As the operating temperature
rises, more and more defects become electrically active, which leads to a significant
increase in the dark current [39]. Therefore, if there is a task to increase the operat-
ing temperatures of the detectors, there is a strong incentive to reduce and eliminate
the effect of defects on the dark current. It would seem that the most reliable way to
5 New Trends and Approaches in the Development of Photonic IR Detector Technology 109

solve this problem is to improve growth and post-growth processing techniques.


However, it turned out that this approach is not optimal. Experiments and simula-
tions have shown that significantly better results can be achieved with new
approaches to band-gap engineering of various compound semiconductors, which
allows the development of new detector architectures that are less sensitive to the
presence of defects. New emerging strategies include barrier structures such as nBn
detectors, low-dimensional structures such as T2SLs, photon trapping detectors,
and multistage/cascade infrared devices [35, 49].

5.3 Quantum Well Infrared Photodetectors

In their principle of operation, quantum well infrared photodetectors (QWIPs) are


fundamentally different from the previously considered bulk detectors [13]. If the
operation of QWIPs is based on quantum-scale physical effects, then bulk detectors
operate on larger scale effects. In 1987, Levine et al. [40] demonstrated the convinc-
ing merit of this approach on the example of structures based on GaAlAs superlat-
tices. It was shown that efficient IR photodetectors for 10 μm spectral range can be
developed on the basis of wide-gap semiconductors. QWIPs in their most basic
form are a periodic repetition of layers of two materials with dissimilar band gaps
(see Fig. 5.1a). A material with a lower band gap is usually called a well layer, and
a material with a higher band gap is called a barrier layer. The well layers are doped
such that without illumination there will be carrier electrons present in the ground
energy state. These carriers are then excited by the incident photons into an energy
state near the edge of the conduction band of the barrier material, where an applied
voltage moves the carrier from the well to the contacts (Fig. 5.1b). In principle,
QWIPs are man-made extrinsic photoconductors, such as Ge:Au or Si:As, in which
quantum wells replace impurity atoms. The ability to vary the binding energy of

Fig. 5.1 (a, b) A band diagram of a quantum well in a QWIP. Carriers are excited by incident
photons from the ground energy state of the well (E1) to the excited energy state of the well (E2).
(c) Photonic mechanisms of excitation of the electron subsystem in a multiple quantum wells. (a)
Reprinted from Ref. [13]. Published 2013 by MDPI as open access. (b, c) (Adapted with permis-
sion from [66]. Copyright 2003: Pergamon)
110 G. Korotcenkov and I. Pronin

electrons in QWIPs to match the desired IR response by changing quantum well


depth and width is an important advantage of such structures.
In the early stages of the development of QWIPs, AlGaAs/InGaAs, InGaAs/
GaInP, InGaAs/InP, GaSb/AlGaSb, and AlGaAs/GaAs were considered as the main
materials suitable for these applications [13]. However, after numerous studies, they
came to the conclusion that the most promising system is AlGaAs/GaAs. It was
found that GaAs/AlGaAs QWIR detectors are the most advances due to the almost
perfect natural lattice match between GaAs and AlGaAs. Besides that, GaA/AlGaAs
quantum well devices were able to utilize standard manufacturing and processing
technologies developed for GaAs. They had high uniformity and well-controlled
growth during molecular beam epitaxy on six-inch GaAs wafers, high yield and
hence low cost, and better thermal and radiation resistance. As a result, the unifor-
mity of GaA/AlGaAs-based QWIP devices is very high, and their performance
approaches theoretical limits. Currently, based on such GaAs/AlGaAs structures,
QWIPs have been developed for the spectral ranges of 3–5 μm and 8–14 μm [85].
When comparing QWIPs with other IR detectors, QWIPs based on III-V com-
pounds have a number of positive and negative performance characteristics [13].
Compared to MCT devices, QWIPs can have lower dark currents, higher detectivity,
and higher NETDs. They use widely applied III-V material processing techniques,
making them easier to fabricate than MCTs-based devices. QWIPs also generally
have higher radiation resistance than narrow band gap materials such as MCT and
InSb. However, there are also significant disadvantages. The most obvious disad-
vantage is associated with the use of QWIP quantum constraint. Due to the limita-
tions imposed by quantum mechanics, QWIP can only absorb light falling on it only
when there is quantum limitation along one of the perpendicular axes. This leads to
a big problem, as due to conventional growing methods, most QWIPs are not able
to absorb normally perpendicular incident light. To mitigate this problem, most
QWIP devices use some form of frontal diffusion filter (optical coupler) that redi-
rects normally incident light so that it can be absorbed.
It should also be noted that the maximum photon energy that QWIP can absorb
is limited by the energy difference between the edges of the conduction band of
materials. This means that there must be a very large difference between the edges
of the band to operate at shorter wavelengths. Finding materials that meet this con-
dition can be extremely difficult. For this reason, QWIPs have rarely been used in
the SWIR spectral range [81]. In addition, the optical cross-section absorption is
limited by the dopant concentrations, leading to a low quantum efficiency. Another
important problem of QWIP-based IR detectors is the relatively high level of ther-
mal excitation at T > 40 K, and especially at T > 70 K. These features of QWIP do
not allow it to compete with MCT-based detectors in this temperature range, as well
as in applications where a short accumulation time is required. However, despite
such a comparison, in the ultra-long-wavelength range (VLWIR) and at low tem-
peratures, the QWIP array demonstrates excellent performance, including in condi-
tions with a low background level.
5 New Trends and Approaches in the Development of Photonic IR Detector Technology 111

5.4 Type-II Strained-Layer Superlattice

Strained-layer superlattice (SLS), while initially appearing to have a similar struc-


ture to QWIPs, actually operates using dramatically different physical principles
[13]. The misleading similarities arise from a superlattice with extremely thin layer
thicknesses (approximately single nanometers) as an active absorbing layer.
However, while QWIPs generally utilize relatively thick barrier layers (approxi-
mately ten of nanometers), all active layers in SLS have a thickness of the same
order of magnitude [64]. Like QWIP, the most common SLS structure is interleav-
ing layers of two different materials, but some later devices utilize more complex
heterostructures [8].
Type-II strained-layer superlattice is usually formed from monolayers of InAs
and GaSb, repeated with a certain period [19, 64], although other combinations are
possible such as InAs/GaInSb, InAsSb/GaSb or InAs/InAsSb [8]. However, the
InAs/GaSb combination turned out to be more promising [13]. The layers can be
interleaved more than 300 times to form an IR absorbing region. These structures
are mainly grown on GaSb (100) substrates. Although the lattice mismatch between
InAs and GaSb is less than 1%, InAs deforms when stretched. These quantum struc-
tures are characterized by a broken band gap with type II band alignment, leading to
spatially indirect transitions between hole states localized in GaSb layers and delo-
calized electronic states in InAs layers (see Fig. 5.2a). An important advantage of
these structures is the ability to vary the band gap, as is done in HgCdTe. The effec-
tive band gap of these structures can be changed from 0.3 eV to values below
0.1 eV. It is important that, in contrast to HgCdTe, this is done not by changing the
composition of the compound, but by changing the thickness and composition of
the layers that form the superlattice. However, the band structure in SLS causes low
overlap integral between the electron and hole wave-functions, and hence a low IR
absorption of such structures. In recent years, it has been proposed to improve the

Fig. 5.2 (a) Energy band diagram of type-II superlattice, showing electronic transition. (b) A band
diagram of a quantum cascade detector. (a) Reprinted from Ref. [30]. Published 2013 by IOP as
open access. (b) Reproduced with permission from (Buffaz et al., State of the art of quantum cas-
cade photodetectors. Proc. SPIE 7660 (2010), 76603Q). Copyright 2010: SPIE
112 G. Korotcenkov and I. Pronin

basic structure of SLS by introducing a very thin (a few angstroms) layer of wider
bandgap material (AlSb) as a barrier to electrons of the main charge carriers. This
opens up great opportunities for varying the band gap and improving performance.
Detectors, using type-II SLS, are photovoltaic devices.
It should be noted that the type-II strain layer superlattice is the most advanced
among the new IR technologies. Successful demonstrations of MWIR imaging
arrays, operating at 120 K with efficiencies close to 40% have been reported. There
are also reports of single pixel LWIR detectors based on type-II strain layer super-
lattice. For example, Rehm et al. [64] reported about InAs/GaSb-based SLS
designed for 8–12 μm (LWIR) spectral range. The main difficulties in this technol-
ogy are the formation of ideal boundaries in the superlattice and passivation of
device. Research has shown that dark current, consisting of bulk leakage and sur-
face leakage, is an important problem in SLS-based photodetectors. The bulk com-
ponent depends on the quality of the formed SLS, while the surface component
depends on the manufacturing conditions of the device itself, including etching and
passivation. In the shorter wavelength region of the spectrum, it appears that due to
the overlap of the band structure of different materials in devices, the effective band
gap between minibands tends to be quite narrow. This tends to make it difficult to
use these devices for SWIR and shorter wavelength applications [13].

5.5 Multi-Stage or Cascade IR Detectors

In an optimally designed photodiode, the sensitivity and diffusion length are closely
related. In particular, increasing the thickness of the absorber does not always lead
to the desired improvement in the signal-to-noise (S/N) ratio. With a thickness of
absorber much greater than the diffusion length, only a limited fraction of the pho-
togenerated charge carriers contributes to the quantum efficiency, since only charge
carriers that are photogenerated at a distance from the junction less than the diffu-
sion length can be collected. This effect is especially pronounced at high tempera-
tures, when the diffusion length usually decreases. This is precisely the situation
that is realized in HOT detectors. For example, calculations using an uncooled 10.6-­
μm HgCdTe photodiode as an example show that the ambipolar diffusion length is
less than 2 μm, and the absorption depth is ~13 μm. This discrepancy reduces the
quantum efficiency to ~15% with a single pass of radiation through the detector
[24]. To solve this problem in PDs operating at temperatures higher than 40–80 K,
quantum cascade infrared detectors (QCIDs) were proposed with a manufacturing
technology identical to that of QWIPs [49]. The only difference was in the number
of base layers. If a period of a QWIP contains only two layers, each period of a
QCID contains many layers (see Fig. 5.2b). The operation of such QCDs is based
on the principles of multi-stage detection. Gendron et al. [20] first demonstrated the
QCID in 2004. Quantum cascade detectors were originally developed as photoelec-
tric detectors, that is, they operate without an applied bias. QCIDs contain several
discrete absorbers that form a series of cascade stages. In such structures with
5 New Trends and Approaches in the Development of Photonic IR Detector Technology 113

carefully selected layer thicknesses incident light excites a carrier from the ground
energy state of the absorbing well to an excited energy state. The carrier then tun-
nels through the barrier layers into the adjacent wells. This continues until the car-
rier reaches the ground state of the next period of the superlattice. Thus, the
photoexcited electrons are transported from one active well to the next one by pho-
non emission through cascaded levels. It is important to note that the thickness of
each step can be less than the diffusion length, while the total thickness of all
absorbers can be comparable or even greater than the diffusion length. As a result,
we were able to drastically reduce the generated dark current and increase the oper-
ating temperatures of the IR detectors [23]. Thus, QCIDs are a good photovoltaic
alternative to QWIPs.
Currently, several options for the implementation of cascade IR detectors have
been developed. They are usually grouped into two main classes: (i) so called inter-
subband (IS) unipolar quantum cascade IR detectors, and (ii) interband (IB) ambi-
polar QCIDs. To describe the performance of IS QCIDs, it is convenient to use the
formalism originally developed for QWIPs detectors [75]. Well-established semi-
conductor material systems are currently available to implement IS QCIDs, such as
InGaAs/AlAsSb (near IR), InGaAs/InAlAs (mid IR), and GaAs/AlGaAs (long IR
up to THz) [9, 25]. These detectors are cryogenically cooled. As for IB QCIDs, one
of the best candidates for realizing such devices is the InAs/GaSb T2SL material
system [45]. These IB QCIDs were able to operate at temperatures up to room tem-
perature [61]. For example, the T2SL cascade detectors have demonstrated high
operating temperatures, up to 400 K, which cannot be achieved with photodetectors
based on the HgCdTe material [49].
Due to the stringent requirements for layer thickness in the structure used in
QWIP and QCID, such devices are usually grown using molecular beam epitaxy
(MBE) [9]. QCIDs are currently being developed for all spectral bands from SWIR
to VLWIR [9].

5.6 Unipolar/Monovalent Barrier IR Detectors

The term “unipolar barrier” was proposed to describe a barrier that can block the
flow of one type of charge carriers (electrons or holes), but does not impede the flow
of carriers of another type (see Fig. 5.3a, b). Currently, many versions of IR detectors
have been proposed that use such barriers to optimize their parameters [33, 34, 49].
Barrier layers were investigated for both the conduction band (nBn structures) and
the valence band (pBp structures), but the most popular are nBn type detectors
[34, 69]. For the first time, Maimon and Wicks [47] proposed detectors based on nBn
structures. In such nBn structures, the n-type semiconductor on one side of the bar-
rier forms a contact layer through which voltage is applied to the device, while the
narrow-gap n-type semiconductor on the other side of the barrier forms a photon
absorbing layer. For efficient collection of photogenerated carriers, the thickness of
this layer must be comparable to the absorption length of light. Usually it is a few
114 G. Korotcenkov and I. Pronin

Fig. 5.3 Schematic energy band diagram of an ideal nBn detector under (a) zero bias and (b)
illumination and low reverse bias V. (с) Schematic energy band diagram of an nBn HgCdTe detec-
tor with high x value Hg1-xCdxTe alloy as the barrier layer. (Reproduced with permission from [39].
Copyright 2015: AIP Publishing LLC)

microns. The same type of doping in the barrier and active layers is the key to main-
taining a low diffusion limited dark current. Essentially, the nBn detector allows
photogenerated holes to flow to a contact (cathode). Thus, the nBn detector operates
as a minority carrier device, selectively blocking electrons in the contact layer by
channeling holes out of the absorber. The optically generated carriers in the absorber
collect at opposite contacts, where collection efficiency can be improved if the device
is operated with a slight bias. At the same time, the dark current of the main charge
carriers associated with the Shockley-Read-Hall (SRH) processes and the surface
leakage current are blocked. As a result, the noise level is sharply reduced. This is
because in the nBnn the depletion region in the active layer is almost completely
absent, where generating processes are usually activated by the SRH mechanism.
For maximum effect, the barrier must be carefully designed. This means that the
material forming the barrier must have good lattice matching with the surrounding
material and have zero offset in one band, and large offset in the other band. It is
believed that the height of the potential barrier should be sufficient (>nkT) to block
thermally induced electrons and thick enough to prevent tunneling. It was found that
barriers with a thickness of more than 100 nm are sufficient for this. In addition, this
barrier should be located near the minority carrier collector and away from the opti-
cal absorption region. This structure can be optimized by adjusting the bias voltage
and doping profiles. Unfortunately, little or no valence band offset was difficult to
realize using standard infrared detector materials such as InSb and HgCdTe. At the
same time, the absence of a depletion region in absorber offers a way for materials
with relatively poor SRH lifetimes, such as all III-V compounds, to overcome such
disadvantage as large depletion dark currents.
The principles of operation of nBn-based detectors are described in detail in the
literature [34, 47, 74]. Although the idea for nBn-based detectors originated in the
development of detectors based on bulk materials (InAs [47]), the greatest success
in its implementation was achieved with the use of materials based on T2SL [65].
Simulation and experiment have shown that the nBn detector offers two important
advantages: (1) nBn-based detector should exhibit a higher signal-to-noise ratio
than a conventional diode operating at the same temperature, and (2) such detectors
5 New Trends and Approaches in the Development of Photonic IR Detector Technology 115

will operate at a higher temperature than a conventional diode with the same dark
current.
A conventional p-n photodiode architecture can also be optimized through the
formation of unipolar barriers in them [73, 74]. For example, placing a barrier in a
p-type layer blocks the surface leakage current. At the same time, currents associ-
ated with diffusion, generation-recombination processes, trap-assisted tunneling
and interband tunneling cannot be blocked [74]. If the barrier is located in the n-type
region, then the currents generated by the p-n junction, as well as surface leakage
currents, are effectively filtered out (see Fig. 5.4a, b).
Experiment has shown that unipolar barriers can significantly improve the per-
formance of infrared p-n photodiodes, especially in low temperature range [74]. For
example, as shown in Fig. 5.4c, an n-side unipolar barrier InAs-based photodiode
(nBp structure) has an R0A value (detector resistance multiplied by active area) in
the low temperature range six orders of magnitude higher than that of a conven-
tional p-n junction.
Despite the fact that the barrier detector can be implemented in different semi-
conductor materials, including InAs [74], InAsSb [33], InAs/GaSb (T2SL) [63] and
HgCdTe [26, 36], the most promising materials for barrier detector structures are
InAs(InAsSb)/B-AlAsSb and InAs/GaSb due to nearly zero valence band offset
(VBO) with respect to AlAsSb barriers. The research results presented in [31] show
how important it is for nBn detectors to have zero VBO. It was shown that in InAsSb
nBn devices the signal-to-noise ratio increases due to a decrease in the valence
band offset.

Fig. 5.4 (a, b) Band diagrams of a p-side (a) and an n-side (b) unipolar photodiode under bias, (c)
R0A product of a conventional InAs photodiode and a comparable n-side barrier photodiode.
(Reprinted with permission from Ref. [74]. Copyright 2013: Elsevier)
116 G. Korotcenkov and I. Pronin

For materials in which large conduction band offset cannot be achieved, the pBn
architecture may be preferable. The same architecture can also be used in devices
that require operation with zero bias [32]. For information, the traditional nBn struc-
ture requires biased device operation. Because of the inevitable changes in the
growth process, most unipolar/monovalent nBn detectors operate at slightly higher
bias voltage than other devices to overcome the unintended barrier that can arise
between the barrier layer and the rest of the device [56].
As for the barrier structures based on HgCdTe, in contrast to III-V compounds,
in the HgCdTe structures with uniform n-type doping, it is impossible to create
conditions under which there is no valence band offset (VBO ≈ 0 eV) between the
absorber and the barrier. Usually VBO < 200 meV (T = 200 K), depending on both
the composition and doping of HgCdTe layers, (see Fig. 5.3c). This feature of is a
key limiting factor in HgCdTe-based barrier detector performance [27]. Such struc-
tures require a relatively high bias (so-called “turn on” voltage) to be applied to the
device to collect the photogenerated carriers. This leads to strong band-to-band tun-
neling (BTB) and trap-assisted tunneling (TAT) effects due to a high electric field at
the barrier-absorber heterojunction. Proper p-type doping of the heterojunctions
should reduce the VBO in HgCdTe-based barriers structures. However, p-type dop-
ing is a technological challenge, associated with the activation of the dopant after
molecular beam epitaxy (MBE) growth. Metalorganic chemical vapor deposition
(MOCVD) is considered more favorable technology because it allows both in situ
donor and acceptor doping. This seems to be more attractive in terms of the growth
of the pBpn and pBpp HgCdTe barrier structures. In these structures, Bp is a p-type
barrier. Barrier structures with p-type-doped constituent layers, grown by MOCVD,
were presented by Kopytko et al. [37]. A further development strategy for HgCdTe
nBn detectors should be focused on reducing or even eliminating the valence band
offset in the barrier layer, which will lead to a lower operating bias, lower dark cur-
rent, and the ability to operate at higher temperatures. Methods for eliminating the
valence band offset have been proposed by Schubert et al. [76] and have been under-
taken for HgCdTe barrier detectors by appropriate bandgap engineering [1].

5.7 HgCdTe-Based Superlattice

Schulman and McGill [78] as an alternative material for infrared detection first
proposed the HgTe-CdTe superlattice in 1979. HgTe-CdTe superlattice has unique
properties, due to the inversion of the conduction and valence bands in the well
materials. This system unites a direct-gap semiconductor with a symmetry-induced
semimetal. That is why HgTe-CdTe superlattice places it in a different class, com-
pared to other II-VI and III-V superlattices – a type-III system. These three different
types of superlattice are illustrated in Fig. 5.5. Type 1 and II are most common
superlattices. They were reviewed in Sects. 5.3 and 5.4 as QWIP and type-II
strained-layer superlattices.
5 New Trends and Approaches in the Development of Photonic IR Detector Technology 117

Fig. 5.5 The conduction and valance band profiles of the three types of superlattices. Holes are
shown as hollow dots, and electrons as black dots

Predictions by Schulman and McGill [50, 77, 78] suggested that this new mate-
rial would have a number of advantages over the HgCdTe compounds, such as: (1)
The cut-off wavelength would be easier to control than the alloy at longer wave-
lengths, (2) Tunneling current prevalent in HgCdTe would be vastly reduced in the
superlattice, and (3) VLWIR operation would be possible without the need for cool-
ing to extremely low temperatures. It is also important that HgTe and CdTe are
nearly lattice-matched to HgCdTe, guaranteeing the epitaxial growth of high quality
HgCdTe heterostructures.
It was shown that in these structures the cut-off wavelength depends on the HgTe
width. An increase in thickness within 1–6 nm leads to an increase in the cut-off
wavelength from 2 to 100 μm. In addition, the HgTe-CdTe superlattice is character-
ised by a large absorption coefficient due to intrinsically high density of states pres-
ent in the semimetal HgTe. The increased absorption coefficient is important
because it allows the use of thinner absorber layers while maintaining high quantum
efficiency. Thinner layers result in less thermal generation of charge carriers in the
detector and therefore lower diffusion dark currents. Another effect of reducing the
thickness of the absorber layer is the reduction of epilayer growth times and simpli-
fied fabrication technology.
Faurie et al. [18] have grew the first HgTe-CdTe superlattice by MBE in 1982.
However, extensive growth, characterization, and theoretical modeling of the HgTe-­
CdTe superlattice only continued from the 1980s to the mid-nineties. As a result of
the studies carried out, it was found that in the manufacture of high-quality HgTe-­
CdTe superlattices, significant difficulties arose associated with surface passivation
and layer interdiffusion at typical processing temperatures. In particular, Zhou et al.
[91] have found that annealing at 225 °C even for 30 min generates large numbers
of defects, presumably mercury vacancies, which leads to a large number of defects,
observed mainly in HgTe layers and, therefore, reduces the 77 K charge carries
mobility in superlattice by two orders of magnitude. This is why, despite the first
118 G. Korotcenkov and I. Pronin

growth of the HgTe-CdTe superlattices 30 years ago, the number of published stud-
ies of IR photodetectors made from the HgTe-CdTe superlattice is very limited.
Only in recent decades, after understanding the basic problems and improving the
technology for growing HgCdTe-based multilayer structures [88], interest in CdTe-­
HgTe superlattices intended for infrared applications, especially in the VLWIR and
FIR spectral regions, was renewed [3, 21, 29, 39, 91]. For example, Zhou et al. [91]
using MBE technology, fabricated HgTe-HgCdTe superlattice-based photodetec-
tors with cut-off wavelength of about 30 μm at 4 K. The superlattice consisted of
100 periods of 8 nm-thick HgTe wells alternating with 7.7 nm-thick Hg0.05Cd0.95Te
barriers. Aleshkin et al. [3] have shown that due to a small probability of the elec-
tron capture into the QWs, the interband HgTe-CdHgTe QWIPs can exhibit very
high gain in photoconductivity. Akhavan et al. [2], based on their calculations also
found that the electron effective mass in the SL absorber is higher than in the
HgCdTe absorber, which results in a lower tunneling dark current. Based on all of
the above, we can agree with Aleshkin et al. [3], that HgTe-HgCdTe superlattice-­
based IR detectors really have significant potential advantages compared to the con-
ventional HgCdTe photodetectors and the III-V heterostructures. However, it is not
clear when it will be possible to fully realize the advantages of HgTe-CdTe superlat-
tices, since due to intense diffusion at the interfaces, only very low-temperature
processes can be used when growing these structures and fabricating devices.

5.8 Quantum Dot Infrared Photodetectors (QDIPs)

Interest in quantum dots (QDs) arose in the 1970s. These are zero dimensional
semiconductor structures with unique optical and electronic properties as a result of
quantum confinement. The development of quantum dots-based infrared photode-
tectors (QDIPs) has actually been stimulated by the advances in quantum wells IR
photodetectors (QWIP). QDIPs are similar to QWIPs, but they have additional ben-
efits due to the 3D constraint in QDs. The detection mechanism in QDIP is also
based on intersubband transitions between quantized QDs energy levels and con-
tinuous states. IR absorption, especially in LWIR spectral range, can also be done
via plasmonic transitions [44]. One of the manufacturing options for single pixel
QDIP is shown in Fig. 5.6. IR technology based on QDs has become a promising
technology for the development of third generation thermal imaging cameras, and
this technology has developed rapidly over the past decade. QDIPs based on II-VI
compounds are considered in details in the Chap. 17, Vol. 2. Prospects for further
development are considered in [44, 72].
QDIPs have three main advantages over QWIPs [87], which are the following:
1. QDIPs are inherently sensitive to infrared radiation at normal incidence due to
violation of the polarization selection rule. This eliminates the need to manufac-
ture a grating coupler on the surface of imaging arrays;
5 New Trends and Approaches in the Development of Photonic IR Detector Technology 119

Fig. 5.6 (a) Schematic of a QD based pin diode device. (b) IR monolithic array structure based on
CQDs. (a) Reprinted from Ref. [30]. Published 2013 by IOP as open access. (b) Reprinted from
Ref. [72]. Published 2021 by MDPI as open access

Table 5.1 Colloidal QD (CQDs) photodetectors advantages and disadvantages in comparison


with single crystal photodetectors
Advantages Disadvantages
Control of dot synthesis and absorption Inferior chemical stability and electronic passivation
spectrum by ability of QD size-filtering, of the nanomaterials in comparison with epitaxial
leading to highly-uniform ensembles; materials;
Much stronger absorption than in Bipolar, interband (or excitonic) transitions across
Stranski-Krastanov grown QDs due to the CQD bandgap (e.g., electrons hopping among
close-packed of CDs; QDs and holes transport through the polymer)
Considerable elimination of strains contrary to the intraband transitions in the epitaxial
influencing the growth of epitaxial QDs QDs;
by better selection of absorber materials; Insulating behaviour due to slow electron transfer
Reduction of cost fabrication (using e.g., through many barrier interfaces in a nanomaterial;
such solution as spin coating, inject Problems with long term stability due to the large
printing, doctor blade or roll-to-roll density of interfaces with atoms presenting different
printing) compared to epitaxial growth; or weaker binding;
Deposition methods are compatible with a High level of 1/f noise due to disordered granular
variety of flexible substrates and sensing systems
technologies such as CMOS (e.g., direct
coating on silicon electronics for imaging)
Source: Reprinted from Ref. [72]. Published 2021 by MDPI as open access

2. QDIPs have a lower dark current than QWIPs due to weaker thermionic emis-
sion from QDs with three-dimensional quantum confinement of carriers, and
3. discrete energy levels in QDs have no dispersion, which reduces phonon scatter-
ing and can lead to an increase in the carrier lifetime (> 100 ps).
The advantages of QDIPs in comparison with QWIPs include also the ability to
work at elevated temperatures, multicolour detection and low cost [30].
QDIPs have already demonstrated MWIR and LWIR imaging. However, QDIPs
currently suffer from lower quantum absorption efficiency compared to interband
type photodetectors [48]. QDIPs have lower absorption quantum efficiency due to
the small fill factor area of QDs and large inhomogeneous broadening of the self-­
assembled QDs [10]. However, if there are applications in which the incident photon
120 G. Korotcenkov and I. Pronin

flux is large, then in such applications QDIPs can achieve the same characteristics as
interband photodetectors due to very low dark current levels. Other advantages and
disadvantages of IR photodetectors based on colloidal QDs are given in Table 5.1.
Colloidal QD PDs are compared to detectors manufactured by epitaxial techniques.

5.9 Multicolor IR Detectors

Thermal radiation from objects made from a mixture of materials can have spectral
variations similar to color in the visible range. Therefore, for reliable interpretation
of the observed image, it is always preferable to have a detector sensitive in several
spectral ranges. As we indicated earlier in Sect. 5.2, detection in various atmo-
spheric windows has its advantages and disadvantages. Multi-range detection allows
to combine these advantages to achieve the best image quality. Multicolour detector
technology is now required in applications such as remote sensing and imaging,
military, and medical imaging [58]. Of course, it is possible to develop FPAs for
these purposes, which contain several types of pixels, with sensitivity in different
spectral ranges. But this significantly complicates the design and manufacturing
technology. Therefore, an approach where the same pixel has these capabilities is
more efficient.
Currently, technologies have already been developed that implement this possi-
bility. These are two-color and three-color array technologies [28, 58]. Their
description in relation to HgCdTe can be found in [5] and Chap. 19, Vol. 2. It is
important to note that QWIP technology is particularly well suited to complex pixel
designs for multi-color photodetectors array (Fig. 5.7b, c). One approach used to
make two-color pixels is shown in Fig. 5.7a. Selective patterning and etching tech-
niques are used to separately contact the two detectors as indicated in Fig. 5.7a.
Various FPAs have already been developed based on two-color detectors [58].
However, such “simultaneous” two colour photovoltaic detector structures suffer

Fig. 5.7 (a) Schematic diagram of the structure of two-color pixel, and (b, c) Examples of HgCdTe
two-color detectors. Note that the high absorption coefficient of HgCdTe in Band 1 limits the
spectral crosstalk from Band 2 to low values. (Reprinted from Ref. [57]. Published 2002 by PAS
as open access)
5 New Trends and Approaches in the Development of Photonic IR Detector Technology 121

Fig. 5.8 (a) Schematic device structure and (b) top-view SEM image of sequential two-colour
detectors; (b) Photoresponse spectra of a sequential LWIR and MWIR two-colour HgCdTe infra-
red detector with cutoff wavelengths of λ1,cutof = 5.5 μm and λ2,cutoff = 10.5 μm. (Adapted with per-
mission from [82]. Copyright 2006: Springer)

from a major disadvantage: a large pixel size due to the requirement of two contacts
per pixel, which limits the pixel density and thus the array format size. Therefore,
the developers are looking for other approaches to solving this problem [39]. One
such alternative approach is the “sequential” approach. Detectors made in accor-
dance with this approach, are also called a bias-selectable detector [82, 83]. Such
detectors have only one contact per pixel, and their configuration does not prevent
pixel size reduction. Figure 5.8 shows the schematic structure of such a detector.
Typical devices include p-n-n-p or n-p-p-n sandwich structures optimized for a spe-
cific wavelength. The wavelength band can be selected by the bias polarity applied
between the two contacts. Such sequential bi-color HgCdTe-based detectors have
been successfully implemented [82, 83]. Smith et al. [82] demonstrated sequential
two-colour (LWIR/MWIR) HgCdTe FPA (256 × 256 pixels) with an MWIR cutoff
wavelength of 5.5 μm and an LWIR cutoff wavelength of 10.5 μm. The spectral
characteristics of these detectors are shown in Fig. 5.8b. This is undoubtedly an
elegant method, but it suffers from the operational disadvantage of non-­simultaneous
integration.
Another approach that multicolor detection can provide is the integration of a
photodetector with a tunable MEMS filter [17, 39, 54], in the development of which
significant progress has been made in recent years [22, 53]. The MEMS optical fil-
ters are electrostatically actuated Fabry–Perot tunable filters. Figure 5.9 shows the
general concept and an example of optical spectral transmission of tunable MEMS-­
based filter [17, 54]. A Fabry-Perot filter requires that its two mirrors be placed
parallel to each other with a spacing (d), which determines the specific wavelengths
transmitted through the filter. This means that by changing the spacing, or optical
cavity length (d), you can control the wavelength of light transmitted through the
filter. In the tunable optical filter design, this change is carried out by applying a
voltage between two electrodes, which generates an electrostatic attraction that
122 G. Korotcenkov and I. Pronin

Fig. 5.9 (a) Schematic concept and (b, c) simulated transmission spectra with different spacing d
for an (b) MWIR and (c) LWIR tunable MEMS optical filters developed for multi-spectral imaging
applications. (Reprinted with permission from Ref. [39]. Copyright 2015: AIP publishing LLC)

moves the upper mirror relatively to the fixed lower mirror. Thus, by controlling the
voltage applied to the MEMS filter, one can control the spectral sensitivity of the IR
detector, which can be conventional broadband multi-color detector.
As for the problem of integrating MEMS filter technology with HgCdTe FPA,
there are two main approaches: hybridization and monolithic integration. Despite
significant progress, the design and manufacture of monolithic integrated filter/
detector structures are very complex and require a lot of effort. In the case of a
hybrid integrated filter/detector, this integration is much easier, since a large-area
tunable MEMS optical filter can be fabricated separately and then hybridized onto
the FPA or a section of the FPA. Currently, work is underway to transfer the existing
technology [22, 53] to a large area tunable filter design to provide wavelength tun-
ing capabilities for large-area FPAs [17, 54]. The realization of adaptive concepts
offers the potential approach to achieving real-time tunable multiband detection.
Despite the great progress made, fabrication of such HgCdTe-based MPAs with tun-
able spectral characteristics is still a very difficult task [39].

5.10 Photon Trapping Detectors

Another approach that makes it possible to reduce dark currents and, therefore, to
increase the operating temperatures of IR detectors is to reduce the volume of the
active region of the detectors [49]. Experiments and theoretical simulations have
shown that this goal can be achieved using the concept of photon trapping (PT). In
this case, a decrease in the dark current is achieved without degrading the quantum
efficiency. However, a decrease in the volume of detectors’ active region is possible
up to a certain limit, after which photon collection begins to decrease faster than the
noise, and, therefore, the overall performance degrades.
Photon trapping detectors have been demonstrated independently in II-VI [84,
90] and III-V [12, 80] based IR detectors. To improve the photon collection, it was
5 New Trends and Approaches in the Development of Photonic IR Detector Technology 123

Fig. 5.10 (a) Examples of photon trapping HgCdTe microstructures; (b) Schematic cross-section
of optically immersed mesa HgCdTe-based photodiode for operation at 200–300 K; (c) SEM pic-
ture of 2D array of 2D immersion lenses photodetectors manufactured by the Vigo Systems. The
size of the individual lens is 50 × 50 μm; (d) schematic diagram of HgCdTe photodiode with GaAs
lens; (e) Schematic diagram of the meta-lens integrated HgCdTe infrared detector. The meta-lens
forms at the top by etching into the CdZnTe substrate. (a) Reprinted with permission from Ref.
[90]. Copyright 2011: Springer. (c, d) Reprinted from Ref. [59]. Published 2004 by PAS as open
access. (e) Reprinted from Ref. [42]. Published 2020 by Nature as open access

proposed to incorporate 3D photonic structures into the IR photodetector. These


structures can be of various shapes. They can have shapes such as pyramidal, sinu-
soidal, or rectangular. An example of such a structure is shown in Fig. 5.10a.
Theoretical estimates have shown that PT photodetector arrays should have signifi-
cantly better device performance than non-PT PD arrays, especially for small pixel
pitches [79]. In addition, PT structures should have superior resolving capability
compared to non-PT structures. Taking into account the general trend towards
decreasing pixel size in FPAs, it is obvious that this technology is an effective means
124 G. Korotcenkov and I. Pronin

of increasing the quantum efficiency of photoconversion without the use of antire-


flection coatings. It is important to note that the results of theoretical modeling have
been experimentally confirmed. As a result of a decrease in the volume of detector’s
active region, both an improvement in device performance and an increase in the
operating temperature of the detector arrays were observed [84]. For example, Dhar
and Dat [12] by reducing the volume of the detector’s active region, observed a
threefold decrease in the dark current without degrading the quantum efficiency
when using the pyramidal structured diodes in comparison with conventional diodes
with the bulk absorber.
As we see in Fig. 5.10a, photon-trapping structures are usually several microns
in size and require a rather complex process to manufacture them. At the same time,
the experiment showed that to improve the absorption of light in detectors with a
reduced area of the active region, simpler designs can be used, such as mesa-­
structures. Figure 5.10b shows a detector with a mesa geometry design. By choos-
ing the correct geometry of the mesa-structure and using reflectors on the sidewalls,
it is possible to achieve that almost all the radiation incident on the bottom surface
will be absorbed by sensing materials. Baker et al. [6], reported that they reached
90% absorption for HgCdTe FPAs with mesa-structure. They have also shown that
such a simple mesa geometry design provides a reliable way to reduce material
volume while maintaining a high level of quantum efficiency for HgCdTe detectors
and, most importantly, the proposed structure is compatible with modern FPAs fab-
rication technology.
It is important to note that the use of the immersion lenses monolithically inte-
grated with photodiode can also contribute to a significant reduction in the volume
of the detector’s active region (see Fig. 5.10c–e). By concentrating the luminous
flux, a significant improvement in the performance enhancement of the detector can
be achieved. This means that the photosensitive area can be significantly reduced to
achieve the required output signal. This significantly reduces dark currents and
noise levels, and, as we indicated earlier, allows the detector to operate at a higher
temperature [42, 59]. For example, Li et al. [42] reported that compared to the pris-
tine device, the integration of the meta-lens together with the reduction in photosen-
sitive area enhances the detectivity (D*) by 3.2–5.5 times.

5.11 Nano Wire-Based Photodetectors

Many believe that the development of nanowires-based IR photodetectors can also


lead to significant performance improvements [87]. Semiconductor nanowires
(NWs) possess unique photonic and electrical properties due to their unique aniso-
tropic geometry and high surface-to-volume ratio. To date, the technology for syn-
thesizing NWs has been developed for almost all semiconductor compounds,
including complex compounds, and therefore NWs with a band gap varying over a
wide range are available at this time. Semiconductor NWs III-V with a narrow band
gap, due to their excellent transport properties and ease of manufacture, are consid-
ered as promising candidates for creating IR photodetectors (PD). In addition, their
5 New Trends and Approaches in the Development of Photonic IR Detector Technology 125

low capacity results in high operating speeds. Recently, there have been reports of
nanowire-based IR PD, such as InAs, GaAsSb, InGaAs, InPAs and InGaSb [41, 46].
For example, it was reported that In0.65Ga0.35As nanowires have demonstrated sensi-
tivity 6.5 × 103 A/W in the range of 1.1–2.0 μm [51], and for InAs NW photodetec-
tors with a Schottky-Ohmic contact, a sensitivity of 5.3 × 103 A/W was recorded at
a wavelength of about 1.5 μm [16]. However, most research related to nanowire-­
based photodetectors is limited to the visible and ultraviolet regions of the spectrum
(read Chap. 16, Vol. 2). In addition, the use of NWs in the manufacture of devices is
accompanied by significant technological difficulties [38] that significantly limit the
possibility of their use in devices intended for the market.

5.12 New Emerging Nanomaterials for Detection

In recent years, interest in so-called 2D materials has grown significantly [68, 71, 72,
87]. Graphene was the first 2D material to attract the attention of many scientists due
to its suitable properties for nanophotonic applications. Due to linear dispersion near
the Dirac point and various forms of interaction of light with substance, graphene has
a high optical sensitivity in a wide spectral range. However, graphene has low
absorption as an IR PD material. This is due to the short carrier lifetime in graphene
and the zero band gap [71]. In other words, the use of graphene-based photodetectors
is limited by the lower external quantum efficiency and photoresponsivity in com-
parison with traditional photodetectors [52]. At the same time, the use of graphene as
an element of photonic photodetectors based on other IR materials, or as a sensitive
element in a thermal detector can improve the parameters of IR photodetectors [11,
86]. For example, encouraging results have been obtained in the development of a
MEMS thermopile with a graphene sensitive layer [11], and when using hybrid pho-
todetectors based on graphene and HgCdTe [7, 72] and graphene-­QDs [70].
Another very interesting family of 2D materials is the single-layer transition
metal dichalcogenides (TMDC) [71, 89], such as molybdenum disulfide (MoS2) and
tungsten diselenide (WSe2). Unlike graphene, the TMDC family has transitions
from an indirect bandgap to a straight bandgap, which occur as the material thick-
ness decreases from multilayer to monolayer. Moreover, TMDCs can be easily
included into a wide variety of atomic-level controlled heterostructures to achieve
higher IR PD performance. However, despite the large amount of funding and
research invested in 2D materials, there is a very limited set of 2D materials, cover-
ing the infrared region [43]. The most studied TMDCs have a band gap in the range
of 1.0–2.5 eV. In addition, these new materials suffer from high environmental sen-
sitivity and large manufacturing area.
Rogalski et al. [72] formulated other limitations that exist in the way of wide-
spread use of 2D nanomaterials in the development of IR detectors. They con-
cluded that:
• In general, the performance of 2D materials-based infrared detectors is lower
compared to commercially available detectors, especially detectors based on
HgCdTe and emerging III-V compounds, including detectors using T2SL;
126 G. Korotcenkov and I. Pronin

• The improvement in responsivity due to the use of a combination of 2D materials


with bulk materials (hybrid photodetectors) owing to the photogating effect
causes the limited linear dynamic range due to the charge relaxation time, which
leads to a decrease in sensitivity with decreasing optical power;
• Responsivity of hybrid and chemically functionalized 2D material photodetectors
is comparable to that of detectors on the world market; however, a significant
decrease in operating speed (bandwidth) is observed; in general, their response time
(millisecond range and longer) is three orders of magnitude longer compared to
commercially available photodetectors (microsecond range and shorter) [70]; and
• The potential for commercialization will depend not only on detector perfor-
mance, but also on whether high quality 2D materials can be produced on a large
scale at a low cost.

5.13 Summary

As a conclusion, Tables 5.2, 5.3 and 5.4, are given, in which various authors [62, 66,
87] summarize the advantages and disadvantages of different approached and mate-
rials used for the development of IR photodetectors, as well as the current state of
their applicability for fabricating large area FPAs.

Table 5.2 Comparison of infrared detectors


Detector type Advantages Disadvantages
Thermal Thermopile, Light, rugged, reliable, and Low detectivity at high
bolometers, pyroelectric low cost; frequency;
Room temperature Slow response (ms order)
operation
Photon
Intrinsic IV-VI (PbS, PbSe, Available low-gap Very high thermal
PbSnTe) materials; Expansion coefficient;
Well studied; Large permittivity
Easier to prepare;
More stable materials
II-VI (HgCdTe) Easy band-gap tailoring; Non-uniformity over large
Well-developed theory and area;
exp.; High cost in growth and
Multicolour detectors processing;
Surface instability
III-V (InGaAs, InAs, Good material and Heteroepitaxy with large
InSb, InAsSb) dopants; lattice mismatch;
Advanced technology; Long wavelength cutoff
Possible monolithic limited to 7 μm (at 77 K)
integration
Photon Si:Ga, Si:As, Ge:Cu, Very long wavelength High thermal generation;
Extrinsic Ge: Hg, Ge:Au operation; Extremely low temperature
Relatively simple operation
technology
(continued)
5 New Trends and Approaches in the Development of Photonic IR Detector Technology 127

Table 5.2 (continued)


Detector type Advantages Disadvantages
Free PtSi-Si, IrSi-Si Low-cost, high yields; Low quantum efficiency;
carriers Large and close packed 2D Low temperature operation
arrays
Quantum wells
Type I GaAs/AlGaAs, InGaAs/ Matured material growth; High thermal generation;
AlGaAs Good uniformity over large Low quantum efficiency;
area; Complicated design and
Multicolour detectors growth
Type II InAs/InGaSb, InAs/ Low auger recombination Complicated design and
InAsSb rate; growth;
Easy wavelength control Sensitive to the interfaces
Quantum InAs/GaAs, InGaAs/ Normal incidence of light Complicated design and
dots InGaP, Ge/Si normal incidence of light; growth
Low thermal generation
Source: Reprinted with permission from Ref. [66]. Copyright 2003: Pergamon Press

Table 5.3 Comparison of IR detectors based on different materials designed for LWIR
spectral range
Bolometer HgCdTe QWIP Type − II SLs QDIP
Status TRL 9 TRL 9 TRL 8 TRL 2–3 TRL 1–2
Designed for Material for Commercial Research and Research and
applications applications development development
requiring requiring high stage stage
medium to low performance
performance
Advantages Low cost; Close to Very uniform Theoretically Not sufficient
requires no theoretical material; uses better than data to
active cooling; performance; commercial HgCdTe at characterize
uses standard basic material manufacturing >14 μm cutoff; material
Si for IR processes; low uses advantages
manufacturing detectors for cost commercial
equipment the next applications III-V
10–15 years. manufacturing
techniques
Military Weapon sight; Missile Being Is at the stage Very early
system night vision intercept; evaluated for of development stages of
examples goggles; tactical some military and assessment development
missile seekers; ground and air applications of the potential
small UAV born imaging; for military use
sensors; hyper spectral
unattended detection;
ground missile
detectors, etc. seeker;
missile
tracking;
space based
sensing, etc.
(continued)
128 G. Korotcenkov and I. Pronin

Table 5.3 (continued)


Bolometer HgCdTe QWIP Type − II SLs QDIP
Limitations Low Performances Narrow Requires a Narrow
sensitivity; are sensitive bandwith; low significant bandwith; low
long time to changes in sensitivity investment and sensitivity
constants manufacturing additional
process; fundamental
difficult to study
expand to
>14 μm cutoff
Source: Data extracted from [55]
TRL technology readiness level. The highest level of TRL (ideal maturity) achieves value of 10

Table 5.4 Summary of the advantages and disadvantages of the current nanostructure-enhanced
IR photodetector
Nanostructure IR Possibility for FPA
PD Advantages Disadvantage fabrication
Quantum well Mature GaAs growth and Low quantum efficiency; Mature technology
structure fabrication process; Cannot absorb normal available for IR FPA
High uniformity; incident angles; production
Low cost; Poor performance at
Covers MWIR to elevated temperatures
WLWIR and THz
Type II High operating High fabrication and Mature technology
superlattice temperature; processing cost; available for IR FPA
structure Covers SWIR to Low yield in large array production
WLWIR; fabrication
High absorption
coefficient
Quantum dot Sensitive to infrared Low absorption quantum Research on going
irradiation at normal efficiency; for IR FPA
incidence; Large inhomogeneous production
Lower dark current than broadening of the
QWIPs; self-assembled QDs
Higher operating
temperature;
Longer carrier lifetime
Nanowires and Higher light sensitivity; Non-uniformity over Possible for IR FPA
nanopillar Antireflection and light large area fabrication; production
trapping properties; Difficult to form large
Can be integrated with detector arrays;
CMOS technology Difficult to fabricate;
Repeatability issues
Graphene-based High internal quantum Weak light absorption in Possible for IR FPA
photodetector efficiency; a single layer; production
Low cost; Zero bandgap
Ease of processing;
Ultrafast process
(continued)
5 New Trends and Approaches in the Development of Photonic IR Detector Technology 129

Table 5.4 (continued)


Nanostructure IR Possibility for FPA
PD Advantages Disadvantage fabrication
Transition metal Low cost; Low speed; Not possible for IR
dichalcogenides High optical absorption Do not easily achieve FPA production
coefficient; LWIR and MWIR;
Ease of processing No proven technology for
large area fabrication
Colloidal Low cost; Relatively low quantum Research on going
quantum dot Scalable for focal plane efficiency; for IR FPA
arrays; Relatively high dark production
Photoconductive gain current
Source: Reprinted from Ref. [87]. Published 2018 by De Gruyter as open access

Acknowledgments This research was funded by the State Program of the Republic of Moldova,
project 20.80009.5007.02.

References

1. Akhavan ND, Jolley G, Umana-Membreno G, Antoszewski J, Faraone L (2014) Performance


modelling of bandgap engineered HgCdTe-based nBn infrared detectors. IEEE Trans Electron
Dev 61(11):3691–3698
2. Akhavan ND, Umana-Membreno GA, Gu R, Antoszewski J, Faraone L (2019) Interdiffusion
effects on bandstructure in HgTe-CdTe superlattices for VLWIR imaging application. J
Electron Mater 48:6159–6168
3. Aleshkin VY, Dubinov AA, Morozov SV, Ryzhii M, Otsuji T, Mitin V et al (2018) Interband
infrared photodetectors based on HgTe–CdHgTe quantum-well heterostructures. Opt Mater
Exprss 8(5):1349–1358
4. Ashley T, Elliot CT (1985) Non-equilibrium devices for infra-red detection. Electron Lett
21(2):451–452
5. Baker IM (2017) II-VI narrow bandgap semiconductors: optoelectronics. In: Kasap S, Capper
P (eds) Springer handbook of electronic and photonic materials. Springer, pp 867–896
6. Baker I, Maxey C, Hipwood L, Weller H, Thorne P (2012) Developments in MOVPE HgCdTe
arrays for passive and active infrared imaging. Proc SPIE 8542:85421A
7. Bansal S, Das A, Jain P, Prakash K, Sharma K, Kumar N et al (2019) Enhanced optoelectronic
properties of bilayer graphene/HgCdTe-based single- and dual-junction photodetectors in long
infrared regime. IEEE Trans Nanotechnol 18:781–789
8. Brown GJ (2005) Type-II InAs/GaInSb superlattices for infrared detection: an overview. Proc
SPIE 5783:65–77
9. Buffaz A, Carras M, Doyennette L, Nedelcu A, Bois P, Berger V (2010) State of the art of
quantum cascade photodetectors. Proc SPIE 7660:76603Q
10. Chakrabarti S, Stiff-Roberts A, Su X, Bhattacharya P, Ariyawansa G, Perera A (2005)
High-performance mid-infrared quantum dot infrared photodetectors. J Phys D Appl Phys
38:2135–2141
11. Chen S-J, Chen B (2020) Research on a CMOS-MEMS infrared sensor with reduced graphene
oxide. Sensors 20:4007
12. Dhar NK, Dat R (2012) Advanced imaging research and development at DARPA. Proc SPIE
8353:835302
130 G. Korotcenkov and I. Pronin

13. Downs C, Vandervelde TE (2013) Progress in infrared photodetectors since 2000. Sensors
13:5054–5098
14. Elliott CT (1990) Non-equilibrium modes of operation of narrow-gap semiconductor devices.
Semicond Sci Technol 5(1990):S30–S37
15. Elliott CT (2001) Photoconductive and non-equilibrium devices in HgCdTe and related alloys.
In: Capper P, Elliott CT (eds) Infrared detectors and emitters: materials and devices. Kluwer
Academic Publishers, Boston, pp 279–312
16. Fang H, Hu W, Wang P, Guo N, Luo W, Zheng D et al. (2016) Visible light-assisted high-­
performance mid-infrared photodetectors based on single InAs nanowire. Nano Lett
16(10):6416–6424
17. Faraone L (2005) MEMS for tunable multi-spectral infrared sensor arrays. Proc SPIE
5957:59570F
18. Faurie JP, Million A, Piaguet J (1982) CdTe-HgTe multilayer grown by molecular beam epi-
taxy. Appl Phys Lett 41:713–715
19. Gautam N, Myers S, Barve AV, Klein B, Smith EP, Rhiger DR et al (2013) Barrier engi-
neered infrared photodetectors based on type-II InAs/GaSb strained layer superlattices. IEEE
J Quantum Electron 49:211–217
20. Gendron L, Carras M, Huynh A, Ortiz V, Koeniguer C, Berger V (2004) Quantum cascade
photodetector. Appl Phys Lett 85(14):2824–2826
21. Grein CH, Jung H, Singh R, Flatte ME (2005) Comparison of normal and inverted band struc-
ture HgTe/CdTe superlattice for very long wavelength infrared detectors. J Electron Mater
34(6):905–908
22. Gunning W, Lauxtermann S, Durmas H, Xu M, Stupar P, Borwick R et al (2009) MEMS-based
tunable filters for compact IR spectral imaging. Proc SPIE 7298:72982I
23. Hinds S, Buchanan M, Dubek R, Haffouz S, Laframboise S, Wasilewski Z et al (2011) Near-­
room-­temperature mid-infrared quantum well photodetector. Adv Mater 23(46):5536–5539
24. Hinkey RT, Yang RQ (2013) Theory of multiple-stage interband photovoltaic devices and
ultimate performance limit comparison of multiple-stage and single-stage interband infrared
detectors. J Appl Phys 114:104506-1–104506-18
25. Hofstetter D, Giorgetta FR, Baumann E, Yang Q, Manz C, Kohler K (2010) Mid-infrared
quantum cascade detectors for applications in spectroscopy and pyrometry. Appl Phys B
Lasers Opt 100:313–320
26. Itsuno AM, Philips JD, Velicu S (2011) Design and modeling of HgCdTe nBn detectors. J
Electron Mater 40:1624–1629
27. Itsuno AM, Phillips JD, Velicu S (2012) Mid-wave infrared HgCdTe nBn photodetector. Appl
Phys Lett 100:161102
28. Jozwikowski K, Rogalski A (2007) Numerical analysis of three-colour HgCdTe detectors.
Opto-Electron Rev 15(4):215–222
29. Jung HS, Boieriu P, Greain CH (2006) P-type HgTe/CdTe superlattices for very-long wave-
length infrared detectors. J Electron Mater 35:1341–1345
30. Karim A, Andersson JY (2013) Infrared detectors: advances, challenges and new technologies.
IOP Conf Ser Mater Sci Eng 51:012001
31. Khoshakhlagh A, Myers S, Plis E, Kutty MN, Klein B, Gautam N et al (2010) Mid-wavelength
InAsSb detectors based on nBn design. Proc SPIE 7660:76602Z
32. Klem JF, Kim JK, Cich MJ, Hawkins SD, Fortune TR, Rienstra JL (2010) Comparison of nBn
and nBp mid-wave barrier infrared photodetectors. Proc SPIE 7608:76081P
33. Klipstein P (2008) “XBn” barrier photodetectors for high sensitivity and high operating tem-
perature infrared sensors. Proc SPIE 6940:69402U
34. Kopytko M (2014) Design and modeling of high-operating temperature MWIR HgCdTe nBn
detector with n- and p-type barriers. Infrared Phys Technol 64:47–55
35. Kopytko M, Martyniuk P (2016) HgCdTe mid- and long-wave barrier infrared detectors for
higher operating temperature condition. In: Beg A, Akbar NS (eds) Modeling and simulation
in engineering sciences. Intech, London, pp 72–90
5 New Trends and Approaches in the Development of Photonic IR Detector Technology 131

36. Kopytko M, Kebłowski A, Gawron W, Madejczyk P, Kowalewski A, Jozwikowski K (2013)


High-operating temperature MWIR nBn HgCdTe detector grown by MOCVD. Opto-Electron
Rev 21(42):402–405
37. Kopytko M, Jóźwikowski K, Rogalski A (2014) Fundamental limits of MWIR HgCdTe barrier
detectors operating under non-equilibrium mode. Solid State Electron 100:20–26
38. Korotcenkov G (2020) Current trends in nanomaterials for metal oxide-based conduc-
tometric gas sensors: advantages and limitations. Part 1: 1D and 2D nanostructures.
Nanomaterials 10:1392
39. Lei W, Antoszewski J, Faraone L (2015) Progress, challenges, and opportunities for HgCdTe
infrared materials and detectors. Appl Phys Rev 2:041303
40. Levine BF, Choi KK, Bethea CG, Walker J, Malik R (1987) New 10 um infrared detector
using intersubband absorption in resonant tunneling GaAlAs superlattices. Appl Phys Lett
50:1092–1094
41. Li Z, Yuan X, Fu L, Peng K, Wang F, Fu X et al (2015) Room temperature GaAsSb single
nanowire infrared photodetectors. Nanotechnology 26:445202
42. Li F, Deng J, Zhou J, Chu Z, Yu Y, Dai X et al (2020) HgCdTe mid-infrared photo response
enhanced by monolithically integrated meta-lenses. Sci Rep 10:6472
43. Liu F, Shimotani H, Shang H, Kanagasekara T, Zólyomi V, Drummond N et al (2013) High-­
sensitivity photodetectors based on multilayer GaTe flakes. ACS Nano 8:752–760
44. Livache C, Martinez B, Goubet N, Ramade J, Lhuiller E (2018) Road map for nanocrystal
based infrared photodetectors. Front Chem 6:575
45. Lotfi H, Hinkey RT, Li L, Yang RQ, Klem JF, Johnson MB (2013) Narrow-bandgap photo-
voltaic devices operating at room temperature and above with high open-circuit voltage. Appl
Phys Lett 102:211103
46. Ma L, Hu W, Zhang Q, Ren P, Zhuang X (2014) Room-temperature near-infrared photodetec-
tors based on single heterojunction nanowires. Nano Lett 14:694–698
47. Maimon S, Wicks G (2006) nBn detector, an infrared detector with reduced dark current and
higher operating temperature. Appl Phys Lett 89:151109
48. Martyniuk P, Rogalski A (2008) Quantum−dot infrared photodetectors: status and outlook.
Prog Quantum Electron 32:89–120
49. Martyniuk P, Antoszewski J, Martyniuk M, Faraone L, Rogalski A (2014) New concepts in
infrared photodetector designs. Appl Phys Rev 1:041102
50. McGill TC, Wu GY, Hetzler SR (1986) Superlattice: progress and prospects. J Vac Sci Technol
A 4(4):2091–2095
51. Miao J, Hu W, Gu N, Lu Z, Zhou X, Liao L et al (2014) Single InAs nanowire room-­temperature
near-infrared photodetectors. ACS Nano 8:3628–3635
52. Mueller T, Xia F, Avouris P (2010) Graphene photodetectors for high speed optical communi-
cations. Nat Photon 4:297–301
53. Musca CA, Antoszewski J, Winchester KJ, Keating AJ, Nguyen T, Silva KKMBD et al (2005)
Monolithic integration of an infrared photon detector with a MEMS-based tunable filter. IEEE
Electron Dev Lett 26:888–890
54. Musca CA, Antoszewski J, Keating AJ, Winchester KJ, Silva KKMBD, Nguyen T et al (2007)
MEMS-based micro-spectrometers for infrared sensing. In: Proceedings of 2007 IEEE/LEOS
international conference on optical MEMS and nanophotonics, Hualien, 12–16 August 2007,
pp 137–138
55. National Research Council (NRC) (2010) Seeing photons: Progress and limits of visible and
infrared sensor arrays. The National Academies Press, Washington, DC. http://www.nap.edu/
catalog/12896.html
56. Nguyen B-M, Chen G, Hoang AM, Abdollahi Pour S, Bogdanov S, Razeghi M (2011) Effect
of contact doping in superlattice-based minority carrier unipolar detectors. Appl Phys Lett
99:033501
57. Norton P (2002) HgCdTe infrared detectors. Opto-Electron Rev 10(3):159–174
58. Norton PR (2006) Third-generation sensors for night vision. Opto-Electron Rev 14:283–296
132 G. Korotcenkov and I. Pronin

59. Piotrowski J (2004) Uncooled operation of IR photodetectors. Opto-Electron Rev


12(1):111–122
60. Piotrowski J, Rogalski A (2007) High-operating-temperature infrared photodetectors. SPIE,
Bellingham
61. Pusz W, Kowalewski A, Gawron W, Plis E, Krishna S, Rogalski A (2013) MWIR type-II InAs/
GaSb superllatice interband cascade photodetectors. Proc SPIE 8868:88680M
62. Razeghi M (1998) Current status and future trends of infrared detectors. Opto-Electron Rev
6:155–194
63. Razeghi M, Pour SA, Huang EK-W, Chen G, Haddadi B-MN et al (2011) High-operating
temperature MWIR photon detectors based on Type II InAs/GaSb superlattice. Proc SPIE
8012:80122Q
64. Rehm R, Masur M, Schmitz J, Daumer V, Niemasz J, Vandervelde T et al (2013) InAs/GaSb
superlattice infrared detectors. Infrared Phys Technol 59:6–11
65. Rodriguez JB, Plis E, Bishop G, Sharma YD, Kim H, Dawson LR, Krishna S (2007) nBn
structure based on InAs/GaSb type-II strained layer superlattices. Appl Phys Lett 91:043514
66. Rogalski A (2003) Infrared detectors: status and trends. Prog Quant Electron 27:59–210
67. Rogalski A (2009) Infrared detectors for the future. Acta Phys Pol A 116(3):389–405
68. Rogalski A (2019) Graphene-based materials in the infrared and terahertz detector families: a
tutorial. Adv Opt Photon 11(2):314–379
69. Rogalski A, Martyniuk P (2014) Mid-wavelength infrared nBn for HOT detectors. J Electron
Mater 43(8):2963–2969
70. Rogalski A, Kopytko M, Martyniuk P (2019) Two dimensional infrared and terahertz detec-
tors: outlook and status. Appl Phys Rev 6:021316
71. Rogalski A, Kopytko M, Martyniuk P (2020) 2D material infrared and terahertz detectors:
status and outlook. Opto-Electron Rev 28:107–154
72. Rogalski A, Martyniuk P, Kopytko M, Hu W (2021) Trends in performance limits of the HOT
infrared photodetectors. Appl Sci 11:501
73. Savich GR, Pedrazzani JR, Sidor DE, Maimon S, Wicks GW (2011) Dark current filtering in
unipolar barrier infrared detectors. Appl Phys Lett 99:121112
74. Savich GR, Pedrazzani JR, Sidor DE, Wicks GW (2013) Benefits and limitations of unipolar
barriers in infrared photodetectors. Infrared Phys Technol 59:152–155
75. Schneider H, Liu HC (2007) Quantum well infrared photodetectors. Springer, Berlin
76. Schubert EF, Tu LW, Zydzik GJ, Kopf RF, Benvenuti A, Pinto MR (1992) Elimination of het-
erojunction band discontinuities by modulation doping. Appl Phys Lett 60:466–468
77. Schulman JN, Chang Y-C (1985) Hg-CdTe superlattice bang-gap enhancement due to interdif-
fusion. Appl Phys Lett 46(6):571–573
78. Schulman JN, McGill TC (1979) The CdTe/HgTe superlattice: proposal for a new infrared
material. Appl Phys Lett 34(10):663–665
79. Schuster J, Bellotti E (2013) Numerical simulation of crosstalk in reduced pitch HgCdTe
photon-­trapping structure pixel arrays. Opt Express 21(12):14712
80. Sharifi H, Roebuck M, De Lyon T, Nguyen H, Cline M, Chang D et al (2013) Fabrication of
high operating temperature (HOT), visible to MWIR, nCBn photon-trap detector arrays. Proc
SPIE 8704:87041U
81. Sherliker B, Halsall M, Kasalynas I, Seliuta D, Valusis G, Vengris M et al (2007) Room tem-
perature operation of AlGaN/GaN quantum well infrared photodetectors at a 3–4 μm wave-
length range. Semicond Sci Technol 22:1240–1244
82. Smith EPG, Patten EA, Goetz PM, Venzor GM, Roth JA, Nosho BZ et al (2006) Fabrication
and characterization of two-color midwavelength/long wavelength HgCdTe infrared detectors.
J Electron Mater 35:1145–1152
83. Smith EPG, Venzor GM, Gallagher AM, Reddy M, Peterson JM, Lofgreen DD, Randolph
JE (2011) Large-format HgCdTe dual-band long-wavelength infrared focal-plane arrays. J
Electron Mater 40:1630–1636
5 New Trends and Approaches in the Development of Photonic IR Detector Technology 133

84. Smith KD, Wehner JGA, Graham RW, Randolph JE, Ramirez AM, Venzor GM et al (2012)
High operating temperature mid-wavelength infrared HgCdTe photon trapping focal plane
arrays. Proc SPIE 8353:83532R
85. Smuk S, Kochanov Y, Petroshenko MP, Solomitskii D (2014) IRnova long-wavelength infra-
red sensors based on quantum wells. Komponenti Tehnologia 1:20–25. (in Russian)
86. Sood AK, Zeller JW, Ghuman P, Babu S, Dhar NK, Ganguly S et al (2019) Development of
high-performance detector technology for UV and IR applications. Proc SPIE 11151:1115113
87. Tan CL, Mohseni H (2018) Emerging technologies for high performance infrared detectors.
Nano 7(1):169–197
88. Wang C, Wang X, Zhao J, Chang Y, Grein CH, Sivananthan S, Smith DJ (2007) Microstructure
of interfacial HgTe/CdTe superlattice layers for growth of HgCdTe on CdZnTe (211)B sub-
strates. J Crystal Growth 309:153–157
89. Wang QH, Kalantar-Zadeh K, Kis A, Coleman JN, Strano MS (2012) Electronics and opto-
electronics of two-dimensional transition metal dichalcogenides. Nat Nanotechnol 7:699–712
90. Wehner JGA, Smith EPG, Venzor GM, Smith KD, Ramirez AM, Kolasa BP et al (2011)
HgCdTe photon trapping structure for broadband mid-wavelength infrared absorption. J
Electron Mater 40:1840–1846
91. Zhou YD, Becker CR, Selamet Y, Chang Y, Ashokan R, Boreiko RT et al (2003) Far-infrared
detector based on HgTe/HgCdTe superlattices. J Electron Mater 32:608–614
Chapter 6
II-VI Semiconductor-Based Unipolar
Barrier Structures for Infrared
Photodetector Arrays

A. V. Voitsekhovskii, S. N. Nesmelov, S. M. Dzyadukh, D. I. Gorn,


S. A. Dvoretsky, N. N. Mikhailov, G. Y. Sidorov, and M. V. Yakushev

6.1 Introduction

The rapid development of thermal imaging technology requires a radical improvement


in the technology of infrared photodetectors in the mid wave (MWIR, 3–5 μm) and
long wave (LWIR, 8–14 μm) regions of the infrared (IR) range. Today, there is an
urgent need to develop MWIR and LWIR array photodetectors of the third generation,
which are subject to increased requirements for photosensitive elements, in particu-
lar, for operating temperatures, weight, dimensions, and power consumption.
One of the main ways to improve the performance of such photosensitive device
structures is to increase the operating temperature of cooled photosensitive layer in
the photodetectors without losing temperature sensitivity and infrared image qual-
ity. An increase in the cooling temperature makes it possible to use microcryogenic
systems with significantly reduced weight, dimensions and power consumption,
making them cheaper. This trend is directly related to the development and imple-
mentation of new photosensitive semiconductor structures that provide low dark
currents and, as a result, low intrinsic noise. This is achieved through the creation of
semiconductor heterostructures by epitaxial methods, which make it possible to
grow photosensitive structures with a complex architecture of layers – with shut-off,
buffer, barrier and other functional layers. The use of zone engineering methods in
theory makes it possible to eliminate individual components of dark currents and, as
a result, to achieve a significant increase in the photosensitive characteristics of

A. V. Voitsekhovskii · S. N. Nesmelov · S. M. Dzyadukh · D. I. Gorn (*)


National Research Tomsk State University, Tomsk, Russia
e-mail: vav43@mail.tsu.ru; gorn.di@gmail.com
S. A. Dvoretsky · N. N. Mikhailov · G. Y. Sidorov · M. V. Yakushev
Rzhanov Institute of Semiconductor Physics of the Siberian Branch of the RAS,
Novosibirsk, Russia

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 135
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_6
136 A. V. Voitsekhovskii et al.

photodetectors [1, 2]. Currently, work on the creation of high operating temperature
focal plane array (HOT FPA) is being actively carried out by leading manufacturers
of optoelectronic equipment from France, USA, Germany, Great Britain, Israel,
China and other countries.
In the field of creating HOT FPA, two main directions can be distinguished. The
first is the development of p-n photodiode arrays based on high-quality HgCdTe
(MCT) heteroepitaxial structures with p on n architecture (local regions of p-type
conductivity are formed in the base active layer of the electronic type of conductiv-
ity by external doping with an acceptor impurity, for example, arsenic) or n+ on p
(the p-type base active layer is vacancy-doped, and n+ is obtained by radiation dop-
ing, for example, with boron ions) [1, 3]. The second direction is the use of so-called
unipolar xBn barrier structures, where x is a contact semiconductor layer of n- or
p-type conductivity, B is a barrier layer, and n is an absorbing layer of n-type of
conductivity. To date, the most studied are unipolar barrier structures in the nBn
configuration.
At present, research and development of nBn structures for FPAs are carried out
both on the basis of III-V and II-VI materials. In the case of II-VI materials, a semi-
conducting HgCdTe solid solution is used. From a fundamental point of view, MCT
is an ideal material for creating IR detectors. This is due, firstly, to the dependence
of the band gap on the CdTe content, secondly, to large optical absorption coeffi-
cients, which leads to high quantum efficiencies, thirdly, to recombination mecha-
nisms that provide long lifetimes of charge carriers and a relatively high operating
cooling temperature, and, fourthly, an extremely weak dependence of the lattice
constant on the composition, which makes it possible to grow high-quality multi-
layer heterostructures. Due to the listed properties, MCT is widely used in the
development of highly sensitive IR detectors for various spectral regions [2].
This chapter will review the latest advances in the development and fabrication
of unipolar barrier structures based on HgCdTe.

6.2 Basics of Barrier Detectors Based


on II-VI Semiconductors

In 2006, Maimon and Wicks [4] proposed the concept of the so-called unipolar bar-
rier nBn photosensitive structure, which is often structurally compared with the
classical p-n-photodiode, in which the space charge region of the p-n junction is
replaced by a wide-gap barrier B, and the p-region is replaced by n-type contact.
The second n-layer plays the role of an active absorbing region. Due to the introduc-
tion of a wide-gap barrier at a negative bias of the structure (when a negative poten-
tial is applied to the contact layer), a potential barrier is created for the main charge
carriers (electrons in the case of the nBn structure) and the dark currents caused by
the main carriers are suppressed. But this does not create a potential barrier for the
photogenerated minority charge carriers [5]. Thus, due to the exclusion of charge
6 II-VI Semiconductor-Based Unipolar Barrier Structures for Infrared Photodetector… 137

carriers of the same sign from the current transfer process and the suppression of the
generation-recombination process, a decrease in the dark currents of the photode-
tector is achieved. The introduction of a wide gap barrier instead of the space charge
region of the p-n photodiode also makes it possible to reduce the contribution of the
Shockley-Read-Hall generation-recombination mechanism, as well as the surface
leakage mechanism, to the dark current.
The characteristic of the dark current of a photodiode with a p-n junction has two
segments with different slopes, which represent the diffusion and generation-­
recombination components of the dark current. Figure 6.1 shows the dependence
log(I) = f(1/T) of a photodiode with p-n junction and nBn structure. The slope
change point corresponds to the temperature Tc at which the diffusion and generation-­
recombination components of the dark current are equal. In the nBn-type structure,
there is no generation-recombination current; therefore, the dark current consists
only of a diffusion component. Consequently, at the same temperatures, the dark
currents of the nBn structure will be significantly less than that of a standard photo-
diode with p-n junction, and the photoelectric parameters will be much higher. By
eliminating the generation-recombination component of the dark current in barrier
nBn structures, it is possible to increase the threshold photoelectric parameters at
the cryogenic temperature of 77 K or significantly increase the cooling temperature
of photosensitive elements to 100 K and higher, while ensuring dark currents less
than the background current. The noise equivalent temperature difference (NETD)
characterizing the temperature sensitivity will be the same as at 77 K.

Fig. 6.1 Temperature dependence of the dark current log(I) = f(1/T) for a photodiode with p-n
junction and for nBn structure. (Reprinted from [6]. Published 2014 by Springer Nature as
open access)
138 A. V. Voitsekhovskii et al.

FPAs for MWIR and LWIR spectral ranges usually have a hybrid architecture: a
matrix of sensitive elements based on p-n or n-p photodiodes made of HgCdTe is
coupled with a silicon readout circuit. The operating temperatures of the FPAs, at
which the mode of limiting the threshold characteristics by background radiation
noise is realized, are dictated by the generation-recombination mechanisms in
HgCdTe, which determine the magnitude of dark currents (noise). Dark currents are
suppressed by cooling the sensitive elements of photodetectors based on HgCdTe to
sufficiently low temperatures (for example, up to 77 K when detecting in LWIR).
The key condition for creating a HOT detector is to minimize thermal generation in
the active region without reducing the quantum efficiency. Among all the mecha-
nisms of thermal generation-recombination (GR) in narrow-gap semiconductors,
the main role is played by interband Auger processes (Auger 1 and Auger 7, which
have the lowest threshold energies), as well as GR by the Shockley-Reed-Hall
(SRH) mechanism through trap levels. Radiative recombination usually does not
limit the performance of properly designed photodetectors based on HgCdTe [7].
Progress in the creation of IR detectors based on HgCdTe nBn structures is based
on the development of such technologies as molecular beam epitaxy (MBE) and
metal-organic chemical vapor deposition (MOCVD), which make it possible to
grow layers with a precisely controlled distribution of the component composition
and dopant concentration over thickness. This makes it possible to develop new
architectures of device structures that provide an increase in the operating tempera-
ture of the detectors and, in some cases, a simplification of the technological cycle
of their manufacture [8].
At present, it is common to create traditional n on p HgCdTe photodiodes by ion
implantation into an epitaxial film, for example, boron ions (without annealing), as
well as p on n HgCdTe photodiodes by implanting arsenic ions and subsequent
activation annealing. The technology of MCT photodiodes with dual-layer planar
heterojunctions (DLPH) is being actively developed. The use of the nBn architec-
ture for HgCdTe can provide a significant advantage over traditional photodiodes
due to an increase in the quality of the material due to the absence of post-­
implantation defects [9, 10]. Eliminating the need for ion implantation (sometimes
followed by annealing) will simplify the technology for detectors creating.
The concept of a barrier photosensitive structure was first implemented in prac-
tice for detectors based on InAs compounds (group III-V) with type II heterojunc-
tions. The structure consisted only of n-type layers with an undoped InAsSb (or
AlAsSb) barrier [11]. A feature of such heterostructures is the close to zero discon-
tinuity (band offset) of the valence band at the heterojunctions boundaries of the
barrier layer with the contact and absorbing layers at high energy barrier height
(>1 eV) in the conduction band. Among the III-V compounds, the greatest progress
in the creation of nBn structures has been achieved based on the InAsSb solid solu-
tion [12], which combines the advantages of the InSb compound with the possibility
of achieving new properties upon transition to InAsSb epitaxial layers with a barrier
6 II-VI Semiconductor-Based Unipolar Barrier Structures for Infrared Photodetector… 139

based on the InAs1-xSbx/InAs1-ySby heterojunction. It has been theoretically and


experimentally shown that nBn detectors based on III-V compounds with full real-
ization of their potential advantages are capable to compete with traditional HgCdTe
detectors [13], especially in MWIR.
In contrast to the III-V compounds, in the HgCdTe solid solution, which is the
II-VI compound, type I heterojunctions are realized, which are characterized by the
presence of an energy barrier ΔEv for holes in the valence band. The presence of this
barrier adversely affects the sensitivity of the detector structure and the threshold
characteristics of devices and is considered one of the main fundamental obstacles
to the practical implementation of barrier detectors based on HgCdTe. The search
for ways to eliminate the influence of this barrier on the hole photocurrent in such
structures is the subject of most of the works in this direction.
Figure 6.2 shows the layout of layers and a schematic band diagram of a unipolar
barrier n+Bn photosensitive structure, which demonstrates the mechanism of selec-
tive blocking of the electron current of the main charge carriers.
To date, many different architectures of barrier detectors based on HgCdTe have
been proposed, among which unipolar configurations seem to be the most promis-
ing [14, 15].

Fig. 6.2 Schematic structure and band diagram of n+Bn HgCdTe detector under reverse bias and
optical radiation
140 A. V. Voitsekhovskii et al.

6.3 HgCdTe Based nBn Unipolar Barrier Structures

As mentioned above, the practical realization of the potential advantages of barrier


structures based on HgCdTe is significantly limited by the fundamental presence of
a potential barrier for holes formed by the B layer.
The most obvious solution to the problem of the non-zero valence band offset is
to increase the external bias on the photosensitive structure. In [16], a numerical
simulation of the energy diagrams of photosensitive nBnn structure with
Cd0.275Hg0.725Te absorbing layer designed for the MWIR and with complex three-­
layer barrier including central layer with Cd0.6Hg0.4Te and two surrounding layers
with variable composition was carried out. The authors also analyzed the influence
of the composition of the central part of the barrier layer on the values of the detec-
tivity. Figure 6.3 shows the calculated dependences of the detectivity on the magni-
tude of the external bias at the temperature of 200 K for various values of the barrier
layer composition. Here, Nd is the donor impurity concentration in the correspond-
ing layer of the structure.
It is clearly seen from the presented curves that the presence of a barrier for holes
in the valence band determines the low values of the detectivity at low values of the
external bias. As the negative external bias increases, the geometry of the potential
barriers undergoes significant changes for both electrons and holes. As the bias
increases, the height of the barrier for holes becomes smaller, which reduces the
barrier for the minority current and leads to an increase in the detectivity. However,
at too high bias values (~ 0.4 V), the change in the barrier geometry in the

Fig. 6.3 Calculated dependences of the detectivity on the value of the external bias for the barrier
photosensitive nBnn structure at the temperature of 200 K for various values of the composition of
the central layer of the barrier. (Reprinted from [16]. Published 2013 by Polish Academy of
Sciences as open access
6 II-VI Semiconductor-Based Unipolar Barrier Structures for Infrared Photodetector… 141

Fig. 6.4 Energy band diagrams of the structure in (a) the absence and (b) the presence of the bias
for the barrier photosensitive nBnn structure at the temperature of 200 K. (Reprinted from [16].
Published 2013 by Polish Academy of Sciences as open access)

conduction band becomes significant: its shape tends to triangular, the tunnel trans-
parency of the barrier increases, and the shielding efficiency of the majority carrier
current decreases, which leads to subsequent decrease in the values of detectivity
(see Fig. 6.4). Here Ec, Ev are the energies of the edges of the conduction band and
the valence band, Ef is the energy of the Fermi level, while Ef,n and Ef,p are the ener-
gies of non-equilibrium quasi-Fermi levels.
Another important conclusion can be drawn from Fig. 6.3. It can be seen that the
dependence of the detectivity on the composition in the barrier layer is ambiguous
and also depends on the bias voltage. At external biases >0.45 V, barrier layers with
a larger composition provide better device detectivity values, which is explained by
more efficient shielding of the current of majority charge carriers in a structure with
a wider gap barrier. However, at bias voltage <0.45 V another situation is observed:
an increase in the values of detectivity with an increase in the composition of the
barrier at a certain value is replaced by a decrease. This is explained by the fact that
in this range of biases, the values of the detectivity will be significantly affected by
the value of the barrier for holes. An increase in the composition of the barrier
increases the potential barrier both for the majority carriers in the structure and for
the minor ones. It is obvious that there is some optimal ratio between the composi-
tion of the barrier layer and the value of the external bias, which provides the best
values of the detectivity.
In the example described above, the optimization of the parameters of the struc-
ture and operating conditions was carried out only in relation to the external bias
and the composition of the barrier layer. When designing real photosensitive struc-
tures, optimization must be carried out with respect to many structural parameters,
such as composition profiles, thicknesses, as well as profiles and doping levels of
the contact, barrier, and absorbing layers. Since 2011, various scientific groups have
been working on finding the optimal ratios of the parameters of the layers of photo-
sensitive structures.
142 A. V. Voitsekhovskii et al.

The first work on this topic was carried out by a research group from the USA
Anne M. Itsuno, Silviu Velicu, Jun Zhao, Michael Morley, Jamie D. Phillips, Angelo
S. Gilmore from the University of Michigan and EPIR Technologies Inc. In
2011–2012 [14, 17–22], this group carried out fairly extensive theoretical and
experimental studies of barrier structures based on HgCdTe in nBn and other con-
figurations grown by MBE with Riber 32 installation. In this studies structures for
MWIR and LWIR were considered. The optimization of layer parameters (composi-
tions, thicknesses, doping levels) and the selection of the operating bias voltage
were used as a mechanism for improving the performance of the structures.
However, the problem of the presence of a barrier for holes in the valence band has
not been solved. The authors noted the need for additional research aimed at improv-
ing the technology of structure passivation and eliminating surface leakage currents.
The necessity of improving the theoretical model (taking into account tunneling
effects, surface properties, Shockley-Read-Hall generation-recombination mecha-
nisms, inhomogeneities of the stoichiometric composition and dopant concentra-
tion) is noted in order to achieve simulation results that would adequately describe
real structures grown by the MBE.
A number of works [23–37] were carried out by scientific groups from Australia
(The University of Western Australia), China (Shanghai Institute of Technical
Physics, University of Chinese Academy of Sciences and National University of
Defense Technology), another group from the USA (Consultant on Infrared
Detectors and Boston University), as well as groups from France (CEA-Leti-­
Minatec), Turkey (Middle East Technical University), and Iran (Aerospace Research
Institute). All of them carried out mainly theoretical and occasionally experimental
studies aimed at improving the characteristics of the barrier structure based on
HgCdTe by developing efficient architectures and optimizing the parameters of the
structure layers. However, as far as we know, to date, none of these works has
yielded breakthrough and practically realizable results.
Despite a significant number of publications devoted to the theoretical substan-
tiation of the potential advantages of nBn structures, there are few attempts to
implement nBn HgCdTe detectors in practice. For the first time, MWIR nBn detec-
tors based on HgCdTe with the cut-off wavelength of 5.7 μm at 77 K were grown by
MBE (Riber 32 MBE system) on a bulk CdZnTe substrate [20, 22]. It was found
that the current-voltage characteristic (CVC) is determined by the shape of the bar-
rier and depends on the applied bias. In the temperature range of 180–250 K, the
diffusion limitation of the dark current was observed (the current density was 1–3
A/cm2); at lower temperatures, the current was limited by generation through sur-
face traps. The photoresponse depends on the applied bias. According to estimates,
the maximum internal quantum efficiency was 66%. The authors do not give abso-
lute values of the spectral sensitivity, which makes it difficult to assess the quality
of the device. During the second development of the HgCdTe nBn structure using
the MBE method, significantly lower (by about 5 orders of magnitude) values of the
dark current density (3.74 × 10−6 A/cm2 at 77 K and the bias of 0.5 V) were obtained
6 II-VI Semiconductor-Based Unipolar Barrier Structures for Infrared Photodetector… 143

[19]. Such a big scatter in the obtained results indicates the need for systematic
research to optimize the parameters of structures and technological processes.
A scientific group from France also attempted to create MWIR nBn structure
based on MBE HgCdTe with an asymmetric barrier layer and uniform doping [28].
Despite the fact that the quantum efficiency, according to estimates, reached 60%,
large current densities were observed at 77 K (of the order of 10−3 A/cm2). The CVC
had a form that differed significantly from that predicted by the simulation, and
studies using secondary ion mass spectroscopy showed that the actual composition
distribution over the film thickness differed significantly from the planned one.
Attempt was made by Itsuno et al. [14] to create a planar LWIR nBn structure
from HgCdTe, for which high values of the dark current density (about 50 A/cm2 at
77 K) were observed. Illumination resulted in a slight increase in current with
reverse bias. The authors explained the results by the presence of surface leakage
currents. The current density values for fabricated MWIR and LWIR nBn structures
were several orders of magnitude higher than the values predicted by the expression
Rule 07 for the dark current of IR detectors [38], which may indicate high leakage
currents along the perimeter of the structure due to insufficient passivation quality.
As far as it is known from open sources, at present, systematic studies of unipolar
photosensitive MBE nBn structures based on HgCdTe, including experimental
ones, are being carried out by Voitsekhovskii et al. [39–51] at Tomsk State University
(Tomsk, Russia) and Institute of Semiconductor Physics of Siberian Branch of
Russian Academy of Sciences (ISP SB RAS, Novosibirsk, Russia).
For example, in [42], the authors fabricated and studied the MWIR structure, in
which the diffusion limitation of dark current was found. MWIR nBn structures
based on MBE HgCdTe/GaAs were fabricated using the plasma-enhanced atomic
layer deposition (PE-ALD) Al2O3 dielectric to passivate the mesa structure. The
diameter of the mesa structures was in the range from 20 to 500 μm. In a wide tem-
perature range, the dark CVC of fabricated nBn structures based on MBE HgCdTe
were studied. It was found that the reverse current of the nBn structure based on
MBE HgCdTe increases significantly when illuminated by infrared radiation.
According to preliminary estimates, the responsivity of fabricated non-optimized
nBn structures at the temperature of 220 K and the voltage of −1 V is about 0.15 A
W−1. It is shown that the dark current is determined by the volume component,
which is much larger than the surface leakage component. From the temperature
dependences of the dark current, the activation energy is determined, which is close
to the energy of the full band gap of the absorbing layer. In the temperature range of
180–300 K, the dark current values slightly exceed the dark current values for high-­
quality p-n photodiodes based on HgCdTe according to the empirical model Rule
07. It is shown that a diffusion-limited dark current is observed for the studied sam-
ples in the temperature range of 180–300 K at the reverse voltages around −1 V.
Figure 6.5 shows the CVC of this nBn structure at different temperatures in the
dark mode and under illumination.
144 A. V. Voitsekhovskii et al.

Fig. 6.5 CVC of the nBn structure based on MBE HgCdTe in dark mode and under illumination

6.4 HgCdTe Based Unipolar Barrier Structures


with P-Type Layers

Another approach to minimizing the valence band offset is to create a p-type barrier.
This mechanism was described, for example, in [5, 52]. According to theoretical
and experimental studies described in the literature, structures with p-type barriers
demonstrate better performance than structures with n-type barriers.
In [52], it was shown that the barrier for holes can be almost completely elimi-
nated at a certain bias by precise acceptor doping of the barrier region. This approach
is unpromising when using MBE and is possible only when heterostructures are
grown by the MOCVD method, which makes it possible to obtain in situ MCT
material of both donor and acceptor types of conductivity. In addition, the formation
of p-type barrier layer will lead to the appearance of the space charge regions near
the heterointerfaces of the barrier layer, which will cause intense generation accord-
ing to the SRH mechanism. Therefore, the advantage of using a barrier doped with
an acceptor impurity is not obvious.
Figure 6.6 shows the band diagrams calculated by the authors of [52] for photo-
sensitive structures with n- and p-type barriers. It can be seen from the above figures
that in the case of acceptor doping of the barrier layer at the external bias of −0.6 V,
it is possible to almost completely eliminate the barrier for holes.
Since 2013, a group of scientists from Poland (M. Kopytko, A. Jóźwikowska,
K. Jóźwikowski, A. Martyniuk, J. Antoszewski, A. Rogalski and others from
Military University of Technology, Vigo System S.A. and University of Life
Science) has been actively engaged in the problem of creating barrier detectors
based on HgCdTe, including layers of p-type conductivity. In 2014–2018 this group
have published a large number of theoretical, review and analytical works [52–76].
6 II-VI Semiconductor-Based Unipolar Barrier Structures for Infrared Photodetector… 145

Fig. 6.6 Calculated band diagrams of photosensitive structures with (a) n- and (b) p-type barriers
at the temperature of 230 K in the presence and absence of an external bias. (Reprinted with per-
mission from [52]. Copyright 2014: Elsevier)

The main ways to optimize barrier structures considered by this group were the
selection of the parameters of the layers of the barrier structure, as well as the use of
complex multilayer bipolar configurations with p-type barriers. The experimental
samples studied in these works were grown mainly by the MOCVD method on an
Aixtron AIX-200 setup.

6.5 HgCdTe Based nBn Unipolar structures


with Superlattice Barriers

To date, none of the approaches aimed at eliminating the energy barrier for minority
charge carriers in nBn structures based on HgCdTe has made it possible to reduce it
to values that provide acceptable values of the sensitivity of the structure while
maintaining the efficiency of blocking majority charge carriers in the conduction
band. The number of works devoted to this problem in the literature has been sys-
tematically decreasing for several years. This, on the one hand, is due to the fact that
the parametric optimization of HgCdTe nBn structures has exhausted its possibili-
ties in theoretical studies, and, on the other hand, is due to the small number and
non-systematic nature of experimental work related to the fabrication of real nBn
HgCdTe structures by the MBE method and studying their properties.
At present, one of the most promising methods for increasing the photosensitiv-
ity and quantum efficiency of nBn structures based on HgCdTe is the use of a super-
lattice with specified parameters as the barrier layer that effectively blocks the dark
electron current in the conduction band and unimpeded hole current flow in the
valence band. The level of technological development of methods for the epitaxial
growth of HgCdTe nanoheterostructures makes it possible to fabricate such
146 A. V. Voitsekhovskii et al.

structures only by MBE, which ensures the reproducibility of the parameters of


HgCdTe layers of any composition with thicknesses of up to several nanometers.
Theoretically, the implementation of this concept can make it possible to com-
pletely solve the problem of the valence band break at the heterointerfaces of the
barrier layer. On the one hand, the use of a superlattice structure will make it pos-
sible to preserve the unipolarity of the photosensitive device and not resort to tech-
nologically complex and defect-forming operations for creating p-type layers. On
the other hand, due to the possibility of managing the properties of the superlattices,
their use will provide high controllability of the operating characteristics of photo-
sensitive structures without fundamental modification of the technological mode of
its growth. The idea of using a superlattice as barrier layer of nBn structure to elimi-
nate the energy barrier for minority charge carriers was first proposed by Kopytko
et al. [77].
Figure 6.7a shows the calculated band diagram of a similar structure, in which
uniform layers of n- and p-type conductivity act as a barrier [77]. Calculation of the
energy band diagram of the structure, in which the third type superlattice (T3SL)
HgTe/Cd0.95Hg0.05Te plays the role of a barrier, is shown in Fig. 6.7b. The thickness
of the Cd0.95Hg0.05Te layers in all cases considered by the authors was 28 nm, while
the thicknesses of the HgTe layers varied during the calculations. Modeling showed
that in such system with the thickness of HgTe layers of 5 nm, the potential barrier
in the valence band is completely eliminated without the formation of space charge
region (Fig. 6.7b, curves 5/28). Unfortunately, the authors do not provide informa-
tion on how many periods of the superlattice they included in the calculations.
This topic was further developed in the works [78–83] of a group of scientists
from The University of Western Australia (N.D. Akhavan, G. Jolley, G. A. Umana-­
Membreno, J. Antoszewski, L. Faraone et al.). Calculations carried out in [82],

Fig. 6.7 Calculated band diagrams of structures with barriers of n- and p-types (a) and with the
barrier in the form of the superlattice (b) at the temperature of 80 K at zero external bias. (Reprinted
from [77]. Published 2015 by Springer as open access)
6 II-VI Semiconductor-Based Unipolar Barrier Structures for Infrared Photodetector… 147

Fig. 6.8 CVCs of nBn structure including superlattice of 18 periods Hg0.20Cd0.80Te (9 nm) – HgTe
(2 nm) as selective barrier, measured at different temperatures

taking into account the quantum mechanical nature of the superlattice in the barrier
layer, showed that at the thickness of the CdTe barriers and HgTe wells equal to 1.3
and 3.7 nm, respectively, the maximum ratio of the hole current to the electron cur-
rent is reached, if the number of layers in the superlattice exceeds 12. It was also
theoretically shown in the works of this group that a reduction in the barrier in the
valence band is possible with a non-uniform distribution of the composition and
dopant in the barrier layer [31]. However, these studies were predominantly theo-
retical. Only one of the works known to us [79] presents the results of measure-
ments of the dark current of a structure with HgTe/CdTe superlattice as the barrier
layer at the temperature of 155 K, as well as the corresponding calculation. The
experimental structure was grown by MBE at The University of Western Australia.
In [84] Voitsekhovskii et al. have studied the electrophysical properties of nBn
structure with a superlattice grown by the MBE method at the ISP SB RAS by
admittance spectroscopy. The structure included superlattice of 18 periods
Hg0.20Cd0.80Te (9 nm) – HgTe (2 nm) as a selective barrier for majority charge carri-
ers. Figure 6.8 shows the CVC of this structure measured at various temperatures.

6.6 HgCdTe Based NBνN HOT Unipolar Structures

One way to suppress Auger processes that does not require semiconductor cooling
is the creation of non-equilibrium depletion [85]. Dark currents, which determine
the operating temperature of highly sensitive IR detectors based on HgCdTe, are
limited by Auger generation processes. It is not a trivial task to reduce the require-
ments for cooling detectors without compromising performance. The concept of
148 A. V. Voitsekhovskii et al.

HOT structures is based on the suppression of Auger GR processes by reducing the


thermal concentration of charge carriers in the absorbing layer to values that are less
than equilibrium. In HOT detectors, a lightly doped narrow-gap absorbing layer is
used, on one side of which a heterojunction is formed, which ensures the exclusion
of charge carriers, and on the other, the extraction of charge carriers. When a reverse
bias is applied, the concentration of charge carriers in the absorbing layer decreases,
which manifests itself in the suppression of Auger processes. Estimates of the
threshold characteristics show that for HOT MWIR detectors, the mode of limita-
tion by background radiation noise is realized at 203 K (compared to T = 155 K for
an ideal DLPH detector), and at 145 K for LWIR detector (102 K for DLPH detec-
tor). The dark current in such HOT structures, according to calculations, should be
less than in photodiodes with p-n junctions [86], and the operating characteristics at
a given temperature should be better.
In the case of n-HgCdTe MBE, a promising architecture of unipolar barrier
detectors is the NBνN configuration, which makes it possible to increase the operat-
ing temperature of detectors by suppressing Auger GR processes. Theoretically,
NBνN-architecture HgCdTe detectors retain the technological advantages of the
nBn configuration, but can show lower dark currents than p-n photodiodes and nBn
detectors.
The NBνN device contains four n-type layers: a heavily doped cover layer, a
lightly doped barrier layer, a lightly doped absorber layer, and a heavily doped bot-
tom layer. The offset of the conduction band at the boundary of the absorbing and
barrier layers prevents the current of electrons from the covering to the absorbing
layer. Under equilibrium conditions, there is a barrier in the valence band at the
boundary between the barrier and absorbing layers, which blocks the transfer of
holes in the direction of the coating layer. When a reverse bias is applied, the barrier
in the valence band decreases. As a result, holes that appear in the absorbing layer
due to thermal and optical generation are collected in the upper layer. The boundary
between the absorbing and barrier (or covering) layer provides hole extraction, and
the boundary between the absorbing and lower layers provides exclusion. As a
result, thermally generated holes are effectively removed from the absorbing layer,
and the hole concentration becomes much lower than the equilibrium values. To
maintain electrical neutrality in the absorbing layer, the electron concentration also
decreases below the equilibrium value. GR processes by the Auger mechanism are
suppressed due to a decrease in the total carrier concentration in the active layer.
According to the simulation results, the use of the NBνN configuration makes it
possible to reduce the dark current values, increase the detectivity and operating
temperature (compared to nBn and DLPH), and also eliminate the technological
problem of forming p-type regions.
The first work on this topic was carried out by a research group from the USA
Anne M. Itsuno, Silviu Velicu at al [19]. from the University of Michigan and EPIR
Technologies Inc. In [19], the characteristics of IR detectors based on HgCdTe with
NBνN architecture were simulated. The calculated values of the dark current den-
sity of the detectors for the MWIR and LWIR ranges are an order of magnitude (or
more) lower than for the nBn or DLPH detectors in the temperature range from 50
6 II-VI Semiconductor-Based Unipolar Barrier Structures for Infrared Photodetector… 149

to 225 K. The calculated values of the peak detectivity (D*) at the maximum sensi-
tivity for the MWIR detector were 6.0 × 1014 and 2.4 × 1010 cm × W−1 × Hz1/2 at the
temperatures of 95 and 225 K, respectively. The D* values for the LWIR detector
were 2.4 × 1014 and 2.3 × 1011 cm × W−1 × Hz1/2 at 50 and 95 K, respectively.
Detectivity values were estimated from the maximum sensitivity calculated in ear-
lier works [17].

6.7 Summary

Based on the review, we can conclude that unipolar structures based on II-VI com-
pounds are promising and can lead to the creation of IR FPAs with improved char-
acteristics, operating at higher cooling temperatures. A large theoretical background
in the study of nBn structures based on HgCdTe for both MWIR and LWIR shows
some advantages over photodiodes based on p-n junction in HgCdTe. If for the
medium-wavelength IR range the creation of FPAs based on barrier structures of
III-V compounds has reached the practical level and mass production, then the prac-
tical implementation of devices based on II-VI compounds is hindered by a large
number of unsolved fundamental, design and technological problems. The presence
of a barrier for holes in the valence band in nBn structures based on MCT material
requires a number of technological solutions, which are the use of large external
biases, control of the barrier layer parameters, including acceptor doping of the bar-
rier, and the use of multilayer structures with a complex barrier layer design, includ-
ing barriers in the form of superlattices. Therefore, despite a certain theoretical
background, the developed technological implementations of the devices have not
yet made it possible to obtain practically significant results realizing the theoreti-
cally predicted advantages of the nBn MCT structure in the development of
the FPAs.
The authors of one of the latest reviews of the achievements of nBn infrared
detectors [87] conclude that for the case of HgCdTe, the technology of barrier detec-
tors is the most promising for achieving operation at close to room temperature.

Acknowledgments Study was partially supported by the Tomsk State University Development
Program («Priority-2030»).

References

1. Rogalski А (2019) Infrared and terahertz detectors, 3rd edn. Taylor & Francis Group, LLC
2. Kinch MA (2015) The future of infrared; III–Vs or HgCdTe? J Electron Mater 44(9):2969–2976
3. Gu R, Antoszewski J, Lei W, Madni I, Umana-Membrenao G, Faraone L (2017) MBE growth
of HgCdTe on GaSb substrates for application in next generation infrared detectors. J Cryst
Growth 468:216–219
150 A. V. Voitsekhovskii et al.

4. Maimon S, Wicks GW (2006) nBn detector, an infrared detector with reduced dark current and
higher operating temperature. Appl Phys Lett 89:151109
5. Kopytko M, Keblowski A, Gawron W, Madejczyk P (2015) Different cap-barrier design for
MOCVD grown HOT HgCdTe barrier detectors. Opto-Electron Rev 23(2):143–148
6. Rogalski A, Martyniuk P (2014) Mid-wavelength infrared nBn for HOT detectors. J Electron
Mater 43(8):2963–2969
7. Kopytko M, Kębłowski A, Gawron W, Pusz W (2016) LWIR HgCdTe barrier photodiode with
Auger-suppression. Semicond Sci Technol 31(3):035025
8. Kopytko M, Rogalski A (2016) HgCdTe barrier infrared detectors. Prog Quantum
Electron 47:1–18
9. Bubulac LO (1988) Defects, diffusion and activation in ion implanted HgCdTe. J Crystal
Growth 86(1–4):723–734
10. Talipov N, Voitsekhovskii A (2018) Annealing kinetics of radiation defects in boron-implanted
p-Hg1-xCdxTe. Semicond Sci Technol 33(6):065009
11. Pedrazzani JR, Maimon S, Wicks GW (2008) Use of nBn structures to suppress surface leak-
age currents in unpassivated InAs infrared photodetectors. Electron Lett 44(25):1487–1488
12. Soibel A, Keo SA, Fisher A, Hill CJ, Luong E, Ting DZ, Gunapala SD, Lubyshev D, Qiu Y,
Fastenau JM, Liu AWK (2018) High operating temperature nBn detector with monolithically
integrated microlens. Appl Phys Lett 112(4):041105
13. Martyniuk P, Kopytko M, Rogalski A (2014) Barrier infrared detectors. Opto-Electron Rev
22(2):127–146
14. Itsuno AM (2012) Bandgap-engineered mercury cadmium telluride infrared detector struc-
tures for reduced cooling requirements. PhD thesis, University of Michigan
15. Iakovleva NI (2019) Unipolar MCT-based nBn-structure for a MWIR FPA. Appl Phys 3:53–60
16. Martyniuk P, Rogalski A (2013) Theoretical modelling of MWIR thermoelectrically cooled
nBn HgCdTe detector. Bull Polish Acad Sci, Techn Sci 61(1):211–220
17. Itsuno AM, Phillips JD, Velicu S (2011) Design and modeling of HgCdTe nBn detectors. J
Electron Mater 40(8):1624–1629
18. Itsuno AM, Phillips JD, Gilmore AS, Velicu S (2011) Calculated performance of an Auger
suppressed unipolar HgCdTe photodetector for high temperature operation. Proc SPIE
8155:81550J
19. Itsuno AM, Phillips JD, Velicu S (2012) Design of an Auger-Suppressed Unipolar HgCdTe
NBvN photodetector. J Electron Mater 41(10):2886–2892
20. Velicu S, Zhao J, Morley M, Itsuno AM, Phillips JD (2012) Theoretical and experimental
investigation of MWIR HgCdTe nBn detectors. Proc SPIE 8268:82682X
21. Itsuno AM, Phillips JD, Velicu S (2012) Unipolar barrier-integrated HgCdTe infrared detec-
tors. In: Proceeding of 70th device research conference, 18–20 June 2012. University Park,
PA, USA, 12908883
22. Itsuno AM, Phillips JD, Velicu S (2012) Mid-wave infrared HgCdTe nBn photodetector. Appl
Phys Lett 100:161102
23. Akhavan ND, Jolley G, Umana-Membreno GA, Antoszewski J, Faraone L (2015) Theoretical
study of Midwave infrared HgCdTe nBn detectors operating at elevated temperatures. J
Electron Mater 44:3044–3055
24. Ye ZH, Chen YY, Zhang P, Lin C, Hu XN, Ding RJ, He L (2014) Modeling of LWIR nBn
HgCdTe photodetector. Proc SPIE 9070:90701L
25. Akhavan ND, Umana-Membreno GA, Jolley G, Antoszewski J, Faraone L (2014) A method
of removing the valence band discontinuity in HgCdTe-based nBn detectors. Appl Phys Lett
105:121110
26. Akhavan ND, Jolley G, Membreno GU, Antoszewski J, Faraone L (2014) Band-to-band tun-
nelling (BTBT) in HgCdTe-based nBn detectors for LWIR applications. In: Proceedings of the
conference on optoelectronic and microelectronic materials & devices, 14–17 December 2014.
Perth, WA, Australia, 14920679
6 II-VI Semiconductor-Based Unipolar Barrier Structures for Infrared Photodetector… 151

27. Akhavan ND, Jolley G, Umana-Membreno GA, Antoszewski J, Faraone L (2014) Performance
modeling of bandgap engineered HgCdTe-based nBn infrared detectors. IEEE Trans Electron
Devices 61(11):3691–3698
28. Gravrand O, Boulard F, Ferron A, Ballet P, Hassis W (2015) A new nBn IR detection concept
using HgCdTe material. J Electron Mater 44(9):3069–3075
29. Akhavan ND, Jolley G, Umana-Membreno GA, Antoszewski J, Faraone L (2015) Design
of Band Engineered HgCdTe nBn detectors for MWIR and LWIR applications. IEEE Trans
Electron Devices 62(3):722–728
30. Ting DZ, Soibel A, Khoshakhlagh A, Gunapala SD (2017) Theoretical analysis of nBn infra-
red photodetectors. Opt Eng 56(9):091606
31. Akhavan ND, Umana-Membreno GA, Gu R, Antoszewski J, Faraone L (2018) Delta doping in
HgCdTe based unipolar barrier photodetectors. IEEE Trans Electron Devices 65(1):4340–4345
32. He J, Hu W (2018) Numerical simulation of HgCdTe nBn longwavelength infrared detector.
In: Proceedings of international conference on numerical simulation of optoelectronic devices
(NUSOD), 05–09 November 2018. Hong Kong, China, 18310399
33. Uzgur F, Kocaman S (2019) A dual-band HgCdTe nBn infrared detector design. Proc SPIE
11129:1112903
34. Uzgur F, Kocaman S (2019) Barrier engineering for HgCdTe unipolar detectors on alternative
substrates. Infrared Phys Technol 97:123–128
35. Sharma I, Srivastava T, Kaushik R, Goyal A (2019) Design and analysis of HgCdTe infrared
photodetector. In: Proceedings of the international conference on signal processing and com-
munication (ICSC), 07–09 March 2019. Noida, India, 19256588
36. He J, Wang P, Li Q, Wang F, Gu Y, Shen C, Lu Chen P, Martyniuk A, Rogalski XC, Wei L, Hu
W (2010) Enhanced performance of HgCdTe long-wavelength infrared photodetectors with
nBn design. IEEE Trans Electron Devices 67(5):2001–2007
37. Shaveisi M, Aliparast P (2021) Dark current evaluation in HgCdTe-based nBn infrared
detectors. In: 2021 Iranian international conference on microelectronics (IICM2021), 22–24
December 2021. IEEE. https://doi.org/10.1109/IICM55040.2021.9730154
38. Tennant WE, Lee D, Zandian M, Piquette E, Carmody M (2008) MBE HgCdTe technology:
a very general solution to IR detection, described by «Rule 07», a very convenient heuristic. J
Electron Mater 37(9):1406–1410
39. Voitsekhovskii AV, Nesmelov SN, Dzyadukh SM, Dvoretsky SA, Mikhailov NN, Sidorov GY
(2019) Admittance characteristics of nBn structures based on HgCdTe grown by molecular
beam epitaxy. Russ Phys J 62(5):818–826
40. Voitsekhovskii AV, Nesmelov SN, Dzyadukh SM, Dvoretsky SA, Mikhailov NN, Sidorov
GY, Yakushev MV (2019) Admittance dependences of the mid-wave infrared barrier structure
based on HgCdTe grown by molecular beam epitaxy. Mater Res Express 6:116411
41. Voitsekhovskii AV, Nesmelov SN, Dzyadukh SM, Dvoretsky SA, Mikhailov NN, Sidorov GY
(2019) Current-voltage characteristics of nBn structures based on mercury cadmium telluride
epitaxial films. Russ Phys J 62(6):1054–1061
42. Voitsekhovskii AV, Nesmelov SN, Dzyadukh SM, Dvoretsky SA, Mikhailov NN, Sidorov GY,
Yakushev MV (2020) Diffusion-limited dark currents in mid-wave infrared HgCdTd-based
nBn structures with Al2O3 passivation. J Phys D Appl Phys 53:53 055107
43. Voitsekhovskii AV, Nesmelov SN, Dzyadukh SM, Dvoretsky SA, Mikhailov NN, Sidorov GY
(2019) Electrical properties of nBn structures based on HgCdTe grown by molecular beam
epitaxy on GaAs substrates. Infrared Phys Technol 102:103035
44. Izhnin II, Voitsekhovskii AV, Nesmelov SN, Dzyadukh SM, Dvoretsky SA, Mikhailov NN,
Sidorov GY, Yakushev MV (2022) Admittance of barrier nanostructures based on MBE
HgCdTe. Appl Nanosci 12:403–409
45. Voitsekhovskii AV, Nesmelov SN, Dzyadukh SM, Dvoretsky SA, Mikhailov NN, Sidorov GY,
Yakushev MV (2020) Admittance of barrier structures based on mercury cadmium telluride.
Russ Phys J 63(3):432–445
152 A. V. Voitsekhovskii et al.

46. Voitsekhovskii AV, Nesmelov SN, Dzyadukh SM, Dvoretsky SA, Mikhailov NN, Sidorov GY,
Yakushev MV (2020) Admittance of metal-insulator-semiconductor devices based on HgCdTe
nBn structures. Semicond Sci Technol 35:055026
47. Voitsekhovskii AV, Nesmelov SN, Dzyadukh SM, Dvoretsky SA, Mikhailov NN, Sidorov GY,
Yakushev MV (2020) Impedance of mis devices based on nBn structures from mercury cad-
mium telluride. Russ Phys J 63(6):907–916
48. Voitsekhovskii AV, Nesmelov SN, Dzyadukh SM, Dvoretsky SA, Mikhailov NN, Sidorov
GY, Yakushev MV (2021) Admittance of MIS structures based on nBn systems of epitaxial
HgCdTe for detection in the 3-5 μm spectral range. Tech Phys Lett 47(6):616–619
49. Voitsekhovskii AV, Nesmelov SN, Dzyadukh SM, Dvoretsky SA, Mikhailov NN, Sidorov GY,
Yakushev MV (2021) An experimental study of the dynamic resistance in surface leakage
limited nBn structures based on HgCdTe grown by molecular beam epitaxy. J Electron Mater
50:4599–4605
50. Voitsekhovskii AV, Nesmelov SN, Dzyadukh SM, Dvoretsky SA, Mikhailov NN, Sidorov GY,
Yakushev MV (2021) Dark currents of unipolar barrier structures based on mercury cadmium
telluride for long-wave IR detectors. Russ Phys J 64(5):763–769
51. Burlakov ID, Kulchitsky NA, Voitsekhovskii AV, Nesmelov SN, Dzyadukh SM, Gorn DI
(2021) Unipolar semiconductor barrier structures for infrared photodetector arrays (review). J
Commun Technol Electron 66(9):1084–1091
52. Kopytko M (2014) Design and modelling of high-operating temperature MWIR HgCdTe nBn
detector with n- and p-type barriers. Infrared Phys Technol 64:47–55
53. Kopytko M, Keblowski A, Gawron W, Madejczyk P, Kowalewski A, Jozwikowski K (2013)
High-operating temperature MWIR nBn HgCdTe detector grown by MOCVD. Opto-Electron
Rev 21(4):402–405
54. Kopytko M, Jozwikowski K (2013) Numerical Analysis of Current–Voltage Characteristics of
LWIR nBn and p-on-n HgCdTe Photodetectors. J Electron Mater 42(11):3211–3216
55. Kopytko M, Jozwikowski K, Rogalski A (2014) Fundamental limits of MWIR HgCdTe barrier
detectors operating under non-equilibrium mode. Solid State Electron 100:20–26
56. Martyniuk P, Gawron W (2014) Barrier detectors versus homojunction photodiode. Metrol
Meas Syst XXI (4):675–684
57. Martyniuk P (2014) HOT HgCdTe infrared detectors. In: IEEE Xplore. Numerical simulation
of optoelectronic devices. https://doi.org/10.1109/NUSOD.2014.6935412
58. Martyniuk P (2015) HOT mid-wave HgCdTe nBn and pBp infrared detectors. Opt Quant
Electron 47:1311–1318
59. Martyniuk P, Gawron W, Pusz W, Stanaszek D, Rogalski A (2014) Modeling of HOT (111)
HgCdTe MWIR detector for fast response operation. Opt Quant Electron 46:1303–1312
60. Kopytko M, Jozwikowska A, Jozwikowska K, Martyniuk A, Antoszewski J, Faraone L (2014)
Numerical analysis of fluctuation phenomena in HOT HgCdTe barrier detectors. In: IEEE
Xplore. 2014 conference on optoelectronic and microelectronic materials & devices. https://
doi.org/10.1109/COMMAD.2014.7038687
61. Martyniuk P, Gawron W, Stanaszek D, Pusz W, Rogalski A (2014) Theoretical model-
ling of mercury cadmium telluride mid-wave detector for high temperature operation. IET
Optoelectron 8(6):239–244
62. Qiu WC, Jiang T, Cheng XA (2015) A bandgap-engineered HgCdTe PBπn long-wavelength
infrared detector. J Appl Phys 118:124504
63. Kopytko M, Jozwikowski K (2015) Generation-recombination effect in MWIR HgCdTe bar-
rier detectors for high-temperature operation. IEEE Trans Electron Devices 62(7):2278–2284
64. Kopytko M, Keblowski A, Gawron W, Martyniuk P, Madejczyk P, Jozwikowski K, Kowalewski
A, Markowska O, Rogalski A (2015) MOCVD grown HgCdTe barrier detectors for MWIR
high-operating temperature operation. Opt Eng 54(10):105105
65. Chen Y, Ye Z, Zhang P, Hu X, Ding R, He L (2016) A barrier structure optimization for widen-
ing processing window in dual-band HgCdTe IRFPAs detectors. Opt Quant Electron 48:294
6 II-VI Semiconductor-Based Unipolar Barrier Structures for Infrared Photodetector… 153

66. Martyniuk P, Gawron W, Madejczyk P, Kopytko M, Grodecki K, Gomulka E (2017) Theoretical


utmost performance of (100) mid-wave HgCdTe photodetectors. Opt Quant Electron 49:20
67. Madejczyk P, Gawron W, Martyniuk P, Keblowski A, Pusz W, Pawluczyk J, Kopytko M,
Rutkowski J, Rogalski A, Piotrowski J (2017) Engineering steps for optimizing high tempera-
ture LWIR HgCdTe photodiodes. Infrared Phys Technol 81:276–281
68. Kopytko M, Keblowski A, Madejczyk P, Martyniuk P, Piotrowski J, Gawron W, Grodecki K,
Jozwikowski K, Rutkowski J (2017) Optimization of a HOT LWIR HgCdTe photodiode for fast
response and high detectivity in zero-bias operation mode. J Electron Mater 46(10):6045–6055
69. Martyniuk P, Gawron W, Mikolajczyk J (2017) The development of the room temperature
LWIR HgCdTe detectors for free space optics communication systems. In: Proceedings of
the SPIE 10437, advanced free-space optical communication techniques and applications III,
104370G (6 October 2017). https://doi.org/10.1117/12.2278628
70. Jozwikowski K, Piotrowski J, Jozwikowska A, Kopytko M, Martyniuk P, Gawron W,
Madejczyk P, Kowalewski A, Markowska O, Martyniuk A, Rogalski A (2017) The numerical-­
experimental enhanced analysis of HOT MCT barrier infrared detectors. J Electron Mater
46:5471–5478
71. Kopytko M (2017) Theoretical performance of mid wavelength HgCdTe(100) heterostructure
infrared detector. Solid State Electron 137:102–108
72. Kopytko M, Gomolka E, Michalczewski K, Martyniuk P, Rutkowski J, Rogalski A (2018)
Investigation of surface leakage current in MWIR HgCdTe and InAsSb barrier detectors.
Semicond Sci Technol 33:125010
73. Kopytko M, Gawron W, Keblowski A, Stępien D, Martyniuk P, Jozwikowska K (2018)
Numerical analysis of HgCdTe dual-band infrared detector. Opt Quant Electron 51:62
74. Martyniuk P, Madejczyk P, Kopytko M, Henig AM, Grodecki K, Gawron W, Rutkowski J
(2018) Theoretical simulation of the thermoelectrically cooled HgCdTe LWIR detector for fast
response operating under unbiased conditions. IET Optoelectron 12(4):161–167
75. Kopytko M, Gawron W, Keblowski A, Stepien D, Martyniuk P, Jozwikowski K (2019)
Numerical analysis of HgCdTe dual-band infrared detector. Opt Quant Electron 51:62
76. He J, Li Q, Wang P, Wang F, Yue G, Shen C, Luo M, Yu C, Lu Chen X, Chen W, Lu WH (2020)
Design of a bandgap-engineered barrier-blocking HOT HgCdTe long-wavelength infrared ava-
lanche photodiode. Opt Express 28(22):33556–33563
77. Kopytko M, Wrobel J, Jozwikowska K, Rogalski A, Antoszewski J, Akhavan ND, Umana-­
Membreno GA, Faraone L, Becker CR (2015) Engineering the bandgap of unipolar HgCdTe-­
based nBn infrared photodetectors. J Electron Mater 44(1):158–166
78. Benyaya J, Martyniuk P, Kopytko M, Antoszewski J, Gawron W, Madejczyk P (2015) nBn
HgCdTe infrared detector with HgTe/CdTe SLs barrier. In: IEEE Xplore, 2015 interna-
tional conference on numerical simulation of optoelectronic devices (NUSOD). https://doi.
org/10.1109/NUSOD.2015.7292881
79. Benyahia D, Martyniuk P, Kopytko M, Antoszewski J, Gawron W, Madejczyk P, Rutkowski J,
Gu R, Faraone L (2016) nBn HgCdTe infrared detector with HgTe(HgCdTe)/CdTe SLs barrier.
Opt Quant Electron 48:215
80. Gu R, Lei W, Antoszewski J, Madni I, Umana-Menbreno G, Faraone L (2016) Recent progress
in MBE grown HgCdTe materials and devices at UWA. Proc SPIE 9819:98191Z
81. Akhavan ND, Umana-Membreno GA, Antoszweski J, Faraone L (2016) Self consistent carrier
transport in band engineered HgCdTe nBn detector. In: IEEE Xplore. 2016 International con-
ference on numerical simulation of optoelectronic devices (NUSOD). https://doi.org/10.1109/
NUSOD.2016.7547060
82. Akhavan ND, Umana-Membreno GA, Gu R, Asadnia M, Antoszewski J, Faraone L (2016)
Superlattice barrier HgCdTe nBn infrared photodetectors: validation of the effective mass
approximation. IEEE Trans Electron Devices 63(12):4811–4818
83. Akhavan ND, Umana-Membreno GA, Gu R (2018) Optimization of superlattice barrier
HgCdTe nBn infrared photodetectors based on an NEGF approach. IEEE Trans Electron
Devices 65(2):591–598
154 A. V. Voitsekhovskii et al.

84. Izhnin II, Kurbanov KR, Voitsekhovskii AV, Nesmelov SN, Dzyadukh SM, Dvoretsky SA,
Mikhailov NN, Sidorov GY, Yakushev MV (2020) Unipolar superlattice structures based on
MBE HgCdTe for infrared detection. Appl Nanosci 10:4571–4576
85. Ashley T, Elliott CT (1985) Nonequilibrium devices for infra-red detection. Electron Lett
21(10):451–452
86. Schaake HF, Kinch MA, Chandra D, Aqariden F, Liao PK, Weirauch DF, Wan C-F, Scritchfield
RE, Sullivan WW, Teherani JT, Shih HD (2008) High-operating-temperature MWIR detector
diodes. J Electron Mater 37(9):1401–1405
87. Shi Q, Zhang S-K, Wang J-L, Chu J-H (2022) Progress on nBn infrared detectors. J Infrared
Millim Waves 41(1):139–150
Chapter 7
Infrared Sensing Using Mercury
Chalcogenide Nanocrystals

Emmanuel Lhuillier, Tung Huu Dang, Mariarosa Cavallo, Claire Abadie,


Adrien Khalili, John C. Peterson, and Charlie Gréboval

7.1 Introduction

II-VI NCs are bright, narrow-band emitters, that achieve many interesting proper-
ties through quantum confinement. Commercially, they have been used as a light
down converter for displays, but NCs have also been applied in ways that make
central use of the materials, such as in single photon emitters [1], in light emitting
diodes [2] (LED), or as biolabels [3].
Solar cells have been the main application of NC materials as light sensors.
Thanks to their size-tunable band gap, lead sulfide (PbS) NCs can have the optimal
band gap for single junction solar cells, around 1.3 eV. In addition, they may be
tuned to the near-infrared part of the solar spectrum, a region that remains unad-
dressed by organic solar cells, the other low-cost alternative to silicon-based solar
cells. The interest in NC-based solar cells has also been motivated by multi-exciton
generation [4–6]. In a semiconductor, high-energy photons may result in the genera-
tion of more than one electron-hole pair if carrier multiplication occurs before their
thermalization. In NCs, the threshold energy for this process to occur is drastically
reduced compared to the bulk material. This motivate infrared sensing based on nar-
rower band gap lead chalcogenides NCs.
Thanks to their small or absent bulk band gap, mercury chalcogenides also offer
a nice playground as infrared optoelectronic building blocks. Mercury chalcogenide
NCs [7] offer a broadly tunable absorption from the visible range up to the THz
(Fig. 7.1) while being synthetically similar to cadmium chalcogenides, the prepara-
tion of which has been extensively studied. Initially, the interest in HgTe

E. Lhuillier (*) · T. H. Dang · M. Cavallo · C. Abadie · A. Khalili · C. Gréboval


Sorbonne Université, CNRS, Institut des NanoSciences de Paris, INSP, Paris, France
e-mail: el@insp.jussieu.fr
J. C. Peterson
The James Franck Institute, The University of Chicago, Chicago, IL, USA

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 155
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_7
156 E. Lhuillier et al.

Fig. 7.1 Size tunability of


HgTe NCs absorption
edge. Wavelength of the
lowest energy absorption
feature (interband edge for
small size and intraband
peak for largest size) as a
function of particle size for
HgTe NCs

nanocrystals had been driven by their use as broadband optical amplifiers [8], how-
ever after the collapse of the telecom bubble in the early 2000s, the use of this mate-
rial has shifted toward detection [9]. Bulk mercury chalcogenides are already used
for infrared sensing, as in the alloy HgCdTe, where the Cd content is used to tune
the band gap. In NCs, the quantum confinement plays the same role, see Fig. 7.1.
Note that this chapter is mainly focused on HgTe since, among mercury chalcogen-
ides, it has generated the most interest by far. Integration of HgS and HgSe was
extremely limited until the observation of intraband absorption in the doped form of
these NCs. A section is dedicated to intraband devices at the end of the chapter.
Though the first report on the preparation of monodisperse NCs was published in
1993 [10], it took nearly 10 years before NCs progressed from being optically active
materials to building blocks for optoelectronics [11]. One major challenge to this
effect is adressing the initial poor charge transport properties of NC films, as dis-
cussed in Sect. 7.2 of this chapter. Surface ligands are required to obtain monodis-
perse NC solution with well-passivated surfaces, but these often act as a barrier to
charge transport, leading to poor charge carrier mobility. Thus, a pristine film of
NCs typically presents a low mobility (<10−6 cm2.V−1.s−1), seemingly limiting the
application of these materials. It is only as the field has matured that such challenges
have been overcome, and NC device performances have increased.

7.2 Enabling Transport and Photoconduction


in Nanocrystal Films

NCs are typically colloidally synthesized: grown in solution in the presence of


ligands. These ligands are necessary. During the synthesis, they reduce the avail-
ability of the NC surface, slowing particle growth. Once the NCs are synthesized,
7 Infrared Sensing Using Mercury Chalcogenide Nanocrystals 157

these ligands also aid colloidal stability and prevent aggregation. They also help to
electronically passivate the surface through hybridization with dangling bonds.
However, from a device perspective, ligands act as energetic barriers to charge
transport. The barrier width roughly matches the ligand length, while the height is
typically a few eV. As a result, the NCs are only weakly coupled with one another.
Macroscopically, this leads to a low carrier mobility. Typical values for films of
pristine NCs (i.e., without further processing) fall in the 10−6 cm2.V−1.s−1 range lim-
iting their use in optoelectronic devices. Due to this weak local coupling, charge
transport in a nanocrystal film may be modeled as a hopping process, where charge
wave functions are localized to individual NCs, which tunnel through the barrier
formed by the ligands with some hopping time. Since single particle devices remain
uncommon [12], and the electrode spacing ranges from 100 nm to a few 10 s of μm
while the particle size is typically 5 to 20 nm, charges must tunnel through any-
where from 10 to a few thousand barriers. To improve the charge transport, the bar-
rier must be modified to minimize its height and width. This is the purpose of the
ligand exchange procedure.

7.2.1 Solid-State Ligand Exchange

A film of deposited NCs may be treated by a solution of short ligands in the solid
state to improve the film conductivity. NCs are initially spread onto a surface using
different methods such as spin-coating, dip-coating or even drop-casting.
Complementary methods, such as spray-coating [13, 14], inkjet printing [9], or
electrophoretic deposition [15] have also been developed, but these will not be
described here. Synthetized NCs are often capped with long, organic ligands (acid,
amine, thiol and phosphine are the most typical ligands at this stage) that bond to the
NC surface, see Fig. 7.2a. These ligands typically include an alkyl chain, commonly
12–18 carbons long. This results in a typical length of the ligand of 1.5–2 nm,
depending on the exact number of carbons and possible ligand interdigitation.
This preformed film of NCs is dipped into a solution of shorter ligands in order
to maximize nanoparticle coupling. The shorter the ligand is, the stronger the NC
coupling will be, leading to higher mobility. The Law’s group (Fig. 7.2e) has been
able to measure a scaling law that relates the ligand length and the mobility, and the
relationship is approximately exponential. Also, note that since the nature of the
ligand is not changed (acid in this case), one may extrapolate the mobility to zero
ligand length estimated to be around a few cm2 V−1 s−1. With EDT, the most com-
mon ligand, the mobility is rather in the 10−4–10−1 cm2 V−1 s−1 range, see Fig. 7.2e,
f. To further increase the carrier mobility, it is necessary to also play on the height
of the tunnel barrier and consequently on the nature of the ligand. This can be done
by introducing inorganic ligands [18] or very short ions (Cl− [19], S2− [20], As2S3
[17, 21]…), see Fig. 7.2f.
This solid-state ligand exchange procedure relies on the diffusion of the new
ligand from the solution to the inner part of the NC film and on the reverse
158 E. Lhuillier et al.

Fig. 7.2 Principle and consequence of the solid-state ligand exchange. (a) Schematic of a NC thin
film deposited onto a substrate and capped with their native ligands. (b) Schematic of the previous
film once the film is exposed to short ligands. (c) Schematic of the previous film after the ligand
exchange process. (d) Infrared absorption spectrum in the region of the C-H bond resonance before
and after ligand exchange. (e) Mobility as a function of the ligand length for a thin film of PbSe
NCs capped with various thiol ligands. Part (e) is adapted with permission from Ref. [16].
Copyright 2010: American Chemical Society. (f) Mobility as a function of temperature for a HgTe
NC thin film capped with organic and inorganic ligands. Part (f) is adapted with permission from
Ref. [17]. (Copyright 2013: WILEY-VCH Verlag GmbH & Co. KGaA)

procedure for the native ligand. As a result, this process is only efficient for thin
films (20–50 nm).
The polarity and general solvent affinity of the native ligand drive the efficiency
of the ligand exchange. Solvents in which the native ligands are more soluble will
lead to more efficient extraction. Care must be taken to select a solvent in which the
NCs themselves are not soluble. The completion of the procedure can easily be fol-
lowed by infrared spectroscopy (see Fig. 7.2d) by tracing the height of the reso-
nance associated with the C-H bond appearing around 2900–3000 cm−1 (the doublet
at 3.4 μm in Fig. 7.2d). The magnitude of this peak is drastically reduced at the end
of the procedure as the free ligands are removed, and the ligand length is reduced.
Another consequence of the ligand exchange is a contraction of the film. A 5 nm
NC, capped with 1.5 nm long ligands, presents an initial equivalent size of 8 nm,
which shrinks to 6 nm once capped with EDT. This 20% diameter reduction poten-
tially leads to a 60% volume reduction, and consequently, cracks may be formed,
see Fig. 7.2c. The presence of these cracks reduces the film quality and makes the
percolation path for the carrier to reach the electrodes even longer. This, in addition
to the diffusive character of the solid-state ligand exchange, requires that the film
7 Infrared Sensing Using Mercury Chalcogenide Nanocrystals 159

deposition should be conducted through a multilayer procedure. New layers may


cover cracks from the previous layers, improving the overall film quality. Each layer
will become poorly soluble in the non-polar solvents used to deposit the NCs once
the ligands are exchanged, which allows this multilayer approach. The procedure is
typically repeated around 10 times to form a 200 nm thick film. Beyond this, the
film quality may suffer.

7.2.2 Ink Preparation

In addition to the solid-state ligand exchange, the procedure can also be conducted
in the liquid phase. Ions (from HgCl2 and mercaptoethanol in the case of HgTe NCs
[22, 23]) are dissolved in a polar solvent such as dimethyl formamide. This ion solu-
tion is mixed with the NCs dissolved in their immiscible non-polar solvent. As the
two solutions are stirred, the native ligands are removed, and the ions cap the NC
surface. The NCs are now charged and thus soluble in the polar solvent. As short
ligands are often polar, this method can result in an even thinner barrier than the
solid-state method, further increasing carrier transport. Mobilities of 1 cm2 V−1 s−1
are commonly obtained [20] (Fig. 7.3b) with record values of above 100 cm2 V−1 s−1
[25], albeit with an additional sintering step. This improved mobility has led to an
increased specific detectivity (D*) in photoconductive devices [24], see Fig. 7.3c.
The constraint of this approach is that the boiling point of a polar solvent is gener-
ally high, requiring the preparation of highly concentrated solutions of NCs (>200 g.
L−1 typically) for practical removal. By doing so, however, it is possible to deposit
a film of NCs with a tunable thickness in the 100–600 nm range (see Fig. 7.3a)
through a single step of deposition, which considerably eases the fabrication pro-
cess when compared to the solid-state ligand exchange.

Fig. 7.3 Benefit of a thin film prepared from an ink. (a) SEM image with false colors of a diode
including a HgTe NC thin film obtained from a single deposition step. Part (a) is adapted with
permission from Ref. [23]. Copyright 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
(b) Transfer curve (i.e., drain current as a function of the applied gate bias under constant drain
source bias) for a HgTe NC thin film obtained from an ink solution. (c) Specific detectivity for a
HgTe NC thin film obtained from an ink solution and from the same nanocrystal film processed
from a solid-state ligand exchange strategy. Parts (b) and (c) are adapted with permission from Ref.
[24]. (Copyright 2010 American Chemical Society)
160 E. Lhuillier et al.

7.3 From Proof of Concept to High Performances Sensors

Now that the strategies to transform a NC solution into a conductive film have been
established, the next step is to design a device with a geometry that allows for
photoconduction.

7.3.1 Photoconductive Devices

While in the case of CdSe, the first demonstration of charge conduction in a NC


array started around the year 2000 [11], the investigation in the case of mercury
chalcogenides started in 2007 with the pioneering work from the Heiss [9] and Kim
[26, 27] groups. Planar photoconductive geometry is certainly the easiest strategy to
probe transport. This geometry appears more tolerant to defects (to NC aggregation
in particular) than the vertical geometry discussed later. A simple implementation is
the use of interdigitated electrodes, which can be seen in Fig. 7.4a. At room tem-
perature, the measured IV curve for a HgTe NC film is often linear, as in the case of
a gold electrode. This is due to the fact that the Schottky barrier remains weak for
both electrons and holes, because of the narrow band-gap nature of HgTe NCs.
Upon illumination, an increase in the current at a given bias is observed (i.e., a rise
of the IV curve slope, see Fig. 7.4b). For a band gap at 6000 cm−1 (720 meV), the
current modulation induced by the light can be large (factor of 10) but decreases as
the band gap is reduced (just a few percent for a 2000 cm−1 (250 meV)) band gap).
These numbers have to be considered as order of magnitude only, since the exact
current modulation depends on the incident light power, the device geometry and
the thin film preparation. In general, the magnitude of the photocurrent does not
scale linearly with the incident light irradiance in NC films. Generally, a higher
power is associated with a weaker response (relative to the input), even in a regime
where the illumination leads to less than 1 carrier per particle, see Fig. 7.4c.
Note that in photoconductive mode, the noise [30] has always been reported to
be 1/f noise [31–33] limited, see Fig. 7.4d. In this case, the Hooge’s empirical law
[32] provides a scaling law to connect the noise magnitude (SI) to the current flow,
I2
SI  . Here, α is a proportionality constant, f is the signal frequency, I is the
Nf
current and N is the number of free charge carriers. It is worth pointing out that the
α value is typically significantly larger than the value measured in the bulk [30]. The
value of the responsivity (Rλ) in this device configuration ranges from 10 μA·W−1 to

Fig. 7.4 (continued) Responsivity of a HgTe NC thin film under illumination (λ = 1.55 μm) as a
function of the laser power used for illumination for various operating temperature. (d) Noise cur-
rent spectral density for a HgTe NC thin film as a function of the signal frequency and for various
operating temperature. (e) Specific detectivity (D*) as a function of the electrode spacing for a
HgTe NC thin film (measured at 200 K for 1.55 μm illumination and for a signal at 1 kHz). (f)
Schematic of a nanotrench device used to probe photoconduction at the few tens of nm scale (i.e.,
few NC size). (Parts (b) and (f) are adapted from Ref. [29])
7 Infrared Sensing Using Mercury Chalcogenide Nanocrystals 161

Fig. 7.4 Photoconductive device from a HgTe NC thin film. (a) Schematic of a HgTe NC thin film
deposited onto interdigitated electrodes. Parts (a) is adapted with permission from Ref. [28].
Copyright 2019: American Chemical Society. (b) I-V curves for a HgTe NC (λcut-off = 2 μm) thin
film under dark condition and under illumination (λ = 1.55 μm) at room temperature. (c)Fig. 7.4
162 E. Lhuillier et al.

100 mA·W−1, depending on the exact geometry, film thickness, and doping level. An
interesting recent development is to reduce the size of the device down to the range
of the the carrier diffusion length, see Fig. 7.4f. Chu et al. [29] have demonstrated
an increase in performance with this geometry, see Fig. 7.4e. Here a specific detec-
tivity above 1012 Jones (for 2.5 μm cut-off wavelength, at 1 kHz and 200 K) is
achieved for a device with a 50 nm electrode spacing. Such high performances are
due to photoconductive gain. As the transit time becomes shorter than the carrier
lifetime, one carrier will recirculate several times during the lifetime of the other to
ensure the film neutrality. This process, called photogating, enables the collection of
more than one carrier per absorbed photon. This process is not specific to HgTe NCs
and is very commonly observed in nanocrystal films, especially those with small
electrode spacing [12, 34].

7.3.2 Phototransistor

A direct evolution of the photoconductor is the phototransistor. One may take an


existing photoconductor and add a gate to tune the carrier density. The gate is used
to bring the material close to intrinsic, reducing the dark current. While for epitaxi-
ally grown semiconductor-based devices, the introduction of a gate is generally
incompatible with the growth process, the field-effect transistor is the most common
strategy to probe carrier mobility in a nanocrystal array [35, 36]. Owing to the low
mobility of a NC film, the Hall effect signal remains weak, and the field-effect mea-
surement appears to be a better tool to probe the nature and density of the majority
carriers.

7.3.2.1 Gating Technology

Traditionally, the gating of a NC film is obtained from a dielectric gate. By far, the
use of commercially available Si/SiO2 wafers is the most common strategy, see
Fig. 7.5a. The SiO2 layer is used as a dielectric [17, 37]. Upon the application of a
gate bias, charges are generated on the dielectric surface. To screen this charge, the
electrode will inject some charge with the NC array as in a capacitor. Though easy
to implement, the silica gating suffers from the limited dielectric constant of SiO2
(εSiO2 = 3.9) necessitating the a large gate bias. An alternative is to use a high-κ mate-
rial such as alumina. In the case of HgTe NCs, this has been tested with success by
Kim et al. [38, 39] and, more recently, by Chee et al. [40, 41].
When even stronger gating is required (to reduce the operating bias or to gener-
ate more charge per particle), alternative methods such as quantum paraelectric
[42], ionic glass [28, 43, 44] and electrolyte [45–47] have also been used to generate
a gate effect on a HgTe NC film. While most of gating methods are surface methods,
7 Infrared Sensing Using Mercury Chalcogenide Nanocrystals 163

Fig. 7.5 Schematic of a NC-based field-effect transistor. (a) Schematic of a back-gate transistor.
(b) Schematic of a top-gated electrolytic transistor. Parts (a, b) are adapted with permission from
Ref. [28]. (Copyright 2019: American Chemical Society)

Table 7.1 Gate technology for a HgTe NC film


Quantum
Gate technology Dielectric paraelectric Ionic glass Electrolyte
Material SiO2, Al2O3 SrTiO3 LaF3 Li based electrolyte
associated with in liquid or ion gel
this technology matrix
Basic concept Capacitance Divergence of the Ionic gating Ionic gating
formation dielectric constant in
the vicinity of the
Curie temperature

Temperature 4–300 K Below 100 K for 180–260 K 300 K


range SrTiO3
Sweep-rate range Fast Fast (several V·s−1)Intermediate Slow (1 mV·s−1)
(several V·s−1) (0.1 V·s−1)
Subthreshold 3400 mV/ 1200 mV/ 152 mV/decade
slope decade decade
Gate voltage < 60 V Up to 200 V without Up to 10 V at <3V
range (dielectric breakdown 200 K (electrochemical
breakdown) stability of the
electrolyte)

the use of an electrolyte where the ions can diffuse in the NC array enables the gat-
ing of thick films, see Fig. 7.5b. Table 7.1 summarizes the different gate technolo-
gies reported for HgTe NC film, along with their benefits and limitations.
164 E. Lhuillier et al.

7.3.2.2 Advantages of Phototransistors

The main benefit of the phototransistor compared to the bare photoconductor is the
tunability of the dark current. Note that with the size, HgTe nanocrystal experiences
a crossover from a p-type behavior for the smallest NC to an n-type behavior for the
largest particle [48, 49]. For a band gap around 2–2.5 μm, the material is ambipolar,
conducting both electrons and holes, as shown in Fig. 7.6a. The use of gating allows
the material to be operated in a state where it acts as an intrinsic semiconductor,
minimizing the dark current and improving reproducibility over photoconductors. It
is worth pointing out that the responsivity (Rλ) is also modulated by the gate bias, as
shown in Fig. 7.6b. However, the response time of the device is also affected by a
gate bias; a bias yielding a larger response will also result in a slower response. In
other words, the gate application has only a moderate effect on the gain bandwidth
product. On the other hand, it is clear that the dark current modulation (by several
orders of magnitude) is much stronger than the one observed in the responsivity
(roughly 1 decade). This means that the detectivity (D*) can be improved by gating.
Chen et al. [37] used this concept to design a 2 μm cut-off phototransistor with a
2 × 1010 jones detectivity at room temperature. Using a SrTiO3 quantum paraelectric
gate operated at low temperature (30 K), Gréboval et al. [42] have demonstrated a
2.5 μm cut-off sensor with a detectivity reaching 1012 jones. These high perfor-
mances come at the price of high driving voltage (>100 V) and very low operating
temperature.

Fig. 7.6 Performances of a HgTe NC-based phototransistor. (a) Transfer curve (dark current as a
function of the applied gate bias) for a HgTe NC thin film gated by various gate materials. (b)
Responsivity as a function of the applied gate bias for a HgTe NC thin film on a LaF3 substrate used
as a back-gate. (c) Turn-on and -off time in response to a 1.5 μm pulse of light as a function of the
applied gate bias for a HgTe NC thin film on a LaF3 substrate used as back-gate. Parts (a–c) are
adapted with permission from Ref. [28]. (Copyright 2019: American Chemical Society)
7 Infrared Sensing Using Mercury Chalcogenide Nanocrystals 165

7.4 Beyond the Control of the Dark Current

As the operation of a phototransistor involves the modulation of the charge carrier


density, this concept may readily be applied to control the doping profile in devices.
For example, it is possible to form a p-n junction through gating without material
modification. The simplest way to design a p-n junction from a field-effect transistor
is to use a single gate device and operate in a regime where the drain-source bias is
greater compared to the gate bias. This generates a homogeneous doping where the
gate-induced charge might be opposite for the drain and the source electrodes. In the
case of HgTe NC by Noumbé et al. [43], they proposed to use a HgTe NC film
coupled to graphene electrodes and a LaF3 ionic gate to achieve this. In particular,
they demonstrate that the introduction of a graphene electrode, thanks to the mate-
rial’s work function tunability and to the partial transparency to the gate-induced
electric field, enables the generation of a planar p-n junction, which is then used to
enhance charge dissociation.
Later on, to gain greater control over the carrier density landscape, Chee et al.
[40] proposed a dual gate geometry, see Fig. 7.7a, b. Using two independent gates,

Fig. 7.7 Gate induced planar p-n junction. (a) Schematic of a dual gated HgTe NC thin film. (b)
False color microscopy image of the device depicted in part (a). (c) (resp (d)) IV curve while the
device is operated in the transistor (resp diode) mode VG1 = VG2 (resp. VG1 = −VG2). Parts (a–d) are
adapted with permission from Ref. [40]. Copyright 2021: American Chemical Society. (e)
Photocurrent under 0 V drain source bias, for a dual gated device which drain and source elec-
trodes are made of graphene as a function of the two gates biases. (f) Photocurrent in response to
a 1 ns long pulse of light at 1573 nm as a function of time for a planar p-n junction and for a vertical
geometry diode, also made of HgTe NC with the same cut-off wavelength. Parts (e–f) are adapted
with permission from Ref. [41]. (Copyright 2021: American Chemical Society)
166 E. Lhuillier et al.

it is possible to generate a p-n junction between the drain source electrodes. By


carefully choosing the bias over the two gates, the device can behave as a traditional
field-effect transistor (VG1 = VG2, see Fig. 7.7c) or as a diode (VG1 = -VG2, see
Fig. 7.7d). The latter mode can then operate the device under a zero-drain source
bias to minimize the dark current, see Fig. 7.7e. A key advantage of this approach
compared to the vertical diode, described in the next paragraph, is its lower capaci-
tance, enabling faster response times. Currently, the time response of a vertical
geometry diode remains above 100 ns, while for the planar configuration
time responses as short as 3 ns have been reported [41], see Fig. 7.7f. On the other
hand, a key limitation of the phototransistor remains its low absorption. Gate effect,
for all gate types except for electrolyte, requires thin films (t < 200 nm and even
thinner), which lead to weak light absorption and thus reduced responsivity (Rλ is in
the 1–10 mA·W−1 range usually).

7.4.1 Photodiode

To date, the vertical photodiode geometry has led to the best-performing devices
based on HgTe NCs. The body of work dedicated to this type of device is much
smaller than for PbS NCs, and there is certainly still room for device improvement.
The first proposed photodiodes were very close to solar cells. The Heiss’ group
[50] proposed a stack based on HgTe NC layer surrounded by P3HT and TiO2 as the
hole and electron transporting layers, respectively. A similar device has been tested
by Jagtap et al. [51] in which HgTe NCs are coupled to a TiO2 layer. Though the
device operated as a diode (i.e. it had an asymmetric I-V curve), the performance
was modest. The use of TiO2 is certainly related to the overall weak response.
Indeed, the narrow band gap of HgTe means the TiO2 filters not only the hole dark
current but also the photoelectron current. Thus, alternative charge transport layers
have to be explored. Jagtpap et al. [48] have proposed the introduction of a unipolar
barrier made of larger band gap HgTe NCs, but the performances of the resulting
devices were still low (D* ≈ 108 Jones at room T). The first diode operated in the
background limited regime has been reported by Guyot-Sionnest et al. [52] by
stacking a CaF2 substrate coated with a NiCr layer, a HgTe NC layer behaving as an
absorbing layer and finally connected to a silver electrode. A key breakthrough in
the design of HgTe NC-based diodes has been obtained by Ackerman et al. [53]
where the authors have introduced a hole extraction layer made of Ag2Te NC that is
cation exchanged with mercury, forming a layer of Ag-doped HgTe/an alloy of
AgHgTe with an unclear stoichiometry. A schematic of such a diode is shown in
Fig. 7.8a, while the band diagram is shown in Fig. 7.8b. The associated IV curve of
such a device is clearly rectifying, as shown in Fig. 7.8c. This layer seems to be
well-suited to design a diode using both SWIR (short-wave infrared) [54, 55] and
MWIR (mid-wave infrared) [53] HgTe NCs. Though for the mid-wave, the
7 Infrared Sensing Using Mercury Chalcogenide Nanocrystals 167

Fig. 7.8 HgTe NC-based vertical geometry photodiode. (a) Schematic of a HgTe NC-based verti-
cal geometry photodiode which hole transport layer is based on Ag2Te NCs. (b) Energy profile
determined by photoemission for the device depicted in part (a). The dashed line represents the
Fermi level. (c) IV curve in the dark and under illumination by a blackbody for an extended short-­
wave operating diode based on the structure depicted in part (a). Part (c) is adapted with permis-
sion from Ref. [53]. Copyright 2018: American Chemical Society. (d) Specific detectivity as a
function of the cut-off wavelength for the device depicted in part (a) and its comparison with
commercial device performances. Part (d) is adapted with permission from Ref. [54]. Copyright
2020: AIP Publishing. (e) Time-response to a 1 ns pulse of light from the device depicted in part
(a). Parts (a) and € adapted with permission from Ref. [55]. (Copyright 2021: WILEY-VCH Verlag
GmbH & Co. KGaA)
168 E. Lhuillier et al.

performance appears higher when the material is also coupled to an electron trans-
port layer made of Bi2Se3 [56]. Fabrication of such diodes also considerably benefits
from the development of the liquid phase ligand exchange. This process leads to the
formation of a photoconductive ink that can easily be deposited as thick layers
(200–500 nm thick film) with a single spin coating step [23, 55]. In the short-wave
infrared, the responsivity of such diodes reaches 0.5 A·W−1 and can even be above
1 A.W−1 if the diode is coupled to a light resonator [55, 57]. This corresponds to a
specific detectivity in the high 1010 Jones at room temperature in the extended short-­
wave infrared, see Fig. 7.8d [54]. In the MWIR, D* reaching 4 × 1011 Jones is
achieved at 80 K [57]. The time-response of this diode is generally measured in the
100 ns to 1 μs range depending on the actual pixel size, see Fig. 7.8e.
Considerable progress has been made in the performance of HgTe-based photo-
diodes. As vertical geometries have required thin devices, these photodiodes have
benefited greatly from the introduction of light management strategies, as discussed
in the next section. The main challenge going forward is the transfer of such diodes
in focal plane arrays.

7.4.2 Performance Comparison

To compare the performances of devices obtained using various geometries, we


have summarized in Table 7.2 and Fig. 7.9 the performance of HgTe NC-based
devices. Since most of the devices obtained operate in the short-wave infrared, we
have chosen to focus on this spectral range. At present, MWIR devices have much
lower performance than their SWIR counterparts, particularly at room-temperature
where these devices have the potential to lead to low-cost, high-performance sen-
sors. Therefore, MWIR detectors will not be discussed here.
Note that intraband absorbing materials (mostly HgS and HgSe) are discussed
separately later in this chapter. Table 7.2 provides typical figures of merit for light
sensors based on HgTe NCs and correlates them with the device type and geom-
etry. Figure 7.9 shows that the responsivities of these devices are now commonly
in the 0.2–2 A.W−1 range, which corresponds to external quantum efficiencies
between 10 and 100%. Devices with gains larger than unity have also been
reported in the photoconductive geometry and with high detectivity (>1012 jones
for 200 K operation and 2.5 μm cut-off wavelength), meaning that high-­performing
devices are not limited to photodiodes. However, it remains that photodiodes con-
stitute most of the best-performing devices. The number of results in the near-IR
is weak and device performances are generally poor compared to those reported
for PbS NC-based [66] devices or InGaAs technology. The toxicity of the material
and the current level of performance for HgTe NC based devices limit their scope
of application to a spectral range where it has no highly developed competitor (i.e.
above 1.7 μm).
7 Infrared Sensing Using Mercury Chalcogenide Nanocrystals 169

Table 7.2 Figures of merit for SWIR HgTe NC-based light sensors
Cut-­
off λ Operating Rλ Response Toper
(μm) mode (A.W−1) time D* (Jones) (K) Specific feature References
2.5 PC 0.1 10 μs 3.5 × 1010 230 As2S3 surface [17]
chemistry
2.5 PC 1000 20 μs 2 × 1012 200 Nanotrench [29]
2.5 PC 150 1.5 ms 6 × 108 80 HgTe decorated [58]
graphene channel
2.4 PC 0.9 264 μs 8 × 109 300 Spray coating with [59]
decay time patterning
2.4 PC 0.22 2.2 ms 3.5 × 108 300 Multicolor pixel [60]
2.5 PT 6.5 × 10−3 10 μs 109 220 Graphene [43]
electrode
2.5 PT 2.0 × 10−3 14 μs 1012 30 STO [42]
gate+resonator
2 PT <0.5 ≈10 μs 3 × 1010 300 SiO2 back gate [37]
2.4 PT 1 1.5 μs 1010 300 Hybrid polymer: [61]
HgTe
2.5 PT – 15 ms – 300 Doped-graphene/ [62]
HgTe
2.5 PT 0.08 10 μs >108 250 Planar pn junction [40]
<1010 based on dual gate
2.5 PD 2.5 × 10−3 370 ns 3 × 109 300 HgTe ink [23]
2.5 PD 0.25 260 ns 3 × 1010 300 Flexible substrate [56]
decay time (without
cavity)
7.5 × 1010
(with
cavity)
2.2 PD 1 1.4 μs 6 × 1010 300 HgCl2 treatment [54]
decay time
1.8 PD 0.13 110 ns 2 × 1010 300 With resonator [55]
2.5 PD 0.6 – 4 × 1011 85 Resonator grating [57]
+ fabry perot
2.4 PD 0.45 13 ns 1010 300 Si/graphene/HgTe [63]
2.5 PD 0.28 2.5 μs 6 × 1010 300 Bi2Se3/HgTe/ [64]
Ag2Te
2 PD 0.8 170 ns 9 × 1011 200 CdSe/HgTe/Ag2Te [65]
9 × 1010 300

7.5 Light Management in HgX Nanocrystal Films

The design of NC-based light sensors faces a trade-off. On the one hand, thick films
are required to absorb most of the incident light. The absorption depth of HgTe NC
with 2.5 μm cut-off wavelength is several μm, and a 200 nm thick film only absorbs
10% of the incident light. On the other hand, these films exhibit low carrier mobility
170 E. Lhuillier et al.

Fig. 7.9 HgTe NC based performances for short wave infrared sensing. Specific detectivity as a
function of device responsivity for HgTe NC-based device operated in the SWIR

and so have a short diffusion length [29] (<50 nm). As a result, charge collection is
not efficient over several μm. In addition, the fabrication of such thick films using
NCs remains extremely challenging. So, in order to help the low absorption of the
thinner films, light management strategies have been developed in order to “focus”
the light on a thin semiconductor slab Here, we would like to review some of the
recent developments in this direction [67].

7.5.1 Enhancement of Absorption

Many of the concepts that are being explored for NCs were first demonstrated in
epitaxially grown semiconductors. The use of a grating to tune the light-matter cou-
pling is well-established for quantum well-based infrared detectors (QWIP). Here,
the grating is etched on the top of the pixel to break the selection rule forbidding
absorption under normal incidence. The first attempt to tune light-matter coupling
in a HgTe NC array was made by Chen et al. [68], where they introduced gold
nanorods in a diode geometry, see Fig. 7.10a. The gold nanorods have one reso-
nance in the visible (around 530 nm) and a second one in the near-infrared, its
energy is driven by the rod length. The latter resonance was spectrally matched with
the NC absorption. In the vicinity of a rod’s tip, the electromagnetic field is locally
enhanced, boosting light absorption. When coupling excitons and plasmons in this
way, the absorption enhancement must occur within the nanocrystal rather than in
the metal used to induce the plasmonic behavior; otherwise, substantial thermal
losses result. Therefore, in their design, the authors have also been careful not to
directly locate the HgTe NCs on the top of the gold in order to avoid any exciton
7 Infrared Sensing Using Mercury Chalcogenide Nanocrystals 171

Fig. 7.10 Strategy to enhance the light absorption in HgTe NC-based devices. (a) Schematic of
HgTe NC based photodiode coupled to gold nanorods. Part (a) is adapted with permission from
Ref. [68]. Copyright 2014: American Chemical Society. (b) Microscopy image of interdigitated
electrode in which the substrate between the electrode is functionalized by a plasmonic array. Part
(b) is adapted with permission from Ref. [69]. Copyright 2017: AIP publishing. (c) Schematic of
a HgTe NC film coupled to a guided more resonator. Part (c) is adapted with permission from Ref.
[70]. (Copyright 2019: American Chemical Society)

quenching. While this first strategy is based on an all-colloidal-material approach,


most of the following efforts have focused on incorporating conventional materials
through microfabrication. Note that in order to avoid any damage to the NCs, much
of the fabrication (lithography, annealing step) is conducted before the NC deposi-
tion. This prevents interparticle sintering and possible oxidation/ligand removal.
Hopefully, in the future, such constraints will be removed with the design of more
stable NCs.
Yifat et al. [69] proposed to functionalize the area between the interdigitated
electrodes by an array of metallic dots, the period is chosen to generate a resonance
matching the NC absorption band edge, see Fig. 7.10b. By doing so, they observe a
change in the photocurrent spectrum and a three times enhancement of the absorp-
tion around the plasmonic resonance. Later on, Chu et al. [70] proposed an evolu-
tion of this strategy where an optical grating is directly used as a set of interdigitated
electrodes, see Fig. 7.10c. The benefit of this strategy is the reduction of electrode
spacing which generates photoconductive gain. While the impact on the absorption
is likewise a factor of 3, the enhancement of the photocurrent is considerably
greater: between 100 and 1000, depending on the material and the cut-off wave-
length. The same approach was later extended to other device geometries, such as
phototransistors [42] and vertical diodes [55]. Devices based on exciton-plasmon
coupling generally present a low-quality factor, which is appealing for broadband
applications such as imaging, but not for specific spectroscopic applications such as
gas sensing, where narrow band enhancement is desirable. Tang et al. [71] have
proposed to couple a HgTe NC film to a Bragg mirror and the obtained resonance is
much narrower (down to 30 cm−1). The same group has also demonstrated that verti-
cal Fabry-Perot interference can be used to enhance light absorption [57], a tech-
nique which is compatible with flexible substrates [56].
While much effort has been focused on the enhancement of the absorption, noise
reduction similarly increases the device performance. Zhu et al. [72] proposed an
approach based on size reduction. Noise in sensors scales like the device electrical
172 E. Lhuillier et al.

volume. Reducing the device size while maintaining the absorption unchanged will
thus result in greater signal-to-noise ratio. To reach this goal, they have placed HgTe
NCs in a small waveguide where the gold edges act as the device electrodes. As
devices move to pixel arrays with small pixel pitch, such approaches will become
necessary.

7.5.2 Spectral Shaping

As a first approximation, one may assume that the response spectrum of a detector
is the same as the absorption spectrum of the active layer. There are, of course,
always optical effects arising from the material interfaces as well as scattering from
non-uniformity and charge transport effects. When parts of a device are smaller than
the wavelength of the light absorbed, the geometry of the device becomes particu-
larly important in determining the shape of the response spectrum. For example,
Chu et al. [70] demonstrated a 500 cm−1 shift of the spectral response of a HgTe NC
film at 2.5 μm by tuning the period of the grating in a structure such as the one
depicted in Fig. 7.10c.
We have already mentioned that the absorption cross-section can be focused on
a very narrow band using a Bragg mirror. For imaging applications, it is also inter-
esting to achieve broadband absorption enhancement. The latter can be obtained by
combining several broad individual resonances, see Fig. 7.11a. Gréboval et al. [42],
for instance designed a grating that shows two shifted guided-mode resonances.
Tang et al. [57] chose to combine a Fabry-Perot mode with a metallic grating to
reach the same goal.

Fig. 7.11 Spectral shaping of the photoresponse through a coupling to a resonator. (a) Absorption
spectrum of a HgTe NC film coupled to three resonances to achieve broadband absorption in the
near- and short- wave infrared. Part a is adapted with permission from Ref. [42]. Copyright 2021:
American Chemical Society. (b) Photocurrent spectra of HgTe thin film integrated in the device
depicted as inset for various applied biases. (c) Energy of the absorption peak as a function of the
applied bias for various temperature for the device of part (b). Parts (b) and (c) are adapted with
permission from Ref. [73]. (Copyright 2021: American Chemical Society)
7 Infrared Sensing Using Mercury Chalcogenide Nanocrystals 173

All the devices discussed so far have had a single, unchanging response spec-
trum. Active reconfigurability would be even more interesting. This can be achieved
at the device level [74], for example, by stacking diodes. Tang et al. [64] proposed
to stack a SWIR and a MWIR diode where the spectral response of the full device
is determined by the bias polarity. Dang et al. [73] demonstrated another form of
active device, showing a shift of the spectral response with bias. Interestingly, the
spectral response shows a blueshift with bias (Fig. 7.11b, c), allowing the exclusion
of the Stark effect as a possible mechanism to explain the observed bias-tunable
photoresponse.
To observe their bias-tunable photoresponse, they included the NC layer into a
cavity, as shown in the inset of Fig. 7.11b. The role of the cavity is to generate a
spatially inhomogeneous electromagnetic map. Then the charge collection length
will be tuned by the bias. Under a low electric field, the charges are collected only
in the very vicinity of the electrode, while under a larger field, longer distance
charge collection can occur. A tunable photoresponse can be obtained if the spectral
response within the cavity differs close and far from the electrode.

7.6 From Single Pixel to Focal Plane Array

Now that single-element detectors have reached a certain maturity, a new challenge
will be to transfer the technology to focal plane arrays. NCs present several poten-
tial advantages compared to epitaxially grown semiconductors. They are inexpen-
sive to synthesize, and there is no need to epitaxially match them to an expensive
wafer. In the case of InGaAs, the III-V layer is responsible for 1/3 of the final cost.
Moreover, NCs are not restricted to backside illumination through an absorbing
substrate, so they can be used for broadband detection (i.e., from band edge to ultra-
violet). Thus, an infrared NC-based camera is also responsive in the visible range,
contrary to InGaAs grown on InP. The solution-processed nature of NCs may also
ease the hybridization to read-out circuit. Epitaxially grown semiconductors are
generally coupled to a CMOS circuit with the use of an array of Indium bumps, see
Fig. 7.12a. NCs can be deposited directly on the top of the read-out-circuit, see
Fig. 7.12b. Aside from removing a complex step of fabrication, NCs are also highly
promising to design smaller pixel pitch closer to the diffraction limit and improve
the image quality. Using near-infrared or short-wave infrared absorbing PbS NCs, a
pixel pitch as small as 1.8 μm have been reported [77]. In the case of HgTe, both
SWIR [78] (2 μm cut-off, see Fig. 7.12c) and MWIR [79] (5 μm cut-off - see
Fig. 7.12d) focal plane arrays have been reported. Bossavit et al. [76] recently
reported a basic proof of concept of an all-nanocrystal-based communication setup
where a HgTe NC light emitting diode (LED) is imaged using a HgTe NC-based
focal plane array, see Fig. 7.12e–g. To date, only few performances have been
reported. Buurma et al. [79] reported a sub-100 mK NETD (noise equivalent tem-
perature difference) for their MWIR sensor, but 8 years later, there has been little
progress, and updated data will be of utmost interest.
174 E. Lhuillier et al.

Fig. 7.12 Imaging using a HgTe NC thin film as the active layer. (a) Schematic of an epitaxially
grown semiconductor coupled to a read-out circuit. (b) Schematic of a HgTe NC thin film coupled
to a read-out circuit. Parts (a) and (b) are adapted with permission from Ref. [7]. Copyright 2021:
American Chemical Society. (c) extended short-wave infrared image obtained from a HgTe NC
thin film. (d) Mid-wave infrared image obtained from a HgTe NC thin film. Part (d) is adapted
from Ref. [75]. (e) Image of a camera which active layer is based on HgTe NCs imaging an HgTe
NC based LED (λ = 1.3 μm). (f) (resp (g)) Obtained image from the setup depicted in part (e) while
the two top left pixels are off (resp on). Parts (e–g) are adapted with permission from Ref. [76].
(Copyright 2021: WILEY-VCH Verlag GmbH & Co. KGaA)

7.7 Intraband Device

The first sections of this chapter have only been focused on HgTe NCs. The synthe-
sis and device fabrication using HgTe NCs are far more advanced compared to other
mercury chalcogenides (HgS and HgSe). Nevertheless, as stated in the volume
1 Chap. 6, HgS and HgSe NCs have demonstrated intraband absorption resulting
from their doped character [80, 81]. This absorption is the 0D counterpart of the
intersubband transition in quantum well structures. A clear signature of this intra-
band absorption is an infrared peak appearing in the mid- or long- wave infrared, see
Fig. 7.13a, which is clearly offset compared to the interband transition appearing at
higher energy. For a small size (5 nm) of HgS and HgSe NCs, this peak overlaps
well with the 3–5 μm band. Though intraband has been mostly observed in heavy
metal-containing materials (HgS [82], HgSe [74, 83–85], HgTe, PbS [86]), the fact
that it has also been observed in Ag2Se [87–91] may raise clear expectation to
achieve mid-infrared absorption in wider band gap materials, even greener materials.
Interestingly the intraband transition not only leads to absorption but also to
some photoconduction, first observed by Deng et al. [92]. However, an array of
7 Infrared Sensing Using Mercury Chalcogenide Nanocrystals 175

Fig. 7.13 Intraband photodetection. (a) Absorption spectrum of HgSe NCs presenting inter and
intraband absorption. (b) Current as a function of temperature for HgSe NC film and for a HgSe/
HgTe NC mixture. (c) Schematic of an intraband photodiode based on a HgSe/HgTe NC mixture
as an active layer. (d) Specific detectivity as a function of the applied voltage under various operat-
ing conditions. The inset shows the 1/f noise limited noise current spectral density. (e) Absorption
and photocurrent spectrum under different biases at 80 K from the device depicted in part (c).
(Parts (b–e) are adapted from Ref. [74])
176 E. Lhuillier et al.

intraband particles generally presents poor mid-infrared performance. The high


doping required to observe the intraband absorption generates an increase of the
dark current, a low current activation energy (i.e., much lower than the energy of the
intraband transition), and a time response that is also generally slow (ms or even s).
The slow time response of intraband devices is due to a bolometric effect. Two strat-
egies have been proposed to solve this problem. Deng et al. [92] observed that for a
doping of exactly n = 2 carriers per nanocrystal, the response is faster (down to
200 ns) and the dark current is reduced (by a factor 30). In this case, the doping
magnitude has to be tuned by carefully choosing the proper capping ligand. An
alternative approach has been proposed by Livache et al. [74], where the author
designed a structure close to a dye-sensitized solar cell. The intraband absorption is
obtained through HgSe NCs mixed with HgTe NCs (HgSe/HgTe core shell leads to
a similar result [93]). Using this strategy, the activation energy of transport is con-
siderably increased, see Fig. 7.13b, and the time response also shortened. To date,
the best performances using intraband materials have led to a detectivity in the 109
Jones [74, 92] for operation at 80 K (5 μm cut off wavelength), see Fig. 7.13d. There
has only been one report of an intraband photodiode thus far, see Fig. 7.13c.
One of the technological issues raised by mid-infrared absorption, such as the
one resulting from intraband transition is the lighting through the substrate and the
design of partially transparent electrodes at wavelengths where ordinary transparent
conductive oxide absorption becomes strong. One proposed strategy relies on a sap-
phire substrate coupled with a metallic grid. The obtained diode offers a reconfigu-
rable response with bias where the weight of the intraband compared to the interband
absorption (coming from other layers) can be strongly tuned, see Fig. 7.13e.

7.8 Conclusion

Over the last decade, there have been considerable progress in the field of IR sens-
ing using colloidal nanocrystals. The level of device complexity has been signifi-
cantly raised, particularly in the achievement of carrier density control and the
design of light-matter coupling. Devices are no longer limited to a single pixel, and
focal plane array integration is now a reality. Any advancements in single-element
detection will become future directions for focal plane array development.
A next-level challenge relates to modeling, which is mandatory to conduct ratio-
nal design of device. To date, too many material-related parameters (doping level,
band bending, trap cross section and density, carrier relaxation pathway and dynam-
ics…) remain unknown to accurately simulate the electrical behavior of a complete
device. This is in addition to the general problem of the reproducibility of such
parameters. Thus, progress at the device level cannot be decorrelated from a deeper
material investigation.
7 Infrared Sensing Using Mercury Chalcogenide Nanocrystals 177

References

1. Pietryga JM, Park Y-S, Lim J, Fidler AF, Bae WK, Brovelli S et al (2016) Spectroscopic
and device aspects of nanocrystal quantum dots. Chem Rev 116:10513–10622. https://doi.
org/10.1021/acs.chemrev.6b00169
2. Wood V, Bulović V (2010) Colloidal quantum dot light-emitting devices. Nano Rev 1:5202.
https://doi.org/10.3402/nano.v1i0.5202
3. Dubertret B, Skourides P, Norris DJ, Noireaux V, Brivanlou AH, Libchaber A (2002) In vivo
imaging of quantum dots encapsulated in phospholipid micelles. Science 298:1759–1762.
https://doi.org/10.1126/science.1077194
4. Schaller RD, Klimov VI (2004) High efficiency carrier multiplication in PbSe nanocrystals:
implications for solar energy conversion. Phys Rev Lett 92:186601. https://doi.org/10.1103/
PhysRevLett.92.186601
5. Semonin OE, Luther JM, Choi S, Chen H-Y, Gao J, Nozik AJ et al (2011) Peak external pho-
tocurrent quantum efficiency exceeding 100% via MEG in a quantum dot solar cell. Science
334:1530–1533. https://doi.org/10.1126/science.1209845
6. Pijpers JJH, Ulbricht R, Tielrooij KJ, Osherov A, Golan Y, Delerue C et al (2009) Assessment
of carrier-multiplication efficiency in bulk PbSe and PbS. Nat Phys 5:811–814. https://doi.
org/10.1038/nphys1393
7. Gréboval C, Chu A, Goubet N, Livache C, Ithurria S, Lhuillier E (2021) Mercury chalcogenide
quantum dots: material perspective for device integration. Chem Rev 121:3627–3700. https://
doi.org/10.1021/acs.chemrev.0c01120
8. Rogach A, Kershaw SV, Burt M, Harrison MT, Kornowski A, Eychmüller A et al (1999)
Colloidally prepared HgTe nanocrystals with strong room-temperature infrared lumines-
cence. Adv Mater 11:552–555. https://doi.org/10.1002/(SICI)1521-­4095(199905)11:7<552::
AID-­ADMA552>3.0.CO;2-­Q
9. Boeberl M, Kovalenko M, Gamerith S, List-Kratochvil E, Heiss W (2007) Inkjet-printed nano-
crystal photodetectors operating up to 3 μm wavelengths. Adv Mater 19:3574–3578. https://
doi.org/10.1002/adma.200700111
10. Murray CB, Norris DJ, Bawendi MG (1993) Synthesis and characterization of nearly mono-
disperse CdE (E = sulfur, selenium, tellurium) semiconductor nanocrystallites. J Am Chem
Soc 115:8706–8715. https://doi.org/10.1021/ja00072a025
11. Leatherdale CA, Kagan CR, Morgan NY, Empedocles SA, Kastner MA, Bawendi MG (2000)
Photoconductivity in CdSe quantum dot solids. Phys Rev B 62:2669–2680. https://doi.
org/10.1103/PhysRevB.62.2669
12. Wang H, Lhuillier E, Yu Q, Zimmers A, Dubertret B, Ulysse C et al (2017) Transport in a
single self-doped nanocrystal. ACS Nano 11. https://doi.org/10.1021/acsnano.6b07898
13. Kramer IJ, Minor JC, Moreno-Bautista G, Rollny L, Kanjanaboos P, Kopilovic D et al (2015)
Efficient spray-coated colloidal quantum dot solar cells. Adv Mater 27:116–121. https://doi.
org/10.1002/adma.201403281
14. Zhang S, Chen M, Mu G, Li J, Hao Q, Tang X (2021) Spray-stencil lithography enabled large-­
scale fabrication of multispectral colloidal quantum-dot infrared detectors. Adv Mater Technol
7:2101132. https://doi.org/10.1002/admt.202101132
15. Lhuillier E, Hease P, Ithurria S, Dubertret B (2014) Selective electrophoretic deposition of
CdSe Nanoplatelets. Chem Mater 26:4514–4520. https://doi.org/10.1021/cm501713s
16. Liu Y, Gibbs M, Puthussery J, Gaik S, Ihly R, Hillhouse HW et al (2010) Dependence of
carrier mobility on nanocrystal size and ligand length in PbSe nanocrystal solids. Nano Lett
10:1960–1969. https://doi.org/10.1021/nl101284k
17. Lhuillier E, Keuleyan S, Zolotavin P, Guyot-Sionnest P (2013) Mid-infrared HgTe/As2S3
field effect transistors and photodetectors. Adv Mater 25:137–141. https://doi.org/10.1002/
adma.201203012
178 E. Lhuillier et al.

18. Kovalenko MV, Scheele M, Talapin DV (2009) Colloidal nanocrystals with molecu-
lar metal chalcogenide surface ligands. Science 324:1417–1420. https://doi.org/10.1126/
science.1170524
19. Tang J, Kemp KW, Hoogland S, Jeong KS, Liu H, Levina L et al (2011) Colloidal-quantum-­
dot photovoltaics using atomic-ligand passivation. Nat Mater 10:765–771. https://doi.
org/10.1038/nmat3118
20. Nag A, Kovalenko MV, Lee J-S, Liu W, Spokoyny B, Talapin DV (2011) Metal-free inorganic
ligands for colloidal nanocrystals: S2−, HS−, Se2−, HSe−, Te2−, HTe−, TeS32−, OH−, and NH2− as
surface ligands. J Am Chem Soc 133:10612–10620. https://doi.org/10.1021/ja2029415
21. Yakunin S, Dirin DN, Protesescu L, Sytnyk M, Tollabimazraehno S, Humer M et al (2014)
High infrared photoconductivity in films of arsenic-sulfide-encapsulated Lead-sulfide nano-
crystals. ACS Nano 8:12883–12894. https://doi.org/10.1021/nn5067478
22. Lan X, Chen M, Hudson MH, Kamysbayev V, Wang Y, Guyot-Sionnest P et al (2020) Quantum
dot solids showing state-resolved band-like transport. Nat Mater 19:323–329. https://doi.
org/10.1038/s41563-­019-­0582-­2
23. Martinez B, Ramade J, Livache C, Goubet N, Chu A, Gréboval C et al (2019) HgTe nanocrys-
tal inks for extended short-wave infrared detection. Adv Opt Mater 7:1900348. https://doi.
org/10.1002/adom.201900348
24. Chen M, Lan X, Tang X, Wang Y, Hudson MH, Talapin DV et al (2019) High carrier mobility
in HgTe quantum dot solids improves mid-IR photodetectors. ACS Photonics 6:2358–2365.
https://doi.org/10.1021/acsphotonics.9b01050
25. Dolzhnikov DS, Zhang H, Jang J, Son JS, Panthani MG, Shibata T et al (2015) Composition-­
matched molecular “solders” for semiconductors. Science 347:425–428. https://doi.
org/10.1126/science.1260501
26. Kim S, Kim T, Im SH, Seok SI, Kim KW, Kim S et al (2011) Bandgap engineered monodis-
perse and stable mercury telluride quantum dots and their application for near-infrared photo-
detection. J Mater Chem 21:15232–15236. https://doi.org/10.1039/C1JM12436F
27. Seong H, Cho K, Kim S (2008) Photocurrent characteristics of solution-processed HgTe
nanoparticle thin films under the illumination of 1.3 μm wavelength light. Semicond Sci
Technol 23:5
28. Gréboval C, Noumbe U, Goubet N, Livache C, Ramade J, Qu J et al (2019) Field-effect tran-
sistor and photo-transistor of narrow-band-gap nanocrystal arrays using ionic glasses. Nano
Lett 19:3981–3986. https://doi.org/10.1021/acs.nanolett.9b01305
29. Chu A, Gréboval C, Prado Y, Majjad H, Delerue C, Dayen J-F et al (2021) Infrared photocon-
duction at the diffusion length limit in HgTe nanocrystal arrays. Nat Commun 12:1794. https://
doi.org/10.1038/s41467-­021-­21959-­x
30. Liu H, Lhuillier E, Guyot-Sionnest P (2014) 1/f noise in semiconductor and metal nanocrystal
solids. J Appl Phys 115:154309. https://doi.org/10.1063/1.4871682
31. Balandin AA (2013) Low-frequency 1/f noise in graphene devices. Nat Nanotechnol
8:549–555. https://doi.org/10.1038/nnano.2013.144
32. Hooge FN (1994) 1/f noise sources. IEEE Trans Electron Devices 41:1926–1935. https://doi.
org/10.1109/16.333808
33. Lai Y, Li H, Kim DK, Diroll BT, Murray CB, Kagan CR (2014) Low-frequency (1/f) noise
in nanocrystal field-effect transistors. ACS Nano 8:9664–9672. https://doi.org/10.1021/
nn504303b
34. Konstantatos G, Sargent EH (2007) PbS colloidal quantum dot photoconductive photodetec-
tors: transport, traps, and gain. Appl Phys Lett 91:173505. https://doi.org/10.1063/1.2800805
35. Talapin DV, Murray CB (2005) PbSe nanocrystal solids for n- and p-channel thin film field-­
effect transistors. Science 310:86–89. https://doi.org/10.1126/science.1116703
36. Hetsch F, Zhao N, Kershaw SV, Rogach AL (2013) Quantum dot field effect transistors. Mater
Today 16:312–325. https://doi.org/10.1016/j.mattod.2013.08.011
37. Chen M, Lu H, Abdelazim NM, Zhu Y, Wang Z, Ren W et al (2017) Mercury telluride quan-
tum dot based phototransistor enabling high-sensitivity room-temperature Photodetection at
2000 nm. ACS Nano 11:5614–5622. https://doi.org/10.1021/acsnano.7b00972
7 Infrared Sensing Using Mercury Chalcogenide Nanocrystals 179

38. Kim H, Cho K, Kim D-W, Lee H-R, Kim S (2006) Bottom- and top-gate field-effect thin-film
transistors with p channels of sintered HgTe nanocrystals. Appl Phys Lett 89:173107. https://
doi.org/10.1063/1.2364153
39. Kim D-W, Jang J, Kim H, Cho K, Kim S (2008) Electrical characteristics of HgTe nanocrystal-­
based thin film transistors fabricated on flexible plastic substrates. Thin Solid Films
516:7715–7719. https://doi.org/10.1016/j.tsf.2008.04.044
40. Chee S-S, Gréboval C, Magalhaes DV, Ramade J, Chu A, Qu J et al (2021) Correlating struc-
ture and detection properties in HgTe nanocrystal films. Nano Lett 21:4145–4151. https://doi.
org/10.1021/acs.nanolett.0c04346
41. Gréboval C, Dabard C, Konstantinov N, Cavallo M, Chee S-S, Chu A et al (2021) Split-gate
photodiode based on graphene/HgTe Heterostructures with a few nanosecond Photoresponse.
ACS Appl Electron Mater. https://doi.org/10.1021/acsaelm.1c00442
42. Gréboval C, Chu A, Magalhaes DV, Ramade J, Qu J, Rastogi P et al (2021) Ferroelectric gating
of narrow band-gap nanocrystal arrays with enhanced light–matter coupling. ACS Photonics
8:259–268. https://doi.org/10.1021/acsphotonics.0c01464
43. Noumbé UN, Gréboval C, Livache C, Chu A, Majjad H, Parra López LE et al (2020)
Reconfigurable 2D/0D p–n graphene/HgTe nanocrystal Heterostructure for infrared detection.
ACS Nano 14:4567–4576. https://doi.org/10.1021/acsnano.0c00103
44. Gréboval C, Noumbé UN, Chu A, Prado Y, Khalili A, Dabard C et al (2020) Gate tunable verti-
cal geometry phototransistor based on infrared HgTe nanocrystals. Appl Phys Lett 117:251104.
https://doi.org/10.1063/5.0032622
45. Liu H, Keuleyan S, Guyot-Sionnest P (2012) n- and p-type HgTe quantum dot films. J Phys
Chem C 116:1344–1349. https://doi.org/10.1021/jp2109169
46. Livache C, Izquierdo E, Martinez B, Dufour M, Pierucci D, Keuleyan S et al (2017)
Charge dynamics and Optolectronic properties in HgTe colloidal quantum Wells. Nano Lett
17:4067–4074. https://doi.org/10.1021/acs.nanolett.7b00683
47. Martinez B, Livache C, Goubet N, Jagtap A, Cruguel H, Ouerghi A et al (2018) Probing charge
carrier dynamics to unveil the role of surface ligands in HgTe narrow band gap nanocrystals. J
Phys Chem C 122:859–865. https://doi.org/10.1021/acs.jpcc.7b09972
48. Jagtap A, Martinez B, Goubet N, Chu A, Livache C, Gréboval C et al (2018) Design of a
Unipolar Barrier for a nanocrystal-based short-wave infrared photodiode. ACS Photonics
5:4569–4576. https://doi.org/10.1021/acsphotonics.8b01032
49. Goubet N, Jagtap A, Livache C, Martinez B, Portalès H, Xu XZ et al (2018) Terahertz HgTe
nanocrystals: beyond confinement. J Am Chem Soc 140:5033–5036. https://doi.org/10.1021/
jacs.8b02039
50. Günes S, Neugebauer H, Sariciftci NS, Roither J, Kovalenko M, Pillwein G et al (2006)
Hybrid solar cells using HgTe nanocrystals and Nanoporous TiO2 electrodes. Adv Funct Mater
16:1095–1099. https://doi.org/10.1002/adfm.200500638
51. Jagtap A, Goubet N, Livache C, Chu A, Martinez B, Gréboval C et al (2018) Short wave
infrared devices based on HgTe nanocrystals with air stable performances. J Phys Chem C
122:14979–14985. https://doi.org/10.1021/acs.jpcc.8b03276
52. Guyot-Sionnest P, Roberts JA (2015) Background limited mid-infrared photodetection
with photovoltaic HgTe colloidal quantum dots. Appl Phys Lett 107:253104. https://doi.
org/10.1063/1.4938135
53. Ackerman MM, Tang X, Guyot-Sionnest P (2018) Fast and sensitive colloidal quantum
dot mid-wave infrared photodetectors. ACS Nano 12:7264–7271. https://doi.org/10.1021/
acsnano.8b03425
54. Ackerman MM, Chen M, Guyot-Sionnest P (2020) HgTe colloidal quantum dot photodi-
odes for extended short-wave infrared detection. Appl Phys Lett 116:083502. https://doi.
org/10.1063/1.5143252
55. Rastogi P, Chu A, Dang TH, Prado Y, Gréboval C, Qu J et al (2021) Complex optical index
of HgTe nanocrystal infrared thin films and its use for short wave infrared photodiode design.
Adv Opt Mater 9:2002066. https://doi.org/10.1002/adom.202002066
180 E. Lhuillier et al.

56. Tang X, Ackerman MM, Shen G, Guyot-Sionnest P (2019) Towards infrared electronic eyes:
flexible colloidal quantum dot photovoltaic detectors enhanced by resonant cavity. Small
15:1804920. https://doi.org/10.1002/smll.201804920
57. Tang X, Ackerman MM, Guyot-Sionnest P (2018) Thermal imaging with Plasmon resonance
enhanced HgTe colloidal quantum dot photovoltaic devices. ACS Nano 12:7362–7370. https://
doi.org/10.1021/acsnano.8b03871
58. Grotevent MJ, Hail CU, Yakunin S, Bachmann D, Calame M, Poulikakos D et al (2021)
Colloidal HgTe quantum dot/graphene phototransistor with a spectral sensitivity beyond 3 μm.
Adv Sci 8:2003360. https://doi.org/10.1002/advs.202003360
59. Cryer ME, Halpert JE (2018) 300 nm spectral resolution in the mid-infrared with robust,
high responsivity flexible colloidal quantum dot devices at room temperature. ACS Photonics
5:3009–3015. https://doi.org/10.1021/acsphotonics.8b00738
60. Cryer ME, Browning LA, Plank NOV, Halpert JE (2020) Large Photogain in multicolor
nanocrystal photodetector arrays enabling room-temperature detection of targets above 100
°C. ACS Photonics 7:3078–3085. https://doi.org/10.1021/acsphotonics.0c01156
61. Dong Y, Chen M, Yiu WK, Zhu Q, Zhou G, Kershaw SV et al (2020) Solution processed
hybrid polymer: HgTe quantum dot phototransistor with high sensitivity and fast infrared
response up to 2400 nm at room temperature. Adv Sci 7:2000068. https://doi.org/10.1002/
advs.202000068
62. Tang X, Lai KWC (2019) Graphene/HgTe quantum-dot photodetectors with gate-tunable infra-
red response. ACS Appl Nano Mater 2:6701–6706. https://doi.org/10.1021/acsanm.9b01587
63. Tang X, Chen M, Kamath A, Ackerman MM, Guyot-Sionnest P (2020) Colloidal quantum-­
dots/graphene/silicon Dual-Channel detection of visible light and short-wave infrared. ACS
Photonics 7:1117–1121. https://doi.org/10.1021/acsphotonics.0c00247
64. Tang X, Ackerman MM, Chen M, Guyot-Sionnest P (2019) Dual-band infrared imaging
using stacked colloidal quantum dot photodiodes. Nat Photonics 13:277–282. https://doi.
org/10.1038/s41566-­019-­0362-­1
65. Rastogi P, Izquierdo E, Gréboval C, Cavallo M, Chu A, Dang TH et al (2022) Extended short-­
wave photodiode based on CdSe/HgTe/Ag2Te stack with Unity internal efficiency. submitted
66. Vafaie M, Fan JZ, Morteza Najarian A, Ouellette O, Sagar LK, Bertens K et al (2021)
Colloidal quantum dot photodetectors with 10-ns response time and 80% quantum efficiency
at 1,550 nm. Matter 4:1042–1053. https://doi.org/10.1016/j.matt.2020.12.017
67. Chen M, Lu L, Yu H, Li C, Zhao N (2021) Integration of colloidal quantum dots with photonic
structures for optoelectronic and optical devices. Adv Sci 8:2101560. https://doi.org/10.1002/
advs.202101560
68. Chen M, Shao L, Kershaw SV, Yu H, Wang J, Rogach AL et al (2014) Photocurrent enhance-
ment of HgTe quantum dot photodiodes by Plasmonic gold Nanorod structures. ACS Nano
8:8208–8216. https://doi.org/10.1021/nn502510u
69. Yifat Y, Ackerman M, Guyot-Sionnest P (2017) Mid-IR colloidal quantum dot detec-
tors enhanced by optical nano-antennas. Appl Phys Lett 110:041106. https://doi.
org/10.1063/1.4975058
70. Chu A, Gréboval C, Goubet N, Martinez B, Livache C, Qu J et al (2019) Near Unity absorption
in nanocrystal based short wave infrared photodetectors using guided mode resonators. ACS
Photonics 6:2553–2561. https://doi.org/10.1021/acsphotonics.9b01015
71. Tang X, Ackerman MM, Guyot-Sionnest P (2019) Acquisition of Hyperspectral Data with col-
loidal quantum dots. Laser Photonics Rev 13:1900165. https://doi.org/10.1002/lpor.201900165
72. Zhu B, Chen M, Zhu Q, Zhou G, Abdelazim NM, Zhou W et al (2019) Integrated Plasmonic
infrared photodetector based on colloidal HgTe quantum dots. Adv Mater Technol 4:1900354.
https://doi.org/10.1002/admt.201900354
73. Dang TH, Vasanelli A, Todorov Y, Sirtori C, Prado Y, Chu A et al (2021) Bias tunable spectral
response of nanocrystal array in a plasmonic cavity. Nano Lett 21:6671–6677. https://doi.
org/10.1021/acs.nanolett.1c02193
74. Livache C, Martinez B, Goubet N, Gréboval C, Qu J, Chu A et al (2019) A colloidal quantum
dot infrared photodetector and its use for intraband detection. Nat Commun 10:2125. https://
doi.org/10.1038/s41467-­019-­10170-­8
7 Infrared Sensing Using Mercury Chalcogenide Nanocrystals 181

75. Lhuillier E, Guyot-Sionnest P (2017) Recent progresses in mid infrared nanocrystal opto-
electronics. IEEE J Sel Top Quantum Electron 23:6000208. https://doi.org/10.1109/
JSTQE.2017.2690838
76. Bossavit E, Qu J, Abadie C, Dabard C, Dang TH, Izquierdo E et al (2021) Optimized infra-
red LED and its use in an all-HgTe nanocrystal-based active imaging setup. Adv Opt Mater
2101755. https://doi.org/10.1002/adom.202101755
77. SWIR cost cut: Imec achieves 1.82 μm pixels | Imaging and Machine Vision Europe. https://
www.imveurope.com/feature/swir-­cost-­cut-­imec-­achieves-­182-­m-­pixels. Accessed 15
Nov 2021
78. Chu A, Martinez B, Ferré S, Noguier V, Gréboval C, Livache C et al (2019) HgTe nanocrys-
tals for SWIR detection and their integration up to the focal plane Array. ACS Appl Mater
Interfaces 11:33116–33123. https://doi.org/10.1021/acsami.9b09954
79. Buurma C, Pimpinella RE, Ciani AJ, Feldman JS, Grein CH, Guyot-Sionnest P (2016) MWIR
imaging with low cost colloidal quantum dot films. In: Optical sensing, imaging, and photon
counting: nanostructured devices and applications 2016. SPIE, p 993303
80. Kim J, Choi D, Jeong KS (2018) Self-doped colloidal semiconductor nanocrystals with
intraband transitions in steady state. Chem Commun 54:8435–8445. https://doi.org/10.1039/
C8CC02488J
81. Jagtap A, Livache C, Martinez B, Qu J, Chu A, Gréboval C et al (2018) Emergence of intraband
transitions in colloidal nanocrystals. Opt Mater Express 8:1174–1183. https://doi.org/10.1364/
OME.8.001174
82. Jeong KS, Deng Z, Keuleyan S, Liu H, Guyot-Sionnest P (2014) Air-stable n-doped colloidal
HgS quantum dots. J Phys Chem Lett 5:1139–1143. https://doi.org/10.1021/jz500436x
83. Robin A, Livache C, Ithurria S, Lacaze E, Dubertret B, Lhuillier E (2016) Surface control of
doping in self-doped nanocrystals. ACS Appl Mater Interfaces 8:27122–27128. https://doi.
org/10.1021/acsami.6b09530
84. Lhuillier E, Scarafagio M, Hease P, Nadal B, Aubin H, Xu XZ et al (2016) Infrared
Photodetection based on colloidal quantum-dot films with high mobility and optical absorp-
tion up to THz. Nano Lett 16:1282–1286. https://doi.org/10.1021/acs.nanolett.5b04616
85. Jeong J, Yoon B, Kwon Y-W, Choi D, Jeong KS (2017) Singly and doubly occupied higher
quantum states in nanocrystals. Nano Lett 17:1187–1193. https://doi.org/10.1021/acs.
nanolett.6b04915
86. Ramiro I, Özdemir O, Christodoulou S, Gupta S, Dalmases M, Torre I et al (2020) Mid- and
long-wave infrared optoelectronics via Intraband transitions in PbS colloidal quantum dots.
Nano Lett 20:1003–1008. https://doi.org/10.1021/acs.nanolett.9b04130
87. Qu J, Goubet N, Livache C, Martinez B, Amelot D, Gréboval C et al (2018) Intraband mid-­
infrared transitions in Ag2Se nanocrystals: potential and limitations for hg-free low-cost
Photodetection. J Phys Chem C 122:18161–18167. https://doi.org/10.1021/acs.jpcc.8b05699
88. Hafiz SB, Scimeca MR, Zhao P, Paredes IJ, Sahu A, Ko D-K (2019) Silver selenide col-
loidal quantum dots for mid-wavelength infrared Photodetection. ACS Appl Nano Mater
2:1631–1636. https://doi.org/10.1021/acsanm.9b00069
89. Hafiz SB, Al Mahfuz MM, Ko D-K (2021) Vertically stacked Intraband quantum dot devices
for mid-wavelength infrared Photodetection. ACS Appl Mater Interfaces 13:937–943. https://
doi.org/10.1021/acsami.0c19450
90. Hafiz SB, Al Mahfuz MM, Lee S, Ko D-K (2021) Midwavelength infrared p–n hetero-
junction diodes based on Intraband colloidal quantum dots. ACS Appl Mater Interfaces
13:49043–49049. https://doi.org/10.1021/acsami.1c14749
91. Sahu A, Khare A, Deng DD, Norris DJ (2012) Quantum confinement in silver selenide semi-
conductor nanocrystals. Chem Commun 48:5458–5460. https://doi.org/10.1039/C2CC30539A
92. Deng Z, Jeong KS, Guyot-Sionnest P (2014) Colloidal quantum dots Intraband photodetec-
tors. ACS Nano 8:11707–11714. https://doi.org/10.1021/nn505092a
93. Goubet N, Livache C, Martinez B, Xu XZ, Ithurria S, Royer S et al (2018) Wave-function engi-
neering in HgSe/HgTe colloidal Heterostructures to enhance mid-infrared photoconductive
properties. Nano Lett 18:4590–4597. https://doi.org/10.1021/acs.nanolett.8b01861
Chapter 8
Graphene/HgCdTe Heterojunction-Based
IR Detectors

Shonak Bansal, M. Muthukumar, and Sandeep Kumar

8.1 Introduction

Infrared (IR) detection technology discovery dates back to the early eighteenth cen-
tury (February 11, 1800), yet the first IR detectors (IRDs) were developed late twen-
tieth century [1]. Detection and sensing of IR radiation is tremendously important in
various domains such as military, medical, imaging, and geological sciences includ-
ing NASA earth science. Nowadays, IRDs are utilized in diverse products and have
profound applications in optical communications, remote control, night vision, motion
detection, satellite remote sensing, gas detection, fire alarming, biomedical and ther-
mal imaging, navigational aids, missile guidance, telecommunications, spectroscopy,
security, chemical analysis, and agriculture, etc. [2–8]. The imaging and night vision
of thermal sources require IRDs with a spectral sensitivity of about 10 μm, whereas
IR spectroscopy of molecules and gases requires a spectral sensitivity of about
2-15 μm [9]. To achieve this epitaxially grown narrow bandgap mature photosensitive
semiconductor materials are required which includes; Si, Ge [10–12], II–VI (HgCdTe,
CdZnTe, CdSeTe) [2, 13–16], III–V (InAsSb, InSb, GaAs, InAs) [2, 17–19], and IV–
VI (PbSnTe) [2]. These photosensitive materials absorb the incident light energy cor-
responding to their energy bandgaps which result in the net photocurrent. The
fundamental properties of narrow bandgap materials (such as high electron mobility,
high optical absorption coefficient, and low thermal generation rate), along with the
bandgap engineering, make these alloys almost ideal for a wide range of IRDs. Thus,
these materials provide an exceptional degree of freedom in the development of IRDs.

S. Bansal (*)
Electronics and Communication Engineering Department, Chandigarh University,
Gharuan, India
e-mail: shonakk@gmail.com
M. Muthukumar · S. Kumar
ICAR-Central Institute for Subtropical Horticulture, Lucknow, Uttar Pradesh, India

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 183
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_8
184 S. Bansal et al.

The bandgap energy tunability results in IR detector applications that cover all IR
spectral regimes. Generally, IR spectral regime can be differentiated into the near-IR
(NIR: 0.75–1.1 μm), short-wave IR (SWIR: 1–3 μm), mid-wave IR (MWIR: 3–5 μm),
long-wave IR (LWIR: 8–12 μm), very long-wave IR (VLWIR: 12–30 μm), and far IR
(FIR: 30–1000 μm) ranges [17, 20]. These spectral regimes have different conven-
tions, for example, 1.55 μm is considered as SWIR, in the Department of Defense
community, but is considered NIR for the astronomy community. The spectral regimes
ranging from MWIR to FIR are widely used for astronomy and free-space communi-
cations because the high transparency of the atmosphere at these spectral regimes
permits lossless transmission. The same spectral regimes are also used for military
applications including natural resources management and environmental monitoring,
cloud properties, sea surface temperatures, forest fires, and volcanic activities [17, 21,
22]. Nowadays, numerous military airplanes are furnished with high performance IR
cameras for scanning the battlefield during poor visibility conditions.
Rapid developments are being made in evolving cost-effective narrow bandgap
semiconductor detectors with enhanced sensitivity and longer wavelengths. Till
date, different IR detector’s architectures, namely, p-n [7, 23–27], p-i-n [4, 27–29],
dual-band detector [5], metal-semiconductor-metal [28], avalanche photodetector
(APD) [30], carbon nanotube detector [31], and nanowire photodetector [31] have
been reported. Similarly, novel design architectures, namely, quantum dots [31, 32],
quantum well [33], and Schottky-barrier photoemissive detector [34–41] based
IRDs are developed with improved performances. However, such device fabrication
processes are more expensive and include sophisticated equipment that are not com-
mercially feasible.
The excellent optoelectronic properties of graphene and HgCdTe makes these
materials a favorable choice for the development of high performance IRD to sup-
port and further advance a variety of applications. Initially, graphene has been pro-
posed as a replacement for the indium tin oxide (ITO) as transparent electrodes for
optical devices such as LCD and LED devices [28]. The high carrier mobility, wide
spectral regime from ultraviolet (UV) to IR due to linear dispersion nearby the
Dirac point of graphene, and the outstanding light absorption properties of other
semiconductor materials enables graphene an excellent candidate choice for the
development of next-generation optoelectronic devices such as photodetectors. This
chapter will present firstly the key properties of HgCdTe and graphene materials,
challenges, developments, and the modeling approach on graphene/HgCdTe based
IRDs in the subsequent sections.

8.2 Graphene/HgCdTe Heterojunction Based IRDs

8.2.1 Early Generation IRD Technologies

The detector technologies that were initially developed uses the compound semi-
conductor materials. PbS was the first material used for the IRD in SWIR spectral
regime for military applications [17, 42]. Later, the detection range was extended up
8 Graphene/HgCdTe Heterojunction-Based IR Detectors 185

to MWIR spectral regime by employing compound semiconductor materials such


as InSb, PbSe, and PbTe, etc. The development of semiconductor alloy materials on
groups II-VI, III-V, and IV-IV was started in the 1950’s. The better stability, low
leakage current, low thermal generation rate, relatively high absorption coefficient,
and tunable bandgap make the HgCdTe most well-regarded semiconductor material
after Si and GaAs for the development of high-performance broadband IR and THz
detectors in military applications to cover a spectral regime from the NIR to the FIR
[4, 17, 43–46]. The reason for this is that HgCdTe can be tuned to the desired IR
spectral regime by varying the Cd concentration.

8.2.2 Existing and Next Generation Technologies, Challenges,


and Prospects of Effective IR Detection

HgCdTe can be epitaxially grown as a chemical ternary alloy of HgTe and CdTe.
HgTe is a semi-metal chemical compound with high conductivity and zero bandgap,
whereas CdTe is a semiconductor with a bandgap of about 1.6 eV. These alloys have
almost the same lattice constant, which allows the growth of HgCdTe in any com-
position without defects introduced by lattice mismatch. The extremely small
change of lattice constant with Cd composition makes it possible to grow high-­
quality layers and heterostructures. Mixing these two materials allows obtaining
any energy bandgap between 0 and 1.6 eV [17]. Furthermore, the growth of HgCdTe
material layer can be controlled by metal-organic chemical vapor deposition
(MOCVD), molecular beam epitaxy (MBE), and liquid phase epitaxy (LPE) [47–
50]. LPE is the most widely used for industrial applications for many years, while
MBE, a vapor phase epitaxy (VPE) process leads to high accuracy in the deposition
of detector material structures resulting in good lateral homogeneity, high quantum
efficiency, and abrupt doping profiles in HgCdTe-based IR detectors [51, 52]. The
bandgap tunability of HgCdTe resulted into the numerous HgCdTe-based high per-
formance IRDs with different configurations such as p-n [25, 53, 54], p-i-n [4, 29],
APD [30], and dual-band IRD [5, 55] at cryogenic temperatures. The need for cryo-
genic cooling facilities to reduce thermally generated dark current adds a significant
amount of energy/power consumption, cost, and weight to such IRDs, thereby
requiring a bulky imaging instrument. HgCdTe technology development continues
to be primarily for military-related security or safety applications for the detection
of a condition or an object. The main motivations for replacing HgCdTe are the
technological disadvantages of this material. Apart from this high Auger recombi-
nation process results in high dark currents [4, 41, 56]. Therefore, it is essential to
design and develop highly efficient HgCdTe based IRD above cryogenic tempera-
ture operations for low weight, size, power, and cost space applications. Nevertheless,
it remains the leading IRD technology for applications with multiple spectral
regimes, where cost is not a major issue. HgCdTe based IRD performance at higher
operating temperatures is significantly affected by the dislocations and defects aris-
ing from lattice mismatches, ensuring residual stress that further reduces lifetimes
186 S. Bansal et al.

and minimizes the mobility of the carriers. Conversely, methods such as thermal
cycle annealing (TCA) reduces the dislocation density down to the saturation limit,
resulting in the improved high temperature operations of HgCdTe based IRDs. The
utilization of epitaxially growth thin substrates can further reduce the dislocations
towards the surface. Moreover, for HgCdTe epitaxial growth the Si (112) orientation
is preferable choice due to Hg defect control, consumption, and doping issues [22].
The synthesis of novel 2D nanomaterials are emerging as enabling technologies
of the future as they provide extended spectral regime in the photonic devices [57].
Graphene is the choicest and most favorable material for the development of high-­
operating detectors in the UV to IR, and THz regimes due to its excellent electrical
and optical properties [58, 59]. It is utilized to form a suitable heterojunction with
other semiconductor materials due to enormous mobility (~105 cm2/Vs), high sensi-
tivity, high current-density carrying capacity (~109 A/cm2), high Fermi velocity
(108 cm/s), better strength, flexibility, high thermal conductivity, low resistivity, tun-
able Fermi-level, broadband light absorption, and so on [60–62]. The high carrier
mobility in graphene permits it to be used in developing an efficient high-speed
charge carrier collector. Ultrafast carrier multiplication (CM) has also been observed
in bilayer graphene (BLG) recently [7, 27]. The graphene has been successfully
composited with different materials, namely, Si [63], ZnO [64–66], CdS [67] (in
UV spectral regime), CdSe [68], GaN [69] (in visible spectral regime), and Si [10,
66, 70], Ge [12], GaAs [71], PbS [72, 73], and HgCdTe [7, 26, 27, 29, 34] (in IR
spectral regime). The heterostructure of graphene with other materials demonstrates
a low dark current, small parasitics, low power dissipation, high response speed, and
higher breakdown voltage than that of conventional homostructures. Such hetero-
junctions offer a high electric field beneficial for fast separation of the photogene-
rated carriers without application of any external bias, enabling self-powered and
higher performance detectors [74]. Till date, numerous graphene-based heterostruc-
tures are reported with improved performances [63, 64, 67, 69, 71]. Additionally,
such heterostructures are utilized as a fundamental component of high-performance
nanoelectronic devices to estimate the device performance such as response time,
open-circuit voltages in solar cells, the ON/OFF ratios, and ON-state current [74].
The graphene layer helps in efficient carrier separation due to the tunable Fermi-­
level [61] and also acts as an antireflection coating to reduce the light reflection by
~80% in the IR spectral regime [75].
The graphene can be synthesized by the processes such as chemical vapor depo-
sition (CVD), oxidation reduction growth process, epitaxial growth on silicon car-
bide, and liquid phase stripping. Out of these, the CVD is a widely used process for
fabricating graphene layers with better structural integrity and crystalline quality on
copper substrates without any contamination issues [76]. But copper and other
metallic substrates are not suitable for many applications as they need the employ-
ment of graphene sheets directly on semiconductors or metal oxide [51]. Therefore,
there is a need for the improvement in effective processes over which graphene can
be effectively and directly transported onto any desirable substrate, while also
avoiding wrinkles, cracks, and any kind of contamination.
8 Graphene/HgCdTe Heterojunction-Based IR Detectors 187

However, the photodetectors utilizing single layer of graphene suffer from the
poor photodetection response due to the zero bandgap and small optical absorption
(~2.3%) of graphene [19, 58, 63, 77–80]. Several efforts including the use of gra-
phene quantum dots, a few graphene layers, inducing bandgap in graphene are made
to improve the detection ability [3, 72, 73, 81–84]. The chemical doping in graphene
shifts the Fermi-level, creating a bandgap [85]. Furthermore, the integration of high
carrier mobility graphene layer with HgCdTe, the detector’s operational capabilities
and performance can be further improved.

8.2.3 Graphene/HgCdTe Detector Fabrication


and Operating Principle

MBE growth method allows the high crystalline quality HgCdTe material on
CdZnTe or Si substrates with precise control of the detector material structure
parameters. On the other hand, HgCdTe is an expensive technology primarily due to
the fact that it can be grown on a close lattice-matched CdZnTe substrate. The IR
detection performance of HgCdTe epitaxially grown on Si substrate is also analo-
gous to that of HgCdTe deposited on CdZnTe substrate and is thermally compatible
with the current readout integrated circuit (ROICs) [22, 86, 87]. CdZnTe substrate
is preferred as HgCdTe/CdZnTe interface offers less interface trap charge as com-
pared to CdTe, Si, and Ge substrates [54] and does not require a buffer layer between
the absorber and substrate. But, CdZnTe substrates present cost-related challenges
that require the search for an alternate viable substrate for HgCdTe growth. GaAs is
high-quality, inexpensive, and easily available substrate material. The lattice mis-
match of HgCdTe with Si and GaAs layer involves the utilization of an extra CdTe
buffer layer [52, 86]. Using the MBE method, n-type doping of 5 × 1014 − 2 × 1015
and up to 1019 cm−3 were obtained by using CdZnTe [88] and Si [86] substrates,
respectively. On the other hand, by using the MOCVD method, the donor and
acceptor doping levels of 5 × 1014 − 5 × 1017 cm−3 were achieved. The HgCdTe
absorbing layer is epitaxially grown on the buffer layer with the desired Cd compo-
sition and thickness. Then the graphene layer is grown on the absorbing layer fol-
lowed by deposition of the top metal contacts.
In terms of material structure, the graphene/HgCdTe detector is mostly compos-
ted of the three principal layers:
(i) Gate: Substrate layer is used as gate terminal which provides electrical field
aiding carrier transport,
(ii) Absorber/active layer: The absorber layer (composed of HgCdTe) is the active
optical layer where the process of carrier photogeneration takes place, and
(iii) Channel: High mobility, low noise graphene channel transfers the photogene-
rated carriers to the electrical readouts.
188 S. Bansal et al.

IR detector operation process can be divided into three main phases;


(i) Carrier generation and separation: In the first step the incident IR photons are
transmitted to HgCdTe absorber layer resulting in the production of electron-­
hole pairs, or excitons,
(ii) Carrier transport and injection: The carriers are transported through the
absorber and get injected into graphene channel, and
(iii) Carrier transport in graphene channel: Injected carriers transported to and col-
lected by ROIC.

8.2.4 Graphene/HgCdTe Heterojunction Based IRD Structures


and Operation

The graphene/HgCdTe based detector technology involves the incorporation of


HgCdTe materials along with graphene, which allows for higher IR detection per-
formance as compared with detectors using HgCdTe material solely. The graphene
layer conformably covers a narrow bandgap, lightly doped HgCdTe active/absorb-
ing layer wherein the photogeneration of carrier takes place to form a heterojunction
IRD. The lightly doped HgCdTe generates total dark-current and photo-current den-
sities. The bandgap and doping concentration of the active region is tuned to allow
better absorption of IR radiations suppressing thermally generated carriers.
Furthermore, the active layer estimates the carrier lifetime, quantum efficiency,
photogeneration rate, and which mutually govern the overall performance of the
detector. The intrinsic interfacial barrier between the graphene and HgCdTe active
layer is designed to efficiently minimize the recombination of photogenerated car-
riers in the detector. The graphene functions as low noise, high mobility channel
that drifts away the photogenerated carriers from the active layer to the electrical
contacts, and consequently to the ROIC for electrical readout before their recombi-
nation. This will further contribute to the IR performance of graphene/HgCdTe
detectors compared to that of conventional HgCdTe detectors [7, 22, 27, 86].
HgCdTe is a material with a relatively high dielectric constant (~14) which makes it
an amenable substrate for providing relatively high electric conductance by trans-
ferring or coupling with graphene film at room temperature. It was experimentally
demonstrated that the integration of 5-10 graphene layers onto HgCdTe substrate
resulted in 25 times higher electrical conductance than that of HgCdTe and 80%
optical transmittance in the IR regime at 77 and 300 K which was slightly lower
than that on the SiO2/Si substrate in the visible regime [89]. This experimental study
proposed that graphene is considered a good candidate for transparent electrodes as
a replacement of metal electrodes while developing better and cost-effective
HgCdTe based IRDs with smooth surfaces without any crack or folding.
Figure 8.1a, b show the schematic structures of the BLG composite Hg1–xCdxTe
based p+- n− single heterojunction and p+-n−-n+ dual junctions IRDs, respectively [7,
26, 27]. For the single-heterojunction IRD formation, highly doped thin p+-BLG
8 Graphene/HgCdTe Heterojunction-Based IR Detectors 189

Fig. 8.1 The schematic of BLG/Hg1–xCdxTe based (a) p+-n− single-heterojunction and (b) p+-n−-n+
dual-junction IR detectors

was employed on the lightly doped wide n−-Hg1–xCdxTe active layer as shown in
Fig. 8.1a. For the p+-BLG/n−-Hg0.7783Cd0.2217Te/n+-Hg1–xCdxTe dual-junction LWIR
detectors formation, the device dimensions were kept constant same as that of the
single-heterojunction IRD (Fig. 8.1a), except the inclusion of a thin heavily doped
n+-Hg1–xCdxTe below active layer as illustrated in Fig. 8.1b. The p+-doping of the
BLG can be achieved through the chemical agents via the CVD technique [7, 90].
The BLG allows the opening of bandgap in addition to other excellent optoelec-
tronic properties [7, 91, 92]. In order to analyze the effect of bandgap tuning on the
detector performance, the Cd composition in n+-Hg1–xCdxTe was optimized to
0.2217 and 0.32 in the p+-BLG/n−-Hg0.7783Cd0.2217Te/n+-Hg1–xCdxTe dual-junction
LWIR detectors. This resulted in the formation of two different dual-junction
devices, named, p+-BLG/n−-Hg0.7783Cd0.2217Te/n+-Hg0.7783Cd0.2217Te (consisting of
one hetero- and one homojunction) and p+-BLG/n−-Hg0.7783Cd0.2217Te/n+-
Hg0.68Cd0.32Te (active layer was sandwiched between two different bandgap materi-
als to form the dual heterojunctions). The dual heterojunctions device helps to
reduce the thermal generation and parasitic impedances. The optimized
 Cd compo-
sition in n+-Hg1–xCdxTe facilitated high
 built-in electric field Efield at n− +
-n hetero-
junction. Such a large built-in Efield drifts the photogenerated carriers to the
electrodes, resulting in higher effective photocurrent through the heterojunction
which further reduces the tunneling current in reverse bias condition. The bandgap
and doping concentration of active layer was selected for optimum absorption of IR
radiations suppressing thermally generated carriers.
In real fabricated devices, the ohmic contacts can be made by using gold/molyb-
denum (Au/Mo) and palladium (Pd) with lower contact resistance [93, 94]. The IR
radiations were incident from the BLG cladding window over the narrow bandgap
active layer. The BLG was used as a light absorber. During the IR illuminations on
the p+-BLG/n−-Hg1–xCdxTe heterojunction, the hot photocarriers
 transport from the
active layer to the BLG layer due to the built-in Efield [95], resulting in the change
in graphene conduction.
190 S. Bansal et al.

The energy bandgap diagram and electric field profile for BLG/Hg1–xCdxTe based
p+-n− heterojunction, p+-n−-n+ dual-junction (one hetero- and one homojunction),
and p+-n−-n+ dual-heterojunction IRDs under 0 V bias or self-powered mode with
IR illumination are shown in Fig. 8.2a–c. The IRDs show a unique band-alignment
of the BLG/Hg1–xCdxTe heterostructure, suggesting an ohmic contact for the elec-
trons. The small bandgap in BLG due to doping will change the work function of
BLG resulting in the shifting of Fermi-level towards the valence band. This makes
graphene different than the other 2D materials [90].
Under IR illumination, the electron-hole pairs are produced in the bandgap of a
lightly doped active layer as demonstrated in Fig. 8.2a–c for IRDs. The carriers
were separated by the built-in electric field at p+-BLG/n−-Hg1–xCd xTe
  heterojunc-
tion, and the external electric field aligns the net electric field Ep − n towards its
direction. Consequently, photogenerated carriers move in the opposite direction
providing net photocurrent to the external circuit.
Figure 8.2b, c show the energy bandgap diagram wherein Cd composition of the
lightly doped active layer (x = 0.2217) is tuned for the LWIR operation. In p+-n−-n+
dual-junction IRDs under IR illumination, both the p+-BLG and n−-Hg1–xCdxTe lay-
ers absorb the incident photons and generate electron-hole pairs as shown in

Fig. 8.2 The energy bandgap diagram and electric field profile of BLG/HgCdTe based (a) p+-n−
heterojunction; (b) p+-n−-n+ dual-junction (one hetero- and one homojunction), and (c) p+- n−-n+
dual-heterojunction IR detectors under 0 V bias or self-powered mode demonstrating light absorp-
tion and carrier multiplication process. Here, ECB, EF, and EVB represent conduction band, Fermi-­
level, and valence band energies, respectively; M– and K– are the points of Brillouin zone in BLG
resulting in a higher CM effect. (Idea from [63])
8 Graphene/HgCdTe Heterojunction-Based IR Detectors 191

 
Fig. 8.2b, c. The built-in electric field Ep − n and En − n at the p+-BLG/n−-Hg1–x
Cdx = 0.2217Te and n−- Hg1–xCdx = 0.2217Te/n+-Hg1–xCdxTe junction separates the photo-
generated carriers. The external field dominates over the junction and results in the
drift of photogenerated carriers towards n+-Hg1–xCdxTe and p+-BLG, respectively,
thereby resulting in a net photocurrent. Due to the existence of dual heterojunctions
a huge built-in electric field exists at the n−-Hg1–xCdx = 0.2217Te/n+-Hg1–xCdx = 0.32Te
hetero-interface which effectively separates the photogenerated carriers so as to
produce a better photoresponse. It is evident that under IR illumination, the photo-
generated carriers accumulate in the potential well, raising the Fermi-level and
increasing the conductivity of the detector [7, 27]. The various electrical and optical
characteristic parameters were computed and analyzed using computer simulations.
The results were further validated by an analytical model based on drift-diffusion,
tunneling, and Chu’s methods [7, 96–98], suggesting potential applications in next-­
generation high-performance, ultra-low-power, and cost-effective IRDs for opto-
electronics devices. The dark current density was reduced due to the lower
thermo-generation rate of BLG and Hg1–xCdxTe heterostructure, offering improved
performance. The heterostructure of BLG with Hg1–xCdxTe offered excellent photo-
detection in the IR regime along with the near room temperature external quantum
efficiency (QEext) of more than 100% attributed to the hot carrier multiplication
mechanism in graphene [99–105].
In the VLWIR spectral regime, the p+-BLG/n−-Hg1–xCdx = 0.1867Te heterojunction
photodetector [7] offers the photosensitivity of 4.8 × 1014, response time of 9.4 ps,
3-dB cut-off frequency of 36.16 GHz, and QEext of 89% at −0.5 V and 77 K. Contrarily
in the LWIR spectral regime, p+-BLG/n−-Hg1–xCdx = 0.2217Te heterojunction photode-
tector [26, 27] demonstrated the photosensitivity of 6.9 × 105, rise (fall) time of 7.4
(5.7) ps, 3-dB cut-off frequency of 73 GHz, and QEext of 54.45% with −0.5 V bias
at 77 K. Moreover, the carrier multiplication and external biasing effects in p+-BLG/
n−-Hg1–xCdxTe heterojunction resulted in the near room temperature QEext of
~3337.70 and ~ 892% for the VLWIR and LWIR detector, respectively.
The dual-heterojunction detector exhibited low dark current density and higher
temperature operation due to the lower thermal generation rate than the single-­
heterojunction based detector [27]. The p+-BLG/n−-Hg1–xCdx = 0.2217Te/n+-Hg1–x
Cdx = 0.32Te dual-heterojunction LWIR detector showed rapid photocurrent switching
with rise (fall) time of ~0.05 (0.013) ps. Under self-powered mode, the photocapaci-
tance of 7.21 × 10−13 F/μm was obtained for the dual-heterojunction detector sug-
gesting the effective photogeneration of carriers and partial screening of the electric
field. The highest QEext of 85.8% at 10.6 μm cut-off wavelength at 77 K, was
achieved for the dual-heterojunction detector
 [27]. Such improved performance was
due to the existence of a huge built-in Efield at n−-n+ heterojunction. Moreover, the
temperature-dependent QEext was observed to be ~101%, for dual-heterojunction
LWIR detector at near room temperature due to the carrier multiplication
effect in BLG.
Inspired by the physical properties of BLG such as hot carrier injection, tunable
bandgap, and carrier multiplication, a new device comprising a BLG over the top of
192 S. Bansal et al.

Fig. 8.3 (a) The schematic of BLG/CdTe/Hg1–xCdxTe heterojunction based IRD, (b) the schematic
of p+-BLG/CdTe/n−-Hg1–xCdxTe heterojunction based IRD with micro-holes, and (c) the band dia-
gram of BLG/CdTe/HgCdTe heterojunction based IRD demonstrating light absorption and carrier
multiplication process

a thin CdTe passivation layer on a HgCdTe absorbing layer (Fig. 8.3a) is proposed
by using a quantum mechanical approach [106]. Using the same design architecture
as shown in Fig. 8.3a, a p+-BLG/CdTe/n−-Hg1–xCdxTe heterojunction detector design
that uses photon-trapping micro-holes in addition to the carrier multiplication effect
in graphene is proposed and is shown in Fig. 8.3b [107]. As shown in Fig. 8.3c, the
CdTe layer allows the passage of electrons in the conduction bands but shows a
significant barrier for the holes due to the existence of a huge Schottky barrier. This
confirms the transport of carriers from ambipolar to unipolar. Using the non-­
equilibrium Green’s function (NEGF) it was found that the electrons are the domi-
nant charge carriers over holes which reduces the carrier recombination, increases
carrier lifetimes, and hence results in improved performance in comparison to con-
ventional HgCdTe based IRDs.
Figure 8.4 shows the graphene/HgCdTe heterojunction based MWIR on a Si
substrate having a CdTe buffer layer deposited on it [22,
86].
 The Si substrate act as
the electrode terminal that provides an electric field Efield in the vertical direction,
resulting in transport of the photogenerated electron-hole pairs in the vertical
8 Graphene/HgCdTe Heterojunction-Based IR Detectors 193

Fig. 8.4 The schematic of graphene/HgCdTe heterostructure based IRD on Si substrate: (a)
Generation of carriers due to incident IR radiations and their separation due to the built-in electric
field, (b) Transportation and injection of photogenerated carrier into graphene, and (c)
Transportation of photogenerated carriers in graphene horizontally

direction as shown in Fig. 8.4a. The IR radiations were incident through the sub-
strate and buffer layers into the active
 region where they are absorbed to yield pho-
togenerated charge carriers. The Efield separates the photogenerated charge carriers
and further suppresses Auger recombination process in the active layer so as to
reduce the loss of photogenerated charge carriers. The photogenerated electrons
drifts from the active layer to the graphene layer and then injected into it (Fig. 8.4b).
The injected electrons into the graphene layer are transported in the lateral direction
and are collected by the ROIC terminal as illustrated in Fig. 8.4c. This indicates that
the graphene/HgCdTe heterojunction provides the improved performance than that
of conventional HgCdTe based IRDs.

8.2.5 Graphene/HgCdTe Heterojunction Based IRD


Modelling Approach

The general modeling approach is based on numerous components that together


constitute a comprehensive model for the optoelectronic technology. The state-of-­
art optoelectronic technology research focusses on improving the detection capa-
bilities by utilizing the novel structures. Therefore, analytical modeling and
194 S. Bansal et al.

simulation tools became a crucial part of the design and development procedure.
The simulation tools are highly recommended in the electronic industry to analyze
the optoelectronic device characterization and optimization. The software’s are gen-
erally used to save time and cost of an experiment before the real device fabrication.
The simulation tools depend on the basic semiconductor physics equations and
mechanism that provide an insight view of the device to understand the device phys-
ics. The goal is to develop accurate electrical and optical behavior of the device.
A solution of fundamental semiconductor equations including continuity and
Poisson’s equations [108] for electrons and holes is used to understand the electrical
behavior of semiconductor devices, including the energy bandgap and doping pro-
file, two and three dimensional effects, heterojunctions, non-equilibrium operation,
and interface, surface, and contact effects [109]. The drift-diffusion model is most
widely used for the simulation of carrier transport in semiconductors. Boltzmann
transport theory is used to approximate the current density equations by the conven-
tional drift-diffusion model, expressed them as a function of electric field and car-
rier concentrations, consisting of drift and diffusion components [108]. The reason
for choosing the drift-diffusion charge transport model is its inherent simplicity and
the better accuracy of this model [110].
The modeling approach includes modularly building the complete detector simu-
lation platform from different materials models as data is made available from
experiments and device characterizations. These including individual materials
models for graphene, HgCdTe, and the graphene/HgCdTe heterounction. The
Silvaco TCAD software was utilized to design and evaluate the optoelectronic per-
formances of BLG/HgCdTe based IR detectors [7, 27]. The computer simulations
for optoelectronic characteristics of the p+-BLG/n−-Hg1–xCdxTe IR detector were
carried out by solving the continuity, carrier transport, and Poisson’s equations with
optimized boundary conditions based on the Boltzmann’s transport model [5, 108].
To reproduce the carrier transport, a drift-diffusion approach was implemented to
the degenerate semiconductor and parabolic shape of the conduction band [111].
The Newton-Richardson iteration method and concentration-dependent Analytic
model were used to estimate the carrier mobility in the detector [5, 54]. Further to
solve these equations, in order to characterize the carrier lifetime, and the dark cur-
rent density in the p+-BLG/n−-Hg1–xCdxTe IR detector, Shockley-Read-Hall, Auger,
optical, surface recombination rates, and standard tunneling mechanism models
were considered. The doping and carrier densities were evaluated using Fermi-­
Dirac statistics [5, 54, 108, 112]. Accordingly, the total dark current density as a
function of applied voltage and ambient temperature arise due to the existence of
diffusion, generation-recombination, and tunneling components [7].
The optical characterizations were performed by coupling the optical and basic
semiconductor equations. The optical absorption coefficient of Hg1–xCdxTe material
within the Kane region was calculated by Chu’s empirical relation [96, 97]. To com-
pute the optical characteristics, wavelength-dependent complex refractive indices
for both the BLG [79, 113] and Hg1–xCdxTe [54, 98] materials were described in IR
8 Graphene/HgCdTe Heterojunction-Based IR Detectors 195

regime. However, the total QEext of the detector comprises of neutral p   QEext p ,


neutral n   QEext n , and the depletion (QEext)dep regions as given by [7]:


QEext  CM V ,T    QEext p   QEext n   QEext dep 


  (8.1)
where, CM(V,T) represents the voltage (V) and temperature (T) dependent hot car-
rier multiplication factor [7, 99, 101], that can be estimated from the simulated cur-
rent density-voltage characteristics under dark and illumination conditions at
different temperatures as [30]:

J light V ,T   J dark V ,T 
CM V ,T  
J light V  0, T   J dark V  0, T 
(8.2)
Jdark and Jlight represents the dark current and photocurrent density, respectively.

8.3 Conclusion and Future Prospects

Graphene is considered as the most favorable material choice for the development
of high-operating detectors in the UV to IR, and THz regimes. A high-performance
graphene/HgCdTe detector utilizes the properties of both materials for the develop-
ment of IR detectors. This chapter presented the overview of the graphene/HgCdTe-­
based IR detectors, their fabrication, principles, optimization procedures, and the
modeling approach. The graphene functions as high carrier mobility channel to drift
the photogenerated charge carriers away before they can recombine. Therefore, the
hetero-interface between the graphene layer and HgCdTe active layer acts as a tun-
able rectifier that minimizes the recombination of photogenerated carriers in the
detector. The presented study in this chapter suggest that the successful integration
of graphene film on to HgCdTe layer can develop the high-performance IR detec-
tors as compared to HgCdTe detectors. The room temperature operation capability
of graphene/HgCdTe IR detectors can be beneficial in various NASA earth Science
applications. Furthermore, with recent developments and advancements made in
nanoscience and nanotechnology, novel fabrication designs and development of
heterostructures of desired interests are done with the use of novel alloy combina-
tions. Utilization of latest technological advancements in materials sciences, it is
possible to refine and fine tune the existing detector technology with cost-effective
alternatives and this could be scaled up towards use in different applications viz.,
development of sensors, lasers, optoelectronic devices like IR detectors and IR light
sources.
196 S. Bansal et al.

References

1. Rogalski A, Kopytko M, Martyniuk P, Hu W (2020) Comparison of performance limits of


HOT HgCdTe photodiodes with 2D material infrared photodetectors. Opto-Electron Rev
28(2):82–92. https://doi.org/10.24425/opelre.2020.132504
2. Rogalski A (2003) Infrared detectors: status and trends. Prog Quantum Electron
27(2–3):59–210. https://doi.org/10.1016/S0079-­6727(02)00024-­1
3. Ryzhii V, Ryzhii M (2009) Graphene bilayer field-effect phototransistor for terahertz and
infrared detection. Phys Rev B Condens Matter Mater Phys 79(24):245311-1-245311–8.
https://doi.org/10.1103/PhysRevB.79.245311
4. Saxena PK (2011) Modeling and simulation of HgCdTe based p+-n-n+ LWIR photodetector.
Infrared Phys Technol 54(1):25–33. https://doi.org/10.1016/j.infrared.2010.10.005
5. Saxena PK (2017) Numerical study of dual band (MW/LW) IR detector for performance
improvement. Def Sci J 67(2):141–148. https://doi.org/10.14429/dsj.67.11177
6. Zhuge F, Zheng Z, Luo P, Lv L, Huang Y, Li H, Zhai T (2017) Nanostructured mate-
rials and architectures for advanced infrared photodetection. Adv Mater Technol
2(8):1700005-1-1700005–26. https://doi.org/10.1002/admt.201700005
7. Bansal S, Sharma K, Jain P, Sardana N, Kumar S, Gupta N, Singh AK (2018) Bilayer gra-
phene/HgCdTe based very long infrared photodetector with superior external quantum effi-
ciency, responsivity, and detectivity. RSC Adv 8(69):39579–39592. https://doi.org/10.1039/
c8ra07683a
8. Yao J, Yang G (2020) 2D material broadband photodetectors. Nanoscale 12(2):454–476.
https://doi.org/10.1039/c9nr09070c
9. Grotevent MJ, Hail CU, Yakunin S, Bachmann D, Calame M, Poulikakos D, Kovalenko MV,
Shorubalko I (2021) Colloidal HgTe quantum dot/graphene phototransistor with a spectral
sensitivity beyond 3 μm. Adv Sci 8:1–7. https://doi.org/10.1002/advs.202003360
10. Amirmazlaghani M, Raissi F, Habibpour O, Vukusic J, Stake J (2013) Graphene-Si
Schottky IR detector. IEEE J Quantum Electron 49(7):589–594. https://doi.org/10.1109/
JQE.2013.2261472
11. Assefa S, Xia F, Vlasov YA (2010) Reinventing germanium avalanche photodetector for nano-
photonic on-chip optical interconnects. Nature 464(7285):80–84. https://doi.org/10.1038/
nature08813
12. Zeng LH, Wang MZ, Hu H, Nie B, Yu YQ, Wu CY, Wang L, Hu JG, Xie C, Liang FX, Luo
LB (2013) Monolayer graphene/germanium Schottky junction as high-performance self-­
driven infrared light photodetector. ACS Appl Mater Interfaces 5(19):9362–9366. https://doi.
org/10.1021/am4026505
13. Norton P (2002) HgCdTe infrared detectors. Opto-Electron Rev 10(3):159–174
14. Rogalski A (2004) Toward third generation HgCdTe infrared detectors. J Alloys Compd
371(1–2):53–57. https://doi.org/10.1016/j.jallcom.2003.06.005
15. Wijewarnasuriya PS, Chen Y, Brill G, Zandi B, Dhar NK (2010) High-performance long-­
wavelength infrared HgCdTe focal plane arrays fabricated on CdSeTe compliant Si substrates.
IEEE Trans Electron Devices 57(4):782–787. https://doi.org/10.1109/TED.2010.2041511
16. Wang J, Chen X, Hu W, Wang L, Lu W, Xu F, Zhao J, Shi Y, Ji R (2011) Amorphous HgCdTe
infrared photoconductive detector with high detectivity above 200 K. Appl Phys Lett
99(11):113508–1–113508–3. https://doi.org/10.1063/1.3638459
17. Rogalski A (2005) HgCdTe infrared detector material: history, status and outlook. Rep Prog
Phys 68(10):2267–2336. https://doi.org/10.1088/0034-­4885/68/10/R01
18. Yoon J, Jo S, Chun IS, Jung I, Kim HS, Meitl M, Menard E, Li X, Coleman JJ, Paik U,
Rogers JA (2010) GaAs photovoltaics and optoelectronics using releasable multilayer epi-
taxial assemblies. Nature 465(7296):329–333. https://doi.org/10.1038/nature09054
19. Miao J, Hu W, Guo N, Lu Z, Liu X, Liao L, Chen P, Jiang T, Wu S, Ho JC, Wang L, Chen
X, Lu W (2015) High-responsivity graphene/InAs nanowire heterojunction near-infrared
8 Graphene/HgCdTe Heterojunction-Based IR Detectors 197

photodetectors with distinct photocurrent on/off ratios. Small 11(8):936–942. https://doi.


org/10.1002/smll.201402312
20. Long M, Wang P, Fang H, Hu W (2018) Progress, challenges, and opportunities for 2D mate-
rial based photodetectors. Adv Funct Mater 29(19):1803807-1-1803807–28. https://doi.
org/10.1002/adfm.201803807
21. Tan CL, Mohseni H (2018) Emerging technologies for high performance infrared detectors.
Nano 7(1):169–197
22. Sood AK, Zeller JW, Ghuman P, Babu S, Dhar NK, Ganguly S, Ghosh AW, Dupuis RD
(2019) Development of high-performance detector technology for UV and IR applications.
In: Proceedings of SPIE 11151, sensors, systems, and next-generation satellites XXIII, vol
11151, p 1115113-1-1115113–11. https://doi.org/10.1109/IGARSS.2019.8897813
23. Saxena PK, Chakrabarti P (2008) Analytical simulation of HgCdTe photovoltaic detector
for long wavelength infrared (LWIR) applications. Optoelectron Adv Mater Rapid Commun
2(3):140–147
24. Dwivedi ADD (2011) Analytical modeling and atlas simulation of p+-Hg0.78Cd0.22Te/
nHg0.78Cd0.22Te/CdZnTe homojunction photodetector for lwir free space optical communica-
tion system. J Electron Devices 9:396–404
25. Bansal S, Sharma K, Jain P, Gupta N, Singh AK (2018) Atlas simulation of a long-infrared
P+-N homojunction photodiode. In: 2018 6th edition of international conference on Wireless
Networks & Embedded Systems (WECON), Rajpura (near Chandigarh), India, pp 19–22.
https://doi.org/10.1109/WECON.2018.8782077
26. Bansal S, Jain P, Kumar N, Kumar S, Sardana N, Gupta N, Singh AK (2018) A highly effi-
cient bilayer graphene HgCdTe heterojunction based p+-n photodetector for long wavelength
infrared (LWIR). In: 2018 IEEE 13th Nanotechnology Materials and Devices Conference
(NMDC), Portland, OR, USA, pp 1–4. https://doi.org/10.1109/NMDC.2018.8605848
27. Bansal S, Das A, Jain P, Prakash K, Sharma K, Kumar N, Sardana N, Gupta N, Kumar
S, Singh AK (2019) Enhanced optoelectronic properties of bilayer graphene/HgCdTe based
single- and dual-junction photodetectors in long infrared regime. IEEE Trans Nanotechnol
18:781–789. https://doi.org/10.1109/TNANO.2019.2931814
28. Song S, Wen L, Chen Q (2015) Graphene composites based photodetectors. In: Sadasivuni K,
Ponnamma D, Kim J, Thomas S (eds) Graphene-based polymer nanocomposites in electron-
ics. Springer International Publishing, pp 193–222
29. Bansal S, Sharma K, Soni K, Gupta N, Ghosh K, Singh AK (2017) Hg1−xCdxTe based p-i-n IR
photodetector for free space optical communication. In: 2017 Progress In Electromagnetics
Research Symposium-Spring (PIERS), St Petersburg, Russia, pp 544–547. https://doi.
org/10.1109/PIERS.2017.8261800
30. Singh A, Shukla AK, Pal R (2017) Performance of graded bandgap HgCdTe avalanche
photodiode. IEEE Trans Electron Devices 64(3):1146–1152. https://doi.org/10.1109/
TED.2017.2650412
31. Shin D, Choi S-H (2018) Graphene-based semiconductor heterostructures for photodetec-
tors. Micromachines 9(7):1–29. https://doi.org/10.3390/mi9070350
32. Asgari A, Razi S (2010) High performances III-nitride quantum dot infrared photodetector
operating at room temperature. Opt Express 18(14):14604–14615. https://doi.org/10.1364/
OE.18.014604
33. Hao MR, Yang Y, Zhang S, Shen WZ, Schneider H, Liu HC (2014) Near-room-temperature
photon-noise-limited quantum well infrared photodetector. Laser Photonics Rev
8(2):297–302. https://doi.org/10.1002/lpor.201300147
34. Bansal S, Prakash K, Sardana N, Kumar S, Sharma K, Jain P, Gupta N, Singh AK (2019)
Bilayer graphene/HgCdTe based self-powered mid-wave IR nBn photodetector. In: 2019
IEEE 14th Nanotechnology Materials and Devices Conference (NMDC), Stockholm,
Sweden, pp 1–4. https://doi.org/10.1109/NMDC47361.2019.9083985
35. Martyniuk P (2015) HOT mid-wave HgCdTe nBn and pBp infrared detectors. Opt Quant
Electron 47(6):1311–1318. https://doi.org/10.1007/s11082-­014-­0044-­7
198 S. Bansal et al.

36. Craig AP, Thompson MD, Tian Z-B, Krishna S, Krier A, Marshall ARJ (2015) InAsSb-­
based nBn photodetectors : lattice mismatched growth on GaAs and low- frequency
noise performance. Semicond Sci Technol 30(10):105011-1-105011–7. https://doi.
org/10.1088/0268-­1242/30/10/105011
37. Haddadi A, Dehzangi A, Chevallier R, Adhikary S, Razeghi M (2017) Bias-selectable
nBn dual-band long-/very long-wavelength infrared photodetectors based on InAs/
InAs1-xSbx/AlAs1-xSbx type-II superlattices. Sci Rep 7(3339):1–7. https://doi.org/10.1038/
s41598-­017-­03238-­2
38. Nguyen TD, Kim JO, Kim YH, Kim ET, Nguyen QL, Lee SJ (2018) Dual-color short-­
wavelength infrared photodetector based on InGaAsSb/GaSb heterostructure. AIP Adv
8(2):025015-1-025015–7. https://doi.org/10.1063/1.5020532
39. Madejczyk P, Gawron W, Keblowski A, Mlynarczyk K, Stepien D, Martyniuk P, Rogalski
A, Rutkowski J, Piotrowski J (2020) Higher operating temperature IR detectors of the
MOCVD grown HgCdTe Heterostructures. J Electron Mater:1–10. https://doi.org/10.1007/
s11664-­020-­08369-­3
40. Casalino M, Sirleto L, Iodice M, Saffioti N, Gioffr̀ M, Rendina I, Coppola G (2010) Cu/p-Si
Schottky barrier-based near infrared photodetector integrated with a silicon-on-insulator
waveguide. Appl Phys Lett 96(24):241112-1-241112-1–3. https://doi.org/10.1063/1.3455339
41. Mohammadian M, Saghai HR (2015) Room temperature performance analysis of bilayer
graphene terahertz photodetector. Optik-Int J Light Electron Optics 126(11–12):1156–1160.
https://doi.org/10.1016/j.ijleo.2015.03.021
42. Corsi C (2010) History highlights and future trends of infrared sensors. J Mod Opt
57(18):1663–1686. https://doi.org/10.1080/09500341003693011
43. Devarakonda V, Dwivedi ADD, Pandey A, Chakrabarti P (2020) Performance analysis of
N+-CdTe∕n0-Hg0.824675Cd0.175325Te∕p+-Hg0.824675Cd0.175325Te n−i−p photodetector operating at
30 μm wavelength for terahertz applications. Opt Quant Electron 52(340):1–19. https://doi.
org/10.1007/s11082-­020-­02450-­1
44. Akhavan ND, Umana-Membreno GA, Gu R, Asadnia M, Antoszewski J, Faraone L (2016)
Superlattice barrier HgCdTe nBn infrared photodetectors: validation of the effective mass
approximation. IEEE Trans Electron Devices 63(12):4811–4818. https://doi.org/10.1109/
TED.2016.2614677
45. Kopytko M, Keblowski A, Gawron W, Kowalewski A, Rogalski A (2014) MOCVD grown
HgCdTe barrier structures for hot conditions. IEEE Trans Electron Devices 61(11):3803–3807.
https://doi.org/10.1109/TED.2014.2359224
46. Akhavan ND, Umana-membreno GA, Gu R, Antoszewski J, Faraone L (2018) Optimization
of superlattice barrier HgCdTe nBn infrared photodetectors based on an NEGF approach.
IEEE Trans Electron Devices 65(2):591–598. https://doi.org/10.1109/TED.2017.2785827
47. Reine MB (2001) HgCdTe photodiodes for IR detection : a review. Proc SPIE Int Soc Opt
Eng 4288:266–277
48. Vasilyev VV, Ovsyuk VN, Sidorov YG (2003) IR photodetectors based on MBE-grown MCT
layers. Proc SPIE Int Soc Opt Eng 5065:39–46
49. Chorter P, Tribolet P, Pelletan C (2001) High performance HgCdTe SWIR detectors develop-
ment at Sofradir. Proc SPIE Int Soc Opt Eng 4369:698–712
50. Piotrowski J, Orman Z, Nowak Z, Pawluczyk J, Pietrzak J, Piotrowski A, Szabra D (2005)
Uncooled long wave infrared photodetectors with optimized spectral response at selected
spectral ranges. Proc SPIE Int Soc Opt Eng 5783:616–624. https://doi.org/10.1117/12.606244
51. Sood AK, Zeller JW, Ghuman P, Babu S, Dhar NK, Jacobs RN, Chaudhary LS, Efstathiadis
H, Ganguly S, Ghosh AW, Ahmed SZ, Tonni FF (2022) Doping and transfer of high mobility
graphene bilayers for room temperature mid-wave infrared photodetectors. In: 21st century
nanostructured materials – physics, chemistry, classification, and applications in industry and
biomedical [Working title]. IntechOpen, London
8 Graphene/HgCdTe Heterojunction-Based IR Detectors 199

52. Gawron W, Sobieski J, Manyk T, Kopytko M, Madejczyk P, Rutkowski J (2021) MOCVD


grown HgCdTe Heterostructures for medium wave infrared detectors. Coatings 11(5):1–13.
https://doi.org/10.3390/coatings11050611
53. Saxena PK, Chakrabarti P (2009) Computer modeling of MWIR single heterojunction photo-
detector based on mercury cadmium telluride. Infrared Phys Technol 52(5):196–203. https://
doi.org/10.1016/j.infrared.2009.07.009
54. Dwivedi ADD (2011) Analytical modeling and numerical simulation of P+-Hg0.69Cd0.31Te/
n-Hg0.78Cd0.22Te/CdZnTe heterojunction photodetector for a long-wavelength infrared free
space optical communication system. J Appl Phys 110(4):043101-1-043101–10. https://doi.
org/10.1063/1.3615967
55. Bellotti E, D’Orsogna D (2006) Numerical analysis of HgCdTe simultaneous two-color
photovoltaic infrared detectors. IEEE J Quantum Electron 42(4):418–426. https://doi.
org/10.1109/JQE.2006.871555
56. Piotrowski A, Madejczyk P, Gawron W, Kłos K, Pawluczyk J, Rutkowski J, Piotrowski J,
Rogalski A (2007) Progress in MOCVD growth of HgCdTe heterostructures for uncooled
infrared photodetectors. Infrared Phys Technol 49(3):173–182. https://doi.org/10.1016/j.
infrared.2006.06.026
57. Bablich A, Kataria S, Lemme MC (2016) Graphene and two-dimensional materials for opto-
electronic applications. Electronics 5(1):1–16. https://doi.org/10.3390/electronics5010013
58. Xia F, Mueller T, Lin YM, Valdes-Garcia A, Avouris P (2009) Ultrafast graphene photodetec-
tor. Nat Nanotechnol 4(12):839–843. https://doi.org/10.1038/nnano.2009.292
59. Rogalski A, Kopytko M, Martyniuk P (2019) Two-dimensional infrared and terahertz
detectors: outlook and status. Appl Phys Rev 6(2):021316-1-021316–23. https://doi.
org/10.1063/1.5088578
60. Boruah BD, Ferry DB, Mukherjee A, Misra A (2015) Few-layer graphene/ZnO nanowires
based high performance UV photodetector. Nanotechnology 26(23):235703-1-235703–7.
https://doi.org/10.1088/0957-­4484/26/23/235703
61. Cheng CC, Zhan JY, Liao YM, Lin TY, Hsieh YP, Chen YF (2016) Self-powered and
broadband photodetectors based on graphene/ZnO/silicon triple junctions. Appl Phys Lett
109(5):053501-1-053501–5. https://doi.org/10.1063/1.4960357
62. Awasthi S, Gopinathan PS, Rajanikanth A, Bansal C (2018) Current–voltage characteristics
of electrochemically synthesized multi-layer graphene with polyaniline. J Sci Adv Mater
Devices 3(1):37–43. https://doi.org/10.1016/j.jsamd.2018.01.003
63. Wan X, Xu Y, Guo H, Shehzad K, Ali A, Liu Y, et al. (2017) A self-powered high-­performance
graphene/silicon ultraviolet photodetector with ultra-shallow junction: breaking the limit of
silicon? npj 2D Mater Appl 1(4):1–8. https://doi.org/10.1038/s41699-­017-­0008-­4
64. Dhar S, Majumder T, Mondal SP (2016) Graphene quantum dot-sensitized ZnO nanorod/
polymer Schottky junction UV detector with superior external quantum efficiency, detectiv-
ity, and responsivity. ACS Appl Mater Interfaces 8(46):31822–31831. https://doi.org/10.1021/
acsami.6b09766
65. Nie B, Hu JG, Luo LB, Xie C, Zeng LH, Lv P, et al. (2013) Monolayer graphene film on
ZnO nanorod array for high-performance schottky junction ultraviolet photodetectors. Small
9(17):2872–2879. https://doi.org/10.1002/smll.201203188
66. Bansal S, Prakash K, Sharma K, Sardana N, Kumar S, Gupta N, Singh AK (2020) A highly
efficient bilayer graphene/ZnO/silicon nanowire based heterojunction photodetector with
broadband spectral response. Nanotechnology 31(40):405205–1–405205–10. https://doi.
org/10.1088/1361-­6528/ab9da8
67. Spirito D, Kudera S, Miseikis V, Giansante C, Coletti C, Krahne R (2015) UV light detection
from CdS nanocrystal sensitized graphene photodetectors at kHz frequencies. J Phys Chem
C 119(42):23859–23864. https://doi.org/10.1021/acs.jpcc.5b07895
68. Gao Z, Jin W, Zhou Y, Dai Y, Yu B, Liu C, Xu W, Li Y, Peng H, Liu Z, Dai L (2013) Self-­
powered flexible and transparent photovoltaic detectors based on CdSe nanobelt/graphene
Schottky junctions. Nanoscale 5(12):5576–5581. https://doi.org/10.1039/c3nr34335a
200 S. Bansal et al.

69. Lin F, Chen SW, Meng J, Tse G, Fu XW, Xu FJ, Shen B, Liao ZM, Yu DP (2014) Graphene/GaN
diodes for ultraviolet and visible photodetectors. Appl Phys Lett 105(7):073103-1-073103–5.
https://doi.org/10.1063/1.4893609
70. Yu X, Dong Z, Liu Y, Liu T, Tao J, Zeng Y, Yang JKW, Wang QJ (2016) A high performance,
visible to mid-infrared photodetector based on graphene nanoribbons passivated with HfO2.
Nanoscale 8(1):327–332. https://doi.org/10.1039/C5NR06869J
71. Luo L-B, Hu H, Wang X-H, Lu R, Zou Y-F, Yu Y-Q, Liang F-X (2015) A graphene/GaAs
near-infrared photodetector enabled by interfacial passivation with fast response and high
sensitivity. J Mater Chem C 3(18):4723–4728. https://doi.org/10.1039/C5TC00449G
72. Sun Z, Liu Z, Li J, Tai GA, Lau SP, Yan F (2012) Infrared photodetectors based on CVD-grown
graphene and PbS quantum dots with ultrahigh responsivity. Adv Mater 24(43):5878–5883.
https://doi.org/10.1002/adma.201202220
73. Konstantatos G, Badioli M, Gaudreau L, Osmond J, Bernechea M, De Arquer FPG, Gatti F,
Koppens FHL (2012) Hybrid graphene-quantum dot phototransistors with ultrahigh gain. Nat
Nanotechnol 7(6):363–368. https://doi.org/10.1038/nnano.2012.60
74. Periyanagounder D, Gnanasekar P, Varadhan P, He JH, Kulandaivel J (2018) High perfor-
mance, self-powered photodetectors based on a graphene/silicon Schottky junction diode. J
Mater Chem C 6(35):9545–9551. https://doi.org/10.1039/c8tc02786b
75. Fan G, Zhu H, Wang K, Wei J, Li X, Shu Q, Guo N, Wu D (2011) Graphene/silicon nanowire
Schottky junction for enhanced light harvesting. ACS Appl Mater Interfaces 3(3):721–725.
https://doi.org/10.1021/am1010354
76. Tai L, Zhu D, Liu X, Yang T, Wang L, Wang R, Jiang S, Chen Z, Xu Z, Li X (2018) Direct
growth of graphene on silicon by metal-free chemical vapor deposition. Nano-Micro Lett
10(20):1–9. https://doi.org/10.1007/s40820-­017-­0173-­1
77. Mueller T, Xia F, Avouris P (2010) Graphene photodetectors for high-speed optical commu-
nications. Nat Photonics 4(5):297–301. https://doi.org/10.1038/nphoton.2010.40
78. Gan X, Shiue RJ, Gao Y, Meric I, Heinz TF, Shepard K, Hone J, Assefa S, Englund D (2013)
Chip-integrated ultrafast graphene photodetector with high responsivity. Nat Photonics
7(11):883–887. https://doi.org/10.1038/nphoton.2013.253
79. Pospischil A, Humer M, Furchi MM, Bachmann D, Guider R, Fromherz T, Mueller T (2013)
CMOS-compatible graphene photodetector covering all optical communication bands. Nat
Photonics 7(11):892–896. https://doi.org/10.1038/nphoton.2013.240
80. Rogalski A (2019) Graphene-based materials in the infrared and terahertz detector families: a
tutorial. Adv Opt Photon 11(2):314–379. https://doi.org/10.1364/aop.11.000314
81. Liu N, Tian H, Schwartz G, Tok JBH, Ren TL, Bao Z (2014) Large-area, transparent, and
flexible infrared photodetector fabricated using P-N junctions formed by N-doping chemi-
cal vapor deposition grown graphene. Nano Lett 14(7):3702–3708. https://doi.org/10.1021/
nl500443j
82. Ryzhii V, Ryzhii M, Mitin V, Otsuji T (2010) Terahertz and infrared photodetection using
p-i-n multiple-graphene-layer structures. J Appl Phys 107(5):054512-1-054512–7. https://
doi.org/10.1063/1.3327441
83. Ryzhii M, Otsuji T, Mitin V, Ryzhii V (2011) Characteristics of p-i-n terahertz and
infrared photodiodes based on multiple graphene layer structures. Jpn J Appl Phys
50(7):070117-1-070117–6. https://doi.org/10.1143/JJAP.50.070117
84. Pykal M, Jurečka P, Karlický F, Otyepka M (2016) Modelling of graphene functionalization.
Phys Chem Chem Phys 18(9):6351–6372. https://doi.org/10.1039/C5CP03599F
85. Zhao S, Xue J (2012) Tuning the band gap of bilayer graphene by ion implantation: insight from
computational studies. Phys Rev B Condens Matter Mater Phys 86(16):165428-1-165428–10.
https://doi.org/10.1103/PhysRevB.86.165428
86. Sood AK, Zeller JW, Ghuman P, Babu S, Dhar NK, Ganguly S, Ghosh A (2021) Development
of high-performance graphene-HgCdTe detector Technology for mid-wave Infrared
Applications. In: Proceedings of SPIE 11530, infrared sensors, devices, and applications XI,
vol 11530, p 115300I-1-115300I–11. https://doi.org/10.1117/12.2572904
8 Graphene/HgCdTe Heterojunction-Based IR Detectors 201

87. Sood AK, Zeller JW, Welser RE, Puri YR, Lewis J, Mto D, Street NR (2015) Development of
GaN/AlGaN UVAPDs for ultraviolet sensor applications. Int J Phys Appl 7(1):49–58
88. Vilela MF, Olsson KR, Rybnicek K, Bangs JW, Jones KA, Harris SF, Smith KD, Lofgreen
DD (2014) Higher dislocation density of arsenic-doped HgCdTe material. J Electron Mater
43(8):3018–3024. https://doi.org/10.1007/s11664-­014-­3180-­8
89. Xu W, Gong Y, Liu L, Qin H, Shi Y (2011) Can graphene make better HgCdTe infrared detec-
tors? Nanoscale Res Lett 6(1):250. https://doi.org/10.1186/1556-­276X-­6-­250
90. Patel K, Tyagi PK (2017) P-type multilayer graphene as a highly efficient transparent con-
ducting electrode in silicon heterojunction solar cells. Carbon 116:744–752. https://doi.
org/10.1016/j.carbon.2017.02.042
91. McCann E, Koshino M (2013) The electronic properties of bilayer graphene. Rep Prog Phys
76(5):056503-1-056503–28. https://doi.org/10.1088/0034-­4885/76/5/056503
92. Schmitz M, Engels S, Banszerus L, Watanabe K, Taniguchi T, Stampfer C, Beschoten
B (2017) High mobility dry-transferred CVD bilayer graphene. Appl Phys Lett
110(26):263110–1–263110–5. https://doi.org/10.1063/1.4990390
93. Liu D, Lin C, Zhou S, Hu X (2016) Ohmic contact of Au/Mo on Hg1−xCdxTe. J Electron
Mater 45(6):2802–2807. https://doi.org/10.1007/s11664-­016-­4375-­y
94. Song SM, Park JK, Sul OJ, Cho BJ (2012) Determination of work function of graphene under
a metal electrode and its role in contact resistance. Nano Lett 12(8):3887–3892. https://doi.
org/10.1021/nl300266p
95. Suhail A, Pan G, Jenkins D, Islam K (2018) Improved efficiency of graphene/Si Schottky
junction solar cell based on back contact structure and DUV treatment. Carbon 129:520–526.
https://doi.org/10.1016/J.CARBON.2017.12.053
96. Chu J, Mi Z, Tang D (1992) Band-to-band optical absorption in narrow-gap Hg1-xCdxTe semi-
conductors. J Appl Phys 71(8):3955–3961. https://doi.org/10.1063/1.350867
97. Chu J, Li B, Liu K, Tang D (1994) Empirical rule of intrinsic absorption spectroscopy in Hg1-­
xCdxTe. J Appl Phys 75(2):1234–1235. https://doi.org/10.1063/1.356464
98. Liu K, Chu JH, Tang DY (1994) Composition and temperature dependence of the refractive
index in Hg1−xCdxTe. J Appl Phys 75(8):4176–4179. https://doi.org/10.1063/1.356001
99. Zhang BY, Liu T, Meng B, Li X, Liang G, Hu X, Wang QJ (2013) Broadband high photores-
ponse from pure monolayer graphene photodetector. Nat Commun 4(1811):1–11. https://doi.
org/10.1038/ncomms2830
100. Tielrooij KJ, Song JCW, Jensen SA, Centeno A, Pesquera A, Zurutuza Elorza A, Bonn M,
Levitov LS, Koppens FHL (2013) Photoexcitation cascade and multiple hot-carrier genera-
tion in graphene. Nat Phys 9(4):248–252. https://doi.org/10.1038/nphys2564
101. Lee YK, Choi H, Lee H, Lee C, Choi JS, Choi CG, Hwang E, Park JY (2016) Hot carrier
multiplication on graphene/TiO2 Schottky nanodiodes. Sci Rep 6(27549):1–9. https://doi.
org/10.1038/srep27549
102. Ploetzing T, Winzer T, Malic E, Neumaier D, Knorr A, Kurz H (2014) Experimental veri-
fication of carrier multiplication in graphene. Nano Lett 14(9):5371–5375. https://doi.
org/10.1021/nl502114w
103. Johannsen JC, Ulstrup S, Crepaldi A, Cilento F, Zacchigna M, Miwa JA, et al. (2015)
Tunable carrier multiplication and cooling in graphene. Nano Lett 15(1):326–331. https://
doi.org/10.1021/nl503614v
104. Kadi F, Winzer T, Knorr A, Malic E (2015) Impact of doping on the carrier dynamics in gra-
phene. Sci Rep 5(16841):1–7. https://doi.org/10.1038/srep16841
105. Winzer T, Knorr A, Malic E (2010) Carrier multiplication in graphene. Nano Lett
10(12):4839–4843. https://doi.org/10.1021/nl1024485
106. Ganguly S, Tonni FF, Ahmed SZ, Ghuman P, Babu S, Dhar NK, Sood AK (2021) Dissipative
quantum transport study of a bi-layer graphene-CdTe-HgCdTe Heterostructure for MWIR
photodetector. In: IEEE research and applications of photonics in defense conference
(RAPID), pp 1–2. https://doi.org/10.1109/RAPID51799.2021.9521427
202 S. Bansal et al.

107. Ahmed SZ, Tonni FF, Ganguly S, Ghosh AW, Ghuman P, Babu S, Dhar NK, Sood AK
(2020) Using novel properties of graphene for designing efficient infrared photodetectors. In:
Graphene & 2D materials international conference and exhibition, p 34
108. ATLAS user’s manual version 5.20.2.R, SILVACO International, Santa Clara, CA, USA. 2016
109. Rogalski A (2010) Infrared detectors, 2nd edn. CRC Press
110. Dwivedi ADD, Pranav A, Gupta G, Chakrabarti P (2015) Numerical simulation of HgCdTe
based simultaneous MWIR/LWIR photodetector for free space optical communication. Int J
Adv Appl Phys Res 2(1):37–45. https://doi.org/10.15379/2408-­977X.2015.02.01.5
111. Ancona MG (2010) Electron transport in graphene from a diffusion-drift perspective. IEEE
Trans Electron Devices 57(3):681–689. https://doi.org/10.1109/TED.2009.2038644
112. Dwivedi ADD, Chakrabarti P (2007) Modeling and analysis of photoconductive detec-
tors based on Hg1-xCdxTe for free space optical communication. Opt Quant Electron
39(8):627–641. https://doi.org/10.1007/s11082-­007-­9122-­4
113. Bruna M, Borini S (2009) Optical constants of graphene layers in the visible range. Appl
Phys Lett 94(3):031901-1-031901–3. https://doi.org/10.1063/1.3073717
Part II
II–VI Semiconductors–Based Detectors
for Visible and UV Spectral Regions
Chapter 9
CdTe-Based Photodetectors and Solar
Cells

Alessio Bosio

9.1 Introduction

Cadmium telluride has been a known compound since the mid-1800s. In 1879 it
was prepared by the French chemist Margottet by making Te react with metals at
red-heat. In 1888 the enthalpy of formation of the compound was obtained and this
favored the production of the material in crystalline form [1]. The ease of prepara-
tion has been a known feature of CdTe from the very beginning and the basic method
of preparation has changed little to date.
For a long time, the interest for CdTe was purely academic and the only reported
use was as a dye, but in 1946 a publication appeared in which the great photosensi-
tivity of CdTe was highlighted, especially towards the β and γ radiation [2]. The
material was readily proposed as a γ-ray detector if coupled with suitable amplifiers
(scintillators). In the 1950s, photoelectric cells, based on CdTe were already inves-
tigated [3]. These studies highlighted that the Cd and Te stoichiometric excess had
great influence in the operation of these devices [4]. In particular, the heat treat-
ments carried out after the material growth assumed great importance and several
detailed investigations into the behavior of CdTe single crystal were made. During
the same years, the role of dopants was clarified and explained [5] in an organic
model, opening to the opportunity to produce new devices such for example as pho-
todiodes, photoconductive sensors, infrared windows, image intensifiers, camera

This chapter is dedicated to my wife Eugenia and our sons Marco and Silvia for all their forbear-
ance during its writing.

A. Bosio (*)
Department of Mathematical, Physical and Computer Sciences, University of Parma v.le delle
Scienze, Parma, Italy
e-mail: alessio.bosio@unipr.it

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 205
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_9
206 A. Bosio

sensors, photovoltaic cells, X-ray dosimeters, γ-ray (with scintillators) and


ultraviolet-­visible (Uv-Vis) detectors, etc...
Since only some applications found great interest, in this chapter we will describe
only some of these, leaving out the detectors sensitive outside the visible light range,
which will be broadly described in another chapter (IR, X- and γ-ray), focusing on
some special uses/devices and especially, on photodetectors exploiting the photo-
voltaic effect (solar cells).
Effectively, considering its direct forbidden band, this material is considered
very suitable for the manufacture of solar cells. Because of the direct energy gap,
the absorption edge is very sharp allowing more than 90% of the incident light to be
absorbed in a thickness of a few micrometers. Moreover, its direct bandgap of
1.5 eV, at room temperature, perfectly matches the requirement for highly efficient
sunlight energy conversion [6]. The maximum photocurrent that can be produced by
a CdTe-based solar cell under a standard global spectrum light with a power density
of 100 mW/cm2 is 30.5 mA/cm2, allowing a theoretically predicted maximum pho-
tovoltaic conversion efficiency (PCE) of around 32% in the Shockley−Queisser
limit and 30.5% if reflectance is considered [7].
Despite these remarkable properties, it was quickly realized that high conversion
efficiencies were limited by intrinsic defects that formed in II-VI heterojunctions.
Not surprisingly, in the early 1980s efficiencies of around 10% were achieved in
homo- and hetero-junction-based devices using CdTe single crystal [8–11]. In fact,
the higher reported efficiency of 13.4% concerns a n-ITO/p-CdTe single crystal
buried homojunction [12]. On the contrary, all the II-VI heterojunctions show
appreciable lattice mismatch (except for n-CdSe/p-CdTe); in particular, for the
n-CdS/p-CdTe system the lattice mismatch is around 10%. The defects introduced
by the lattice mismatch were responsible for the recombination losses associated
with the junction interface and the conversion efficiency was limited to a value not
exceeding 10%. However, it was not clear why the homojunctions behaved like the
heterojunctions, not having the problem of the lattice mismatch and therefore the
associated defects.
An initial explication to this remark was given considering the severe difficulties
to obtain a low-resistance ohmic contacts to p-CdTe and this problem was common
to every device based on this material. In the same years, the scientific community
observed, not without surprise, that solar cells based on n-CdS/p-CdTe heterojunc-
tions were made by means of the thin-film technology [13]. Soon, this technology
took over, achieving excellent results in terms of conversion efficiency. In fact, the
10% efficiency value was overcome in 1982 [14] followed by an efficiency of 15.8%
in 1993 [15]. In 2001, the National Renewable Energy Laboratory (NREL) reported
an efficiency of 16.5% [16]. Since then, other notable advances have been made and
in 2016 a solar cell based on a CdTe(1 − x)(S, Se)x thin film exhibited a world effi-
ciency record of (22.1 ± 0.5) % [17]. Nowadays, commercial modules with an effi-
ciency of (19 ± 0.9) % are available on the market [18].
Since after several years, a convincing explanation of many of the outstanding
points has been given, in the following, we will report the most important clarifica-
tions to involve the reader into the so varied and apparently contradictory world of
9 CdTe-Based Photodetectors and Solar Cells 207

the CdTe-based devices. So, we will discover together how the problem of ohmic
contact on the p-type CdTe was worked out, how the grain boundaries in polycrys-
talline thin films were passivated and how the defects formed in the heterojunctions
were made harmless, allowing to reach very high efficiencies comparable to the
higher ones obtained with single- and/or multi-crystalline silicon [18].

9.2 Noteworthy Applications

9.2.1 Infrared Window

Considering a prohibited energy gap of 1.5 eV, this material was immediately con-
sidered particularly suitable as a window material for infrared radiation. In the
1960s there was a great demand for transparent materials in the IR, which allowed
the development of CO2 power lasers. The first system made available on the market
was based on vacuum pressed CdTe powder (Irtran 6 by Kodak), but it was only
with the advent of the monocrystalline material that a great boost took place. CdTe
single crystals, grown by Bridgman method, exhibited a very flat transmittance
spectrum up to 40 μm (see Fig. 9.1), which is mandatory for use as window/modula-
tor in high power CO2 lasers, considering the CdTe low thermal conductivity [19] as
well as low mechanical strength. However, it is quite clear, that the extremely flat
transmission (up to ≈ 40 μm) is not easily matched with any other material with
similar energy gap (GaAs), chemical bond or lattice structure.
Thanks to the great demand of IR windows for military use, in the 80–90s the
CdTe production was gradually replaced by two other materials of the II-VI family:

Fig. 9.1 Optical transmission of a CdTe polished single crystal grown by Bridgman method
between 2.5 and 40 μm. In the inset a sketch of the crystal ingot without any anti-reflecting coating
(ARC). Parallelism: 3 arcmin; Flatness: 2 waves at 633 nm
208 A. Bosio

ZnSe and ZnS. These two semiconductors, although not having the same character-
istics as CdTe, have been widely used, since it is possible to grow them in form of
single crystals with the chemical vapor deposition (CVD) technique [20]. This tech-
nique allows to obtain crystals of excellent quality with growth rates much higher
than those typical of the Bridgman method. Today, polycrystalline CdTe IR win-
dows are still available when it is necessary to have good transmittance in the far
infrared and, at the same time, to be completely opaque in the visible region of the
solar spectrum (solar blind feature). This is a characteristic of CdTe which, together
with the transparency width in the IR spectrum, is not reached by any other semi-
conductor currently available on the market. On the contrary, the high cost places
the use of this material only in niche applications [21].

9.2.2 Electro-Optical Modulator

In order to exploit the excellent electro-optical characteristics of CdTe, both in form


of bulk crystalline material or as an epitaxial thin film, the material must be free of
any segregation, with a limited concentration of impurities and with very little crys-
talline stress. Under these conditions there aren’t any phenomena of birefringence,
scattering or absorption. Furthermore, CdTe crystallizes in the zincblende phase
described by the space group 43m ; this group highlights that no center of symme-
try is present, allowing the CdTe system to exhibit the linear electro-optic effect
(Pockels effect). By applying an electric field through the CdTe crystal, it is observed
a variation in the refractive index with a linear behavior with respect to the applied
voltage (see Fig. 9.2).

Fig. 9.2 Schematic of the Pockels effect in cubic non-centrosymmetric media (CdTe). (a) trans-
verse geometry; (b) longitudinal geometry; in both cases the electric field E is parallel to the
(001) direction of the cubic crystal
9 CdTe-Based Photodetectors and Solar Cells 209

This anisotropic variation is typically described by a tensor, but in case of the


43m symmetry, the description of the electro-optical properties can be reduced to
the r41 tensorial element only. As a consequence of the change of the refractive
index, a phase modification of the electromagnetic wave, propagating inside the
crystal, is induced by the applied electrical potential.
The induced phase change ∆ (λ), as a function of the wavelength, in terms of the
applied voltage V is (Eq. 9.1):

2 n3   r41   VL
    ,
d
(9.1)
where λ, n(λ), r41(λ), V, L and d, are the wavelength, the refraction index, the electro-­
optical coefficient, the applied bias, the crystal length along the direction of light
propagation, the distance between the electrical contacts respectively. Therefore, a
CdTe-based modulator is well-described in terms of n3(λ) · r41(λ) and, considering
λ = 10.6 μm, we have: n3(10.6) · r41(10.6) = 1.2 · 10−11 m/V. Frequently, modulators
are specified in terms of the voltage acquired to achieve a phase change ∆(λ) = π
corresponding to a maximum retardance of a half-wave. The voltage producing a
π phase shift is given by Eq. (9.2):

d
V  ,
2 n3   r41    L
(9.2)
At a fixed λ = 10.6 μm and d/L = 1 (i.e., the distance between contacts is equal to
the light path length) Vπ ≈ 53 kV, while for low power modulators d = 2.5 · 10−2 cm,
L = 6 cm (d/L = 4·10−3) and Vπ ≈ 200 V, which is a very intriguing result from the
device point of view. For this reason, CdTe can be considered a good candidate for
the assembly of electro-optical modulators (Pockels cells). Nowadays, very effi-
cient Pockels cells are made with a lot of different non-linear materials (LiNbO3,
LiTaO3, NH4H2PO4 etc.…), but in mid-infrared applications, CdTe is the best suited
material, thanks to its very low absorption coefficient from the band edge at ≈
0.83 μm up to ≈ 35 μm (see Table 9.1). In this wavelength range the optical absorp-
tion coefficient of CdTe is lower than that of many other materials, since the funda-
mental lattice vibration located at ≈ 70 μm (≈ 141 cm−1) is sufficiently far away
from the wavelength of interest (≈ 40 μm). The transmittance in the mid-infrared
region is also due to the low reflectivity since the reststrahlen band is sufficiently far
from the upper limit of the wavelength range. Moreover, anti-reflecting coatings
that give 99% transmission are possible in this wavelength range.
At high light power CdTe crystals exhibit a non-linear behavior, which occurs in
generation of harmonics and frequency intermixing; the driving parameter is the

non-linear susceptibility coefficient. In this case, the  polarization density P
becomes a function of the square of the electric field E:
  
P     0 (    E       E 2   ,
1 2
(9.3)
210 A. Bosio

Table 9.1 Electrooptical properties of some different materials; the values of the linear
electrooptical coefficient and of the refractive index are referred to the specified wavelength
Symmetry ri, j
Material (system) λ [μm] Δλ [μm] [pm/V] n References
SiO2 32 (trigonal) 0.633 0.19– r11=0.20 1.542 [22]
Silicon dioxide 2.30 r41=0.93
GaAs 43 m (cubic) 1.150 1.00– r41=1.51 3.43 [23]
Gallium arsenide 11.0
CdTe 43 m (cubic) 10.60 0.83– r41=6.80 2.6 [24]
Cadmium telluride 35.0
KH2PO4 (KDP) 42 m 0.633 0.20– r41=8.60 1.507 [25]
Potassium dihydrogen (tetragonal) 1.50 r63=10.6
phosphate
KD2PO4 (KD*P) 42 m 0.633 0.20– r41=8.80 1.493 [26]
Potassium dideuterium (tetragonal) 2.10 r63=26.4
phosphate
NH4H2PO4 (ADP) 42 m 0.633 0.18– r41=8.50 1.521 [27]
Ammonium dihydrogen (tetragonal) 1.53 r63=24.1
phosphate
LiNbO3 3 m (trigonal) 0.633 0.33– r13=8.60 2.286 [28]
Lithium niobate 4.50 r22=3.40
r33=30.8
r51=28.0
LiTaO3 3 m (trigonal) 0.633 0.28– r13=8.20 2.176 [29]
Lithium tantalate 4.00 r22=0.50
r33=35.0
r15=20.0
CdS 6 mm 10.60 0.60–14 r13=2.45 2.226 [23]
Cadmium sulphide (hexagonal) r33=2.75
r42=1.70
LiIO3 6 (hexagonal) 0.633 0.30– r13=4.10 1.883 [30]
Lithium iodate 5.5 r23=6.40
r41=1.40
r51=3.30
Abbreviations: n refractive index, ri, j linear electrooptical coefficient, λ wavelength, Δλ transpar-
ency range


stopping the expression for P   to the second order. The quantity χ(2) is the
so-called second-order non-linear optical susceptibility and ε0 is the permittivity of
free space. In particular, χ(2) is a second order tensor, which could be greatly simpli-
fied by means of the crystal symmetry. CdTe crystals, at a wavelength of 10.6 μm
(0.117 eV), exhibit a non-linear susceptibility coefficient χ(2) = 1.7 × 10−10m/V [31]
that is one of the greater values among the harmonic generator materials, compara-
ble with χ(2) = 1.88 × 10−10m/V typical value for GaAs. From the symmetry of the
CdTe crystal, the most efficient second harmonic generation occurs for the input
electromagnetic wave along (110) with the electric field polarized in (110).
It is worth remembering here, that the power of the second harmonic (at fre-
quency 2ω) is proportional to the square of the non-linear susceptibility coefficient
9 CdTe-Based Photodetectors and Solar Cells 211

χ(2) as well as to the square of the power density of the incident light beam (at a
frequency ω). For example, KDP, KD*P and ADP exhibit a non-linear susceptibility
coefficient of about (0.5 ÷ 0.6) × 10−10m/V [32]. With the same incident light power,
by using CdTe or GaAs crystals as second harmonic generators, a power density 10
times greater, could obtained.

9.2.3 UV-Vis-Photodetector

Many CdTe-based devices are realized using bulk crystals, but where it is important
to exploit large areas, such as in solar cells, the preferred technology is thin film.
Furthermore, low-dimensional semiconductor nanostructures, such as quantum dots
(QDs), nanowires (NWs), nanorods (NRs), nanotubes (NTs) and nanobelts (NBs)
[33–37] are considered the most sensitive and fast responsive materials for photon
sensors due to their large surface-to-volume ratio and direct pathway for charge
transport. Nanostructures are very attractive due to their unique optical properties,
which are generated by the quantum confinement effect. In fact, by modifying the
size of the nanoparticles (NPs) it is possible to change the charge carrier confine-
ment while band gap engineering can be achieved by changing the distance between
the NPs. CdTe nanostructures have been synthetized via many techniques, including
physical vapor deposition (PVD), close-spaced deposition (CSS), chemical vapor
deposition (CVD), molecular beam epitaxy (MBE) and RF magnetron sputtering
(RFMS) [38–42].
CdTe-based photodetectors have been made with different structures such as
Schottky barriers, p-n junctions, pure photoconductive metal-semiconductor-metal
devices and field-effect transistors (FETs). Among these deposition techniques and
device configurations it is hard to find the best combination, since each of them has
its advantages, but also some drawbacks. In fact, for a high-performance photode-
tector a proper spectral range, high-signal-noise ratio, high responsivity and fast
response are important requirements, but the most characterizing parameter is
undoubtedly the photo-gain, defined as the number of photogenerated carriers per
incident photon collected at the external contacts. This is the case of CdTe-based
photoconductive detectors which are highly sensitive thanks to the presence of deep
traps in the semiconductor layer, which causes a long recombination lifetime τ for
one type of charge carriers. As it is well explained by Dan et al. in [43] the photo-­
gain depends on the ratio between the charge trapping lifetime τ and the transit time
τt and a long recombination lifetime leads to a long response time, which restrains
the photodetector application.
It’s a matter of fact that a high photo-gain is coupled with a low response speed
and a good photodetector, depending on its use, is the result of the best compromise
between these two parameters. From this point of view, nanostructures have an
advantage over bulk materials or thin films since, at the nanoscale, it is virtually
easier to extract the photogenerated carriers, to obtain large photogain and
212 A. Bosio

contemporarily fast response time, defined as the rise time from dark current to 90%
of the maximum current. Effectively, CdTe-based nanowires and nanoribbons
devices show response time on the order of 1 and 3.3 s [44, 45], respectively and a
photogain in the range of 103–104, while CdTe thin film photodetector exhibits aver-
age response times, on the order of tenth of seconds. So long response times are
generally due to the detrapping lifetime of the deep traps. It has been proposed to
remove the deep traps keeping the shallow traps could maintain the high gain and,
at the same time, a fast response speed. In fact, the shallow traps can be thermally
activated, promoting the release of the charge carriers more quickly. As an example
of this technique, a CdTe QDs-based photodetector capped with poly(3-hex-­
ylthiophene) (P3HT), with enhanced behavior through trap engineering is reported
in [46]. In that case the Cdi2+ deep traps (trap depth of 0.64 eV), on the surface of
CdTe QDs, were passivated by P3HT coordination, while shallow traps can still
activate a high photogain of 50 and a short response time of 2 μs. For the same
device, without any surface passivation of the QDs, the exhibited response time was
about 1–2 s. CdTe microwire (μW)-based ultraviolet photodetectors reported in
[47], not having any passivation of the deep-traps, showed a pohotogain of ≈ 50 and
an average response time of about 7.7 s.
In [45] a CdTe-based Field Effect Transistor, by using single-crystalline NW,
was realized. Exploiting the good properties of the CdTe single crystal an average
photogain of 250 and a response time of 0,7 s were obtained. Despite an excellent
crystalline quality, the photoresponse was just sufficient, indicating that the
­performance of this device could be enhanced if a deep traps engineering will be
adopted.
Other photodetector structures are realized for obtaining high response speed, as
in the case of CdS-CdSxTe1-x-CdTe core-shell NB detector [48]. This promising
device exploits the built-in potential which is formed at the interface between the
three materials. In CdTe- and CdS-based NB detectors, photoresponse is principally
ruled by defects heavily influenced by absorption and desorption of oxygen atoms
on the surface of the nanoparticles, in the above-mentioned core-shell NB-based
devices, the photogenerated carriers are quickly separated by the built-in potential,
naturally formed at the CdS-CdSxTe1-x and CdSxTe1-x-CdTe interfaces. The absorp-
tion/desorption of atoms, on the surface of the nanoparticles is very slow, if com-
pared with the separation of charge carriers due to the interface potential. As a
result, a response time of the order of 11 μs for the rise time and 23 μs for the fall
time is obtained. For comparison, CdS NB photodetector exhibits typical response
time of 16 and 367 μs respectively. By the way, this core-shell NB detector shows
an excellent photogain of about 4.7 × 103 mainly due to the simultaneous absorption
of photons, with energy larger than the energy gap of CdTe, by all the three constitu-
ent materials. Table 9.2 summarizes important results obtained by some photodetec-
tor described above.
9 CdTe-Based Photodetectors and Solar Cells 213

Table 9.2 Typical parameters of different CdTe-based photodetectors. All the reported values are
optimized by varying wavelengths and voltages as indicated in the reported references
Responsivity Photoresponse rise/ Spectral
Photodetector [A/W] Photogain decay time [s] range [nm] References
CdTe – NR 780 2400 1.1/3.3 400–800 [44]
CdTe – NW FET 80.1 250 0.7/1 400– [45]
CdTe – QD 13.6 50 0.2·10−6/− 390– [46]
P3HT-capped
CdTe – μW – 50 7.7/0.06 365– [47]
CdTe core-shell 1520 4700 11·10−6/23·10−6 355–785 [48]
NB

Fig. 9.3 (a) maximum efficiency (ηmax) vs energy gap (Eg) for two different atmosphere absorp-
tion; AM 1 = sea level, sun at zenith; AM 0 = outside the earth atmosphere; AM = air mass. The
vertical dashed lines indicate different semiconductors industrially used as absorbers in solar cells.
(Adapted from Ref. [49]. With the permission of AIP Publishing). (b) The absorption coefficient α
of different semiconductors as a function of energy

9.3 Solar Cells

CdTe, the most commercially successful TF technology, finds its fortune in some
physical and chemical peculiarities:
1. direct energy band gap of 1.45 eV near the maximum of the solar spectrum
(Fig. 9.3a) [49];
2. absorption coefficient in the visible part of the solar spectrum in the range (104 ÷
105) cm−1, which means that 1 μm thick layer is sufficient to capture all visible
light (Fig. 9.3b);
3. it shows a high enthalpy of formation (100 kJ mol−1), which means great thermo-
dynamic stability;
214 A. Bosio

4. it sublimates and evaporates congruently by means of the equilibrium reaction


1
CdTe  Cd  Te2 . The high stability and congruent evaporation allow growth
2
by very different preparation methods;
5. it naturally grows with inherent stoichiometry defects making it moder-
ately p-type;
6. post-growth treatment reduces defects, increases crystalline quality by making
grain boundaries electrically inactive.
Historically, the best performance has been achieved with heterojunctions where the
n-type partner was cadmium sulfide (CdS). A few different attempts have been
made when p-type CdTe single crystals were paired with In2O3 [12], ZnO [9] or a
very thin n-type CdTe layer [11], obtaining a 13.8% maximum efficiency.
Surprisingly, CdTe-based solar cells, produced using thin-film technology, show
higher efficiencies than those made using monocrystalline materials (see Table 9.3).
It is a matter of fact that the success of this material was achieved by exploiting
one of its best features, namely the possibility of producing a complete solar cell by
using the thin film technology. Total thin-film CdTe/CdS heterogiunction resulted in
6% conversion efficiency as it has been known since 1972 [13]. However, the psy-
chological limit of 10% efficiency was exceeded in the 1980s only after the applica-
tion of a heat treatment in a chlorine atmosphere to the stacked CdTe/CdS layers
[14]. In the following 10 years, devices with efficiency close to 17% were optimized
by developing new front and back contacts [16, 50].

Table 9.3 Representative data for single crystal and all thin film CdTe-based solar cells
Open-circuit Short-circuit
voltage Voc current density Energy conversion
Type of cell [V] [mA/cm2] efficiency, h [%] References
CdTe single crystal
Buried homojunction: 890 20.0a 13.4 [12]
n-ITO/p-CdTe
Heterojunction: 540 19.5a 8.8 [9]
n-ZnO/p-CdTe
CdTe homojunction: 820 21.0a 10.7 [11]
p-CdTe(CSVT)/n-CdTe
single crystal
Thin films
All thin film CdTe-­ 6.0 [13]
based cells
CdS and CdTe (low 750 17.0b 10.5 [14]
T-CSS)
CdS (CBD) – CdTe 843 25.1a 15.8 [15]
(high T-CSS)
CdS (CBD) – CdTe 845 25.9a 16.5 [16]
(low T-CSS)
CSVT Close-Spaced Vapor Transport, CSS Close-Spaced Sublimation
a
Under simulated AM 1.5 solar illumination at 100 mW/cm2
b
Under simulated AM 2 solar illumination at 75 mW/cm2
9 CdTe-Based Photodetectors and Solar Cells 215

These noceably succcess was followed by a period in which many researchers


have dedicated their activities to the technology transfer from laboratory to indus-
trial scale of the production processes, slowing down the achievement of ever higher
device efficiencies. It was increasingly realized that some difficulties lay mainly in
the stability over time of the back contact and in the real impossibility of extrinsi-
cally doping the polycrystalline thin films of CdTe. Around 2010, PCE started to
increase again, quickly reaching values close to 20%. The continuous increase in
conversion efficiency of that devices, was principally due to the optimization of the
anti-reflective coating (ARC), the transparent electrical contact and the window
layer, in addition to a careful choice of the glass substrate. This resulted in a consid-
erable increase in photocurrent, going from 26.1 mA/cm2 for a cell with 17.6%
efficiency [16] to a photocurrent of 28.59 mA/cm2 for a cell with an efficiency of
19.6% corresponding only to a slightly increase in photovoltage [51]. Unfortunately,
there are no details in the literature regarding these impressive results. Nevertheless,
it immediately appeared evident that the increase in photocurrent is not only due to
an excellent optimization of light collection, but also to an accurate management of
the energy gap of the CdTe to extend the absorption of light to longer wavelengths.
CdTe(1-X)(S,Se)X compounds show energy bandgap values lower than CdTe when
X ≤ 0.05, corresponding to a maximum increase in the cut-off wavelength of
approximately 15 nm. The corresponding increase in the photocurrent can be esti-
mated at 1 mA/cm2. Furthermore, the use of CdTe(1-X)(S,Se)X alloy, by reducing the
lattice mismatch between the window and the absorber materials in the region of the
metallurgical junction, results in improved charge transport properties since fewer
recombination centers and killer levels due to interface states are present.
In 2015, a solar cell with a thin CdTe(1-X)(S,Se)X intermixed layer in the junction
region showed a world record efficiency of (22.1 ± 0.5)%, exhibiting the following
parameters measured with the AM1.5G spectrum (1000 W/m2) at 25 °C:
Voc = 0.8872 V, Isc = 31.69 mA/cm2, fill factor = 0.785 over a designated exposed
area of 0.4798 cm2 [51, 52]. Nowadays, the most used and accepted architecture of
the CdTe thin film solar cell is sketched in Fig. 9.4.
The electrical in-series connections of adjacent cells, monolithically integrated
inside the production process, by means of a robotic laser scribing, made possible a
fully automated large-scale in-line production process. In 2002, by exploiting the
electrodeposition technique for CdTe-deposition and laser scribing for the elecrical
in-serie connections a 11% efficient module was realized [53]. In 2010–2011, the
number of factories able to produce tens of megawatts/years of CdTe-based mod-
ules were about 10 units. The CdTe films deposition was mainly based on the close-­
spaced vapor transport (CSVT) or close-spaced sublimation (CSS) techniques and
large-area modules, exhibiting efficiency in the 10% to 12% range was obtained
[54]. In the following years, a 14.4% efficient device was reported, which became
16.1% at the beginning of 2013. Immediately after (2014), a 17.5% efficient module
was obtained, followed by a world efficiency record of (18.6 ± 0.5)% in 2015. The
photovoltaic parameters of such a module, taken under the global AM 1.5 spectrum
(1000 W/m2) at a cell temperature of 25 °C, are: Voc = 110.6 V, Isc = 1.533 A and fill
factor = 0.742 over a designated illumination area of 7038 cm2 [51]. Nowadays,
216 A. Bosio

Fig. 9.4 The CdTe/CdS solar cell in superstrate configuration (light enters through the substrate).
(Reprinted from “A. Bosio, S. Pasini, N. Romeo, The History of Photovoltaics with Emphasis on
CdTe Solar Cells and Modules, Coatings 2020, 10, 344”. Published 2020 by MDPI as open access)

Table 9.4 Photovoltaic Conversion Efficiency (PCE) of the best solar cells and commercial
modules collected for different technologies
Solar cells Modules
Parameter PCE, % Area, cm2 Producer PCE, % Area, cm2 Producer
CdTe 22.1 0.4798* First solar 19.0 23,573* First solar
CIGS 23.35 1043* Solar Frontier 18.6 10,858** Miasolé
CIS 15.4 100.0*** IPE – – –
CZTS 12.6 0.4209** IBM – – –
Cu2S 10.0 ≈1.0 IEC – – –
GaAs 6-J 47.1§ 0.099 NREL – – –
GaAs 3-J 37.9 1047** Sharp 31.2 968* Sharp
GaAs SJ 32.8 1000** LG electronics – – –
Si single- crystal 26.7 79* Kaneka 24.4 13,177* Kaneka
Si multi-crystal 23.2 247.79*** Trina solar 20.4 14,818** Hanwa Q CellsH
The confirmed PCE data are measured under the global AM 1.5 spectrum (1000 W/m2) at 25 °C
(IEC 60904–3: 2008, ASTM G-173-03 global)
*(da) = designated illumination area; ** (ap) = aperture area; *** (t) = total area; § 6-J = six-­
junction with concentrator (143X)

commercial modules with an efficiency close to 19.0% are commonly available on


the market [55].
In today’s photovoltaic panorama, the technology based on CdTe shows conver-
sion efficiencies comparable to those of CIGS and of both single- and poly-­crystalline
Si, the most used material in photovoltaic modules production. An overview of the
best performing solar cells and commercial modules is shown in Table 9.4.
A CdTe-based solar cell is typically made with very thin overlapping stacked
layers, arranged to form a high-quality heterojunction. When high-temperature
deposition techniques are used, generally p-type CdTe thin films are naturally
9 CdTe-Based Photodetectors and Solar Cells 217

obtained, forcing to select an n-type partner to make the p-n junction. The high-­
efficiency CdTe-based solar cells are manufactured in a superstrate configuration,
which means that light passes through the substrate and the front contact is as trans-
parent and conductive as possible. On the other hand, the back-contact, which is
generally opaque, must ensure the ohmicity with p-type CdTe to efficiently collect
all the photogenerated carriers.
To date, the most widely used system is the CdS/CdTe heterojunction, where
CdS constitutes the “window”, and CdTe is the “absorber”. A pressing demand for
this device is that the window layer must be as transparent as possible, while the
absorber layer must be thick enough to fully absorb visible light. A thickness of a
few microns is enough for CdTe, while thin films up to a few tens of nm are used for
CdS. At the operating temperature, the concentration of free carriers of both CdS
and CdTe films ensures that the electric field falls mainly into the absorber material,
so that all the photogenerated electron-hole pairs can be separated and pushed
through the material, enhancing the collection of the charge carriers. The electrodes
complete the device, ensuring the passage of the photocurrent. Now, we will review
the main requirements the individual materials should have.

9.3.1 The Substrate

When realized in superstrate configuration, high efficiency CdTe-based solar cells


and modules normally make use of the so-called “soda-lime glass” (SLG), which is
the common window glass. Since sunlight passes through the substrate, the glass
must be as transparent as possible to minimize the parasitic absorption of the visible
light. Inexpensive minerals such as trona, sand, and feldspar are typically used in
place of pure chemicals, making soda lime glass cost-effective. Unfortunately, start-
ing from minerals, some impurities are inadvertently introduced, such as Fe2O3 and
MnO2, which are responsible for a decrease in transparency in the visible part of the
solar spectrum. For this reason, iron-free glasses are normally used in CdTe technol-
ogy, obtaining 8% more transparency at short wavelengths (λ <300 nm) [56].
Attempts to employ flexible substrates have been made using polymers, such as
Dupont’s polyimide, which enables high-speed roll-to-roll technology to produce
large-area and very light cost-effective photovoltaic devices. Figure 9.5 shows the
optical transparency of standard SLG and polyimide, which are comparable at long
wavelengths, but the transmittance of polyimide below 530 nm is not sufficient for
photovoltaic application, despite using thin foils (typical thickness 7 μm). This
results in a photocurrent loss of at least 3 mA/cm2. For this reason, together with the
limitation of the process temperature, solar cells made with this polymer have not,
until now, presented an efficiency greater than 14% [57].
218 A. Bosio

Fig. 9.5 Transmittance spectra of commonly used substrates such as soda-lime glass (SLG) and
polyimide in CdTe-based production process of solar cells and modules

Fig. 9.6 Transmittance spectra of SLG (3.3 mm thick) covered with only ITO or FTO and SLG
covered with ITO or FTO coupled with ZnO as a high-resistivity transparent (HRT) layer.
(Reprinted from “A. Bosio, S. Pasini, N. Romeo, The History of Photovoltaics with Emphasis on
CdTe Solar Cells and Modules, Coatings 2020, 10, 344”. Published 2020 by MDPI as open access)

9.3.2 The Front Contact

In thin-film photovoltaic technology, one of the most stringent needs is the use of
transparent and conductive electrical contacts, capable of passing light and to be
electrically conductive to effectively collect the photogenerated charge carriers
without introducing unnecessary sheet resistance. For this purpose, high-­
performance transparent and conductive oxides (TCOs), exhibiting transparencies
to visible light close to 90% (see Fig. 9.6) and electrical conductivity up to
104 Ω−1·cm−1 are normally used. These two opposing requirements are achieved by
considering sufficiently high energy gaps (larger than 3 eV) in heavily doped
9 CdTe-Based Photodetectors and Solar Cells 219

degenerate semiconductor characterized by the Fermi level inside the conduction


band. This particularly condition promotes the Burstein–Moss effect widening the
energy gap and increasing, consequently, the transparency [58].
Near-degenerate semiconductors exhibit the typical free-carrier absorption in the
near-infrared (NIR) making some of these TCOs not widely used in large-scale PV
production. To overcome this drawback, semiconductors exhibiting high mobility
of the charge carriers are commonly used, thus obviating the need to be degenerate
to obtain high electrical conductivities.
The most common TCOs used in CdTe technology, are: Sn-doped In2O3 (ITO),
F-doped SnO2 (FTO), Al-doped ZnO (AZO) and Cd2SnO4 (CTO). All these TCOs
are generally coupled with high-resistivity buffer layers (HRT) for reducing shunt
effects caused by pinholes in active layers and to hinder the diffusion of impurities
from the TCO layers or from the substrate. Un-doped SnO2, In2O3 and ZnO are HRT
layers commonly coupled with ITO, while FTO is generally paired with pure SnO2
and CTO is combined with Zn2SnO4 (ZTO) [59]. The typical double layer structure
of these TCOs modifies the chemical and physical interaction between the front
contact and the window layer, as seen when ZnO and CdS, heat-treated at high tem-
perature, form a mixing layer, which changes the optical properties of both films.

9.3.3 The Window Layer

The most widely used n-type partner with CdTe is cadmium sulfide (CdS), which
exhibits n-type conduction due to stoichiometric defects, such as sulfur vacancies,
which form during film growth. With an energy gap of 2.42 eV, it allows sunlight to
pass up to a wavelength of 512 nm, cutting the wavelengths of the near ultraviolet.
With a typical dark resistivity of the order of (106–107) Ω·cm, CdS is not particularly
suited as window material in solar cells because, in an efficient p/n junction, the
electric field must principally fall into the p-type region (CdTe). This important
requirement is satisfied if the spatial density of the p-type carrier in CdTe is signifi-
cantly less than the density of the n-type carriers in CdS. Under sunlight this condi-
tion is satisfied thanks to the photoconductivity of CdS which helps to distribute the
electric field into the p-type film. Considering the transparency, since nothing can be
done about the absorption coefficient, very thin films, with a thickness of only
100 nm, give optical density suitable to be used as window layer in CdTe-based
solar cells.
RF sputtering, chemical bath deposition (CBD) and high vacuum thermal evapo-
ration (HVTE) belong to low-temperature (L-T) processes, while close-spaced sub-
limation (CSS) and close-spaced vapor transport (CSVT) are high-temperature
(H-T) deposition procedures. As a consequence, CdS is normally deposited at a
substrate temperature lower or higher than 200 °C. The choice of the deposition
technique is crucial, since the quality of the solar cell relies on the interaction
between the active layers, which in turn strongly depends on the deposition tem-
perature of all the layers. The CBD process produces high-quality pin-hole free CdS
220 A. Bosio

films characterized by a very high density and compactness. A heat-treatment at


400 °C is needed to remove the Cd and S excesses naturally present in the CBD
deposited films. While this deposition method offers excellent results, in industrial
manufacturing, sputtering and H-T processes are generally preferred as CBD is a
low-speed process producing a large amount of waste, which must be costly
recycled.
Among the L-T technique, sputtering is the best suited for large scale in-line
production. This technique is considered an L-T deposition technique even though
the growing film surface is continuously bombarded with impinging electrons and
atoms, which exchange their kinetic energy and promote the typical effects of high
temperatures. Indeed, RF sputter deposition is not suitable for producing CdS films
with the quality and chemical stability suitable for use in CdTe-based solar cells
(see Fig. 9.7). Only reactive RF sputtering, which introduces oxidizing atoms into
the process chamber, such as gases containing fluorine or oxygen, produces high
quality CdS films that can form a good CdS/CdTe heterojunction. What happens in
the sputter discharge when a hydrofluorocarbon or oxygen gas is introduced into the
process gas (Ar) is described in [60].
Between the H-T deposition techniques, CSS is the most used. CdS films
obtained with this process show superior quality, even if depositions carried out in
pure Ar provide low density films with many pinholes. If oxygen is added to the
process chamber, the deposition equilibrium is substantially altered, growth slows
down, since grain boundaries of the growing film are decorated with oxides such as
CdSO4 and CdSO3. As a result, a denser film without pinholes is obtained, but the
surface of the CdS film is covered with an oxide layer. For this reason, high effi-
ciency solar cells are obtained when CSS-deposited CdS films in Ar + O2 atmo-
sphere, are heat-treated at high temperature (400 °C) in presence of hydrogen.

Fig. 9.7 Transmittance spectra of sputtered CdS film: deposited in pure argon or deposited in
argon + CHF3
9 CdTe-Based Photodetectors and Solar Cells 221

9.3.4 The Absorber Layer

The Close-Spaced Sublimation of CdTe is made possible since at high temperature,


CdTe dissociates into its elements Cd and Te, which can recombine on the substrate
to form the CdTe film. Generally, with this technique the CdTe films are deposited
in the temperature range of (500–600) °C in an Ar atmosphere at a pressure in the
(1–100) mbar range. If soda lime glass is used a temperature of 520 °C cannot be
exceeded.
CdTe film thickness of 2 μm is more than enough to absorb all the visible light,
since CdTe exhibits an absorption coefficient in the range of (104/105) cm−1. This
thickness is optimal both for optical and electrical requirements, even though it’s
very difficult to obtain CSS-deposited pinhole-free films with proper qualities and
compactness. For this reason, oxygen is generally added to the inert gas in the CSS
or CSVT deposition chamber. Due to the presence of CdO and TeO2 species, a
greater superficial diffusion of the incoming Cd and Te atoms is expected. The pres-
ence of oxygen increases the number of nucleation sites promoting a denser growth
of the CdTe film (see Fig. 9.8). In these conditions a small quantity of CdTeO3 is
formed, which has an appreciated effect of passivation of the grain boundaries [61].
As already mentioned, CdS/CdTe junction can never work because a 9.7% lattice
mismatch between the two materials generates too many interface defects, which
can capture the photogenerated carriers crossing the junction. It is now established
that a good way to overcome this drawback is to create in the junction region a
mixed compound, namely CdSXTe(1-X), between the two active materials. In this way
the lattice mismatch is progressively adapted, and the number of defects at the inter-
face can be significantly reduced. Since the CSS-deposition of CdTe on top of CdS
occurs at a high temperature, the beneficial intermixing between these materials
begins to take place. Moreover, under AM 1.5 G solar light a solar cell reaches its
maximum efficiency when the energy gap of the absorber is 1.34 eV. The energy
gap of CdTe is a little bit wider, being 1.5 eV, but could be adjusted exploiting the

Fig. 9.8 Atomic Force Microscope (AFM) image of the surface of a CdTe film deposited by CSS;
(a) deposition performed in pure Ar, (b) in Ar + 10%O2
222 A. Bosio

favorable Cd-Te-Se phase diagram. In fact, according to [62] a mixed CdTe(1 − X)SeX
compound with X ≈ 0.4, exhibits an energy gap Eg = 1.4 eV. With the CdTe0.6Se0.4
alloy a better collection of the solar light spectrum, extended to a longer wave-
length, is obtained. Nowadays, an engineered profile of the energy gap of the CdS/
CdTe(1 − X)SeX is beneficial to improve the photocurrent maintaining the high photo-
voltage typical of the CdTe absorber.
Similarly, to what is done in CIGS-based cells, the small gap material is used
only in a thin surface layer of the absorber in such a way to not compromise the typi-
cal photovoltage of the CdTe bulk. Besides, when a CdS/CdTe(1 − X)SeX heterojunc-
tion is made, it was found that the thickness of the CdS layer can be drastically
reduced or the layer can be even avoided, decreasing the parasitic absorption in the
high-energy region of the solar spectrum and obtaining an increasing photocurrent
[63]. Additional improvements can be acquired if the CdS is replaced with MgZnO,
a material with which CdTe and CdS/CdTe(1 − X)SeX have a better alignment to the
energy band. This material is characterized by a better transparency in the high-­
energy region with respect to CdS [64, 65]. Furthermore, a greater passivation effect
of the interface defects due to the presence of Se is effective for a longer lifetime of
the charge carrier in CdS/CdTe(1 − X)SeX compared to CdTe [66]. By adopting these
methods, solar cells with an efficiency of over 22% were realized, including the
world record [51].
High-efficiency solar cells have also been made on flexible polymeric substrates,
where a L-T process is essential. CdTe and CdS films, deposited by HVTE at a
temperature of about 200 °C with a typical thickness of (2/3) μm, show excellent
smoothness and compactness with a characteristic grain size of (100/500) nm [67].
The size of the crystallites suggests that the mobilities and lifetime of the charge
carriers are very low, unsuitable for forming an efficient p-n junction capable of
collecting the photogenerated carriers. Never, as in this case, is a post-deposition
treatment necessary, the so-called “chlorine treatment”. This annealing in a chlorine
atmosphere can reduce stacking defects, mismatch dislocations and grain boundar-
ies, thereby increasing the crystal grain size [68].
Among the L-T techniques, electrodeposition was originally presented in the
1970s [69, 70], but only in the early 1980s was an electrodeposited thin-film CdTe-­
based solar cell developed, showing a remarkable efficiency of 8% [71–73].
Normally, the electrodeposited CdTe layer exhibits n-type conductivity. As a result,
an innovative SLG/FTO/n-CdS/n-CdTe/p-CdTe/Au layer sequence was imple-
mented. The main improvement is represented by a buried homojunction near the
back-contact. The presence of a thin p-type CdTe film close to the back contact
allows a better control of the junction position within the absorber layer, with an
increase of the barrier height and a consequent good collection of the photogene-
rated electron-hole couples [74].
When the CdS and CdTe layers are deposited by sputtering, the electro-optical
properties of the films strongly depend on the sputtering parameters, as well as on
the sputtering power density, on the Ar and reactive gas pressure, on the polarization
voltage, on the substrate temperature and on the target-substrate distance. These
9 CdTe-Based Photodetectors and Solar Cells 223

numerous varieties of process parameters allow a very fine tuning of the physical
characteristics of the growing film. Taking advantage of the excellent coverage and
high density of the sputter-deposited films, a 2 μm thick layer of CdTe is used to
realize very thin solar cells [75]. By appropriately regulating the chlorine treatment,
it would be possible to produce CdTe thin films with crystalline grains large enough
to obtain high efficiency solar cells. Following this philosophy, a 14% efficient solar
cell was made in the early 2000s [76].

9.3.5 The Heat Treatment in Chlorine Atmosphere

It is commonly accepted by the scientific community that chlorine treatment


increases the grain size of the CdTe film, improves the quality of the grain boundar-
ies and promotes the intermixing of CdS and CdTe layers at their interface. If this
treatment is not performed, the short-circuit current of the solar cell is very low
therefore the efficiency is very poor. The treatment is generally carried out by
depositing a CdCl2 film on top of the CdTe layer by evaporation or by dipping the
CdTe layer in a solution of CdCl2-methanol. The chlorine-treatment is completed
keeping the CdTe + CdCl2 system at a temperature in the range of (350–400) °C in
air or in an inert gas such as Ar at atmospheric pressure [77, 78]. Recently, alterna-
tive Cl2-containing compounds were proposed such as MgCl2, NaCl and NH4Cl
[79–84]. Cd-free salts have the great advantage of being much more environmen-
tally sustainable than CdCl2. Moreover, another way to perform a Cd-free chlorine
treatment is to use a halogen gas capable of releasing chlorine or chlorine-­containing
radicals at the processing temperature. These gases belong to the chlorine-based
Freon family or to the same family of chlorine, such as HCl or Cl2, but hydrochloric
acid or chlorine are very aggressive gases making the CdTe treatment extremely
critical [85, 86]. Since 2011, the industrial use of chlorinated Freon gases has been
banned in Europe because they are ozone depleting substances (ODS). For this rea-
son, other chlorinated hydrocarbons, effective in the treatment of CdTe and not
considered ODS, were considered. Possible candidates have been identified among
liquid chlorinated hydrocarbons (LCHY).
It is assumed that the reaction on the CdTe surface during the treatment is:
1. if Cl2 is supplied by means of chlorine-based salt

CdTe  s   CdCl2  2Cd  g   1 / 2Te2  g   Cl2  g   CdCl2  s   CdTe  s  (9.4)


2. if Cl2 is supplied by means of chlorine-based gas

CdTe  s   2Cl2  g   CdCl2  g   TeCl2  g   2Cl2  g   CdTe  s  (9.5)


After the treatment the CdTe surface morphology is completely changed due to an
increase in the size of the small grains (see Fig. 9.9a, b). The morphology of the
224 A. Bosio

surfaces shown in Fig. 9.9b presents the typical mesa-like structure resulting from
the etching of the surface by chlorine and/or hydrogen chloride. The crystalline
grains tend to coalescence, becoming more compact with narrow boundaries.
From a stoichiometric point of view, a CSS-deposited CdTe film shows a Te-rich
surface if Cl2 is supplied by means of chlorine-based gas. In this case, the peculiar-
ity is that this treatment does not introduce Cd on the surface of the CdTe film from
outside, leaving the Cd-vacancies not compensated; this fact cannot happen if a
cadmium salt such as CdCl2 is used. This important feature could be exploited by
carrying out the back-contact directly on the surface of the CdTe layer since, to
obtain good ohmic contact with p-type CdTe, a Te-rich surface is required.

9.3.6 The Back-Contact

In the production of high efficiency CdS/CdTe solar cells, the realization of a stable
over time ohmic back-contact has long been critical. Commonly, many manufactur-
ing processes use a Cu-containing compound, such as the Cu-Au alloy, Cu2Te,
ZnTe:Cu and Cu2S [87–90]. Excluding the Cu-Au bilayer, the electrical contact is
terminated with a sputtering deposition of a metallic layer such as Mo, Ni-V or with
a graphite paste. The diffusion of copper reduces the resistivity of the CdTe layer,
resulting effective in the formation of a very low resistance ohmic contact suitable
for high-performance solar cells.
Devices made with non-Cu-containing contacts exhibit high series resistance,
mainly due to the non-ohmic contact. For this reason, very high efficiency solar
cells have been realized, exploiting the presence of a Cu-based compound in the
back contact.

Fig. 9.9 SEM image of the surface morphology of a CSS-deposited CdTe film at a substrate tem-
perature of 520 ° C. The thickness of the CdTe film is 6 μm. (a) as deposited, (b) after heat-­
treatment in LCHY atmosphere. (Reprinted from “A. Bosio, S. Pasini, N. Romeo, The History of
Photovoltaics with Emphasis on CdTe Solar Cells and Modules, Coatings 2020, 10, 344”.
Published 2020 by MDPI as open access)
9 CdTe-Based Photodetectors and Solar Cells 225

Typically, a chemical attack is performed in Br2-methanol or in a mixture of


HNO3/HPO3 acids (N-P), which leaves the CdTe surface Te-rich. A Te-rich surface
favors the correct band alignment between the valence band of the CdTe and the
working function of the metal contact and/or the electronic affinity of a degenerated
semiconductor, through the formation of a low resistance tunneling barrier. Since
copper is very reactive with Te, a thin layer of Cu2Te is formed at the interface
between CdTe and the back contact, which limits the diffusion of copper into the
grain boundaries of the absorber film. Unfortunately, Cu2Te is not a stable com-
pound, triggering the usual diffusion of Cu atoms into the grain boundaries. To
overcome this problem, the amount of copper is carefully controlled. Several solu-
tions are proposed; one of the most reliable methods makes use of a very thin layer
of Cu (1/2 nm thick) or by means of a buffer layer, which acts as a filter for the
diffusion of Cu atoms. In both cases the formation of a layer of CuXTe is necessary
to ensure a good contact able to stop the diffusion of Cu. It was discovered that
CuXTe behaves as a stable material only if x ≤ 1.4. This condition is obtained by
placing a buffer layer, based on M2Te3, between the very thin layer of Cu and the
surface of CdTe (where M = Sb, Bi, As) [91, 92]. Two different phenomena could
take place:
(i) M atoms can bind tellurium excess, forming a M2Te3 or CuXTe interface,
confirming the fundamental role of a Te-rich surface of the CdTe film.
(ii) A heat treatment in air of the whole system, carried out at a temperature of
200 °C for 15 min, favors the formation of an oxide compound into the grain
boundaries of the CdTe layer, which contributes to the passivation of the grain
boundaries by removing their ability to recombine charge carriers. A similar
result is obtained if a ZnTe film is used as a buffer layer [93].
This is a general recipe; if respected, an efficient and stable over-time solar cell is
regularly produced, which is the basis of an industrial production of large-sized
modules based on CdTe.

9.4 Conclusion

Since its discovery, CdTe has immediately proved to be a very promising material
for the realization of electro-optical devices thanks to the fact that good-quality
single crystals can be easily obtained. Historically, many devices have been obtained
thanks to the material transparency in the mid infrared also shown by a few cm thick
monocrystalline sheets.
The greatest success of this material is due to its use as an absorber material in
solar cells. What is most surprising is that solar cells totally made with thin film
technology are much more performing than the same devices built with bulk CdTe
single crystal. Recently, CdTe thin film solar cells achieved efficiencies of 22.1% by
exploiting a sophisticated engineering of the energy gap of the absorber simply add-
ing a little amount of Se at the interface with the n-type partner.
226 A. Bosio

This is a very mature technology since fully automated in-line production


machine could produce large-area, 19% efficiency module every 1 minute.
Moreover, the main actors on the market assure that at their end-of- life these
modules are completely recyclable implementing a virtuous circular loop.

References

1. Fabre M (1888) Tellurures métalliques cristallisés. Ann Chim Phys Ser 6:110–120
2. Frerichs R (1947) The photo-conductivity of “incomplete phosphors”. Phys Rev 72:594
3. Schwarz E (1948) New photoconductive cells. Nature 162:614
4. Jenny DA, Bube RH (1954) Semiconducting cadmium telluride. Phys Rev 96:1190
5. De Nobel D (1959) Phase equilibria and semiconducting properties of cadmium telluride.
Philips Res Repts 14:361-399 and 430-492
6. Lofersky JJ (1956) Theoretical considerations governing the choice of the optimum semicon-
ductor for photovoltaic solar energy conversion. J Appl Phys 27:777
7. Kato Y, Fujimoto S, Kozawa M, Fujiwara H (2019) Maximum efficiencies and performance-­
limiting factors of inorganic and hybrid perovskite solar cells. Phys Rev Appl 12:024039
8. Werthen JG, Fahrenbruch AL, Bube RH, Zesch JC (1983) Surface preparation effects on effi-
cient indium-tin-oxide-CdTe and CdS-CdTe heterojunction solar cells. J Appl Phys 54:2750
9. Aranovich JA, Golmayo D, Fahrenbruch AL, Bube RH (1980) Photovoltaic properties of ZnO/
CdTe heterojunctions prepared by spray pyrolysis. J Appl Phys 51:4260
10. Courreges F, Farhenbruch AL, Bube RH (1980) Surface preparation effects on efficient
indium-tin-oxide-CdTe and CdS-CdTe heterojunction solar cells. J Appl Phys 51:2175
11. Arroyo J, Marfaing Y, Cohen-Solal G, Triboulet R (1979) Electric and photovoltaic properties
of CdTe p-n homojunctions. Sol Energy Mater 1(1–2):171–180
12. Nakazawa T, Takamizawa K, Ito K (1987) High efficiency indium oxide/cadmium telluride
solar cells. Appl Phys Lett 50:279
13. Bonnet D, Rabenhorst H (1972) New results on the development of thin film p-CdTe/n-CdS
heterojunction solar cell. In: Proceedings of 9th IEEE Photovoltaic Specialists Conference.
Silver Spring, 2–4 May 1972, pp 129–132
14. Tyan YS, Albuerne EA (1982) Efficient thin-film CdS/CdTe solar cells. In: Proceedings of
16th IEEE Photovoltaic Specialist Conference. San Diego, September 27–30, 1982,
pp 794–800
15. Ferekides C, Britt J, Ma Y, Killian L (1993) High efficiency CdTe solar cells by close spaced
sublimation. In: Proceedings of 23rd IEEE photovoltaic specialists conference. Louisville,
10–14 May, 1993, pp 389–393
16. Wu X, Keane JC, Dhere RG, DeHart C, Albin DS, Duda A et al (2002) 16.5% Efficiency CdS/
CdTe polycristalline thin-film solar cells. In: Proceedings of the 17th European Photovoltaic
Solar Energy Conference and Exhibition. Munich, 22–26 October, 2002, pp 995–1000
17. NREL. [Online]. Available: https://www.nrel.gov/pv/cell-­efficiency.html. Accessed 29
Dec 2021
18. Green MA, Dunlop ED, Hohl-Ebinger J, Yoshita M, Kopidakis N, Hao X (2020) Solar cell
efficiency tables (version 56). Prog Photovolt Res Appl 28:629–638
19. Pollok DB (1969) Thermal expansion values for installation of an lrtran-6 window. Appl Opt
8(4):837–838
20. Taylor JB, Boland R, Gowac E, Stupik P, Tricard M (2007) Recent advances in high-­
performance window fabrication, “Defense and Security Symposium”. In: Proceedings of
window and dome technologies and materials XIII, 1–2 May 2013. Baltimore, pp 8708–8740
21. EdmundOptics. [Online]. Available: https://www.edmundoptics.com/f/cadmium-­telluride-­
cdte-­windows/39727/. Accessed 02 Jan 2022
9 CdTe-Based Photodetectors and Solar Cells 227

22. Long X-C, Myers RA, Brueck SRJ (1994) Measurement of the linear electro-optic coefficient
in poled amorphous silica. Opt Lett 19(22):1819–1821
23. Sugie M, Tada K (1976) Measurements of the linear Electrooptic coefficients and analysis of
the nonlinear susceptibilities in cubic GaAs and hexagonal CdS. Jpn J Appl Phys 15(3):421
24. Chenault DB, Chipman RA, Lu S-Y (1994) Electro-optic coefficient spectrum of cadmium
telluride. Appl Opt 33(31):7382–7389
25. Huang H, Lin ZS, Chen CT (2008) Mechanism of the linear electro-optic effect in potassium
dihydrogen phosphate crystals. J Appl Phys 104:073116
26. Sliker TR, Burlage SR (1963) Some dielectric and optical properties of KD2PO4. J Appl
Phys 34:1837
27. Onaka R, Ito H (1976) Pockels effect of KDP and ADP in the ultraviolet region. J Phys Soc
Jpn 41:1303–1309
28. Haibeiflavor, 2022. [Online]. Available: https://www.haibeiflavor.com/material-­factory-­
optical-­l inbo3-­l n-­s aw-­filter-­q -­s witch-­e lectro-­o ptic-­d immer-­s ingle-­c rystal-­s ubstrate-­
manufacturer-­from-­china_p242.html. Accessed 26 Jan 2022
29. Kurt MZ (2003) Electrooptical properties of LiTaO3. Ferroelectrics 296:127–137
30. Lovett DR (2017) Tensor properties of crystals. Taylor & Francis Group, New York
31. Akitt DP, Johnson C, Coleman PD (1970) Nonlinear susceptibility of CdTe. IEEE J Quantun
Electron QE-6(8):196–499
32. Wagnière GH, Woźniak S (2017) Nonlinear optical properties: noncentrosymmetric crystals.
In: Lindon JC, Tranter GE, Koppenaal DW (eds) Encyclopedia of spectroscopy and spectrom-
etry, London. Academic, pp 375–387
33. Olson J, Rodriguez Y, Yang L, Alers G, Carter S (2010) CdTe Schottky diodes from colloidal
nanocrystals. Appl Phys Lett 96:242103
34. Baines T, Papageorgiou G, Hutter O, Bowen L, Durose K, Major J (2018) Self-catalyzed CdTe
wires. Nano 8:274
35. Wang X, Wang J, Zhou M, Wang H, Xiao X, Li Q (2010) CdTe nanorods formation via
nanoparticle self-assembly by thermal chemistry method. J Cryst Growth 312:2310–2314
36. Wang X, Xu Y, Zhu H, Liu R, Wang H, Li Q (2011) Crystalline Te nanotube and Te nanorods-­
on-­CdTe nanotube arrays on ITO via a ZnO nanorod templating-reaction. CrystEngComm
13:2955–2959
37. Lee S, Yu Y, Perez O, Puscas S, Kosel T, Kuno M (2010) Bismuth-assisted CdSe and CdTe
nanowire growth on plastics. Chem Mater 22:77–84
38. Kret S, Szuszkiewicz W, Dynowska E, Domagala J, Aleszkiewicz M, Baczewski L, Petroutchik
A (2008) MBE growth and properties of ZnTe- and CdTe-based nanowires. J Korean Phys Soc
53:3055–3063
39. Kulkarni R, Rondiya S, Pawbake A, Waykar R, Jadhavar A, Jadkar V, Bhorde A, Date A,
Pathan H, Jadkar S (2017) Structural and optical properties of CdTe thin films deposited using
RF magnetron sputtering. Energy Procedia 110:188–195
40. Yang G, Jung Y, Chun S, Kim D, Kim J (2013) Catalytic growth of CdTe nanowires by closed
space sublimation method. Thin Solid Films 546:375–378
41. Salim H, Patel V, Abbas A, Walls J, Dharmadasa I (2015) Electrodeposition of CdTe thin films
using nitrate precursor for applications in solar cells. J Mater Sci Mater Electron 26:3119–3128
42. Consonni V, Rey G, Bonaim J, Karst N, Doisneau B, Roussel H, Renet S, Bellet D (2011)
Synthesis and physical properties of ZnO/CdTe core shell nanowires grown by low-cost depo-
sition methods. Appl Phys Lett 98:96–99
43. Dan Y, Zhao X, Chen K, Mesli A (2018) A photoconductor intrinsically has no gain. ACS
Photonics 5(10):4111–4116
44. Xie X, Kwok S-Y, Lu Z, Liu Y, Cao Y, Luo L, Zapien JA, Bello I, Lee C-S, Lee S-T, Zhang
W (2012) Visible–NIR photodetectors based on CdTe nanoribbons. Nanoscale 4:2914–2920
45. Shaygan M, Davami K, Kheirabi N, Baek CK, Cuniberti G, Meyyappand M, Lee J-S (2014)
Single-crystalline CdTe nanowire field effect transistors as nanowire-based photodetector.
PhysChemChemPhys 16:22687
228 A. Bosio

46. Wei H, Fang Y, Yuan Y, Shen L, Huang J (2015) Trap engineering of CdTe nanoparticle for
high gain, fast response, and low noise P3HT:CdTe nanocomposite photodetectors. Adv Mater
27:4975–4981
47. Park H, Yang G, Chun S, Kim D, Kim J (2013) CdTe microwire-based ultraviolet photodetec-
tors aligned by a non-uniform electric field. Appl Phys Lett 103:051906
48. Tang M, Xu P, Wen Z, Chen X, Pang C, Xu X, Meng C, Liu X, Tian H, Raghavan N, Yang
Q (2018) Fast response CdS-CdSxTe1-xCdTe core-shell nanobelt photodetector. Sci Bull
63:1118–1124
49. Lofersky JJ (1956) Theoretical considerations governing the choice of the optimum semicon-
ductor for photovoltaic solar energy conversion. J Appl Phys 27(7):777–784
50. Britt J, Ferekides C (1993) Thin-film CdS/CdTe solar cell with 15.8% efficiency. Appl Phys
Lett 62:2851–2852
51. Green M, Hishikawa I, Dunlop E, Levi D, Hohl-Ebinger J, Yoshit M, Ho-Baillie A (2019)
Solar cell efficiency tables (version 54). Prog Photovolt Res 27:565–575
52. “First Solar Establishes New World Record for CdTe Efficiency,” 23 February 2016. [Online].
Available: https://www.solarpowerworldonline.com/2016/02/24939/. Accessed 13 Mar 2022
53. Cunningham D, Rubcich M, Skinner D (2002) Cadmium telluride PV module manufacturing
at BP Solar. Prog Photovolt Res Appl 10:159–168
54. Bosio A, Menossi D, Mazzamuto S, Romeo N (2011) Manufacturing of CdTe thin film photo-
voltaic modules. Thin Solid Films 519:7522–7525
55. FirstSolar.com, “FirstSolar.com, Series 6 Datasheet,” October 2021. [Online]. Available:
http://www.firstsolar.com/-­/media/First-­Solar/Technical-­Documents/Series-­6-­Datasheets/
Series-­6-­Datasheet.ashx. Accessed 13 March 2022
56. Bosio A, Rosa G (2018) Past present and future of the thin film CdTe/CdS solar cells. Sol
Energy 175:31–43
57. Salavei A, Artegiani E, Piccinelli F, di Mare S, Menossi D, Bosio A, Romeo N, Romeo A
(2015) Flexible CdTe solar cells on polyimide and flexible glass substrates. In: Proceedings of
the 31st European Photovoltaic Solar Energy Conference. Hamburg, 14–18 September, 2015,
pp 1356–1357
58. Moss T (1954) The interpretation of the properties of indium antimonide. Proc Phys Soc
B 67:775
59. Wu X, Ribelin R, Dhere RG, Albin DS, Gessert TA, Asher S, Levi DH, Mason A, Moutinho
HR, Sheldon P (2000) High-efficiency Cd2SnO4/Zn2SnO4/ZnxCd1-xS/CdS/CdTe polycrys-
talline thin-film solar cells. In: Record of the Twenty-Eighth IEEE Photovoltaic Specialists
Conference. Anchorage, 15–22 Sept. 2000, pp 470–474
60. Romeo N, Bosio A, Canevari V (2003) The role of CdS preparation method in the performance
of CdTe/CdS thin film solar cell. In: Proceedings of the 3rd world conference on photovoltaic
energy conversion, WCPEC-3. Osaka, 11–18 May, 2003, pp 469–470
61. Bosio A, Romeo A, Romeo N (2011) Polycrystalline CdTe thin films solar cells. In: Thin
film solar cells: current status and future trends. Nova Science Publishers Inc., New York,
pp 161–200
62. Wei S, Zhang S, Zunger A (2000) First-principles calculation of band offsets, optical bowings,
and defects in CdS, CdSe, CdTe, and their alloys. J Appl Phys 87:1304–1311
63. Mia M, Swartz C, Paul S, Sohal S, Grice C, Yan Y, Holtz M, Li J (2018) Electrical and optical
characterization of CdTe solar cells with CdS and CdSe buffers—a comparative study. J Vac
Sci Technol B 36:052904
64. Kephart J, McCamy J, Ma Z, Ganjoo A, Alamgir F, Sampath W (2016) Band alignment of front
contact layers for high-efficiency CdTe solar cells. Sol Energy Mater Sol Cells 157:266–275
65. Munshi A, Kephart J, Abbas A, Shimpi T, Barth K, Walls J, Sampath W (2018) Polycrystalline
CdTe photovoltaics with efficiency over 18% through improved absorber passivation and cur-
rent collection. Sol Energy Mater Sol Cells 176:9–18
9 CdTe-Based Photodetectors and Solar Cells 229

66. Fiducia T, Mendis B, Li K, Grovenor C, Munshi A, Barth K, Sampath W, Wright L, Abbas A,


Bowers J (2019) Understanding the role of selenium in defect passivation for highly efficient
selenium-alloyed cadmium telluride solar cells. Nat Energy 4:504–511
67. Romeo A, Bätzner D, Zogg H, Vignali C, Tiwari A (2001) Influence of CdS growth process
on structural and photovoltaic properties of CdTe/CdS solar cells. Sol Energy Mater Sol Cells
67:311–321
68. Moutinho H, Al-Jassim M, Abulfotuh F, Levi D, Dippo P, Dhere R, Kazmerski L (1997)
Studies of recrystallization of CdTe thin films after CdCl2 treatment. In: Proceedings of the
twenty sixth IEEE photovoltaic specialists conference. Anaheim, September 29–October 3,
1997, pp 431–434
69. Danaher W, Lyons L (1978) Photoelectrochemical cell with cadmium telluride film.
Nature 271:139
70. Kröger F (1978) Cathodic deposition and characterization of metallic or semiconducting
binary alloys or compounds. J Electrochem Soc 125:2028–2034
71. Ortega-Borges R, Lincot D (1993) Mechanism of chemical bath deposition of cadmium
sulfide thin films in the ammonia-thiourea system in situ kinetic study and modelization. J
Electrochem Soc 140:3464–3473
72. Fulop G, Taylor R (1985) Electrodeposition of semiconductors. Ann Rev Mater Sci 15:197–210
73. Basol B (1984) High-efficiency electroplated heterojunction solar cell. J Appl Phys 55:601–603
74. Ojo A, Dharmadasa I (2016) 15.3% efficient graded bandgap solar cells fabricated using elec-
troplated CdS and CdTe thin films. Sol Energy 136:10–14
75. Paudel N, Wieland K, Compaan A (2012) Ultrathin CdS/CdTe solar cells by sputtering. Sol
Energy Mater Sol Cells 105:109–112
76. Compaan A, Gupta A, Lee S, Wang S, Drayton J (2004) High efficiency, magnetron sputtered
CdS/CdTe solar cells. Sol Energy 77:815–822
77. Zanio K (1978) Cadmium telluride, semiconductors and semimetals. Academic, New York
78. Ferekides CS, Marinskiy D, Viswanathan V, Tetali B, Palekis V, Selvaraj P, Morel D (2000)
High efficiency CSS CdTe solar cells. Thin Solid Films 361-362:520–526
79. Bayhan H (2004) Investigation of the effect of CdCl2 processing on vacuum deposited CdS/
CdTe thin film solar cells by DLTS. J Phys Chem Solids 65:1817–1822
80. Hiie J (2003) CdTe:CdCl2:O2 annealing process. Thin Solid Films 431-432:90–93
81. Niles D, Waters D, Rose D (1998) Chemical reactivity of CdCl2 wet-deposited on CdTe films
studied by X-ray photoelectron spectroscopy. Appl Surf Sci 136:221–229
82. Williams L, Major J, Bowen L, Keuning W, Creatore M, Durose K (2015) A comparative study
of the effects of nontoxic chloride treatments on CdTe solar cell microstructure and stoichiom-
etry. Adv Energy Mater 5:1500554–1500563
83. Potlog T, Ghimpu L, Gashin P, Pudov A, Nagle T, Sites J (2003) Influence of annealing in
different chlorides on the photovoltaic parameters of CdS/CdTe solar cells. Sol Energy Mater
Sol Cells 80:327–334
84. Potter M, Halliday D, Cousins M, Durose K (2000) A study of the effects of varying cadmium
chloride treatment on the luminescent properties of CdTe/CdS thin solar cells. Thin Solid
Films 361-362:248–252
85. Zhou T, Reiter N, Powell R, Sasala R, Meyers P (1994) Vapor chloride treatment of polycrys-
talline CdTe/CdS films. In: Proceedings of the 1st IEEE photovoltaic specialists conference.
Waikoloa, pp 103–106. https://doi.org/10.1109/WCPEC.1994.519818
86. Qu Y, Meyers P, Mc Candless B (1996) HCl vapor post-deposition heat treatment of CdTe/CdS
films. In: Proceedings of the 25th IEEE photovoltaic specialists conference, Washington, DC,
pp 1013–1016. https://doi.org/10.1109/PVSC.1996.564303
87. Albright S, Jordan J, Akerman B, Chamberlain R (1989) Developments on CdS/CdTe photo-
voltaic panels at photon energy. Inc Sol Cells 27:77
88. Gessert T, Mason A, Sheldon P, Swartzlander A, Niles D, Coutts T (1996) Development of
Cu-doped ZnTe as back-contact interface layer for thin-film CdS/CdTe solar cells. J Vac Sci
Technol 14:806
230 A. Bosio

89. Uda H, Ikegami S, Sonomura H (1990) Compositional change of the Au–Cu2Te contact for
thin-film CdS/CdTe solar cells. Jpn J Appl Phys 29:495
90. McCandless B, Qu Y, Birkmire RA (1994) Treatment to allow contacting CdTe with different
conductors. In: Proceedings of the first world conference on photovoltaic energy conversion.
Waikoloa, pp 107–110. https://doi.org/10.1109/WCPEC.1994.519819
91. Du M (2009) First-principles study of back-contact effects on CdTe thin-film solar cells. Phys
Rev B 80:205322
92. Durose K, Boyle D, Abken A, Ottley C, Nollet P, Degrave S, Burghelman M, Wendt R, Bonnet
D (2002) Key aspects of CdTe/CdS solar cells. Phys Status Solidi 229:1055–1064
93. Bosio A, Ciprian R, Lamperti A, Rago I, Ressel B, Rosa G, Stupard M, Weschke E (2018)
Interface phenomena between CdTe and ZnTe:Cu back contact. Sol Energy 176:186–193
Chapter 10
CdSe – Based Photodetectors
for Visible-­NIR Spectral Region

Hemant Kumar and Satyabrata Jit

10.1 Introduction

The discovery of the photoelectric effect by Albert Einstein in 1905 was the first
stepping stone toward the creation of the new and emerging area of research in
optoelectronics and photonics [1]. The photodetectors belong to a class of optoelec-
tronic devices which are used to convert optical signals of certain wavelengths into
an electrical signal for various applications including optical communications and
computing, optoelectronics integrated circuits and optical imaging systems.
Photodetectors are designed to detect lights of wavelengths belonging to one or
more regions of the electromagnetic spectrum namely the ultraviolet (UV) region
(i.e. wavelengths below 400 nm), visible region (i.e. wavelengths in between 400
and 700 nm) and infrared (IR) region (i.e. wavelengths above 700 nm but below
1 mm) [2].
The photodetection mechanism primarily depends upon the accurate detection of
the wavelength (or energy) of incident photons on the photodetectors. Based on the
spectral width of the incident photons to be detected by the photodetectors, the pho-
todetectors are broadly divided into two types: narrowband and wideband photode-
tectors. When the photodetectors are designed to detect the incident photons of a
single wavelength, they are known as narrowband photodetectors. The photo
response characteristics of such type of photodetectors possess a dominant peak

H. Kumar
Department of Electronics and Communication Engineering, Jaypee Institute of Information
Technology, Noida, India
e-mail: hemant.kumar@jiit.ac.in
S. Jit (*)
Department of Electronics Engineering, Indian Institute of Technology (Banaras Hindu
University), Varanasi, India
e-mail: sjit.ece@iitbhu.ac.in

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 231
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_10
232 H. Kumar and S. Jit

centered around the wavelength of the incident photons to be detected with the full-­
width at half-maximum (FWHM) <= 100 nm [3]. On the other hand, the wideband
photodetectors are designed to detect the incident photons of different wavelengths,
may be ranging from UV to IR. Such photodetectors have a larger FWFH value
>100 nm. The wideband photodetectors thus can’t discriminate the incident photons
on the basis of their wavelength or energy. The wideband or narrowband nature of
the photodetectors primarily depends on the properties of the photoactive materials
used for the light detection in the devices. Traditionally, Si is used for wideband
photodetectors for detecting all the photons with wavelengths ranging from 400 to
1100 nm [4]. Since it is not possible to distinguish two incident photons of different
energies or wavelengths by such detectors, external optical filters are connected
with such devices for the same [5].
It is well known that different materials have different types of absorption spec-
tral characteristics. Thus, the absorption properties of photoactive materials used in
the photodetectors play vital roles in determining the spectral characteristics of the
detectors. This chapter is devoted to discuss some Cadmium Selenide (CdSe)
material-­based photodetectors for their operation in the visible – NIR region [6].
CdSe belongs to the II-VI group semiconductors which carries exciting properties
for optoelectronic applications [6]. CdSe exists in three crystalline forms: wurtzite,
zinc blend, and rock-salt [7]. However, the wurtzite structure of CdSe is widely
explored for optoelectronic applications due to its higher stability than the other
structures [6]. CdSe is inherently an n-type material. The p-type CdSe is not stable
and very difficult to achieve due to the self-compensation effect [6]. The absorption
band edge of the bulk CdSe lies at ~1.7 eV (~729 nm) for both the wurtzite and zinc
blend structures at the room temperature [7]. The band gap of the bulk CdSe is suit-
able for photodetectors operating in the visible and near infrared (NIR) regions of
the electromagnetic spectrum [8, 9]. It may be mentioned that photodetectors work-
ing in the visible region find multiple areas of applications such as in artificial eye
[3], imaging [3], machine vision [3], forensic analysis [10], atmospheric contamina-
tion [10], criminology [10], biomedical and diagnostic imaging [3], civil investiga-
tions [10] and detecting heart rate using smartwatches [3]. Similarly, the
photodetectors operating in the NIR region are widely used for medical diagnosis
and biomedical imaging without tampering with the biological tissues [11].
Various CdSe nanostructures (2D, 1D, 0D) are projected as potential materials
for developing filter-less photodetectors [11]. Luo et al. [12] have reported CdSe
nano-ribbons (1D) based photodetectors with a photo response around 650 nm cor-
responding to the red region of the visible spectrum without using any external fil-
ter. They [12] have used hollow gold nanospheres (HGN) on CdSe nano-ribbons
(NRs) to enhance the absorption of CdSe nano-ribbons by multiple folds at the red
region by using localized surface plasmon resonance (LSPR) effect. The device
structure and TEM image of HGN-on-CdSe NRs surface are shown in Fig. 10.1a, b,
respectively. The I-V and photo response characteristics of the aforementioned
CdSe NRs based photodetector with and without HGN decorations are compared in
Fig. 10.1c, d. In a different work, Kumar et al. [13] have explored optical resonance
mechanism in the Pd/CdSe QDs (0D) based photodetector to demonstrate an
10 CdSe – Based Photodetectors for Visible-NIR Spectral Region 233

Fig. 10.1 (a) Device schematic, (b) TEM image of decorated HGN on CdSe NRs, (c) I-V response
of the fabricated detector with and without HGN, and (d) spectral response of the device with or
without HGNs. (Adapted from Ref. [12]. Published 2010 by Optica Publishing as open access)

Fig. 10.2 Photo response of CdSe QD based Schottky device with (a) Au electrode and (b) Pd
electrode. (Reprinted with permission from Ref. [13]. Copyright 2019: IEEE)

enhanced photo-response at ~400 nm corresponding to the blue region of the visible


spectrum. The photo
response of the fabricated Schottky diode is shown in Fig. 10.2, where Fig. 10.2a
shows the wide spectrum coverage (FWHM = 190 nm) of the device with Au elec-
trode and Fig. 10.2b presents the improved photo response over blue region due to
the resonance with Pd electrode thereby making detector a narrowband
(FWHM = 61 nm) one.
234 H. Kumar and S. Jit

In this chapter, various types of CdSe photodetectors are discussed in details.


Section 10.2 focuses on the discussion of carrier dynamics and their applications in
CdSe Photodetectors. Section 10.3 discusses different CdSe based photoconductors
and Schottky photodiodes. Section 10.4 presents some CdSe heterojunction-based
photodetectors while the Sect. 10.5 focuses on some hybrid photodetectors.
Application of CdSe photodetectors are discussed in Sect. 10.6. Conclusion and
future scopes are discussed in Sect. 10.7.

10.2 CdSe Carrier Dynamics

Charge carrier dynamics is the crucial part for determining the electrical character-
istics of any electronic device. In case of photodetectors, transportation of photo-
generated charge carriers defines the performance parameters such as the
photoconductivity, photoconductive gain, transient response, and responsivity of
the device. Charge carrier dynamics is generally evaluated by the lifetime measure-
ment of the carriers. Lifetime of the carriers plays a vital role in determining the
efficiency of the photodetectors. A photoexcited carrier can relax in the ground state
and allow electron-hole scattering or electron-electron scattering [14]. Other possi-
ble way is that the photoexcited carrier may escape from the original site and relax
on another site (trap site) or may be collected at the electrodes [14]. There should be
one effective carrier lifetime (τeff) to describe the time elapsed before photoexcited
carrier falls back into the ground state [14]. In photodetection operation, the photo-
excited carriers must be collected by the electrodes within τeff period after their
generation. Thus, the lifetime, τeff, of a carrier plays critical role for determining the
responsivity and gain of the photodetectors. For the photoconductive detector one
can write the relation of photocurrent (Iph) in terms of incident flux density (ϕ),
detector’s quantum efficiency (η), and photconductive gain (g) as Iph = eϕηg [14]
where g is the ratio of τeff to carrier transit time (τtr) or the ratio of the total collected
carrier to total excited carriers, for both the thermally generated and photogenerated
carriers [14]. Thus, one can easily notice that the improvement in τeff can improve
the photoconductive gain and hence the photocurrent and the responsivity of the
photoconductors or photodetectors. McGott et al. [18] estimated the carrier lifetime
of CdSexTe1 − x film on a MgyZn1 − yO buffer layer using time resolved photolumines-
cence. They observed improvement in the carrier lifetime only after x = 0.2 mole
fraction of Se [18]. They observed a response with a fast initial decay and a long-­
lived tail indicating the trap-dependent recombination in the film [18]. They further
observed that the long-lived tail was absent for the samples prepared on Al2O3
instead of MgyZn1 − yO due to lower density of surface trap states and better surface
passivation [18]. Vietmeyer et al. [19] performed time-resolved lifetime (TCPSC)
and transient differential absorption (TDA) analyses of CdSe nanowires to assess
the carrier recombination process at 405 nm with 10 MHz repetition rate. They [19]
observed that ~98% of the total signal constituted the short decay component
(~100 ps). There were also other two components of ~450 ps and ~2.5 ns
10 CdSe – Based Photodetectors for Visible-NIR Spectral Region 235

constituting ~1.5% and ~0.5% of the signal, respectively. Mismatch was observed
between the TCPSC (~100 ps) and TDA (~1.5 ns) kinetics due to the fast hole trap-
ping and long-lived electrons in the CdSe nanowire conduction band [19]. The mea-
surement results showed improvements in the photoconductive gain and
photoresponsivity with the increase in the diameter of the CdSe nanowires [19].
Kung et al. [20] investigated the performance of CdSe nanowires based photocon-
ductors with varying diameters. They observed the improvement in the photocon-
ductive gain from g = 0.017 to g = 4.9 corresponding to the increase in the diameter
of the nanowires from 10 nm to 100 nm. The improvement was also observed in the
responsivity with the increase in the diameter of the CdSe nanowires. The results
were consistent with the observations of Vietmeyer et al. [19]. However, they noted
that the improvements in gain and responsivity were at the expense of deteriorating
transient response of the device from 8 μs (at 10 nm diameter) to 8 s (at 100 nm
diameter) [20].
Effect of doping on the carrier lifetime of CdSe QDs were investigated by Straus
et al. [15]. They fabricated a FET by using a CdSe material-based channel deposited
on an Indium film. The combination of CdSe QDs/Indium thin films was annealed
at 300 °C to allow diffusion of Indium into the CdSe QDs. A schematic of the
indium diffusion into the CdSe QDs film with annealing temperature is shown in
Fig. 10.3a. The mobility and lifetime of the carriers were increased with the higher
thickness of indium film due to enhanced diffusion of Indium into the CdSe QDs as
indicated in Fig. 10.3b. Kongkanand et al. [16] assembled CdSe QDs over TiO2
nanoparticles and nanaotubes for photovoltaic applications. The carrier lifetimes of
0.4 ns and 1.3 ns were observed in CdSe/TiO2 nanoparticle and CdSe/TiO2 nano-
tubes, respectively, which were smaller than 4.1 ns carrier lifetime of CdSe QDs
[16]. The decrease in lifetime was attributed to the charge transfer from CdSe QDs

Fig. 10.3 (a) Schematic diagram indicating indium doping via annealing procedure and (b)
Lifetime variation with varying thickness of Indium against quantum yield mobility product.
(Reprinted with permission from Ref. [15]. Copyright 2015: American Chemical Society)
236 H. Kumar and S. Jit

Fig. 10.4 Emission decay of CdSe quantum dots (a) for size 2.6 nm, and (b) for size 3.7 nm. The
experiment also indicates the results with TiO2 nanoparticles and TiO2 nanotubes. (Reprinted with
permission from Ref. [16]. Copyright 2008: American Chemical Society)

Fig. 10.5 (a) Schematic of fabricated CdSe QDs based photodetector with varying ZnO QDs size
and (b) EQE of the detector showing effect of increased temperature on ZnO QDs acting as a
charge transport layer. (Adapted with permission from Ref. [17]. Copyright 2017: IEEE)

to TiO2 which implied that a charge transport layer was required for CdSe QDs
based devices to efficiently collect the carriers before recombination [16].
The detailed emission decays of CdSe QDs of 2.6 nm and 3.7 nm deposited on
the glass substrate, TiO2 nanoparticles, and TiO2 nanotubes are shown in Fig. 10.4a,
b. Kumar et al. [17] fabricated a CdSe colloidal QDs (CQDs) based photodetector
using ZnO CQDs as an electron transport layer (ETL) as shown in Fig. 10.5a.
Evaluation of the photodetector characteristics was carried out for varying particle
sizes of ZnO CQDs obtained by varying the annealing temperature. The size of ZnO
QDs in the deposited thin film was observed to be increased with annealing tem-
perature. ZnO QDs are reported to be 4.42 nm, 8.74 nm, and 13.89 nm were esti-
mated at respective annealing temperatures of 250 °C, 350 °C, and 450 °C. They
[17] observed that the transient response, contrast ratio, responsivity, detectivity,
and external quantum efficiency (EQE) of the detector were deteriorated with the
increased size of the ZnO QDs at higher annealing temperatures. The EQE
10 CdSe – Based Photodetectors for Visible-NIR Spectral Region 237

characteristics of the photodetector with ZnO QDs ETL have been shown in
Fig. 10.5b. Osedach et al. [21] used thermally evaporated spiro-TPD for the charge
transport layer sensitized by CdSe QDs over a 10 μm finger width (or channel
length). The photo response was shown to be very much dominated over the blue
region of the spectrum, and the EQE followed a power square law function of the
bias voltage due to the space charge limitation in spiro-TPD. The EQE peak
observed at 400 nm confirmed the efficient dissociation of excitons at the interface
of spiro-TPD and CdSe QDs.

10.3 Photoconductors and Schottky Photodiodes

Photoconductors are the simplest form of light-detecting devices which work on the
principle of conductivity modulation by incident photons to be detected. The photo-
conductor contains a photoactive material with two electrodes placed on it to con-
nect the device with an external circuit. When light is incident on the active material,
its conductivity is increased (hence the resistivity is decreased) due to excess
electron-­hole pairs generated by absorption of photons in the photoactive material.
When a potential is applied across the two electrodes of the device, the decreased
resistivity increases the current flowing through the external circuit. On the other
hand, the Schottky photodiodes are fabricated by forming rectifying contact on a
semiconductor. The work function the metal must be larger (smaller) than that of an
n-type (p-type) semiconductor for the Schottky contact formation. Since CdSe is
inherently an n-type semiconductor, one needs to select a metal with work function
larger than that of the CdSe for forming the Schottky junction based CdSe photodi-
odes [25].
The photoconductivity of the CdSe largely depends on the trap states which may
lead to increased transient time. Skarman [26] carried out a detailed investigation on
the photoconductivity of CdSe material with the trap states under a varying tem-
perature. The effect of traps was found to be more dominating in single crystalline
CdSe thin films than the polycrystalline films. However, the trap states were shown
to be reduced by using controlled fabrication and activation techniques [26].
Margulis and Sibbett [27] fabricated an ultrafast CdSe based photoconductor using
two Al electrodes placed with a separation distance of 40–100 μm and reported a
transient response of 20 ps. Jiang et al. [22] fabricated a single crystalline CdSe
nano-ribbons based photodetector using Ti/Au electrodes and observed photo
response over 400–710 nm as shown in Fig. 10.6. Presence of various traps at dif-
ferent energy levels (shallow and deep in the bandgap) was observed. The device
showed a transient response of <1 ms. A SiO2 coating over the CdSe nano-ribbons
was used to reduce the trap states and improve the transient response of the device
at the expense of reduced sensitivity [22]. Researchers have also used variational
technique for the efficient charge separation in CdSe. Xu et al. [23] used site-­
controlled growth of aligned CdS/CdSe core-shell nanowalls based photodetectors
at wafer scale without post-growth transfer. The SEM image of the interdigitated
238 H. Kumar and S. Jit

Fig. 10.6 (a) SEM image of CdSe NRs, (b) Sensitivity curve of the photodetector against wave-
length, and (c) I-V characteristics of the device and structure of the device. (Adapted with permis-
sion from Ref. [22]. Copyright 2007: John Wiley and Sons)

Fig. 10.7 (a) SEM image depicting the structure of the photodetector and (b) Transient response
of the detector under 405 nm light at 10 V bias. (Adapted with permission from Ref. [23]. Copyright
2018: John Wiley and Sons)
10 CdSe – Based Photodetectors for Visible-NIR Spectral Region 239

detector structure is shown in Fig. 10.7a. The fabricated photoconductor showed a


high gain of 3800 and photoresponsivity of 1200 A/W with an ultrafast transient
response of 200 ns as shown in Fig. 10.7b [23]. The superior photo response and
gain of the detector were attributed to enhanced charge-separation efficiency of
core-shell nanowall geometry [23].
Some researchers have achieved spectral shift in CdSe photodetectors with effi-
cient charge carrier separation techniques. Das et al. [28] fabricated a CuO/CdSe
core-shell nanowire based photoconductor using Ag electrodes. Note that the band-
gaps of CuO and CdSe are 1.2 eV and 1.7 eV respectively. However, the bandgap of
CuO/CdSe nanowire structures was shown to be 3.96 eV which is much higher than
the bandgaps of both the CdSe and CuO. The high bandgap of CuO/CdSe nanowire
structure was explored for fabricating photodetector to work in deep UV region
with a responsivity of 0.63 A/W at 254 nm wavelength [28]. The high bandgap was
attributed to improved separation of photogenerated electron-hole pairs due to the
Type-II band alignment of CuO/CdSe heterojunction formation.
It is well known that CdSe is inherently an n-type material. Wu et al. [24] studied
the effect of impurity on the photoconductivity of intrinsic and n-type CdSe nano-
belts (NB). They observed that the intrinsic CdSe NBs based device possessed a
high gain while the n-type CdSe NBs based device showed a fast response. The
above behavioral studies of the intrinsic and n-type CdSe NBs were carried out in
both the Schottky Device (CdSe/Ni/Au) and Photoconductive device (CdSe/In/Au)
with a channel length of 2.5 μm. The measured I-V characteristics of the devices
are shown in Fig. 10.8a, while their transient responses are shown in Fig. 10.8b [24].
Ani et al. [29], introduced Cu impurity in CdSe thin films by vacuum annealing at
350 ° C. They observed an improved photoconductive gain due to the sensitizing
effect of Cu centers in the CdSe lattice.

Fig. 10.8 (a) I-V relation of interdigitated Ohmic Device (IOD) and interdigitated Schottky
device (ISD) under dark and under illumination of 633 nm light, inset indicates the FESEM image
of the device, and (b) Transient response of both IOD and ISD device. (Adapted with permission
from Ref. [24]. Copyright 2011: American Chemical Society)
240 H. Kumar and S. Jit

The environmental conditions affect the photoconductivity of the CdSe based


devices. Samanta et al. [8] studied the stabilization of CdSe thin films under air
oxidation and observed stable photocurrent and dark current over a period of
12–25 days in room temperature depending on the thickness of the films. However,
the CdSe thin films showed temperature-dependent quenching of photoconductivity
at 265 °K [8]. When the graphite based Ohmic contact was formed on the films,
photosensitivity of the device was increased with air exposure due to the reduction
of free electrons owing to the oxygen diffusion along the grain boundaries of the
CdSe in the films. The fabricated CdSe thin film based device also exhibited time
varying photosensitivity due to slow recombination states and oxygen assisted con-
version of selenium and cadmium vacancies [8]. An et al. [30] used one-step ther-
mal evaporation method to synthesize hollow tubular CdSe nanotubes for fabricating
Ag electrode based CdSe/Ag Schottky junction photodiodes. Their device showed a
high photoresponsivity of 76 A/W with a high photoconductive gain of 190 at low
operating voltage of −1 V. The performance of the device was observed to be
strongly influenced by the presence of oxygen molecules in the atmosphere. Further,
the dark current of the device was noted to be reduced with the increased content of
oxygen molecules [30].
The photoconductive gain of the CdSe photodetector also depends of the metal
electrodes used in the devices. Mehta et al. [25] achieved greater than unity photo-
conductive gain in Au/CdSe thin film Schottky photodiodes. It was attributed to the
trapping of photogenerated holes in the Schottky junction which resulted in an
injection of electrons in CdSe thin film by reducing the Schottky barrier.
Photoconductive gain of 300 was measured in CdSe thin film-based Schottky device
under an illumination of 800 nm wavelength.

10.4 Heterojunction Based Photodetectors

Heterojunction photodiodes are the photodetectors formed by the junctions of two


different semiconductors. They are widely investigated photodetectors where the
performance parameters can be optimized by modifying the heterojunction inter-
faces of the devices. The generalized structure of the heterojunction photodetectors
is metal-electrode/semiconductor-1/semiconductor-2/metal-electrode where the
semiconductor-1 and semiconductor-2 represent two different types of semiconduc-
tors of different band gap energies and metal electrodes form ohmic contacts with
the semiconductors.
Yuan et al. [31] fabricated a CdSe nanoplates/2D MoS2 heterojunction photodi-
ode for operating in the visible region. The time resolution photoluminescence
spectroscopies showed efficient charge transfer across the heterointerface.
Suppression of photons within CdSe of the CdSe/MoS2 heterojunction is demon-
strated by their Streak camera images at 710 nm shown in Fig. 10.9a, b. The lifetime
measurements of CdSe and CdSe/MoS2 heterojunctions are shown in Fig. 10.9c.
The lifetime of carriers in CdSe/MoS2 is only 118.9 ps which is much smaller than
10 CdSe – Based Photodetectors for Visible-NIR Spectral Region 241

Fig. 10.9 Streak camera images recorded at 710 nm (a) CdSe only, (b) CdSe/MoS2 heterostruc-
ture, and (c) Indicating the lifetime observed in 1 dimension indicating efficient charge separation
in CdSe when in contact with MoS2. (Adapted with permission from Ref. [31]. Copyright 2020:
Elsevier)

Fig. 10.10 (a) Schematic of the fabricated device consisting CdS, CdSe, and PbS NPs with WSe2
crystalline flakes, and (b) Photocurrent against wavelength indicating the spectrum selectivity and
tunability with CdSe/WSe2, CdS/WSe2, and PbS/WSe2. (Reprinted from Ref. [32]. Published 2022
by American Chemical Society as open access)

the lifetime of 1602 ps of the carriers of CdSe material. This shows that efficient
charge separation in CdSe may take place when it is in contact with the MoS2 [31].
Ren et al. [33] fabricated CH3NH3PbI3 nanowire arrays/single CdSe NB based het-
erojunction photodiode to operate in 400–800 nm with contrast ratio of 2.152 × 103,
responsivity of 69.11 A/W, EQE of 11,000%, and detectivity of 8.6 × 1012 Jones.
The device showed a rise time (τr) of 0.81 ms and fall time (τf) 0.77 ms. In order to
improve the response, Medda et al. [34] prepared a heterojunction between 2D
CdSe nanoplatelets (NPL) and Au25 nanoclusters (NCs) for photodetection applica-
tion. The DFT based analysis of the CdSe NPL and Au25 NCs heterostructure
showed the replacement of the conduction band of CdSe by Au. The device showed
a detectivity of 2.5 × 1011 Jones and transient response of 200 ms. Luo et al. [35]
explored a type-II heterojunction structure of CdSe/CdTe core-shell configuration
to demonstrate a rise-time of 0.7 s (τr) and fall-time of 0.5 s (τf). The device showed
a detection range extending from UV to NIR region. The optical properties of the
device were improved by applying a compressive load over the device.
CdSe based heterostructures are employed to tune the spectrum coverage of the
photodetectors. Ghods et al. [32] used the SILAR method to synthesize CdS, CdSe,
242 H. Kumar and S. Jit

and PbS nanoparticles directly over WSe2 crystalline flakes to fabricate the hetero-
junction photodetectors shown in Fig. 10.10a. The fabricated CdS/WSe2, CdSe/WSe2,
PbS/WSe2 heterostructures exhibited spectrum selective characteristics as shown in
Fig. 10.10b. The response peaks of CdS/WSe2, CdSe/WSe2 and PbS/WSe2 hetero-
junction photodetectors were observed at 510 nm, 700 nm, and 810 nm, respec-
tively. The reported device exhibited a quantum efficiency of 71% with a rise-­time
of 2.5 ms (τr) and fall-time of 3.5 ms (τf).
CdSe was introduced with other photoactive materials to improve the photo
response of the photodetectors. Li et al. [36] fabricated a photodetector using PbI2
nanosheet with CdSe NB (Nanobelt) to improve the detection capability of PbI2
based photodetectors. The composite structure exhibited a broad spectral response
over 400–730 nm. It was found the PbI2 /CdSe NB interface improved the separa-
tion of the photogenerated e-h pairs, which in turn, improved the performance of the
device. Introduction of the CdSe NB with PbI2 improved the responsivity, detectiv-
ity, and EQE of the device at the expense of degraded transient response.
Villa-Angulo et al. [37] proposed CdSeTe based photodetectors for image sens-
ing applications under adverse conditions. Thermal evaporation technique was used
to fabricate a Glass/ITO/CdS/CdSe/CdSe0.4Te0.6/CdSe/Ag based quantum-well het-
erostructure for detecting of photons in the range of 500–1100 nm wavelengths. The
detector showed good responsivity and specific detectivity for wavelengths of
500–1100 nm with a power conversion efficiency (PCE) of 10.4%. Villa-Angulo
et al. [37] observed that CdSeTe based detectors could work satisfactorily in the
infra-red region under adverse weather conditions resulted from fog, rain, and
snow fall.

10.5 CdSe-Organic Hybrid Photodetectors

Organic semiconducting materials are extensively explored for fabricating large-­


area optoelectronic devices due to their flexible nature and cost-effective fabrication
methods. However, the low light absorption coefficient and poor carrier mobility of
the organic materials have restricted their use for optoelectronic device applications.
Some researchers have explored hybrid heterojunctions formed between an organic
semiconductor and an inorganic semiconductor to utilize the benefits of both the
materials. Wang et al. [38] fabricated a hybrid photodetector using P3HT polymer
and CdSe nanowires to explore low ionization potential of organic molecules of
P3HT and high electron affinity of CdSe semiconductor. The I-V curve of the
reported device is shown in Fig. 10.11a. They fabricated the photodetector on Si,
PET, and printing paper to demonstrate the possibility of achieving a CdSe based
flexible photodetector shown in the inset of Fig. 10.11a. The photodetector was
shown to work over a broad spectral region of 400–710 nm as shown in Fig. 10.11b.
The Si substrate-based detector showed a transient response of 0.1 s. The PET and
printing paper substrate-based detectors showed transient response of about 10 ms.
The latter two devices exhibited detection characteristics with 180° bending. They
10 CdSe – Based Photodetectors for Visible-NIR Spectral Region 243

Fig. 10.11 (a) I-V characteristics of hybrid photodetector over PET substrate with and without
illumination, inset shows the optical image of flexible detector, (b) Photo response of the flexible
detector against different wavelengths, and (c) I-t curve of the flexible device under a constant bias
of 3 V evaluating hybrid detector under different bending curvatures. (Adapted with permission
from Ref. [38]. Copyright 2013: John Wiley and Sons)

observed a stable performance under different bending curvatures as shown in


Fig. 10.11c.
Shih et al. [39] used CdSe colloidal quantum dots (CQDs) and rGO based het-
erojunction photodetector using low cost spin coating method. The fabricated detec-
tor was shown to have a very high ON/OFF current ratio of ~2195 for −1 V reverse
bias voltage. The device showed a broadband spectral characteristic ranging from
350 nm (UV) to 900 nm (NIR). A huge factor of improvement in the figure of merit
was achieved by decorating rGO with CdSe QDs. The EQE and detectivity of the
photodetector are shown in Fig. 10.12a, b. The CdSe CQD based device is shown in
the inset of the Fig. 10.12b.
Oertel et al. [40] fabricated the first hybrid structure photodetector using CdSe
QDs and PEDOT:PSS. They used the solution processed CdSe QDs to grow a
200 nm film on the PEDOT:PSS polymer composite. The charge transport proper-
ties and exciton dissociation efficiency of the CdSe QDs thin film were improved by
using n-butylamine. The result showed a blue shift in the band gap of the CdSe QDs
thereby extending its photo response towards the visible region of the spectrum with
244 H. Kumar and S. Jit

Fig. 10.12 (a) External Quantum efficiency of the detector with and without CdSe QDs and (b)
Detectivity with and Without CdSe QDs, also shows synthesized CdSe QDs, and Spin Coated
Large area rGO/CdSe QDs based photodetector in the inset. (Adapted from Ref. [39]. Published
2020 by IEEE as open access)

a cut-off at 600 nm. A contrast ratio of 100 of the hybrid heterojunction photodetec-
tor was observed at ~514 nm. Ramar et al. [41] fabricated a hybrid heterostructure
photodetector using CdSe QDs and P3HT-PC71BM. The device showed a spectral
response in 350–800 nm wavelengths. The P3HT:CdSe:PC71BM active layer based
device showed a detectivity of 10 times higher than the device with the heterojunc-
tion of P3HT: PC71BM and P3HT:CdSe. This shows that CdSe QDs can be mixed
with organic polymers to improve the spectral broadening, responsivity and detec-
tivity of the hybrid photodetectors. Malik et al. [42] improved the photo response of
a MEH-PPV and VOPcPhO heterojunction photodetector by mixing CdSe QDs in
the materials. Authors observed that the introduction of CdSe QDs not only
improved the absorption in the region of 400–560 nm but also improved the tran-
sient response two times faster. Apart from using CdSe QDs, other CdSe nanostruc-
tures can also be used. Dutta et al. [43] fabricated a hybrid heterojunction
photodetector using 2D CdSe nanoplatelets (NPL) and phenothiazine (PTZ) hetero-
structure materials as the active layer in the device. They observed photolumines-
cence quenching and shortening of average decay time of CdSe QDs when PTZ was
mixed with 2D CdSe NPL. Further, the recombination of excitons in the CdSe
NPL. Photocurrent for the PTZ/CdSe NPL was shown to be ten-fold higher than
that of the only CdSe NPL based photodetectors. The photodetector also showed an
EQE of 40%, detectivity of 4 × 1011 Jones, and responsivity of 160 mA/W.

10.6 Application of CdSe Based Photodetectors

CdSe based photodetectors are actively used in polarization sensitive detectors.


Polarization is used to distinguish the artificial objects in complicated environ-
ments. Polarization sensitive devices rarely operate in UV region due to difficulty in
10 CdSe – Based Photodetectors for Visible-NIR Spectral Region 245

fabricating anisotropic materials and optical elements for polarization modulation.


Ge et al. [46] developed a polarization sensitive device by combining an electron-­
multiplying charge-coupled device (EMCCD) with a polarized luminescence down-
shifting material. They used CdSe@CdS-dot-in-rods (CdSe Dots in CdS rods) for
the downscaling the UV light into visible light where the down converted visible
light was allowed to be absorbed by an Si detector EMCCD (electron-multiplying
charge-coupled device). Singh at al [47] also fabricated a polarization sensitive pho-
todetector using solution processed CdSe nanowires with a bandgap of
1.75 eV. They observed a consistent polarization sensitivity in single as well as
ensembled CdSe NWs. The fabricated detector showed a wide spectral response
ranging from 350 to 720 nm. Though the authors claimed the device as a photocon-
ductor but they did not provide the details of the metallic contacts used in the
device [47].
CdSe based detectors integrated with Field Effect Transistors (FET) are used for
imaging systems. Shalev et al. [44] synthesized guided CdSe nanowires on five dif-
ferent planes of sapphire (substrate planes) shown in Fig. 10.13a, b. They fabricated
a CdSe nanowires based FET to examine the electronic and optoelectronic proper-
ties of the material as shown in Fig. 10.13c, d. Cr/Au metallic contacts and Al2O3

Fig. 10.13 (a) Three modes of guided growth with HRSEM image, (b) HRSEM image of
Graphotepitaxial growth of CdSe NWs, (c) SEM image indicating the device with source, drain
and CdSe NWs, (d) I-V curve of the device under different illumination at 473 nm, and (e)
Transient response of the device (at 10 kHz, 2 V bias, and 520 mW/cm2). (Adapted from Ref. [44].
Published 2017 by American Chemical Society as open access)
246 H. Kumar and S. Jit

Fig. 10.14 (a) Experimental setup for capturing a image using external filter, (b) Photocurrent
against time measured using CdSe/mica detector, (c) Full color image obtained using experimental
setup, (d) Red mono color image, (e) Green mono color image, and (f) Blue mono color image.
(Reprinted with permission from Ref. [9]. Copyright 2022: Springer Nature)

gate dielectric oxide were used in the FET to achieve 5 μm active area (between the
interdigitated contacts) of the device [44]. The fabricated detector showed a rise-­
time of 2.3 μs (τr) and fall-time of 2.5 μs (τd) as shown in Fig. 10.13e.
To achieve a complete visible imaging system, Pan et al. [9] fabricated the CdSe
thin film based photodetectors on flexible substrates. The photoconductor was ana-
lyzed over the complete visible spectrum for full-scale imaging using external filters
as shown in Fig. 10.14a, b. The obtained results are shown for full color and mono
color images in Fig. 10.14c–f.
The photo response and transient response were not very good. The only signifi-
cant part of the work was the fabrication of CdSe thin film-based photodetector on
a flexible substrate by using the very expensive molecular beam epitaxy (MBE)
method. Kim et al. [45] reported a flexible full-color detector utilizing all QDs only.
They used CdSe QDs for Red and Green light detectors by varying their sizes
(bandgap tuning) from 7 to 5 nm and used CdS QDs for the Blue region also. They
introduced PbS QDs in CdSe QDs to extend the spectral coverage of the detector
from the visible to infrared region. The detailed illustration of the full-color detector
with its absorption region is shown in Fig. 10.15a, b while the schematics of the
infrared, red, green, and blue pixels are shown in Fig. 10.15c. Working image of the
detector under illumination is shown in Fig. 10.15d, while Fig. 10.15e, f correspond
to the cross-sectional area of the detector obtained from the HRSEM.
10 CdSe – Based Photodetectors for Visible-NIR Spectral Region 247

Fig. 10.15 (a) Full-color detector array fabricated using QDs and Structure of single-pixel, (b)
Absorption characteristics of the QDs involved, and (c) Different devices for different regions of
the spectrum are arranged as per the QDs size, (d) Working image of the detector under illumina-
tion, (e) and (f) corresponds to depicted cross-sectional area HRSEM. (Reprinted from Ref. [45].
Published 2019 by AAAS as open access)

10.7 Summary

The present chapter introduces various state-of-the-art developments in the area of


CdSe based photodetectors operating in a wide range of wavelengths over visible to
NIR region. It is observed that CdSe is a potential material for photodetection appli-
cations due its direct bandgap nature and easy synthesis routes. Various novel CdSe
nanostructures of CdSe are believed to explored effectively for developing new gen-
eration photodetectors for imaging systems. Further, the solution-processed and
low-temperature fabrication processes of CdSe based photodetectors are suitable
for CdSe based flexible photodetectors.
248 H. Kumar and S. Jit

References

1. Singh V (2007) Albert Einstein: his Annus Mirabilis 1905. arXiv:physics/0701240. Accessed
15 Feb 2022 [Online]. Available http://arxiv.org/abs/physics/0701240
2. Baeg K-J, Binda M, Natali D, Caironi M, Noh Y-Y (2013) Organic light detectors: photodiodes
and phototransistors. Adv Mater 25(31):4267–4295. https://doi.org/10.1002/adma.201204979
3. Jansen-van Vuuren RD, Armin A, Pandey AK, Burn PL, Meredith P (2016) Organic photo-
diodes: the future of full color detection and image sensing. Adv Mater 28(24):4766–4802.
https://doi.org/10.1002/adma.201505405
4. Lin Q, Armin A, Burn PL, Meredith P (2015) Filterless narrowband visible photodetectors.
Nat Photonics 9(10):687–694. https://doi.org/10.1038/nphoton.2015.175
5. Bayer BE (1976) Color imaging array. US3971065A. Accessed 15 Feb 2022. [Online].
Available https://patents.google.com/patent/US3971065/en
6. Jin W, Hu L (2019) Review on quasi one-dimensional CdSe nanomaterials: synthesis and
application in photodetectors. Nano 9(10):1359. https://doi.org/10.3390/nano9101359
7. Ninomiya S, Adachi S (1995) Optical properties of cubic and hexagonal CdSe. J Appl Phys
78(7):4681–4689. https://doi.org/10.1063/1.359815
8. Samanta D, Samanta B, Chaudhuri AK, Ghorai S, Pal U (1996) Electrical characterization
of stable air-oxidized CdSe films prepared by thermal evaporation. Semicond Sci Technol
11(4):548–553. https://doi.org/10.1088/0268-­1242/11/4/016
9. Pan W, Liu J, Zhang Z, Gu R, Suvorova A, Gain S et al (2022) Large area van der Waals epi-
taxy of II–VI CdSe thin films for flexible optoelectronics and full-color imaging. Nano Res
15(1):368–376. https://doi.org/10.1007/s12274-­021-­3485-­x
10. Eyring MB, Martin P (2013) Spectroscopy in forensic science. In: Reference module in chemis-
try, molecular sciences and chemical engineering. Elsevier, p B978012409547205455X. https://
doi.org/10.1016/B978-­0-­12-­409547-­2.05455-­X
11. Li K, Lu Y, Fu XL, He J, Lin X, Zheng J et al (2021) Filter-free self-power CdSe/Sb2(S1−x,Sex)3
near infrared narrowband detection and imaging. InfoMat 3(10):1145–1153. https://doi.
org/10.1002/inf2.12237
12. Luo L-B, Xie W-J, Zou Y-F, Yu Y-Q, Liang F-X, Huang Z-J, Zhou K-Y et al (2015) Surface
plasmon propelled high-performance CdSe nanoribbons photodetector. Opt Express
23(10):12979. https://doi.org/10.1364/OE.23.012979
13. Kumar H, Kumar Y, Mukherjee B, Rawat G, Kumar C, Pal BN et al (2019) Effects of optical
resonance on the performance of metal (Pd, au)/CdSe quantum dots (QDs)/ZnO QDs optical
cavity based spectrum selective photodiodes. IEEE Trans Nanotechnol 18:365–373. https://
doi.org/10.1109/TNANO.2019.2907529
14. Kochman B, Stiff-Roberts AD, Chakrabarti S, Phillips JD, Krishna S, Singh J et al (2003)
Absorption, carrier lifetime, and gain in InAs-GaAs quantum-dot infrared photodetectors.
IEEE J Quantum Electron 39(3):459–467. https://doi.org/10.1109/JQE.2002.808169
15. Straus DB, Goodwin ED, Gaulding EA, Muramoto S, Murray CB, Kagan CR (2015) Increased
carrier mobility and lifetime in CdSe quantum dot thin films through surface trap passivation
and doping. J Phys Chem Lett 6(22):4605–4609. https://doi.org/10.1021/acs.jpclett.5b02251
16. Kongkanand A, Tvrdy K, Takechi K, Kuno M, Kamat PV (2008) Quantum dot solar cells.
Tuning photoresponse through size and shape control of CdSe−TiO2 architecture. J Am Chem
Soc 130(12):4007–4015. https://doi.org/10.1021/ja0782706
17. Kumar H, Kumar Y, Rawat G, Kumar C, Mukherjee B, Pal BN et al (2017) Heating effects of
colloidal ZnO quantum dots (QDs) on ZnO QD/CdSe QD/MoOx photodetectors. IEEE Trans
Nanotechnol 16(6):1073–1080. https://doi.org/10.1109/TNANO.2017.2761785
18. McGott DL, Good B, Fluegel B, Duenow JN, Wolden CA, Reese MO Carrier lifetime as
a function of Se content for CdSex Te1-x films grown on Al2O3 and MgZnO. In: 2021 IEEE
48th photovoltaic specialists conference (PVSC), Fort Lauderdale, FL, USA, pp 1301–1303.
https://doi.org/10.1109/PVSC43889.2021.9518914
10 CdSe – Based Photodetectors for Visible-NIR Spectral Region 249

19. Vietmeyer F, Frantsuzov PA, Janko B, Kuno M (2011) Carrier recombination dynam-
ics in individual CdSe nanowires. Phys Rev B 83(11):115319. https://doi.org/10.1103/
PhysRevB.83.115319
20. Kung S-C, Xing W, van der Veer WE, Yang F, Donavan KC, Cheng M et al (2011) Tunable
photoconduction sensitivity and bandwidth for lithographically patterned nanocrystalline cad-
mium selenide nanowires. ACS Nano 5(9):7627–7639. https://doi.org/10.1021/nn202728f
21. Osedach TP, Geyer SM, Ho JC, Arango AC, Bawendi MG, Bulović V (2009) Lateral hetero-
junction photodetector consisting of molecular organic and colloidal quantum dot thin films.
Appl Phys Lett 94(4):043307. https://doi.org/10.1063/1.3075577
22. Jiang Y, Zhang WJ, Jie JS, Meng XM, Fan X, Lee S-T (2007) Photoresponse properties of
CdSe single-nanoribbon photodetectors. Adv Funct Mater 17(11):1795–1800. https://doi.
org/10.1002/adfm.200600351
23. Xu J, Rechav K, Popovitz-Biro R, Nevo I, Feldman Y, Joselevich E (2018) High-Gain 200 ns
photodetectors from self-aligned CdS-CdSe core-shell nanowalls. Adv Mater 30(20):1800413.
https://doi.org/10.1002/adma.201800413
24. Wu P, Dai Y, Sun T, Ye Y, Meng H, Fang X et al (2011) Impurity-dependent photoresponse prop-
erties in single CdSe nanobelt photodetectors. ACS Appl Mater Interfaces 3(6):1859–1864.
https://doi.org/10.1021/am200043c
25. Mehta RR, Sharma BS (1973) Photoconductive gain greater than unity in CdSe films with
Schottky barriers at the contacts. J Appl Phys 44(1):325–328. https://doi.org/10.1063/1.1661881
26. Skarman JS (1965) On the relationship between photocurrent decay time and trap distri-
bution in CdS and CdSe photoconductors. Solid State Electron 8(1):17–29. https://doi.
org/10.1016/0038-­1101(65)90005-­5
27. Margulis W, Sibbett W (1983) Picosecond CdSe photodetector. Appl Phys Lett 42(11):975–977.
https://doi.org/10.1063/1.93820
28. Das B, Sa K, Mahakul PC, Subramanyam BVRS, Das S, Alam I et al (2018) Efficient ultra-
violet photodetector device based on modulated wide band gap Type-II CuO/CdSe core-shell
nanowires. Superlattice Microst 123:234–241. https://doi.org/10.1016/j.spmi.2018.08.021
29. Al-Ani SKJ, Mohammed HH, Al-Fwade EMN (2002) The optoelectronic properties of
CdSe:Cu photoconductive detector. Renew Energy 25(4):585–590. https://doi.org/10.1016/
S0960-­1481(01)00088-­X
30. An Q, Meng X, Xiong K, Qiu Y, Lin W (2017) One-step synthesis of CdSe nanotubes with
novel hollow tubular structure as high-performance active material for photodetector. J Alloys
Compd 726:214–220. https://doi.org/10.1016/j.jallcom.2017.07.336
31. Yuan Y, Zhang X, Liu H, Yang T, Zheng W, Zheng B et al (2020) Growth of CdSe/MoS2 verti-
cal heterostructures for fast visible-wavelength photodetectors. J Alloys Compd 815:152309.
https://doi.org/10.1016/j.jallcom.2019.152309
32. Ghods S, Esfandiar A, Zad AI, Vardast S (2022) Enhanced photoresponse and wavelength
selectivity by SILAR-coated quantum dots on two-dimensional WSe2 crystals. ACS Omega
7(2):2091–2098. https://doi.org/10.1021/acsomega.1c05591
33. Ren W, Tan Q, Wang Q, Liu Y (2021) Hybrid organolead halide perovskite microwire arrays/
single CdSe nanobelt for a high-performance photodetector. Chem Eng J 406:126779. https://
doi.org/10.1016/j.cej.2020.126779
34. Medda A, Dutta A, Bain D, Mohanta MK, De Sarkar A, Patra A (2020) Electronic structure
modulation of 2D colloidal CdSe nanoplatelets by au 25 clusters for high-performance pho-
todetectors. J Phys Chem C 124(36):19793–19801. https://doi.org/10.1021/acs.jpcc.0c04774
35. Luo J, Zheng Z, Yan S, Morgan M, Zu X, Xiang X, Zhou W (2020) Photocurrent enhanced
in UV-vis-NIR photodetector based on CdSe/CdTe core/shell nanowire arrays by
piezo-­phototronic effect. ACS Photonics 7(6):1461–1467. https://doi.org/10.1021/
acsphotonics.0c00122
36. Li C, Li W, Cheng M, Yang W, Tan Q, Wang Q, Liu Y (2021) High sensitive and broadband
photodetectors based on hybrid PbI2 nanosheet/CdSe nanobelt. Adv Opt Mater 9(20):2100927.
https://doi.org/10.1002/adom.202100927
250 H. Kumar and S. Jit

37. Villa-Angulo C (2020) CdSe0.4 Te0.6 quantum well-based photodetector toward imaging vision
sensors. IEEE Sensors J 20(22):13357–13363. https://doi.org/10.1109/JSEN.2020.3006219
38. Wang X, Song W, Liu B, Chen G, Chen D, Zhou C, Shen G (2013) High-performance
organic-inorganic hybrid photodetectors based on P3HT:CdSe nanowire heterojunctions on
rigid and flexible substrates. Adv Funct Mater 23(9):1202–1209. https://doi.org/10.1002/
adfm.201201786
39. Shih Y-H, Chen Y-L, Tan J-H, Chang SH, Uen W-Y, Chen S-L et al (2020) Low-power, large-­
area and high-performance CdSe quantum dots/reduced graphene oxide photodetectors. IEEE
Access 8:95855–95863. https://doi.org/10.1109/ACCESS.2020.2995676
40. Oertel DC, Bawendi MG, Arango AC, Bulović V (2005) Photodetectors based on treated CdSe
quantum-dot films. Appl Phys Lett 87(21):213505. https://doi.org/10.1063/1.2136227
41. Ramar M, Kajal S, Pal P, Srivastava R, Suman CK (2015) Study of binary and ternary organic
hybrid CdSe quantum dot photodetector. Appl Phys A Mater Sci Process 120(3):1141–1148.
https://doi.org/10.1007/s00339-­015-­9293-­y
42. Ashraf Malik H, Aziz F, Asif CM, Raza E, Najeeb MA, Ahmad Z et al (2016) Enhancement
of optical features and sensitivity of MEH-PPV/VOPcPhO photodetector using CdSe quantum
dots. J Lumin 180:209–213. https://doi.org/10.1016/j.jlumin.2016.08.038
43. Dutta A et al (2020) Hybrid nanostructures of 2D CdSe nanoplatelets for high-performance
photodetector using charge transfer process. ACS Appl Nano Mater 3(5):4717–4727. https://
doi.org/10.1021/acsanm.0c00728
44. Shalev E, Oksenberg E, Rechav K, Popovitz-Biro R, Joselevich E (2017) Guided CdSe nanow-
ires parallelly integrated into fast visible-range photodetectors. ACS Nano 11(1):213–220.
https://doi.org/10.1021/acsnano.6b04469
45. Kim J, Kwon S-M, Kang YK, Kim Y-H, Lee M-A, Han K-J et al (2019) A skin-like two-­
dimensionally pixelized full-color quantum dot photodetector. Sci Adv 5(11):eaax8801.
https://doi.org/10.1126/sciadv.aax8801
46. Ge Y, Zhang M, Meng L, Tang J, Chen Y, Wang L, Zhong H (2019) Polarization-sensitive
ultraviolet detection from oriented-CdSe@CdS-dot-in-rods-integrated silicon photodetector.
Adv Opt Mater 7(18):1900330. https://doi.org/10.1002/adom.201900330
47. Singh A, Li X, Protasenko V, Galantai G, Kuno M, Xing H, Jena D (2007) Polarization-­
sensitive nanowire photodetectors based on solution-synthesized CdSe quantum-wire solids.
Nano Lett 7(10):2999–3006. https://doi.org/10.1021/nl0713023
Chapter 11
CdS-Based Photodetectors for Visible-UV
Spectral Region

Nupur Saxena, Tania Kalsi, and Pragati Kumar

11.1 Introduction

Photodetector (PD) is a device that can convert light illuminations either directly
(photoelectric effect) or indirectly (photothermal effect) into electric signals. In for-
mer case, the absorbed photons’ energy generates electron-hole (e-h) pairs and are
detected as change in resistance/current in external circuit like in high speed Si and
GaAs photodetectors [1]. While, in other class of PDs, light induced temporal tem-
perature gradient (dT/dt) is converted into an electrical signal like in pyroelectric
detectors [2]. Further, PDs may also be classified in variety of ways viz.; on the
basis of their basic structures & working mechanism, spectral sensing range & their
applications, architecture etc. Figure 11.1 shows the cataloguing of photooelectric
effect based PDs on the basis of their basic structure & working mechanism whereas
Fig. 11.2 depicts spectral sensing range and field of applications of PDs.
The specific field of application require a PD that detects the light of particular
spectral region. For example, ultra violet (UV) and UV-Vis. PDs are exploited in the
monitoring of ozone layer to UV based skin therapy and missile warning systems to
astronomy whereas Vis. and/or near infrared (NIR) PDs are employed in many com-
mercial consumer electronics to imaging systems such as smart phones, digital
camera, diagnostics, and remote sensing [9–11]. Further, the choice of materials,
photosensitivity in diverse range of spectral wavelength are primarily depend on
their band gaps. For instance, narrow band gap semiconductor materials like PbS,
PbSe, InAs, and GaSb etc. are well suited for the fabrication of NIR to long

N. Saxena
Organisation for Science Innovations and Research, Garha Pachauri, India
T. Kalsi · P. Kumar (*)
Nano-Materials and Device Lab, Department of Nanoscience and Materials, Central
University of Jammu, Rahya-Suchani, Jammu & Kashmir, India
e-mail: pkumar.phy@gmail.com; pkumar.phy@cujammu.ac.in

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 251
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_11
252 N. Saxena et al.

Fig. 11.1 Types of PDs on the basis of basic structures and working mechanism. (Photoconductor
reprint from Ref. [3]. Published 2010 by MDPI as open access, M-S-M PD reprint from Ref. [4].
Published 2014 by Intechopen as open access, MIS PD adapted with the permission from Ref. [5].
Copyright 2015: Elsevier B.V., p-i-n photodiode. Reprinted from Ref. [6], Schottky photodiode
adapted from Ref. [7]. Published 2019 by DE Gruyter as open access, and phototransitor adapted
with the permission from Ref. [8]. Copyright 2014: WILEY-VCH Verlag GmbH & Co)

wavelength infrared (LWIR) PDs. Contrary, wide band gap semiconductor materi-
als viz.: ZnO, ZnS, CdS, GaN, and SiC are usually sense the light in UV to NIR
spectral region.
Among various functional materials CdS is one of the most extensively exciting
and explored material due to its wide range of applications starting from photonic
devices to diagnostic tools including PDs [12–24]. In particular, CdS is a phenom-
enal material for the development of UV-Vis.-NIR PDs because of its bulk band gap
(~2.42 eV at room temperature) lying in visible region of spectrum which can be
either tailored towards UV region by varying the particle size or NIR region by
creating the intermediate energy states in between band gap via very familiar and
easy strategy of impurity incorporation. In addition, solution processability, low
cost synthesis, easy control on shape and size etc. are the other key factors for enor-
mous use of CdS in PDs and other applications. Further, CdS based multispectral
PDs may also be fabricated via synthesis of heterojunction, composite, decoration
with other materials of interest. Till date, various nanostructures of CdS including
quantum dots (QDs), nanowires, nanobelts, thin and thick films and their hybrids
have been employed for the design of diverse configuration of PDs including pho-
toconductive (ohmic), photodiodes (p-n junction, Schottky junction and p-i-n type
photodiodes etc.), and phototransistor as illustrated in Fig. 11.1.
11 CdS-Based Photodetectors for Visible-UV Spectral Region 253

Fig. 11.2 Application areas of PDs for better living on the basis of their spectral range of detec-
tion. (Adapted with the permission from Ref. [1]. Copyright 2017: Wiley-VCH Verlag GmbH)

11.2 Conventional Photodetectors and Features


of Their Functioning

A conventional photoconductor PD has a planar structure consist of photo-sensitive/


active layer between two transverse metal electrodes often in ohmic contacts [25].
In spite of the advantages of the facile fabrication and high employment probability
in flexible optoelectronic devices [11], large gain (G) [26], high external quantum
efficiency (EQE) and responsivity (R) [25], application of such PDs are limited due
to slow photoresponse, low photosensitivity and obvious dielectric hysteresis [11].
A conventional photodiode PD is a vertical structure (except M-S-M type) with
rectification effect similar as photovoltaic devices. The photodiode PDs exhibit
small dark current, quick response and high separation efficiency of e-h pairs espe-
cially at low bias. In specific cases, where high quantum efficiency of PDs is in
demand, a large external bias is applied to trigger avalanche or carrier multiplication
effects that leads to the generation of multiple carriers excited by a single pho-
ton [25].
A metal–semiconductor–metal (M-S-M) PD is a special kind of Schottky barrier
detector with back-to-back Schottky contacts in planner geometry separated by
semiconducting layer in its simplest structure, i.e., two transverse metallic elec-
trodes on a semiconductor material, in contrast to a p–n junction as in a photodiode.
However, the use of interdigited electrode structures (the finger spacing or ring-­
shaped) is a general practice for construction of M-S-M PDs [27]. A practically
254 N. Saxena et al.

important aspect of M-S-M PDs is relatively simple planar structure that is suitable
for monolithic integration in particular with other photonic components on semi-
conductor integrated circuits. Besides, M-S-M PDs offer the advantages of fast
response, low capacitance per unit area, low dark current, low-noise, and high sen-
sitivity [28].
Conventional phototransistor PDs are three-terminal devices consisting of
source, drain, gate, and photoactive channel in which gate voltage controls channel
conductivity and encourage dark current to be lower and higher photocurrent [26].
Inclusion of gate and dielectric layers into the PDs device structure fallouts reduc-
tion in the noise signal, augmentation in the electrical signal, and finally boost up
the responsivity and gain [11]. Figure 11.3 represents the schematic of architectures
along with their energy band diagrams of above discussed PDs.
The different kinds of PDs structures discussed above are though based on dif-
ferent photodetection mechanism their performance is governed and assessed by the
same key parameters known as figures of merit (FOM) or performance parameter.
FOM include photosensitivity (S) [9, 29], photoresponsivity (R) [25, 26], external
quantum efficiency (EQE) [25, 26, 29], specific detectivity (D*) [11, 25, 29], gain
(G) [11, 25], response time/speed (τ), and noise equivalent power (NEP) [11, 25]
and are given by the expressions:

I Ph  I d
S (11.1)
Id

where, Id and Iph are the dark and photocurrent respectively.

Fig. 11.3 Schematic of architectures and corresponding energy band diagrams of diverse PDs. (a,
b) Photoconductors; (c, d) Photodiodes; (e, f) Phototransistors. (Adapted with the permission from
Ref. [26]. Copyright 2020: Wiley-VCH GmbH)
11 CdS-Based Photodetectors for Visible-UV Spectral Region 255

I Ph
R= (11.2)
Pin

where, Pin indicates the light intensity of the source. More precisely, spectral respon-
sivity ‘Rλ’ is given by [11, 29];

I Ph  I d
R
Pin
 AW 
1
(11.3)

EQE measures the number of collected charge carriers per unit incident photons and
is closely related to R value. Therefore, it is the ratio of and can be expressed as;

Nc hc 1.24
EQE  ex    R  R (11.4)
N ip e e

where Nc is the number of photogenerated carriers, Nip is the number of incident


photons, h is Planck’s constant, c is velocity of light, e is electronic charge, and λ is
wavelength of incident light. Additionally, one can also estimate the internal quan-
tum efficiency (IQE) by the equation [25];

N c EQE ex
IQE  in    or (11.5)
Na a a

Where ηa is the light absorption efficiency of photoactive material.

R
D 
IN
A f  Jones or cmHz 1/ 2
W 1  (11.6)

where A is effective area of device, ∆f is electrical bandwidth, and IN is noise cur-


rent. Ordinarily, Id is the dominating contribution to the noise signal, so the equation
can be truncated as [10, 11, 29];

A
D  R
2qI d
 Jones or cmHz 1/ 2
W 1  (11.7)

A f I
NEP  
 N (11.8)
D R

NEP may also be estimated as [26];


256 N. Saxena et al.

I N2
NEP  (11.9)
R

Under these circumstance, electrons drifted to the external circuit may circulate
several times in the channel and result in the photoconductive gain. ‘G’ is deter-
mined by [11, 25];

l l
G  VDS (11.10)
 t L2

where τl is lifetime of trapped carrier say hole, τt is transit time of the other carrier
say electron, L is the channel length, μ is the carrier mobility, and VDS is the applied
bias voltage.
Alternative critical performance parameter is ‘τ’ is defined as in [9–11, 25, 29]
and can be estimated as [26];

1
fc  (11.11)
2

11.3 Fabrication of Photosensitive Devices

11.3.1 CdS PDs Based on Nanostructures

A PD device was fabricated with aligned CdS NTs arrays as active layer and Ag
NWs network as transparent electrodes by An and Meng [30]. The schematic dia-
gram depicting the fabrication process of their PD is shown in Fig. 11.4. The trans-
parent and conducting Ag NWs networks with uniform distribution were produced
by spraying the Ag NWs redispered in ethanol onto the prepared quartz using micro
syringes (Fig. 11.4a). Further, mechanically scratching of the substrate using a razor
blade at a gap about 40 μm was carried out and was cut into two totally disconnected
parts to use the Ag NWs network as electrodes (Fig. 11.4b). Next, the electrodes
produced as aforementioned were turned over and placed on top of the aligned CdS
NTs arrays substrate (Fig. 11.4c, d), producing PDs. The PD based on the single Sn
doped CdS NW was fabricated by photo lithography on Si/SiO2 substrate with In
electrodes [11].
Jin et al. [31] fabricated PD with active layer of CdS flakes by a laser direct writ-
ing system and standard electron-beam lithography (Fig. 11.5a) followed by the
deposition of Cr/Au (10 nm/50 nm) electrodes using the thermal evaporation.
Nawaz et al. [32] fabricated M-S-M PDs of CdS NBs deposited over SiO2/Si and
polyimide (PI) substrates followed by the deposition of Ag electrodes (Fig. 11.5b),
whereas PD with “L” shape electrodes of conductive Ag paste were designed onto
11 CdS-Based Photodetectors for Visible-UV Spectral Region 257

Fig. 11.4 Schematic diagram showing the fabrication process of PD. (a) Deposition of Ag NWs
on quartz substrate, (b) Process of mechanically scratching of the substrate, (c) Turning the elec-
trodes, (d) placing upside down on top of the aligned CdS NTs networks substrate, and (e)
Photographs of final device. (Adapted with the permission from Ref. [30]. Copyright 2016:
Springer)

Fig. 11.5 (a) CdS NFs based PD fabricated by laser directed writing system. (Adopted with the
permission from Ref. [31]. Copyright 2018: WILEY-VCH Verlag GmbH). (b) CdS NBs M-S-M
PD. (Adapted with the permission from Ref. [32]. Copyright 2021: the Royal Society of Chemistry).
(c) Schematic of (i) the pristine film, (ii) the film under exposure to ion beam, (iii) various promis-
ing phenomena as a consequence of passage of ion beam through the material and (iv) the fabri-
cated PD. (Adapted with the permission from Ref. [9]. Copyright 2016: The Royal Society of
Chemistry)

CdS over glass slides [33]. The PD of spray deposited Y:CdS TF on glass was fab-
ricated by making Ag contacts in M-S-M geometry [3]. PD of CdS/ZnO core/shell
NWs were fabricated by coating the paste of CdS/ZnO core/shell NWs in deionized
water on a ceramic substrate (over which silver interdigitated electrodes with both
258 N. Saxena et al.

finger-width and inter-finger spacing of about 200 μm was previously printed) [34].
An another core-shell PD was designed by deposition of two step CVD synthesized
CdS-CdSxTe1- x -CdTe core-shell NBs over SiO2 substrate followed by the construc-
tion of electrodes using silver paste [35].
An entirely different strategy was adopted by Kumar et al. [9] to fabricate PDs in
M-S-M geometry with Ni6+ ion beam irradiated CdS films over Si substrate
(Fig. 11.5c (i & ii)). In this work, the CdS films were grown over Si via PLD and
irradiated under different ion fluences. Ion beam irradiation resulted many processes
(as shown in Fig. 11.5c (iii)), which modified various properties of the films. Finally,
these films were used to design PD (Fig. 11.5c (iv)).

11.3.2 CdS-Based Heterostructures in PDs

Figure 11.6 shows process of PD fabrication and of final PD device developed by


Chakrabarti et al. [36]. The spin coated PEDOT:PSS films on ITO coated glass sub-
strate were patterned using water vapour assisted capillary force lithography. They
have fabricated patterned stamps via Replica Molding of Sylgard 184 against the
polycarbonate part of commercially available optical data discs such as CD and
DVD. The details of using optical data storage discs for patterning can be found
elsewhere [37]. To make the cross linked Sylgard 184 stamp wettable, it was exposed
in a UV ozone chamber for 30 min. For pattern replication, the patterned Sylgard
184 stamp was carefully placed over the film surface (step 2, Fig. 11.6) and was
subsequently transferred for 18 h in a water vapour’s pre-saturated chamber at room
temperature (step 3, Fig. 11.6).
The capillary driven rise of polymer meniscus along the contour of the confining
stamp developed replica of pattern (inset of step 3, Fig. 11.6). After this, the

Fig. 11.6 Schematic diagram for imprinting of PEDOT:PSS film and final device. (Adapted with
the permission from Ref. [36]. Copyright 2018: Institute of Physics)
11 CdS-Based Photodetectors for Visible-UV Spectral Region 259

assembly of patterned film and stamp was taken out from the chamber and heated
under the application of uniform normal load of 4 KPa in a vacuum oven at 80 °C
for 25 min to remove the residual water molecules in the film matrix. In the last step,
stamp was gently peeled off after cooling (step 4, Fig. 11.6). The final architecture
(ITO/PEDOT:PSS/CdS/ZnO/Al) of PD devices were achieved by growing 150 nm
CdS TF over patterned and flat PEDOT:PSS films, followed by the deposition of
ZnO TF over CdS TF layer via rf-sputtering and subsequent deposition of 150 nm
thick layer of Al electrodes over ZnO TF by thermal evaporation method.
An architecture FTO/CdS NRs/SnS NFs/Au was designed by growing CdS NRs
array via hydrothermal over FTO substrate followed by the deposition of SnS NFs
layer and Au electrodes via thermal evaporation using a shadow mask [40]. Whereas,
an architecture FTO/CdS NRs/perovskite/Spiro-OMeTAD/Ag was fabricated by
post annealing of hydrothermally grown CdS NRs array on ITO in air at 200 °C for
different time to make gradient O-CdS NRs array and then spin coated perovskite
layer over gradient O-CdS NRs array followed by the coating of Spiro-OMeTAD
layer. Finally, thermal evaporation was used to deposit the100 nm thick Ag elec-
trode [41]. Perpendicular CdS NWs array of 600 nm length and 50 nm dia was
grown through a hydrothermal route on the flexible substrate of Si micropyramids
(Fig. 11.7a, b (b2)) of bottom edge length ~3–7 μm that were synthesized by chemi-
cal etching process to form the p-Si/n-CdS heterojunction. Further, ITO layer as the
transparent top-electrode and Al as bottom-electrode were deposited on the CdS
NWs and the flexible p-Si substrate respectively to construct a flexible device struc-
ture ITO/CdS NWs/p- Si/Al [38]. An ordered and aligned CdSe@CdS dots in rods
(DIRs) embedded in PVDF composite free standing films were synthesized by sol-­
gel and peel off method [42]. A PD of configuration FTO/TiO2/CdS/Se/Ag was
fabricated by growing 2 nm TiO2 layer by atomic layer deposition (ALD) followed
by microwave assisted hydrothermal growth of CdS layer and CVD growth of Se
layer and 100 nm Ag electrode deposition via thermal evaporation [43]. An archi-
tecture of Al/CdS NPs-CdO/p-si/Al was fabricated by deposition of CdO thin film
using sol-gel spin coating method onto p-Si, followed by CdS NPs structures using
SILAR technique and Al contact by thermal evaporating using physical mask [44].
A position-sensitive array (PSA) PD was designed by the growth of comb-like Sn
doped CdS NSs with cone-shape branched via in situ seeding CVD over Au coated
Si substrate and depositing the Pt electrode using focused ion beam (FIB) technol-
ogy by introducing precursor gas into the Ga ion beam as shown in Fig. 11.7c–h [39].
A PD of configuration FTO/CdS NFs/Ag was fabricated by annealing the as
deposited thin film of CdS NFs over FTO at 300 °C in air for 30 min. Followed by
the pattering of Ag electrodes of thickness 200 nm using mask in thermal evapora-
tion and subsequently wiping out the part of CdS NFs film upon FTO to make the
FTO another transport electrode as illustrated in Fig. 11.8a [45]. CdS core-Au/
MXene-Based (Fig. 11.8b) PD was fabricated by Jiang et al. [46]. First, they
dropped the solution of CdS core-Au NPs in ethanol onto fluorophlogopite substrate
and natural air-drying at room temperature followed by the dropping of colloidal
solution of MXene on both sides of CdS core-Au NPs as the electrodes and at last,
the total device was annealed at 100 °C for 5 min in air. The architecture P
­ ET/ITO/
260 N. Saxena et al.

Fig. 11.7 (a) Schematic structure of the self-powered photodetector (SPPD). (b) Characterizations
(b1) an optical image of flexible Si wafer after etching; (b2) SEM image of the etched Si wafer
surface; (b3, b4) top view (b3) and side view (b4) of the CdS NWs array. (Adapted with the per-
mission from ref. [38]. Copyright 2018: Wiley-VCH Verlag GmbH). (c) Schematic diagram of the
CdS branched PD, (d) SEM image of the fabricated device, Schematic diagram to realize light
position identification in the (e) X-direction, (f) Y-direction of the PD, (g) Optical image of the
laser irradiating at different positions on a branch with wavelength 405 nm, and (h) Resistance
distribution of the branched nanostructure. Above and below simulations correspond to the
branched structures without and with light illumination, respectively. (Adapted with the permis-
sion from Ref. [39]. Copyright 2018: Wiley-VCH Verlag GmbH)

CdS/NiO/Al was developed by deposition of CdS TF onto PET/ITO substrate via


photochemical method followed by the spin coating of hydrothermally synthesized
NiO NPs and depositing the top Al metal contacts (120 nm) on the heterojunction
using a shadow mask (Fig. 11.8c) [47].
The MoTe2 flakes were exfoliated directly on a 300 nm SiO2/Si substrate and
contact electrode for flake was constructed using photolithography, followed by
electron beam evaporation of a 50 nm Au electrode. Subsequently, the Ga-doped
11 CdS-Based Photodetectors for Visible-UV Spectral Region 261

Fig. 11.8 Schematic illustrations (a) CdS NFs based PD. (Adopted with the permission from Ref.
[45] © 2018 Elsevier GmbH). (b) Assembled CdS core-Au/MXene PD. (Adapted with the permis-
sion from Ref. [46]. Copyright 2020: Elsevier B.V). (c) the solution manufacturing process for
CdS, NiO and PD construction. (Adapted with the permission from Ref. [47]. Copyright 2021:
Elsevier B.V)

Fig. 11.9 (a) Experimental process for the fabrication of a 1D Ga-doped CdS NWs/2D MoTe2
flake heterojunction PD. (Adapted with the permission from Ref. [48]. Copyright 2018: WILEY-­
VCH Verlag GmbH). (b) Fabrication process of the stretchable SnO2-CdS interlaced NW PD and
optical image of the as-fabricated stretchable device arrays. (Adapted with the permission from
Ref. [49]. Copyright 2019: Springer Nature)

CdS NWs were placed using a manipulator onto a MoTe2 flake followed by photo-
lithography for the realization of (100 nm/150 nm) Ti/Au electrode to Ga-doped
CdS NW to develop final structure of the p–n heterojunction PD (Fig. 11.9a) [48].
The PD based on stretchable SnO2-CdS interlaced NWs was designed by Li et al.
[49] They have used 3D printed two kinds of polydimethylsiloxane (PDMS) masks
(one for Ag NW electrodes, and other for SnO2 and CdS NW channels). Fabrication
of the SnO2-CdS interlaced NW PD involved the placement of the masks on a
262 N. Saxena et al.

polycarbonate (PC) filter membrane to filter the Ag NW electrode pattern and the
SnO2 and CdS NW channel patterns respectively. After this PC filter membrane
with Ag, SnO2 and CdS NW patterns was placed in a plastic culture dish followed
by the injection of PDMS liquid on the top of the filter membrane and peeled off
thermally cured PDMS substrate from the PC filter membrane with Ag, SnO2 and
CdS NW patterns transferred onto the PDMS matrix. The entire fabrication process
is illustrated here in Fig. 11.9b. A similar method was previously used by Pe et al.
[50] to design PD with mixed cellulose esters (MCE)/CdS NWs/Ag NWs architec-
ture. They drop casted CdS NWs on MCE substrate, covered with PDMS mask
followed by the drop casting of Ag NWs over CdS NWs layer covered by another
mask. The configuration glass/CdSxSe1-x/Al was developed by the deposition of
CdSxSe1-x via thermal co-evaporation over glass followed by the deposition of Al
electrodes [51] and in subsequent year they fabricated Al/p-Si/CdSxSe1-x/Ag PD fol-
lowing the same process [52]. UV lithography followed by thermal evaporation and
lift-off process were used to fabricate two terminal nanodevices of configuration Si/
SiO2/CdS NWs/(10 nm) Ti/(300 nm) Au [53] and n-Si/SiO2/GaTe NFs/Sn:CdS
NWs/(2 nm) Ti/(120 nm) Au [54]. An M-S-M PD based on NCs CdS-porous silicon
(PS):p-Si heterostructure was developed via thermal evaporation of chemically syn-
thesized CdS powder simultaneously on both glass and PS:p-Si (prepared by elec-
trochemical anodization of (100) oriented boron doped p-type Silicon wafer)
substrates followed by the deposition of Al metal electrode of ~20 nm thickness
using interdigitated shadow mask [55]. A photoconductor PD with stacking Si/SiO2/
WS2 NSs/CdS NWs/Cr/Au was designed via CVD growth of CdS NWs/WS2 NSs
over a SiO2(300 nm)/Si substrate followed by the spin coating of MMA and poly-
methyl methacrylate (PMMA) copolymers layers and deposition of Au/Cr elec-
trodes (50 nm/5 nm) using electron beam lithography [56].
Xu et al. [26] developed PD based on self-aligned CdS/CdSe core/shell nanow-
alls (NWAs). Figure 11.10a illustrates schematic of typical experimental steps for
the growth of CdS/CdSe core/shell NWAs. The nanogroves were formed by anneal-
ing as-received M(1010) sapphire at 1600 °C for 10 h followed by selective deposi-
tion of Au NPs. Subsequently, CdS powder was used to grow aligned CdS NWAs by
PVD followed by the covering of polycrystalline Al2O3 using ALD and defining the
area to be etched via photolithography, etching in a buffered oxide etch (BOE) sol-
vent, and lift-off the remained photoresist with acetone. Next, CdSe was grown via
PVD on the sample with selective-etched area so that only the exposed NWs sur-
faces were coated with CdSe layers via a surface-epitaxial growth owing to the
selective protection of the Al2O3 layer and remaining Al2O3 layer was removed by
another etching in the BOE solvent. A PD is designed by sputtering p and n type
Nd:TiO2 and CdS layer onto FTO followed by the formation of connections in the
circuit board as illustrated in Fig. 11.10b [58].
11 CdS-Based Photodetectors for Visible-UV Spectral Region 263

Fig. 11.10 Schematic illustrations (a) experimental steps for the growth of CdS/CdSe core/shell
NWs. (Adapted with the permission from Ref. [57]. Copyright 2018: WILEY-VCH Verlag GmbH).
(b) Device structure, the light is injected from the back of the glass (left) and the real devices with
silver paste connection on the circuit board (in the red circles). (Adapted with the permission from
Ref. [58]. Copyright 2020: IOP Publishing Ltd)

Fig. 11.11 Schematic illustrations (a) single CdS NB FET PD. (Adapted with the permission
from Ref. [59]. Copyright 2019: American Institute of Physics). (b) Ferroelectric side-G single
CdS NW FET PD. (Adapted with the permission from Ref. [60]. Copyright 2016: WILEY-VCH
Verlag GmbH). (c) the fabrication process of the M-WS2 /CdS/WS2 NFs PD. (Adapted with the
permission from Ref. [62]. Copyright 2021: the Royal Society of Chemistry)

11.3.3 CdS-Based Field Effect Transistor (FET) PDs

FET PDs of CVD grown single crystal CdS NBs on a SiO2/Si substrate via electron-­
beam lithography process and metal deposition and lift-off process followed by the
deposition of HfO2 as a passivation layer using ALD (device-I) and exfoliation and
transfer of a thin layer of transparent graphene onto HfO2 as a top gate (G) electrode
(device-II shown in Fig. 11.11a) [59]. The configuration Si/SiO2/CdS/Ti-Au of PD
264 N. Saxena et al.

was developed by syntheses of CdS TF on Si/SiO2 substrate by CBD followed by


the deposition of Ti(10 nm)/Au (90 nm) electrode by sputtering method. Ferroelectric
polymer side-G FET PD (Fig. 11.11b) was fabricated by transferring CVD synthe-
sized CdS NWs onto Si/SiO2 (110 nm) substrate followed by the spin coating of
MMA and PMMA, and defining the drain (D), source (S), and side-G patterns via
electron-beam lithography technique. Subsequently, the Cr/Au (15 nm/50 nm) elec-
trodes (S and D) were fabricated by metal evaporation and lift-off processes fol-
lowed by spin coating of 200 nm of P(VDF-TrFE) (70:30 in mol%) ferroelectric
polymer film on the NW channel to construct side-G electrode [60]. A similar pro-
cess was used to fabricated single CdS NW based FET PD where, photoresist was
spin coated onto Si/SiO2/CdS NWs and photolithography was used to define S, D
and G followed by the deposition of (10/100 nm) Ti/Au via e-beam evaporation
[61]. A FET PD was designed based on PVD synthesized WS2 NFs and SnO2-­
assisted CVD grown CdS micro wires (μWs) heterojunction. The fabrication pro-
cess is illustrated in Fig. 11.11c and involves first, spin coating of photoresist over
Si/SiO2/WS2 NFs followed by photolithography to define S, D, and G and deposi-
tion of (10/60 nm) Ti/Au electrode via e-beam evaporation. Secondly a CdS μWs
were transferred in the middle of the WS2 NFs channel under the assistance of
PMMA followed by mechanical exfoliation and transfer of multi-layered WS2 NFs
to the top of CdS μWs and WS2 using polyvinyl alcohol (PVA) [62].

11.3.4 CdS-Based Self-Powered Photodetectors

The configuration FTO/ZnO NWs/G/CdS NPs of photoelectrochemical (PEC) type


self-powered (SP) PDs was designed by spin coating seed layer of ZnO over FTO
and growth of ZnO NWs via hydrothermal followed by the addition of graphene
dispersion and subsequent annealing at 300 °C for 30 min and attachment of CdS
NPs through SILAR method [63]. The PDs performance was analysed by electro-
chemical workstation using photoanode of polyacrylate encapsulated ZnO NWs/G/
CdS NPs, electrolyte of aqueous solution of Na2SO4, and cathode of Pt wire. As
another major breakthrough in order to design self-powered PDs, a relatively new
phenomenon namely piezophototronic effect was harnessed. The PD device based
on ZnO μWs-CdS NWs core-shell structure was fabricated by Zhang et al. [64] by
fixing tightly the two ends of a single ZnO-CdS wire by silver paste (served as S and
D electrodes) on a polystyrene (PS) substrate (flexible, and robust under repeated
mechanical strains) followed by deposition of polydimethylsiloxane (PDMS) to
package the device (Fig. 11.12a).
The above structure was modified further by the same group [65] to improve the
device performance. This time, branched ZnO-CdS double-shell NW array was
grown via a facile two-step hydrothermal method on the surface of a carbon fiber
(CF/ZnO-CdS) to fabricate PD. In typical synthesis process, CF (black) was first
coated with ITO (layer I-transparent) followed by coating of ZnO seed layer (blue)
for a better adherence and conductivity and hydrothermal growth of highly dense
11 CdS-Based Photodetectors for Visible-UV Spectral Region 265

Fig. 11.12 Schematic illustrations (a) Zno-CdS core-shell NWs (top), the proposed sandwich
model of the device, i.e. two back-to-back Schottky diodes connected to a ZnO core and CdS shell,
respectively, and simulation of the piezopotential distribution in the ZnO core under compressive
strain (bottom). (Adapted with the permission from Ref. [64]. Copyright 2012: American Chemical
Society). (b) synthesis procedure of a branched ZnO-CdS double-shell NWs array on a CF, (c, d)
fabrication of a single CF/ZnO-CdS wire based photodetector, (e) Optical microscopy image of a
typical device. (Adapted with the permission from Ref. [65]. Copyright 2013: American Chemical
Society). (f) Various steps involved in the fabrication of the hybrid heterojunctions of ACCZ along
with the optical image of the device. (Adapted with the permission from Ref. [66]. Copyright
2018: Royal Society of Chemistry)

ZnO NW array (blue) perpendicular to the surface of the CF. Finally, the CdS NW
array (yellow) was grown over ZnO NWs (Fig. 11.12b). Further, top transparent
electrode (D) of ITO (layer II) was deposited for development of a photon sensitive
device with S electrode of CF (Fig. 11.12c). Then two electrodes of a single CF/
ZnO-CdS/ITO wire were fixed on a polystyrene (PS) tightly by silver paste
(Fig. 11.12d) followed by packaging of the device by PDMS. Figure 11.12e illus-
trates microscopic image of the final device. Plasmonic Au-g-C3N4/CdS/ZnO
(ACCZ) based hybrid heterojunction structure was used to design PD on a flexible
platform [66]. The stepwise fabrication process is illustrated in Fig. 11.12f which
involves deposition of rf-sputtered ZnO TF over ITO coated substrate (glass and
PET) followed by growth of CdS TF via PLD, spin coating of Au-C3N4 TF and
deposition of top electrode Al using thermal evaporation method. SPPD based on
CdS N/μWs: Poly(3-hexylthiophene) (P3HT) μWs was fabricated by Yu et al. [67]
In typical synthesis process, CVD grown an ultralong (10 mm) CdS N/μW was
deposited on PS and fixed tightly by epoxy resin through one end followed by oxy-
gen plasma treatment of device for 5 min and drop casting of P3HT. A flexible PD
based on 2D WSe2/1D CdS heterojunction was developed by Lin et al. [68] First,
they transferred PVD grown CdS NWs on PET substrate by lift up and aligned
266 N. Saxena et al.

stamp method and then WSe2 nanosheet was synthesized via mechanical exfoliation
and precisely positioned onto the CdS NW followed by the patterning of contact
electrodes (Cr/Au (15 nm/50 nm)- Ohmic for CdS, and Pd/Au (15 nm/50 nm) for
WSe2) via electron beam lithography. In another attempt to comply with existing
silicon technology and fabricating SPPD based on piezophototronic effect, CdS
NWs were hydrothermally grown on Si micropyramids that were prepared by
chemical etching of p-Si wafers. The top and bottom electrodes (Al and ITO respec-
tively) were deposited by rf-sputtering followed by spin coating of PDMS for pack-
aging the device [69].

11.4 Performance and Figures of Merit (FOM) of CdS Based


UV-Visible Photodetectors

As discussed earlier, CdS based UV-Vis. PDs have been designed and fabricated by
many research groups. The performance and figures of merit (FOM) of these PDs
depend on various factors and they will be discussed in more detail in following
sections.

11.4.1 Self-Powered Photodetectors

Among the most successful and less power consuming PDs are self-powered photo-
detectors (SPPDs). Their parameters are listed in the Table 11.1.
These are numerous effects and phenomena to realize a SPPD. A few of them are
discussed here. High performing SPPDs were designed by perovskite/CdS hetero-
structure. The superior performance of device was driven by gradient energy band
(Fig. 11.13a). It was found that photocurrent first increases with increase in gradient
energy band and then a negative effect was noticed. The enhancement in photocur-
rent results from stronger extraction and separation efficiency of photogenerated
e–h pairs at the interface due to large built-in potential (Vbi = 0.17 V) at CdS10/
perovskite interface than that of CdSx/perovskite (0.07 V), which was estimated by
M–S measurement under dark condition. Besides, the effect of perovskite layer on
device performance was also analysed and found that the photocurrent mainly
depends on perovskite layer rather than CdS. However, the presence of CdS layer
enables to sense the broader range of spectral region starting from 350 to 800 nm
[41]. The performance parameters in this study are summarized here in Table 11.1.
A visible SPPD of SnS NFs covered CdS NRs array was designed by coupling
pyro-electricity and photo-electricity to optimize the cryogenic detecting perfor-
mance. It was found that both the pyroelectric and photoelectric effects jointly con-
tribute in the device performance and this dual effect becomes more significant with
the reduction in temperature. Figure 11.13b, c show the current-time (I-t) response
11 CdS-Based Photodetectors for Visible-UV Spectral Region 267

Table 11.1 FOMs of different SPPDs based on CdS nanostructures (the bias voltage is 0)
EQE (%)/D (10y
Material λ (nm) S (%)/R (A/W) Jones) τr/τf References
SnS/CdS 650 0.0104 3.56 × 1011 30 ms [40]
p-Si/n-CdS NWs 325– 0.00034 – 245 μs/277 μs [38]
1550
CdS NBs 484 1.54 × 106/0.036 2.36 × 1012 40 ms/30 ms [32]
ZnONAs/G/CdS/ 365; .0273 – 5 ms [63]
electro-lyte 475 .0043
n-CdS/p-Se 500 9 × 103/0.040 1.3 × 1013 2 μs/22 ms [43]
Nb:TiO2/CdS 550 0.125 – ~10 ms [58]
CdS/ZnO/ 400– ∼2 × 102/0.017 7.10 × 1011 0.14 s/0.16 s [36]
PEDOT:PSS 550 4 × 103/0.038 5.76 × 1011 0.12 s/0.16 s
Flat ∼1 × 102/0.032 2.65 × 1011 0.14 s/0.17 s
Lp = 350 nm
Lp = 750 nm
ZnO/CdS core/shell 350 2.8 × 103 – <20 ms/20 ms [70]
NRs/PEDOT:PSS 470 1.07 × 103 – <20 ms/20 ms
λ wavelength, S sensitivity (Ilight/Idark or Ilight-Idark/Idark), R responsivity, D detectivity, response time
(τr-rise time, τf -fall time)

Fig. 11.13 Schematic illustrations (a) gradient energy levels, and carrier transport at the gradient-
­O CdS/perovskite interface. (Adapted with the permission from Ref. [41]. Copyright 2019:
WILEY-VCH Verlag GmbH). (b) Time response curves under 365, 405, 532, and 650 nm light
illumination at 0 V, (c) Photodetection capability and Pyroelectric current testing of SnS/CdS
device under illumination of different light power density at temperatures 130 K, performance
parameters of the SnS/CdS heterojunction PD (d) The ratio of pyroelectric current/photocurrent
under different light power densities and temperatures, (e) Responsivity and detectivity at 130 K
and (f) Response times of pyroelectric and photoelectric effect at different temperatures. (Adapted
with the permission from Ref. [40]. Copyright 2020: WILEY-VCH Verlag GmbH)
268 N. Saxena et al.

curves for the SnS/CdS device under different light wavelengths without bias volt-
age and under illumination of different light power density at temperatures 130 K
respectively. It was found that the amplitude of pyroelectric pulse signals increases
with reduction in measurement temperature as one can see (Fig. 11.13c) the ampli-
tude of pyroelectric pulse signals is significantly high (25 nA) with respect to pho-
tocurrent (just 3.5 nA) under the light irradiation of 0.019 mW cm−2. Figure 11.13d
illustrates the reduction in pyroelectric components in total photocurrent as a func-
tion of increasing temperature, whereas the contribution of this component in
responsivity and detectivity of the sensor drastically fall with increase in power of
illuminating light (Fig. 11.13e). In contrast to these, the response time majorly
depends on the component of photoelectric effect induced photocurrent in all tem-
perature range and become more prominent at cryogenic temperatures (Fig. 11.13f)
[40]. The estimated performance parameters of the device are summarized the
table I.
Chakrabarty et al. [36] developed high performance hybrid SPPDs on soft litho-
graphically patterned organic platform and investigated the effect of line width (Lp),
periodicity (λP) and the feature height (hS) of the two grating patterned stamps used
for patterning on the device performance. The morphology of PEDOT:PSS layer
and final device (subsequent deposition of CdS, ZnO & Al) on (1) flat and (2)
Lp = 750 nm; (3) Lp = 350 nm patterned films are illustrated in Fig. 11.14a (1–3) and
11.14b (1–3), respectively. The device fabricated on Lp = 350 nm patterned films
demonstrated enhanced sensing parameters.

Fig. 11.14 AFM images (a) PEDOT:PSS films (1) flat and (2) Lp = 750 nm; (3) Lp = 350 nm pat-
terned film. Inset shows the corresponding optical diffraction patterns, (b) devices (1) flat and
patterned samples with line widths of Lp = 750 nm (2) and Lp = 350 nm (3), respectively. (Adapted
with the permission from Ref. [36]. Copyright 2018: IOP Publishing Ltd)
11 CdS-Based Photodetectors for Visible-UV Spectral Region 269

11.4.2 Photodetectors Based on Nanostructures

The parameters estimated in their study are listed in Table 11.2. The effect of sul-
phur concentration on sensing characteristic of CdS based M-S-M device is studied
by Halge and co-workers [71]. They tested the devices under the illumination of
360 nm, 550 nm and 700 nm light and observed that the photosensing parameters
first improved with reducing Cd/S ratio, attained a maxima and then reduced with
further reduction in Cd/S ratio. They achieved optimum sensing parameter
(Table 11.2) for device S2 (Cd/S = 1.24) under UV illumination. 2D microscale
position-sensitive PDs were designed using highly ordered comb-like CdS NWs
array with cone-shape branches. The position sensitivity of PDs was examined via
variable resistance in different transportation routes and variable optical responses
at different parts of the cone shape branches. They found that the gradient resistance
rises from the trunk to the branch end due to the transportation route, lengthens
along the X-direction and the photoresponse of cone shape branches varies along
the Y-direction (Fig. 11.7h). The transportation route i.e. position/distance depen-
dent resistance in trunk direction is found to decrease under the illumination of
white light with respect to dark in both the cases either the connection position 1 or
6 (Fig. 11.7e) on the trunk to different terminals of branches (Fig. 11.15a, b),
whereas position dependent resistance in branch direction i.e. along Y-axis (changes
on one single conical branch Fig. 11.7f). The branch resistance under dark gradually
decreases from bottom to top (Fig. 11.7g) and photocurrent increase linearly with
increasing distance (Fig. 11.15c). The change in resistance per nm distance (resis-
tance sensitivity) along X and Y direction was estimated ~85 K Ω and 58 Ω respec-
tively [39].
The effect of dopant Sn [5, 53, 54], Cu [74] Y [78], Ce [79], Mg, Al and Mg:Al
[55, 80] on the photo-sensing properties was studied by various researchers and it
was found that single dopant may enhance the device performance up to a certain
concentration of dopant and beyond that device performance may degrade, whereas
at that critical concentration of single dopant introduction of other dopant element
may results in degradation of device performance (Table 11.3). Lin et al. [72] devel-
oped a PD with picowatt sensitivity in UV region (larger than visible region) via
deep ultraviolet (DUV) laser treatment on the surface of CdS. They investigated the
device performance before and after irradiation with a single shot from a KrF laser
at power densities of 0.7, 14, and 140 mJ cm−2 and found that the performance of
devices was improved with increase irradiation in power density and improvement
is maximum under 365 nm illumination (Fig. 11.16a). Figure 11.16b illustrates the
detectivity of the laser-treated PD as a function of the illumination power density at
a bias voltage of 1 mV. Channel lengths (CLs) specific broad spectral photoresponse
properties were investigated by Sharma et al. [81]. They observed linear depen-
dency of rise time with CLs up to 700 nm and non-linearity on further increase in
CLs i.e. for CLs = 7 μm, 45 μm and 350 μm (Fig. 11.16c–f). Further, they were
noticed increase in response time with decrease in CLs down to nanometers
(300 nm).
270

Table 11.2 Comparison of the FOMs of different CdS nanostructure-based photodetectors


Material Bias (V) λ (nm) S (%)/R (A/W) EQE (%)/D (10y Jones) τr/τf References
nc-CdS 5 470 1.02 × 103/ 82 19.5 × 103/5.05 × 1011 183 ms/61 ms [10]
CdS NCs 1 365 7.3 × 105 3.5 × 1016 – [72]
Al:CdS-PS:p-Si −2 400 0.6 180/3.4 × 1013 160 ms/350 ms [55]
Au NPs/CdS QDs 10 390 2/1.27 × 10−4 1.42 × 109 – [73]
CdS QDs/PVA 5 365 20/9.5 × 10−9 39.8 × 10−9 – [74]
CdS:Cu QDs/PVA 24/15 × 10−9 63 × 10−9
Graphene/CdSe/CdS/ZnS QDs 5 365 2/45.77 – 1 s/1 s [75]
O-doped CdS/perovskite 700 0.48 – – [41]
CdS/ZnO core/shell NWs 4 367 0.0011 – ∼26 ms/2.1 ms [34]
468 0.0013
SnO2-CdS interlaced NWs 5 370 7417 – 1.5 s/0.6 s [49]
CdS/WS2 5 ~50 ~1012 – [56]
M-WS2/CdS/F-WS2 1 405 1.5 × 104/∼4.7 3.4 × 1012 13.7 ms/15.8 ms [62]
GaTe/Sn: CdS 2 White 100(f) & 3000(r)/607 – 260 ms/267 ms [54]
CdS/ZnO 5 405 3.8 × 103 – 3.6 s/3.3 s [76]
CdS-CdSxTe1-x-CdTe core-shell NB 5 355 4 × 103/1.5 × 103 4.7 × 105 11 μs/23 μs [35]
ZnO NR/CdS 5 350 146/12.89 – 62.4 s/44.9 s [77]
Ag/CdS/Ag 10 360 104 – 1.4 ms [71]
Ag/Y:CdS/Ag 15 532 55.07/0.83 193.8/4.28 × 1011 78 ms/87 ms [78]
p-NiO/n-CdS – UV-vis. 5/0.04043 13.73/1.4 × 1010 3.5 s [47]
CdS–CdSe core–shell nanowalls 10 405 ~100/1.2 × 103 – 250 ms/330 ms [57]
CdS core-au/MXene 4 405 0.086 1.34 × 1011 – [46]
a
f forward bias, r reverse bias
N. Saxena et al.
11 CdS-Based Photodetectors for Visible-UV Spectral Region 271

Fig. 11.15 The transportation route dependent resistance along X-direction (a) connection posi-
tion 1 on trunk to different terminals of branches, (b) connection position 6 to different terminals
2–5 of branches, and (c) The light position dependent photocurrent on a branch along the
Y-direction. (Adapted with the permission from Ref. [39]. Copyright 2018: Wiley-VCH Verlag
GmbH & Co)

Table 11.3 Parameters of PDs based on doped CdS nanostructures


Bias λ S EQE (%)/D (10y
Material (V) (nm) (%)/R (A/W) Jones) τr/τf Referenes
Sn:CdS NWs 3 405 51.2 – 270 ms/310 ms [11]
O-doped CdS/ 700 0.48 – – [41]
perovskite
Sn-doped 1D CdS 0.05 405 30/1.46 × 101 4.5 × 104 260 ms/260 ms [53]
μWs/NSs 532 20/1.03 × 101 2.4 × 103
650 8/2.66 6.2 × 102
GaTe/Sn: CdS 2 White 100(f) & – 260 ms/267 ms [54]
3000(r)/607
Al:CdS-PS:p-Si -2 400 0.6 180/3.4 × 1013 160 ms/350 ms [55]
CdS QDs/PVA 5 365 20/9.5 × 10−9 39.8 × 10−9 – [74]
CdS:Cu QDs/ 24/15 × 10−9 63 × 10−9
PVA
Ag/Y:CdS/Ag 15 532 55.1/0.83 1948/4.3 × 1011 78 ms/87 ms [78]
CdS 5 532 0.17 41.4/7.1 × 1010 840 ms/910 ms [80]
CdS:Mg(3%) 1.40 327.5/4.1 × 1011 780 ms/880 ms
CdS:Al(3%) 2.13 497.3/5.2 × 1011 750 ms/850 ms
CdS:Mg(3%) 0.31 72.5/1.4 × 1011 820 ms/890 ms
:Al(3%)
CdS:Ce (0%) – 460 560/0.0097 2.51/3.5 × 1010 800 ms/860 ms [79]
CdS:Ce (10%) 5523/0.1440 38.6/2.7 × 1011
CdS:Ce (20%) 331/0.0245 6.2/1.3 × 1011
CdS:Ce (30%) 84/0.0153 4.1/1.9 × 1010
a
a air, v vacuum

11.4.3 CdS Photodetectors Using Piezo-Phototronic Effect

A decade ago, it was found that the performance of PDs can be largely improved
with the use of piezo-phototronic effect. This is a three-way coupling effect of
piezoelectricity, semiconductor and photonic properties in piezoelectric
272 N. Saxena et al.

Fig. 11.16 (a) Responsivity enhancements at different wavelengths for laser-treated CdS devices,
normalized with respect to the responsivity of the device prepared without laser treatment. (b)
Detectivity of the laser-treated CdS device plotted with respect to the illumination power density
at a bias voltage of 1 mV. Inset: Detectivity plotted with respect to the illumination power density
at a bias voltage of 1 V. (Adapted with the permission from Ref. [72]. Copyright 2014: American
Chemical Society). Variation in the rise and decay times as a function of channel lengths (c) and
(d) for the short channels (nm) respectively and (e) and (f) for the long channels (μm) respectively.
(Adapted with the permission from Ref. [81]. Copyright 2015: AIP)

(non-central symmetric) semiconductors like CdS, ZnO etc. The basic measurement
setup for examine the piezo-phototronic effect and the device under various strain
conditions are illustrated in Fig. 11.17a–d. There are various reports utilizing this
effect to improve performance of many optical devices, here only few reports on
PDs are discussed.
Zhang et al. [65] studied piezo-phototronic effect first for ZnO-CdS core-shell
NWs [64] and subsequently for CF/ZnO-CdS core-shell NWs. Here, the various
outcome of these studies are compared. Figure 11.17e, f illustrate the I-V character-
istics of a single ZnO-CdS core-shell NW (device I) and CF/ZnO-CdS core-shell
NW (device II) based device respectively under different tensile and compressive
strains and 548 nm (1.43 mW/cm2) illumination of light. One can see that as strain
changes from compressive to tensile, photocurrent is increased continuously for
both devices. However, the enhancement in photocurrent for device II is more than
an order of 2 as compare to device I. Simultaneously, change of responsivity for
device II under compressive strain and illumination of various wavelengths is much
steeper than the device I as illustrated in Fig. 11.17g, h. It can also be noted that
change in responsivity continuously increases with reduction in strain i.e. tensile to
compressive strain. I-t responses were also studied by Yu et al. [67] for P3HT: CdS
heterojunction based device under varying strain and different illuminations along
with the various parameters that were studied in Refs. [64] and [65]. Such I-t
response curves are illustrated in Fig. 11.17i, j under different strains for +ve and –ve
11 CdS-Based Photodetectors for Visible-UV Spectral Region 273

Fig. 11.17 Schematics of (a) experimental setup for studying the piezo-phototronic effect, (b–d)
the device under various strain conditions. (Adapted with the permission from Ref. [65]. Copyright
2013: American Chemical Society). I-V characteristics of (e) device I and (f) device II under dif-
ferent tensile and compressive strains, excited green light of wavelength 548 nm and power
1.43 mW/cm2. The change of responsivity (g) under compressive strains for device I and (h) tensile
and compressive strains for device II excited by various wavelengths. (Adapted with the permis-
sion from Refs. [64] and [65]. Copyright 2012 and 2013: American Chemical Society). I-t response
of P3HT: CdS heterojunction based device under 530 nm illumination with different (i) tensile
strains and (j) compressive strain. (Adapted with the permission from Ref. [67]. Copyright 2017:
Elsevier Ltd)

piezopotentials and 530 nm illumination respectively. It is evident from Fig. 11.17i,


j, photocurrent increases as a function of increasing piezopotential and device shows
good reproducibility and repeatability under various strains. Along with the above
discussed studies Pal et al. [66] examined the I–V characteristics of the ZnO and
ACCZ based flexible device under normal and concave bending conditions at an
angle of ∼32 ± 2° (Fig. 11.18a and inset of 11.18a). The enhancement in the current
due to the development of a compressive strain induced piezopotential for both
devices can be clearly seen. The enhancement in current is prominent in ACCZ
based device than ZnO based device. Figure 11.18b illustrates the relative change in
current between the bending and relaxed conditions at a constant bias of 2 V for
ACCZ device. It is obvious that for each cycle of strain, the current initially increases
sharply and becomes stable after reaching a peak value, which indicates good repro-
ducibility and repeatability of device under various cycles of strain. In another study
on p-Si/n-CdS heterojunction PD under 650 nm light illumination, Zhao et al. [69]
274 N. Saxena et al.

Fig. 11.18 (a) I–V characteristics of the bare ZnO TF under normal and concave bending condi-
tions. The inset shows the same for the ACCZ and (b) Current modulation under repetitive applica-
tion of bending cycles at 2 V. (Adapted with the permission from Ref. [66]. Copyright 2018: Royal
Society of Chemistry). (c) Response time p-Si/n-CdS heterojunction PD under 442 nm (10 mW/
cm2) light illumination under different strains. (Adapted with the permission from Ref. [69].
Copyright 2019: The Royal Society of Chemistry)

Fig. 11.19 Variation of (a) responsivity, (b) photosensitivity, (c) external quantum efficiency, (d)
specific detectivity, and (e) response time with increase in ion irradiation influence. (Adapted with
the permission from Ref. [10]. Copyright 2018: Elsevier B.V)

examined the effect of strain on response time of device as illustrated in Fig. 11.18c.
It can be noticed that device becomes slower with increase in strain i.e. when stain
changes from compressive to tensile.
An another tool for enhancing the PD performance was explored by Kumar et al.
[4] by using ion beam modified CdS TFs for the fabrication of PD. Ion beam irradia-
tion is a powerful technique to engineer the desired defects by multiple phenomena
occurring simultaneously in the material. They investigated the photosensing char-
acteristics of ion beam irradiated CdS TFs as a function of Ni6+ ion influence and
found that up to a certain ion influence (1 × 1013 ion/cm2) device performance was
improved and then degraded for higher influences under illumination of both UV
and visible light. Figure 11.19a–e shows the variation of different sensing parame-
ters as a function of ion influence under illumination of various light [10].
11 CdS-Based Photodetectors for Visible-UV Spectral Region 275

11.5 Summary

In summary, CdS is enormously used as an active layer for photodetection in


UV-Visible spectral region as well as in low frequency range. This chapter dis-
cussed CdS based PDs with various architectures, structures and working on differ-
ent mechanisms. Various types of nanostructures, thin films, heterostructures,
composites etc. have been utilized to fabricate high performing PDs. Besides,
numerous approaches like doping with different elements, decoration with suitable
materials, and ion beam irradiation etc. were proven to be fruitful in enhancing the
performance and range of PDs fabricated. Moreover, self-powered PDs also have
been realized using multifarious effects like piezoelectric, piezophototronic, pyro-
electric, etc. in CdS and other materials used in architectures. In nut-shell, CdS is an
integral part of UV-Visible PDs fabricated so far with sufficiently high FOM.

Acknowledgments The author TK is thankful to the University Grants Commission (UGC),


New Delhi, India, for providing scholarship under NF-OBC scheme (NFO-2018-19-OBC-
JAM-69666).

References

1. Zhuge F, Zheng Z, Luo P, Lv L, Huang Y, Li H, Zhai T (2017) Nanostructured materials and


architectures for advanced infrared photodetection. Adv Mater Technol 2(8):1700005
2. Song K, Ma N, Mishra YK, Adelung R, Yang Y (2019) Achieving light-induced ultrahigh
pyroelectric charge density toward self-powered UV light detection. Adv Electron Mater
5(1):1800413
3. Liu K, Sakurai M, Aono M (2010) ZnO-based ultraviolet photodetectors. Sensors (Basel)
10(9):8604–8634
4. Masouleh FF, Das N (2014) Application of metal-semiconductor-metal photodetector in high-­
speed optical communication systems. In: Das N (ed) Advances in optical communication.
Intech, pp 87–114
5. Kim H, Kumar MD, Kim J (2015) Highly-performing Ni/SiO2/Si MIS photodetector for NIR
detecting applications. Sensors Actuators A Phys 233:290–294
6. https://www.electrical4u.com/p-­i-­n-­photodiode-­avalanche-­photo-­diode/. PIN photodiode.
Accessed 08 Jan 2022
7. Gao XD, Fei GT, Xu SH, Zhong BN, Ouyang HM, Li XH, Zhang LD (2019) Porous Ag/
TiO2-Schottky-diode based plasmonic hot-electron photodetector with high detectivity and
fast response. Nano 8(7):1247–1254
8. Park S, Kim SJ, Nam JH, Pitner G, Lee TH, Ayzner AL, Wang H, Fong SW, Vosgueritchian
M, Park YJ, Brongersma ML, Bao Z (2015) Significant enhancement of infrared photodetector
sensitivity using a semiconducting single-walled carbon nanotube/C60 phototransistor. Adv
Mater 27(4):759–765
9. Kumar P, Saxena N, Dewan S, Singh F, Gupta V (2016) Giant UV-sensitivity of ion beam
irradiated nanocrystalline CdS thin films. RSC Adv 6(5):3642–3649
10. Kumar P, Saxena N, Singh F, Gupta V (2018) Ion beam assisted fortification of photoconduc-
tion and photosensitivity. Sensors Actuators A Phys 279:343–350
11. Guo S, Wang L, Ding C, Li J, Chai K, Li W, Xin Y, Zou B, Liu R (2020) Tunable optical loss
and multi-band photodetection based on tin doped CdS nanowire. J Alloys Compd 835:155330
276 N. Saxena et al.

12. Bao Q, Li W, Xu P, Zhang M, Dai D, Wang P, Guo X, Tong L (2020) On-chip single-mode CdS
nanowire laser. Light Sci Appl 9:42
13. Wen Z, Liu P, Ma J, Jia S, Xiao X, Ding S, Tang H, Yang H, Zhang C, Qu X, Xu B, Wang
K, Teo KL, Sun XW (2021) High-performance ultrapure green CdSe/CdS core/crown nano-
platelet light-emitting diodes by suppressing nonradiative energy transfer. Adv Electron Mater
7(7):2000965
14. Kumar P, Saxena N, Chandra R, Gao K, Zhou S, Agarwal A, Singh F, Gupta V, Kanjilal D
(2014) SHI induced enhancement in green emission from nanocrystalline CdS thin films for
photonic applications. J Lumin 147:184–189
15. Liu B, Guo J, Hao R, Wang L, Gu K, Sun S, Aierken A (2020) Effect of Na doping on the per-
formance and the band alignment of CZTS/CdS thin film solar cell. Sol Energy 201:219–226
16. Saxena N, Kumar P, Gupta V, Kanjilal D (2018) Radiation stability of CBD grown nanocrys-
talline CdS films against ion beam irradiation for solar cell applications. J Mater Sci Mater
29(13):11013–11019
17. Li J-Y, Li Y-H, Qi M-Y, Lin Q, Tang Z-R, Xu Y-J (2020) Selective organic transformations over
cadmium sulfide-based photocatalysts. ACS Catal 10(11):6262–6280
18. Xie T, Zhong X, Liu Z, Xie C (2020) Silica-anchored cadmium sulfide nanocrystals for the
optical detection of copper(II). Mikrochim Acta 187(6):323
19. Faraz M, Abbasi A, Naqvi FK, Khare N, Prasad R, Barman I, Pandey R (2018) Polyindole/
cadmium sulphide nanocomposite based turn-on, multi-ion fluorescence sensor for detection
of Cr3+, Fe3+ and Sn2+ ions. Sensors Actuators B Chem 269:195–202
20. Saxena N, Kumar P, Gupta V (2019) CdS nanodroplets over silica microballs for efficient
room-temperature LPG detection. Nanoscale Adv 1(6):2382–2391
21. Wang H, Ma J, Zhang J, Feng Y, Vijjapu MT, Yuvaraja S et al (2021) Gas sensing materials
roadmap. J Phys Condens Matter 33(30):303001
22. Harish R, Nisha KD, Prabakaran S, Sridevi B, Harish S, Navaneethan M, Ponnusamy S,
Hayakawa Y, Vinniee C, Ganesh MR (2020) Cytotoxicity assessment of chitosan coated CdS
nanoparticles for bio-imaging applications. Appl Surf Sci 499:143817
23. Kim H-R, Bong J-H, Jung J, Sung JS, Kang M-J, Park J-G, Pyun J-C (2020) An on-chip che-
miluminescent immunoassay for bacterial detection using in situ-synthesized cadmium sulfide
nanowires with passivation layers. Biochip J 14(3):268–278
24. Saxena N, Kumar P, Gupta V (2015) CdS : SiO2 nanocomposite as a luminescence-based wide
range temperature sensor. RSC Adv 5(90):73545–73551
25. Miao S, Cho Y (2021) Toward green optoelectronics: environmental-friendly colloidal quan-
tum dots photodetectors. Front Energy Res 9:1–18
26. Xu K, Zhou W, Ning Z (2020) Integrated structure and device engineering for high perfor-
mance and scalable quantum dot infrared photodetectors. Small 16(47):e2003397
27. Paschotta DR (2008) Metal–semiconductor–metal photodetectors. Laser Physics and
Technology, Wiley-VCH
28. https://www.inup.cense.iisc.ac.in/msm-­photodetectors. Fabrication Of ZnO Based MSM
Photodetectors, 2021. https://www.inup.cense.iisc.ac.in/msm-­photodetectors. 16 Oct 2021
29. Kalsi T, Kumar P (2021) Cd1-xMgxS CQD thin films for high performance and highly selec-
tive NIR photodetection. Dalton Trans 50(36):12708–12715
30. An Q, Meng X (2016) Aligned arrays of CdS nanotubes for high-performance fully nano-
structured photodetector with higher photosensitivity. J Mater Sci Mater 27(11):11952–11960
31. Jin B, Huang P, Zhang Q, Zhou X, Zhang X, Li L, Su J, Li H, Zhai T (2018) Self-limited
epitaxial growth of ultrathin nonlayered CdS flakes for high-performance photodetectors. Adv
Funct Mater 28(20):1800181
32. Nawaz MZ, Xu L, Zhou X, Shah KH, Wang J, Wu B, Wang C (2021) CdS nanobelt-based
self-powered flexible photodetectors with high photosensitivity. Mater Adv 2(18):6031–6038
33. Munde S, Shinde N, Khanzode P, Budrukkar M, Lahane P, Dadge J, Jejurikar S, Mahabole
M, Khairnar R, Bogle K (2018) Nano-crystalline CdS thick films: a highly sensitive photo-­
detector. Mater Res Exp 5:066203
11 CdS-Based Photodetectors for Visible-UV Spectral Region 277

34. Yang Z, Guo L, Zu B, Guo Y, Xu T, Dou X (2014) CdS/ZnO core/shell nanowire-built films
for enhanced photodetecting and optoelectronic gas-sensing applications. Adv Opt Mater
2(8):738–745
35. Tang M, Xu P, Wen Z, Chen X, Pang C, Xu X, Meng C, Liu X, Tian H, Raghavan N, Yang
Q (2018) Fast response CdS-CdSxTe1−x-CdTe core-shell nanobelt photodetector. Sci Bull
63(17):1118–1124
36. Chakrabarty P, Gogurla N, Bhandaru N, Ray SK, Mukherjee R (2018) Enhanced performance
of hybrid self-biased heterojunction photodetector on soft-lithographically patterned organic
platform. Nanotechnology 29(50):505301
37. Roy S, Bhandaru N, Das R, Harikrishnan G, Mukherjee R (2014) Thermally tailored gradi-
ent topography surface on elastomeric thin films. ACS Appl Mater Interfaces 6(9):6579–6588
38. Dai Y, Wang X, Peng W, Xu C, Wu C, Dong K, Liu R, Wang ZL (2018) Self-powered Si/CdS
flexible photodetector with broadband response from 325 to 1550 nm based on pyro-­phototronic
effect: an approach for photosensing below bandgap energy. Adv Mater 30(9):1705893
39. Hao Y, Guo S, Weller D, Zhang M, Ding C, Chai K, Xie L, Liu R (2018) Position-sensitive
array photodetector based on comb-like CdS nanostructure with cone-shape branches. Adv
Funct Mater 29(1):1805967
40. Chang Y, Wang J, Wu F, Tian W, Zhai W (2020) Structural design and pyroelectric property
of SnS/CdS heterojunctions contrived for low-temperature visible photodetectors. Adv Funct
Mater 30(23):2001450
41. Cao F, Meng L, Wang M, Tian W, Li L (2019) Gradient energy band driven high-performance
self-powered perovskite/CdS photodetector. Adv Mater 31(12):e1806725
42. Ge Y, Zhang M, Wang L, Meng L, Tang J, Chen Y, Wang L, Zhong H (2019) Polarization-­
sensitive ultraviolet detection from oriented-CdSe@CdS-dot-in-rods-integrated silicon photo-
detector. Adv Opt Mater 7(18):1900330
43. Wang J, Chang Y, Huang L, Jin K, Tian W (2018) Designing CdS/Se heterojunction as high-­
performance self-powered UV-visible broadband photodetector. APL Mater 6(7):076106
44. Gozeh BA, Karabulut A, Yildiz A, Dere A, Arif B, Yakuphanoglu F (2019) SILAR controlled
CdS nanoparticles sensitized CdO diode based photodetectors. SILICON 12(7):1673–1681
45. Li J, Zhu Y, Li M, Cai H, Ding H, Pan N, Wang X (2018) One-step fabrication of CdS nano-
flake arrays and its application for photodetector. Optik 169:190–195
46. Jiang T, Huang Y, Meng X (2020) CdS core-Au/MXene-based photodetectors: positive deep-
­UV photoresponse and negative UV–vis-NIR photoresponse. Appl Surf Sci 513:145813
47. Reddy KCS, Selamneni V, Rao MGS, Meza-Arroyo J, Sahatiya P, Ramirez-Bon R (2021) All
solution processed flexible p-NiO/n-CdS rectifying junction: applications towards broadband
photodetector and human breath monitoring. Appl Surf Sci 568:150944
48. Lu MY, Chang YT, Chen HJ (2018) Efficient self-driven photodetectors featuring a mixed-­
dimensional van der Waals heterojunction formed from a CdS nanowire and a MoTe2 flake.
Small 14(40):e1802302
49. Li L, Lou Z, Chen H, Shi R, Shen G (2019) Stretchable SnO2-CdS interlaced-nanowire film
ultraviolet photodetectors. Sci China Mater 62(8):1139–1150
50. Pei Y, Pei R, Liang X, Wang Y, Liu L, Chen H, Liang J (2016) CdS-nanowires flexible photo-­
detector with Ag-nanowires electrode based on non-transfer process. Sci Rep 6:21551
51. Moger SN, Mahesha MG (2020) Colour tunable co-evaporated CdSxSe1-x (0 ≤ x ≤ 1) ter-
nary chalcogenide thin films for photodetector applications. Mater Sci Semicond Process
120:105288
52. Moger SN, Mahesha MG (2021) Investigation on spectroscopic and electrical properties of
p-Si/CdSxSe1−x (0≤ x ≤1) heterostructures for photodetector applications. J Alloys Compd
870:159479
53. Zhou W, Peng Y, Yin Y, Zhou Y, Zhang Y, Tang D (2014) Broad spectral response photodetector
based on individual tin-doped CdS nanowire. AIP Adv 4(12):123005
278 N. Saxena et al.

54. Zhou W, Zhou Y, Peng Y, Zhang Y, Yin Y, Tang D (2014) Ultrahigh sensitivity and gain
white light photodetector based on GaTe/Sn:CdS nanoflake/nanowire heterostructures.
Nanotechnology 25(44):445202
55. Das M, Sarmah S, Sarkar D (2019) Photo sensing property of nanostructured CdS-porous sili-
con (PS):p-Si based MSM hetero-structure. J Mater Sci Mater 30(12):11239–11249
56. Gong Y, Zhang X, Yang T, Huang W, Liu H, Liu H, Zheng B, Li D, Zhu X, Hu W, Pan A (2019)
Vapor growth of CdS nanowires/WS2 nanosheet heterostructures with sensitive photodetec-
tions. Nanotechnology 30(34):345603
57. Xu J, Rechav K, Popovitz-Biro R, Nevo I, Feldman Y, Joselevich E (2018) High-gain 200 ns
photodetectors from self-aligned CdS-CdSe core-shell nanowalls. Adv Mater 30(20):e1800413
58. Wang D, Chen H, Min Y, Liang J, Pan F (2020) Tunable p- and n-type Nb:TiO2 and per-
formance optimizing of self-powered Nb:TiO2/CdS photodetectors. Semicond Sci Technol
35(7):075015
59. Peng M, Wu F, Wang Z, Wang P, Gong F, Long M, Chen C, Dai J, Hu W (2019) Enhancement-­
mode CdS nanobelts field effect transistors and phototransistors with HfO2 passivation. Appl
Phys Lett 114(11):111103
60. Zheng D, Fang H, Wang P, Luo W, Gong F, Ho JC, Chen X, Lu W, Liao L, Wang J, Hu W
(2016) High-performance ferroelectric polymer side-gated CdS nanowire ultraviolet photode-
tectors. Adv Funct Mater 26(42):7690–7696
61. Zhao W, Liu L, Xu M, Wang X, Zhang T, Wang Y, Zhang Z, Qin S, Liu Z (2017) Single CdS
nanorod for high responsivity UV-visible photodetector. Adv Opt Mater 5(12):1700159
62. Zhou Y, Zhang L, Gao W, Yang M, Lu J, Zheng Z, Zhao Y, Yao J, Li J (2021) A reasonably
designed 2D WS2 and CdS microwire heterojunction for high performance photoresponse.
Nanoscale 13(11):5660–5669
63. Huang G, Zhang P, Bai Z (2019) Self-powered UV–visible photodetectors based on ZnO/
graphene/CdS/electrolyte heterojunctions. J Alloys Compd 776:346–352
64. Zhang F, Ding Y, Zhang Y, Zhang X, Wang ZL (2012) Piezo-phototronic E ff ect enhanced
visible and ultraviolet photodetection using a ZnO-CdS core-shell micro/nanowire. ACS Nano
6(10):9229–9236
65. Zhang F, Niu S, Guo W, Zhu G, Liu Y, Zhang X, Wang ZL (2013) Piezo-phototronic E ffect
enhanced visible/UV photodetector of a carbon-fiber/ZnO-CdS double-shell microwire. ACS
Nano 7(5):4537–4544
66. Pal S, Bayan S, Ray SK (2018) Piezo-phototronic mediated enhanced photodetection char-
acteristics of plasmonic Au-g-C3N4 /CdS/ZnO based hybrid heterojunctions on a fl exible
platform. Nanoscale 10:19203–19211
67. Yu X-X, Yin H, Li H-X, Zhang W, Zhao H, Li C, Zhu M-Q (2017) Piezo-phototronic effect
modulated self-powered UV/visible/near-infrared photodetectors based on CdS:P3HT
microwires. Nano Energy 34:155–163
68. Lin P, Zhu L, Li D, Xu L, Wang ZL (2018) Tunable WSe2-CdS mixed-dimensional van der
Waals heterojunction with a piezo-phototronic effect for an enhanced flexible photodetector.
Nanoscale 10(30):14472–14479
69. Zhao ZH, Dai Y (2019) Piezo-phototronic effect-modulated carrier transport behavior in dif-
ferent regions of a Si/CdS heterojunction photodetector under a Vis-NIR waveband. Phys
Chem Chem Phys 21(18):9574–9580
70. Sarkar S, Basak D (2015) Self powered highly enhanced dual wavelength ZnO@CdS
core-shell nanorod arrays photodetector: an intelligent pair. ACS Appl Mater Interfaces
7(30):16322–16329
71. Halge DI, Narwade VN, Khanzode PM, Begum S, Banerjee I, Dadge JW, Kovac J, Rana AS,
Bogle KA (2021) Development of highly sensitive and ultra-fast visible-light photodetector
using nano-CdS thin film. Appl Phys A Mater Sci Process 127(6):446
72. Lin KT, Chen HL, Lai YS, Liu YL, Tseng YC, Lin CH (2014) Nanocrystallized CdS beneath
the surface of a photoconductor for detection of UV light with picowatt sensitivity. ACS Appl
Mater Interfaces 6(22):19866–19875
11 CdS-Based Photodetectors for Visible-UV Spectral Region 279

73. Kan H, Liu S, Xie B, Zhang B, Jiang S (2017) The effect of Au nanocrystals applied in CdS
colloidal quantum dots ultraviolet photodetectors. J Mater Sci Mater 28(13):9782–9787
74. Kakati J, Datta P (2015) Schottky junction UV photodetector based on CdS and visible photo-
detector based on CdS:Cu quantum dots. Optik 126(18):1656–1661
75. Al-Alwani AJK, Chumakov AS, Shinkarenko OA, Gorbachev IA, Pozharov MV, Venig S,
Glukhovskoy EG (2017) Formation and optoelectronic properties of graphene sheets with
CdSe/CdS/ZnS quantum dots monolayer formed by Langmuir-Schaefer hybrid method. Appl
Surf Sci 424:222–227
76. Zhang C, Tian W, Xu Z, Wang X, Liu J, Li SL, Tang DM, Liu D, Liao M, Bando Y, Golberg
D (2014) Photosensing performance of branched CdS/ZnO heterostructures as revealed by in
situ TEM and photodetector tests. Nanoscale 6(14):8084–8090
77. Lam KT, Hsiao YJ, Ji LW, Fang TH, Hsiao KH, Chu TT (2017) High-sensitive ultraviolet
photodetectors based on ZnO nanorods/CdS heterostructures. Nanoscale Res Lett 12(1):31
78. Shkir M, Khan ZR, Chandekar KV, Alshahrani T, Ashraf IM, Khan A, Marnadu R, Zargar RA,
Mohanraj P, Revathy MS, Manthrammel MA, Sayed MA, Ali HE, Yahia IS, Yousef ES, Algarni
H, AlFaify S, Sanaa MF (2021) Facile fabrication of Ag/Y:CdS/Ag thin films-based photode-
tectors with enhanced photodetection performance. Sensors Actuators A Phys 331:112890
79. Ibrahim IM, Safi AA, Al-Hardan NHM (2019) Enhancement the sensitivity of CdS nano struc-
ture by adding of rare earth materials. J Phys Conf Ser 1178:012013
80. Kumar KDA, Mele P, Golovynskyi S, Khan A, El-Toni AM, Ansari AA, Gupta RK, Ghaithan
H, AlFaify S, Murahari P (2022) Insight into Al doping effect on photodetector performance
of CdS and CdS:Mg films prepared by self-controlled nebulizer spray technique. J Alloys
Compd 892:160801
81. Sharma A, Kaur M, Bhattacharyya B, Karuppiah S, Singh SP, Senguttuvan TD, Husale S
(2015) Channel length specific broadspectral photosensitivity of robust chemically grown CdS
photodetector. AIP Adv 5(4):047116
Chapter 12
ZnTe-Based Photodetectors for Visible-UV
Spectral Region

Jiajia Ning

12.1 Introduction

In the past few decades, the photoelectronic industry has been well developed to
change the life in the world. Various photoelectronic devices and structures were
developed and utilized, one of them, possessing the ability of transformation from
light to electrical signals, such as photodetectors, has attracted scientists’ intensive
attention based on their potential for light detection. Benefiting from the develop-
ment of semiconductor industry, the current detector based on the different semi-
conductors can detect the light from ultraviolet (UV: 10–400 nm) to the visible (vis:
400–700 nm) to the near-infrared (NIR: 700–1000 nm) to the terahertz (0.1–10
THz), and transfer them to electrical signals [1]. Generally, the whole photodetec-
tion system mainly contains the two parts, the photodetectors and signal processing
chips. The photodetector part is built on the basis of semiconductors [2].
Due to the different band gap in semiconductors, the photodetector can absorb
the light in the different range and excite the electron to conduction band, finally, the
excitons are transferred to electronic signals. The range of light detected is depended
by the ban gap in semiconductors. So various semiconductors are used in photode-
tectors to cover the full range of light (UV to THz). In the whole range of light, the
UV and visible range are more important and closer to our life, which can be used
in flame sensing, ozone sensing, convert communications, air and water purifica-
tion, environmental monitoring, video imaging, materials identification [3–11].
The band gap is an important parameter in photodetectors, depending the
detected range of light. The general way to change the band gap is using semicon-
ductors with different composition. To one semiconductor, the band gap is difficult

J. Ning (*)
Key Laboratory of Physics and Technology for Advanced Batteries, Ministry of Education,
College of Physics, Jilin University, Changchun, China
e-mail: jiajianing@jlu.edu.cn

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 281
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_12
282 J. Ning

to change. In 1980s, the quantum size confinement effect was found in the semicon-
ductors [12–14], the band gap of semiconductors can be tuned via changing the size
of semiconductors [15, 16], which provides a simple way to turn the band gap in
semiconductors. The photodetectors with one semiconductor, covering a broad
region of light, is possible on the basis of quantum size confinement effect. Such as
the full visible region can be detected just used CdSe nanocrystals with the different
size [17].
Based on the intrinsic band gap in semiconductors and the quantum size confine-
ment effect, the developed photodetectors based on semiconductors can cover the
region from violet to infrared range [2, 9]. Among the whole detected region of
light, the photodetectors for visible region the most widely used in our life. The
commercial reason of the photodetectors for the visible region promotes the inves-
tigation and development on visible region materials for photodetectors. Up to now,
CdSe [18], CdS [19], CdTe [20], InP [21], Cd3P2 [22], and Si [23] based photodetec-
tors were developed for the detection of visible region light. However, the toxic
element of Cd limits the further applications of cadmium chalcogenide-based pho-
todetector in our life. The Cd-free semiconductors-based photodetectors were
attracted much attention for their potential in applications. As an important Cd-free
semiconductor, ZnTe is an important candidate for visible region detectors. Herein,
we review the development of ZnTe based photodetector for visible region from
controlled synthesis, optical characterizations and the design of photodetectors.

12.2 Optical Properties of ZnTe

Zinc chalcogenide are the important members in II-VI semiconductors, exhibiting


the widely band gap of 3.8 eV for ZnS, 2.8 eV for ZnSe and 2.2 eV for ZnTe [24].
The widely band gap in ZnS and ZnSe makes them to be used in ultraviolet to deep
blue region [25]. Compared with ZnS and ZnSe, the smaller band gap of 2.2 eV in
ZnTe can extend the optical properties to visible light. The absorption peak of ZnTe
nanocrystals (NCs) shifts to 430 nm with the smaller diameter based on the quan-
tum size confinement effect [26].
Due to the band gap of 2.2 eV in ZnTe, the fluorescence emission for ZnTe NCs
is limited 563 nm. In order to extend the fluorescence emission of ZnTe NCs to the
whole visible range, the alloyed ZnTe with Se to form ZnSeTe [27, 28] or the forma-
tion of type-II heterostructures [29, 30] with another compound are the available
methods. The band gap in alloyed semiconductor can be tuned to smaller than each
of them due to the bowing effect [31]. Alloyed ZnSeTe NCs showed the fluores-
cence emission at 540 nm with the size of 4.3 nm [32].
The fluorescence emission from ZnTe-based NCs was also widely tuned for the
type-II heterostructures, such as type-II core/shell NCs, combining the charge in the
core and shell for recombination [33]. The red shift and larger stokes-shift are
observed in type-II core/shell NCs. On the basis of band gap alignment between
12 ZnTe-Based Photodetectors for Visible-UV Spectral Region 283

Fig. 12.1 (a) the schematic illustration for ZnTe/ZnSe core/shell NCs, (b) the optical properties
for ZnTe/ZnSe core/shell NCs, (c) the absorption spectra of ZnTe/ZnSe core/shell NCs with differ-
ent thickness of ZnSe shell, (d) the photoluminescence emission spectra of ZnTe/ZnSe core/shell
NCs with different thickness of ZnSe shell. (Reproduced with permission from Ref. [34]. Copyright
2010: American Chemical Society)

ZnTe and shell materials, ZnTe/ZnSe core/shell NCs are one of typical type-II het-
erostructures (Fig. 12.1a), coupled with the growth of ZnSe on ZnTe core NCs, an
obvious red shift can be observed in absorption spectra and photoluminescence
emission spectra (Fig. 12.1c, d). The emission can be tuned from blue to orange
region via changing the thickness of ZnSe shell (Fig. 12.1b) [34].
In semiconductor NCs, the optical properties and energy-level structure are also
influenced by the shape and crystal structure of NCs [35]. 1D nanorods were devel-
oped for the special polarized photoluminescence emission [36], which have been
widely used in display, solar energy conversion and optoelectronic devices [37–40].
Based on the 1D ZnTe nanorods, ZnSe dots were further grown on the two tips of
nanorods to form nanodumbbells structure [41]. The formed dumbbells structured
of ZnTe/ZnSe NCs exhibited the optical characters for type-II heterostructures, the
fluorescence emission can tunned between 500 nm to 580 nm via changing the size
of ZnSe dots on the tips of nanorods, and the PLQY in ZnTe/ZnSe nanodumbbells
can reach 40%.
284 J. Ning

12.3 Thin Film ZnTe Based Photodetectors

12.3.1 ZnTe Thin Film with Vacuum Evaporation Method

ZnTe has been attracted much attention for optoelectronic device. Thin film is the
most widely developed for device in industry. The thin film of ZnTe was early pro-
duced by the vacuum evaporation method [42]. ZnTe was deposited on glass or sili-
con substrate, the type of substrate, the deposit temperature and the following
annealing procedure showed the important influence on the properties of ZnTe film.
Gowrish et al. [43] found the photoconductivity of ZnTe thin film is related to the
substrate temperature and post deposition annealing. The ZnTe film, deposited at
elevated temperature, showed faster and improved photoresponse, and the post-­
deposition annealing was found to further enhance the photoresponse of the films.
More than the substrate and the deposition temperature of thin film, the thickness
is another important parameter for thin film structure. Thin films of ZnTe with vary-
ing compositions and thickness are formed on glass substrate at different tempera-
ture [44]. The dark and illuminated conditions photoconductivity exhibits a function
of wavelength of incident radiation, and the maximum photocurrent was obtained at
about thickness of 500 nm irrespective of composition.
In order to modified the ZnTe thin films, the impurities were induced. The impu-
rities of BaF2 and PbCl2 were found to have little influence to the electrical proper-
ties of ZnTe thin films [42]. Furtherly, hydrogen was induced to ZnTe thin film to
passivate the surface of ZnTe film [45]. After the hydrogen passivation of ZnTe film
via hydrogen plasma method, the initial PL emission peaks of 2.06 eV, 1.47 eV,
1.33 eV and 1.06 eV became one strong band edge green emission at 2.37 eV. The
hydrogen passivated ZnTe film also showed an improved photoconductivity at
higher temperature. The investigation on the mechanism of hydrogen treat showed
the passivated deep defect energy level of ZnTe.

12.3.2 ZnTe Thin Film with Wet Chemical Method

Beside the vacuum evaporation method, the wet chemical methods were also devel-
oped to prepare ZnTe thin film. Hamdi and his co-workers [46] developed the modi-
fied Bridgman solvent method to produce ZnTe film at 1980, which showed the
responsivity of 5 × 10−2A/W with the light of 610 nm. At 2000, Bozzini et al. [47]
further developed the electrodeposition method to synthesize ZnTe thin film
(Fig. 12.2). The crystalline structure of the deposited ZnTe film occur the cubic
zinc-blende structure under suitable electrochemical conditions.
12 ZnTe-Based Photodetectors for Visible-UV Spectral Region 285

Fig. 12.2 SEM micrograph of an electrodeposit from the ZnTe-bath on Cu, deposited at 0 mV
(Ag/AgCl) (fibrous appearance-left) and at −200 mV(Ag/AgCl) (rosette appearance-right).
(Reproduced with permission from Ref. [47]. Copyright 2000: Elsevier)

12.4 ZnTe Nanostructures Based Photodetectors

Because of the difficulty in the preparation of dot-shape ZnTe NCs-based photode-


tectors, 1D or two-dimensional (2D) nanostructures are widely used for ZnTe based
photodetectors. As the above discussion, ZnTe can be used for photodetectors
between ultra-violet (UV) and green region. 1D nanowires and 2D nanosheets are
produced by physical vapour deposition (PVD), chemical vapour deposition (CVD)
and metal-organic chemical vapour deposition (MOCVD) method for scale-up
[48–51].

12.4.1 One Dimensional ZnTe Based Photodetectors

1D nanostructures have been studied extensively for their novel properties and
potential applications in nanoscale devices, such as ultra-sensitive nanosensors and
photodetectors [52, 53]. The single ZnTe nanowire-based photodetectors was
286 J. Ning

developed in 2011 [51]. The untreated ZnTe nanowire produced by PVD method
exhibited the diameter of 60–400 nm with cubic zinc-blende crystal structure. One
single ZnTe nanowire with length of 88 μm and width of 330 nm was selected to
build the photodetectors on the SiO2/Si substrate. Ti/Au (2 nm/60 nm) interdigitated
electrodes with 3 μm separation was deposited on the ZnTe nanowires. Compared
with dark condition, this ZnTe nanowire-based photodetector showed 2 times of
photocurrent under the light of 500 nm.
In order to improve the performance of ZnTe nanowire-based photodetectors, the
surface of ZnTe nanowires can be modified with other ligand or compound to pas-
sivate the defects [54, 55]. Growth of the inorganic compound as the shell is the
most efficient way to passivate the surface. ZnO shell was grown on ZnTe nanow-
ires to form ZnTe/ZnO core/shell nanowires [56]. Starting from the initial ZnTe
nanowires, the thickness of ZnO shell can be tuned to increase the ratio of shell for
19% in core/shell nanowires. The performance of ZnTe nanowires-based photode-
tectors greatly improved by increasing the thickness of ZnO shell. The Ipn/Idark,
responsivity, and the photoconductive gain can be improved from 1.95 (A W−1) and
8 × 102% to 199 (A W−1) and 8.12 × 104% with increasing the thickness of ZnO
shell, respectively.
A method to modify the intrinsic properties in semiconductor is inducing the
impurities with controlling amount. The present impurities can increase the density
of charge in semiconductor to form n-type or p-type semiconductor based on the
different type of impurities [57, 58]. During the formation process of ZnTe struc-
tures, Cu, P, Sb and N can act as induced impurities in ZnTe [59–61]. The doped
ZnTe can reveal the enhanced hole mobility or electron mobility. The performance
of 1D ZnTe nanowires in optoelectronic devices can be promoted via the pres-
ent impurities, such as Sb dope ZnTe nanowires and Cu doped ZnTe nanowires.
Ga doped ZnTe nanowires were developed via the simple thermal evaporation
method [62]. The Ga content in the ZnTe nanowires can be tuned from 1.3 to 8.7%,
and the hole mobility and hole concentration will increase from 0.0069 to
0.46 cm2V−1S−1, respectively. The ZnTe:Ga nanowires based photodetectors also
show the high sensitivity to visible light illumination, the responsivity and detectiv-
ity were estimated to be 4.17 × 103 AW−1 and 3.19 × 1013 cmHz1/2 W−1, higher than
other undoped ZnTe nanostructures-based photodetectors.

12.4.2 Two Dimensional ZnTe Based Photodetectors

More than 1D nanostructures, 2D nanostructures are another important candidate


for optoelectronic device. As the most famous 2D nanostructures, graphene has
been widely investigated and used in optoelectronic devices [63–65]. Because of the
quantum size confinement effect in 2D nanostructures, the electron can move in
another two dimensions. Generally, 2D nanostructures exhibit the thickness depen-
dent optical and optoelectronic properties [66, 67].
12 ZnTe-Based Photodetectors for Visible-UV Spectral Region 287

Nanosheet is one type of 2D nanostructures. Wang and his co-workers [68] syn-
thesized ZnTe nanosheet with the cubic zinc-blende structure via PVD method
(Fig. 12.3a, b). The ZnTe nanosheets with single crystal structure had the thickness
of 65 nm and the length of micrometers (Fig. 12.3c), showing the emission band
peaked at 550 nm near-bandgap emission of ZnTe. Figure 12.3d shows the Ids–Vds
curves of the device under various incident power density, which indicating a good
photo conducting behavior. The obtained net photocurrent increases as the incident
power density increased, indicating more electrons and holes are generated under
stronger light illumination condition (Fig. 12.3e). Larger bias voltage can further
enhance the charge transport process. Furthermore, such ZnTe nanosheets based
photodetector shows good reversible switching properties under different bias volt-
ages (Fig. 12.3g), indicating the good stability of the ZnTe devices. The produced

Fig. 12.3 (a) Schematic diagram of PVD system, (b) XRD pattern of ZnTe nanosheets, (c) SEM
image of ZnTe nanosheets, (d) Ids-Vds curves of the photodetectors in the dark (black) and under
520 nm laser illumination at different incident power, (e, f) the net photocurrent and calculated
photoresponsivity as a function of the incident power density under different cource-drain bias
voltage, (g) time dependent Ids of the ZnTe device with the laser (520 nm, 4.77 mW/cm2) switching
on and off under a positive source-drain voltage Vas from 1 to 15 V. (Reproduced with permission
from Ref. [68]. Copyright 2019: Elsevier)
288 J. Ning

ZnTe nanosheets have the p-type conductivity, and ZnTe nanosheet-based photode-
tectors exhibit the high photoresponsivity of 453.9 A/W, excellent stability and
reliability.
Recently, p-type 2D ZnTe nanostructures were developed by You et al. [69]. This
ultrathin ZnTe nanoflakes were controllably synthesized by PVD method. The
monolayer ZnTe flakes-based photodetectors showed a broadband response varying
from the visible to near-infrared range under the dark and illumination with diverse
light intensities (405 nm). The ZnTe based device gave a typical p-type semiconduc-
tor behavior. The large and different photocurrent at the whole range of gate bias
suggested the significant photoresponse. The photocurrent raises gradually with
increase light intensity due to the enhancement of photogenerated carriers. The
ZnTe flakes-based photodetectors also showed the stability after 3 months in air.
This monolayer ZnTe nanoflake show photoresponse properties from the visible to
near-infrared region and the responsivity of 18.3 A W−1 and detectivity of 2.89 × 109
Jones under 405 nm illumination in air, implying the potential application in elec-
tronics and optoelectronics worked in harsh environment.

12.5 ZnTe Based Photodetectors for Terahertz Region

Most of ZnTe-based photodetectors are using the intrinsic band gap in ZnTe. The
energy of light can be absorbed by ZnTe to produce the electron and holes, the
charge can give the single via the transformation or recombination. Another type of
photodetectors with band gap-independent properties can be built with the array of
semiconductor, named as subwavelength structure (SWS), which can enhance ultra-
broad band transmission [70, 71]. The working region of photodetectors is related
to the surface of semiconductor-based SWS, covering the visible-near-infrared,
infrared and terahertz region.
The SWSs on a ZnTe single crystal are designed via a modified reactive ion etch-
ing method to increase the broadband transmission [72]. Large-area polystyrene
(PS) nanoparticle nanolayers are spin-coated on the ZnTe surface by adopting oxy-
gen plasma treatment to improve the surface wettability (Fig. 12.4a). After the etch-
ing of Ch4/H4 and O2 plasma, the SWSs based on etched ZnTe can be tuned via
etching time and radio frequency power (Fig. 12.4b–d).
Finally, the well-define conical SWS arrays were fabricated on the ZnTe crystal
by reactive ion etching over the PS monolayer template, with the size of SWS arrays
customized by optimizing the etching process. The ultra-broadband antireflection
on the surface structured ZnTe crystals in the visible-near-infrared, infrared, and
terahertz regions with transmittance increase of 11.6%, 10.0% and 24.8%, which
are attributed to the decrease of surface Fresnel reflection by SWS, and the transmit-
tance can reach 70% in 0.2–1.0 terahertz (Fig. 12.5). This provides a new strategy
to enhance the terahertz efficiency and detection sensitivity based on ZnTe crystals
by surface engineering.
12 ZnTe-Based Photodetectors for Visible-UV Spectral Region 289

Position 2 Position 1

O2 plasma Spin coating

CH4/H2
O2 plasma plasma O2 plasma

c
ZnTe

Etch rate (nm/min)


20 W 50 W 80 W 16
PS
12
8
4
300 nm 300 nm 300 nm 0
0 50 100 150 200
110 W 140 W 170 W RF power (W)
d
W1

H
300 nm 300 nm 300 nm W4
T

Fig. 12.4 (a) The schematic illustration of fabricating SWSs on ZnTe crystals, (b) SEM image of
SWSs fabricated at a radio frequency power range of 20–170 W with an etching time of 15 min,
(c) etch rates of ZnTe and PS nanoparticles as a function of radio frequency, (d) the schematic
diagram of conical SWS morphology. (Reproduced with permission from Ref. [72]. Copyright
2021: American Chemical Society)

12.6 Heterostructured ZnTe Based Photodetectors

12.6.1 ZnTe-Si Heterostructures Based Photodetectors

As well-known, Si has response to broadband lights whose wavelength ranged from


400 to 1100 nm because of its narrow band gap of 1.12 eV. More and more scientific
researchers have drawn much attention to silicon-based heterojunction materials
because of two aspects: on the one hand, mature silicon microelectronic technology
has laid a solid foundation for silicon-based heterojunction materials; on the other
hand, the development of energy band engineering research, single atomic layer
epitaxy and nanotechnology have provided deep theoretical basis and new technical
means [73–75].
ZnTe could also be combined with other semiconductor to form heterostructures
for photodetectors. ZnTe film can be formed via vacuum evaporation method on
n-type or p-type Si substrate to build photodetectors with p-n junction structure. Lin
290 J. Ning

3.2
Air

Transmission increase (%)


Probe laser VIS-NIR

Effective refractive index


30 IR
24.8%
THz 3.0 H neff
20 20 W

Detector
ZnTe
Pump laser 11.6% 50 W
ZnTe

ZnTe
THz 2.8
80 W
10 110 W
140 W
10.0% 2.6 170 W
800 nm 0.25~2.5 Thz 0
20 50 80 110 140 170 0 40 80 120 160 200
visible infrared terahertz
RF power (W) SWS height (nm)
80
Coverage: 85.2% Coverage: 85.2% 100 Coverage: 85.2%
Coverage: 54.6% 70 Coverage: 54.6% Coverage: 54.6%
Transmittance (%)

Transmittance (%)

Transmittance (%)
70 Before etching Before etching Before etching
80
60
60
60
11.6% 50 10.0% 24.8%
50
40
40 40
20
600 700 800 900 1000 2000 3000 4000 0.5 1.0 1.5 2.0 2.5
Wavelength (nm) Wavenumber (cm–1) Frequency (THz)

Fig. 12.5 (a) Schematic illustration of terahertz generation and detection of the ZnTe crystal by
optical rectification and electro-optic sampling techniques, (b) transmission increase of the SWS
ZnTe compared to the as-grown crystals from visible to terahertz region, (c) the effective refractive
index profile with SWS height, (d-f) the optimal transmittance before and after etching in the
visible-­near infrared region (d), infrared region (e) and terahertz region (f). Reproduced with per-
mission from Ref. [72]. (Copyright 2021: American Chemical Society)

et al. [76] reported the deposition of ZnTe film of silicon substrate via the thermal-­
furnace evaporation method. The crystallinity, charge mobility, carries concentra-
tion and sheet resistance in ZnTe film exhibited the dependent of argon pressure and
deposition temperature. The higher annealing temperature and deposition tempera-
ture is better to increase the grain size of ZnTe on Si substrate. The highest carrier
concentration of 1.9 × 1012 cm−3, the lowest sheet resistance of 3180 ohm/square
and the largest mobility of 5.1 × 103 cm2V−1 S−1 were obtained at an argon pressure
of 100 sccm and a deposition temperature of 580 °C, respectively.
In order to build the p-n junction heterostructure for optoelectronic devices, the
doped ZnTe for p-type semiconductor was deposited on the n-type Si substrate by
using thermal vacuum evaporation technique [77]. The deposited aluminum doped
ZnTe film had the polycrystalline structure with cubic zinc-blende phase, and the
roughness of film increased with increasing the doped amount of Al. With tuning
the amount of impurities in ZnTe film, the optical band gap can be decreased from
2.24 eV to 1.86 eV. The increase of incident lighting intensity and doped Al can rise
the illumination current of ZnTe-Si heterostructures. When the doped amount of Al
is 0.2% in ZnTe films, the heterostructure-based device showed the best value of
specific detectivity and quantum efficiency, originating from the height crystal qual-
ity of ZnTe films.
12 ZnTe-Based Photodetectors for Visible-UV Spectral Region 291

Furtherly, the high-quality ZnTe and TeO2 composite was grown on an n-type
silicon substrate via a modified metal-assisted chemical vapor deposition method
[78]. Self-powered photodetector based on a ZnTe–TeO2 composite/Si heterojunc-
tion with ultra-broadband and high responsivity is obtained. The photodetector
shows ultra-broadband photoresponsivity from UV to NIR lights as ZnTe has a
moderate, direct band gap of 2.26 eV, TeO2 has a wide band gap of 4.0 eV, and Si
has a narrow band gap of 1.12 eV. Upon exposure to 850 nm light at a zero-bias
voltage, the detector shows a high responsivity of 75 mA/W, detectivity of
1.4 × 1013 cm Hz1/2/W, fast response and recovery properties with response and
recovery times both below 0.61 s, respectively.

12.6.2 Heterostructures Based on II-VI Semiconductors

ZnTe can form heterostructures with other numerous II-VI group semiconductor
materials. Their hetero-structures were used to fabricate photodetector with high
sensitivity, large photocurrent gain, good reliability and low response time. Among
these, CdTe and ZnTe with the direct band gap of around 1.50 eV and 2.17 eV cover
the visible and near IR regions, promising candidates for photodetector applica-
tions. ZnTe/CdxZn1-xTe (0.2 ≤ x ≤ 1.0) heterostructures were fabricated by thermal
evaporation method by using CdTe and ZnTe as source materials [79]. ZnTe/
CdxZn1-­x Te heterostructures with low barrier height and good photo response are
explored in this work for the photodetector application. ZnTe/Cd0.8Zn0.2Te showed
higher response in the visible region with improved response at longer wavelengths.
On the basis of band gap alignment of semiconductors in heterostructures, the
heterostructures can be separated to type-I, type-II and quasi type-II. Generally, the
type-II structure is good for charge separation and transfer in optoelectronic device.
ZnTe and ZnSe can form the type-II heterostructures. Averin et al. reported the type-
­II ZnSe/ZnTe/GaAs superlattice for photodetectors [80]. The ZnSe/ZnTe/GaAs
superlattice based-photodetector demonstrates very low dark current, high current
sensitivity and external quantum efficiency. The maximum photoresponse of the
ZnSe/ZnTe/GaAs superlattice based-detector at the wavelength 620 nm corresponds
to current sensitivity 0.22 A/W and external quantum efficiency 44%. Photoresponse
of the ZnSe/ZnTe/GaAs superlattice based-detector shows two peaks of response
located at 620 nm and 870 nm. For the MSM-diode with finger width and gap of
3 μm and 100 × 100 μm2 photosensitive area we have obtained dark current density
10–8 A/cm2 at room temperature.
More than ZnSe, ZnTe can also combine with ZnO and CdSe to form type-II
heterostructures. Rai et al. [81] reported the CdSe/ZnTe core/shell nanowire for
broad band photodetector. CdSe nanowires arrays were first grown on the muscovite
mica substrate. Then ZnTe was further deposited on CdSe nanowires to form
CdSe/ZnTe core/shell nanowires. Photodetection is greatly enhanced by the
292 J. Ning

piezo-phototronic effect. The photodetector performance under UV (385 nm), blue


(465 nm), and green (520 nm) illumination infers a saturation free response with an
intensity variation near two orders of magnitude, where the peak photocurrent
(125 μA) is two orders higher at 0.25-kilogram force compared to no load (0.71 μA).
The resulting (%) responsivity changed by four orders of magnitude. The significant
increase in responsivity is believed to arise from: (1) the piezo-phototronic effect
induced by a change in the Schottky barrier height at the Ag–ZnTe junction, and in
the type-II band alignment at the CdSe–ZnTe interfaces, in conjugation with (2) a
small lattice mismatch between the CdSe and ZnTe epitaxial layers, which lead to
reduced charge carrier recombination.
With similar method, ZnO/ZnTe core/shell nanowires were developed by You
et al. [82]. A self-powered core/shell photodetector was fabricated by sputtering a
uniform p-type ZnTe layer on n-type ZnO nanorod array (Fig. 12.6). By integrating
pyro-electric and photovoltaic effects, the photodetector realizes broadband detec-
tion from 325 nm ultraviolet to 1064 nm near infrared under zero bias. The maxi-
mum responsivity and detectivity reach 196.24 mA/W and 3.47 × 1012 cm Hz1/2/W
for 325 nm laser illumination with power density 2.13 mW/cm2, respectively, which
are improved ten-fold relating to the device responded to photovoltaic effect only.
While the rise and fall time are drastically reduced from 1.2 ms to 62 μs and 1.6 ms
to 109 μs, respectively (Fig. 12.7). Moreover, applied bias voltage and light power
densities also play a significant impact in photovoltaic–pyroelectric coupled effects
of device.

12.7 ZnTe-Based Materials for Solar Cells

Solar cells based on II-VI semiconductors (with II = Zn, Cd, Hg and VI = S, Se, Te)
are the leading candidates for low-cost photovoltaic conversion of solar energy due
to their high absorption coefficients and therefore the low material consumption for
their production. Owing to its wide direct band gap of 2.23 ~ 2.28 eV at room tem-
perature and low electron affinity 3.73 eV [83], zinc telluride, an important member
of II-VI group semiconductors, is a suitable material for several applications, such
as solar cells where it is used as a p-type doped window materials or a top layer for
achieving low contact resistance in photovoltaic heterojunctions, switching devices
and infrared and X-ray detectors [84, 85].
Tanaka et al. [86] developed the Al doped p-type ZnTe for solar cells. An open
circuit voltage of approximately 0.9 V was obtained under 1 × sun AM1.5G condi-
tion in all solar cells, independent of diffusion times, while a short circuit current
dropped down with increasing the diffusion time due to an increased light absorp-
tion in heavily defective Al-diffused layer. These fundamental results provide a
basis for future development of intermediate band solar cells based on ZnTe
materials.
12 ZnTe-Based Photodetectors for Visible-UV Spectral Region 293

Fig. 12.6 (a) Schematic of synthesis process to fabricate vertically aligned ZnO/ZnTe core/shell
nanorod array photodetector device. (b, c) Cross-sectional view SEM images of ZnO nanorods
before and after the ZnTe layer coating, (inset b and c: high-magnification SEM of single nanorod).
(d) EDS spectrum and (e) XPS measurement of ZnO/ZnTe core–shell nanorod array. (f) XRD pat-
terns of bare ZnO and ZnO/ZnTe. (g) Optical absorption spectra of ZnO/ZnTe core–shell nanorods.
(h) Energy band diagram of ITO/ZnO/ZnTe/Al. (Reproduced with permission from Ref. [82].
Copyright 2019: Elsevier)

Tang et al. [87] further doped ZnTe film with oxygen via ion implantation. The
proper concentrations of oxygen ions attributed to the formation of the intermediate
band which was approximately 1.88 eV above the valence band maximum. ZnTe
with high crystalline quality and appropriate concentrations of oxygen ions led to
the improvements of absorption efficiency of the intermediate band. The crystalline
quality and the dose of oxygen concentration are important to achieve better ZnTe:O
intermediate-band photovoltaic materials.
294 J. Ning

Fig. 12.7 Impacts of incident light power densities on photovoltaic–pyroelectric coupled effects
on the ZnO/ZnTe photodetector under different laser illumination at zero bias under frequency of
20 Hz. I-t characteristics of the photodetector under (a) 325, (b) 532 and (c) 1064 nm laser illumi-
nation with various light power densities, respectively. (d–f) The corresponding I(pyro+photo) and
I(photo) as a function of the light power densities. Impacts of bias voltage on the photovoltaic–
pyroelectric coupled effects under 325 nm light illumination at light power density 2.13 mW/cm2
under frequency of 20 Hz. (g) I–t characteristics of the photodetector at different bias voltages. (h)
Ipyro+photo and Iphoto as a function of bias voltage. (i) Enlarged plot of a single output period as
shown in a for 0 V and 1.8 V under 325 nm illuminations. (Reproduced with permission from Ref.
[82]. Copyright 2019: Elsevier)

Beside as the absorber layer in photovoltaic device, ZnTe can also be used as
back-contact in CdTe solar cell. With a cubic zinc-blende structure and a bandgap
of ~2.26 eV, ZnTe has negligible valence band discontinuity with respect to CdTe
(which would not impede hole transport), and a large conduction band offset, which
can be beneficial for electron back reflection to CdTe and hence minimize minority
carrier recombination related losses at the interface to the metal back contact [88,
89]. High p-type doping concentrations achievable with ZnTe can provide a low
resistance Ohmic contact to the metal electrode. Cu-doped ZnTe back contact to
CdTe have been credited with improvement in device efficiency [90, 91]. As an
alternative to ZnTe:Cu, group V (N, As)-doped ZnTe can be used as a back contact
(Fig. 12.8) [92], ZnTe:N has been recognized as a suitable back contact for CdTe
thin film solar cells, showing improved device stability [93, 94].
12 ZnTe-Based Photodetectors for Visible-UV Spectral Region 295

Fig. 12.8 Schematic of (a) baseline CdTe cell, (b) ZnTe:As back-contacted CdTe cell. (Reproduced
from Ref. [92]. Published 2015 by MDPI as open access)

12.8 Summary and Outlook

In summary, recent advances of ZnTe based-photodetectors ranging from synthesis,


device building and performance of device were discussed. On the basis of the band
gap of 2.2 eV in ZnTe, the photodetectors covering the ultraviolet to visible region
are developed, and the detected region can be further extended to near-infrared and
terahertz region with the SWSs on a single ZnTe crystal. However, compared to
other semiconductor-based photodetectors, the performance and investigation of
ZnTe based photodetectors are limited. For example, ZnTe are not stable in air at
room temperature, the shell growth of inorganic compound on the 1D ZnSe nano-
structures can improve the stability and performance of photodetectors. There are
several issues that remain unsolved for ZnTe based photodetectors. Some future
aspects of ZnTe-based photodetectors might be explored in order to accomplish
their widespread applications.
1. Semiconductor nanocrystals have the size and shape-dependent optical and
optoelectronic properties. The precise controlled synthesis of semiconductor
nanostructure can turn the band gap and optoelectronic properties based on the
applications’ requirement. 1D and 2D ZnTe nanocrystals via CVD or PVD
method have the diameter or thickness of tens nanometers. The region of ultra-
violet and blue light in ZnTe-based photodetectors are not be really achieved.
Further controlled synthesis of ZnTe nanocrystals with tuned size to Bohr radius
for strong quantum size confinement effect is necessary to develop.
2. Most of the developed ZnTe-based photodetectors are built on the PVD or CVD
produced ZnTe nanostructures. The colloidal method produced nanocrystals
have tunable shape and size. However, the photodetectors based on colloidal
ZnTe nanocrystals are limited developed because of the complex ink-print
296 J. Ning

t­echnology. The development of the ink-print technology for colloidal ZnTe


nanocrystals can produce high-quality optoelectronic devices, the colloidal ZnTe
nanocrystals-based photodetectors can give the wider tunable working region.

References

1. Rieke G (2003) Detection of light: from the ultraviolet to the Submilli-meter. Cambridge
University Press, Cambridge
2. Teng F, Hu K, Ouyang W, Fang XS (2018) Photoelectric detectors based on inorganic p-type
semiconductor materials. Adv Mater 30:1706262
3. Li PK, Liao ZM, Zhang XZ, Zhang XJ, Zhu HC, Gao JY, Laurent K, Leprince-Wang Y,
Wang N, Yu DP (2009) Electrical and Photoresponse properties of an intramolecular p-n
Homojunction in single phosphorus-doped ZnO nanowires. Nano Lett 9:2513–2518
4. Zou R, Zhang Z, Liu Q, Hu J, Sang L, Liao M, Zhang W (1848) High Detectivity solar-blind
high-temperature deep-ultraviolet photodetector based on multi-layered (l00) facet-oriented
β-Ga2O3 Nanobelts. Small 2014:10
5. Fang Y, Dong Q, Shao Y, Yuan Y, Huang J (2015) Highly narrowband perovskite single-crystal
photodetectors enabled by surface-charge recombination. Nat Photonics 9:679–686
6. Han X, Du W, Yu R, Pan C, Wang ZL (2015) Piezo-Phototronic enhanced UV sensing based
on a nanowire photodetector Array. Adv Mater 27:7963–7969
7. Lin Q, Armin A, Burn PL, Meredith P (2015) Filter less narrowband visible photodetectors.
Nat Photonics 9:687–694
8. Youngblood N, Chen C, Koester SJ, Li M (2015) Waveguide-integrated black phosphorus
photodetector with high responsivity and low dark current. Nat Photonics 9:247–252
9. Chen H, Liu H, Zhang Z, Hu K, Fang XS (2016) Nanostructured photodetectors: from ultra-
violet to terahertz. Adv Mater 28:403–433
10. Yu X, Marks TJ, Facchetti A (2016) Metal oxides for optoelectronic applications. Nat Mater
15:383–396
11. Fang XS, Hu LF, Hui KF, Gao B, Zhao LJ, Liao MY, Chu PK, Bando Y, Golberg D (2011)
New ultraviolet photodetector based on individual Nb2O5 Nanobelts. Adv Funct Mater
21:3907–3915
12. Ekimov AI, Onushchenko AA (1981) Quantum size effect in three-dimensional microscopic
semiconductor crystals. JETP Lett 34:345–349
13. Efros AL, Efros AL (1982) Interband absorption of light in semiconductor sphere. Sov Phys
Semicond 16:772–775
14. Brus LE (1983) A simple model for the ionization potential, electron affinity, and aqueous
redox potentials of small semiconductor crystallites. J Chem Phys 79:5566–5571
15. Rossetti R, Nakahara S, Brus LE (1983) Quantum size effects in the redox potentials, reso-
nance Raman spectra, and electronic spectra of CdS crystallites in aqueous solution. J Chem
Phys 79:1086–1088
16. Brus LE (1984) Electron-electron and electron-hole interactions in small semiconductor crys-
tallites: the size dependence of the lowest excited electronic state. J Chem Phys 80:4403–4409
17. Murray CB, Norris DJ, Bawendi MG (1993) Synthesis and characterization of nearly mono-
disperse CdE (E = S, Se, Te) semiconductor Nanocrystallites. J Am Chem Soc 115:8706–8715
18. Shalev E, Oksenberg E, Rechav K, Popovitz-Biro R, Joselevich E (2017) Guided CdSe nanow-
ires Parallelly integrated into fast visible-range photodetectors. ACS Nano 11:213–220
19. Li LD, Lou Z, Shen GZ (2015) Hierarchical CdS nanowires based rigid and flexible photode-
tectors with ultrahigh sensitivity. ACS Appl Mater Interfaces 7:23507–23514
20. Xie X, Kwok S, Lu Z, Liu Y, Cao Y, Luo L, Zapien JA, Bello I, Lee CS, Lee ST, Zhang WJ
(2012) Visible-NIR photodetectors based on CdTe nanoribbons. Nanoscale 4:2914–2919
12 ZnTe-Based Photodetectors for Visible-UV Spectral Region 297

21. Yan X, Li B, Wu Y, Zhang X, Ren XM (2016) A single crystalline InP nanowire photodetector.
Appl Phys Lett 109:053109
22. Chen G, Liang B, Liu X, Liu Z, Yu G, Xie X, Luo T, Chen D, Zhu M, Shen G, Fan ZY (2014)
High-performance hybrid phenyl-C61-butyric acid methyl Ester/Cd3P2 nanowire ultraviolet-­
visible-­near infrared photodetectors. ACS Nano 8:787–796
23. Michel J, Liu J, Kimerling LC (2010) High-performance Ge-on-Si photodetectors. Nat
Photonics 4:527–534
24. Langer DW, Vesely CJ (1970) Electronic Core levels of zinc chalcogenides. Phys Rev B
2:4885–4892
25. Katre A, Togo A, Tanaka I, Madsen GKH (2015) First principles study of thermal conductivity
cross-over in nanostructured zinc-chalcogenides. J Appl Phys 117:045102
26. Zhang J, Sun K, Kumbhar A, Fang JY (2008) Shape-control of ZnTe nanocrystal growth in
organic solution. J Phys Chem C 112:5454–5458
27. Jang E, Han C, Lim SW, Jo J, Jo DY, Lee SH, Yoon SY, Yang H (2019) Synthesis of alloyed
ZnSeTe quantum dots as bright, color-pure blue emitters. ACS Appl Mater Interfaces
11:46062–46069
28. Lesnyak V, Dubavik A, Plotnikov A, Gaponik N, Eychmüller A (2010) One-step aqueous syn-
thesis of blue-emitting glutathione-capped ZnSe1−xTex alloyed nanocrystals. Chem Commun
46:886–888
29. Fairclough SM, Tyrrell EJ, Graham DM, Lunt PJB, Hardman SJO, Pietzsch A, Hennies F,
Moghal J, Flavell WR, Watt AAR, Smith SM (2012) Growth and characterization of strained
and alloyed type-II ZnTe/ZnSe Core/Shell nanocrystals. J Phys Chem C 116:26898–26907
30. Jin S, Zhang J, Schaller RD, Rajh T, Wiederrecht GP (2012) Ultrafast charge separation from
highly reductive ZnTe/CdSe type II quantum dots. J Phys Chem Lett 3:2052–2058
31. Smith AM, Nie SM (2010) Semiconductor nanocrystals: structure, properties and band gap
engineering. Acc Chem Res 43:190–200
32. Asano H, Tsukuda S, Kita M, Fujimoto S, Omata T (2018) Colloidal Zn(Te, Se)/ZnS Core/
Shell quantum dots exhibiting narrow-band and green photoluminescence. ACS Omega
3:6703–6709
33. Reiss P, Protiere M, Li L (2009) Core/Shell semiconductor nanocrystals. Small 5:154–168
34. Bang J, Park J, Lee JH, Won N, Nam J, Lim J, et al. (2010) ZnTe/ZnSe (Core/Shell) type-II
quantum dots: their optical and photovoltaic properties. Chem Mater 22:233–240
35. Ngo CY, Yoon SF, Fan WJ, Chua SJ (2006) Effect of size and shape on electronic states of
quantum dots. Phys Rev B 74:245331
36. Shabaev A, Efros AL (2004) 1D exciton spectroscopy of semiconductor nanorods. Nano Lett
4:1821–1825
37. Huynh WU, Dittmer JJ, Alivisatos AP (2002) Hybrid nanorod-polymer solar cells. Science
295:2425–2427
38. Nam S, Oh N, Zhai Y, Shim M (2015) High efficiency and optical anisotropy in double-­
heterojunction Nanorod light-emitting diodes. ACS Nano 9:878–885
39. Oh N, Kim BH, Cho SY, Nam S, Rogers SP, Jiang Y, et al. (2017) Double-heterojunction
Nanorod light-responsive LEDs for display applications. Science 355:616–619
40. Pawar AA, Halivni S, Waiskopf N, Ben-Shahar Y, Soreni-Harari M, Bergbreiter S, Banin U,
Magdassi S (2017) Rapid three-dimensional printing in water using semiconductor-metal
hybrid nanoparticles as Photoinitiators. Nano Lett 17:4497–4501
41. Ji B, Panfil YE, Banin U (2017) Heavy-metal-free fluorescent ZnTe/ZnSe Nanodumbbells.
ACS Nano 11:7312–7320
42. Pal U (1993) Dark-and photoconductivity in doped and Undoped zinc telluride films. Semicond
Sic Technol 8:1331–1336
43. Rao GK, Bangera KV, Shivakumar GK (2010) Studies on the photoconductivity of vacuum
deposited ZnTe thin films. Mater Res Bull 45:1357–1360
298 J. Ning

44. Shinde UP, Patil AV, Dighavkar CG, Patil SJ, Kapadnis KH, Borse RY, Nikam PS (2010)
Photoconductivity study as a function of thickness and composition of Zn-Te thin films for
different illumination conditions at room temperature. Optoelectron Adv Mater 4:291–294
45. Bhunia S, Pal D, Bose DN (1998) Photoluminescence and photoconductivity in hydrogen-­
passivated ZnTe. Semicond Sci Technol 13:1434–1438
46. Hamdi H, Valette S (1980) ZnTe extrinsic photodetector for visible integrated optics. J Appl
Phys 51:4739–4741
47. Bozzini B, Baker MA, Cavallotti PL, Cerri E, Lenardi C (2000) Electrodeposition of ZnTe for
photovoltaic cells. Thin Solid Films 361-362:388–395
48. Cui Y, Wei Q, Park H, Lieber CM (2001) Nanowire nanosensors for highly sensitive and selec-
tive detection of biological and chemical species. Science 293:1289–1292
49. Chen RS, Wang SW, Lan ZH, Tsai JT, Wu CT, Chen LC, et al. (2008) On-chip fabrication
of well-aligned and contact-barrier-free GaN nanobridge devices with ultrahigh photocurrent
responsivity. Small 4:925–929
50. Cao Q, Rogers JA (2009) Ultrathin films of single-walled carbon nanotubes for electronics and
sensors: a review of fundamental and applied aspects. Adv Mater 21:29–53
51. Cao YL, Liu ZT, Chen LM, Tang YB, Luo LB, Jie JS, Zhang WJ, Lee ST, Lee CS (2011)
Single-crystalline ZnTe nanowires for application as high-performance green/ultraviolet pho-
todetector. Opt Exp 19:6100–6108
52. Lieber CM, Wang ZL (2007) Functional nanowires. MRS Bull 32:99–108
53. Kind H, Yan HQ, Messer B, Law M, Yang PD (2002) Nanowire ultraviolet photodetectors and
optical switches. Adv Mater 14:158–160
54. Luo J, Zheng Z, Yan S, Morgan M, Zu X, Xiang X, Zhou W (2020) Photocurrent enhanced
in UV-vis-NIR photodetector based on CdSe/CdTe Core/Shell nanowires arrays by piezo-­
Phototronic effect. ACS Photonics 7:1461–1467
55. Tang M, Xu P, Wen Z, Chen X, Pang C, Xu X, Meng C, Liu X, Tian H, Raghavan N, Yang
Q (2018) Fast response CdS-CdSxTe1-x-CdTe Core-Shell Nanobelt photodetector. Sci Bull
63:1118–1124
56. Shaygan M, Davami K, Jin B, Gemming T, Lee J, Meyyappan M (2016) Highly sensitive pho-
todetectors using ZnTe/ZnO Core/Shell nanowire field effect transistors with a tunable Core/
Shell ratio. J Mater Chem C 4:2040–2046
57. Chattopadhyay D, Queisser HJ (1981) Electron scattering by ionized impurities in semicon-
ductors. Rev Mod Phys 53:745–768
58. Grimmeiss HG (1977) Deep level impurities in semiconductors. Annu Rev Mater Sci
7:341–376
59. Luo L, Huang X, Wang M, Xie C, Wu C, Hu J, Wang L, Huang J (2014) The effect of Plasmonic
nanoparticles on the optoelectronic characteristics of CdTe nanowires. Small 13:2645–2652
60. Zhang Q, Zhang J, Utama MIB, Peng B, de la Mata M, Arbiol J, Xiong Q (2012) Exciton-­
phonon coupling in individual ZnTe Nanorods studied by resonant Raman spectroscopy. Phys
Rev B 85:085418
61. Li S, Jiang Y, Wu D, Wang L, Zhong H, Wu B, Lan X, Yu Y, Wang Z, Jie J (2010) Enhanced
p-type conductivity of ZnTe nanoribbons by nitrogen doping. J Phys Chem C 114:7980–7985
62. Luo L, Zhang S, Lu R, Sun W, Fang Q, Wu C, Hu J, Wang L (2015) p-Type ZnTe:Ga Nanowires:
controlled doping and optoelectronic device application. RSC Adv 5:13324–13330
63. Wan X, Huang Y, Chen Y (2012) Focusing on energy and optoelectronic applications: a jour-
ney for graphene and graphene oxide at large scale. ACS Chem Res 45:598–607
64. Chang H, Wu H (2013) Graphene-based nanomaterials: synthesis, properties, and optical and
optoelectronic applications. Adv Funct Mater 23:1984–1997
65. Xie C, Wang Y, Zhang Z, Wang D, Luo L (2018) Graphene/semiconductor hybrid
Heterostructures for optoelectronic device applications. Nano Today 19:41–83
66. Mak KF, Shan J (2016) Photonics and optoelectronic of 2D semiconductor transition metal
dichalcogenides. Nat Photonics 10:216–226
12 ZnTe-Based Photodetectors for Visible-UV Spectral Region 299

67. Wang J, Ardelean J, Bai Y, Steinhoff A, Florian M, Jahnke F, Xu X, Kira M, Hone J, Zhu X
(2019) Optical generation of high carrier densities in 2D semiconductor conductor heterobi-
layers. Sci Adv 5:0145
68. Wang Y, Li H, Yang T, Zou Z, Qi Z, Ma L, Chen J (2019) Space-confined physical vapour
deposition of high quality ZnTe Nanosheets for optoelectronic application. Mater Lett
­
238:309–312
69. You S, Wu Z, Niu L, Chu X, She Y, Liu Z, Cai Y, Liu H, Zhang L, Zhang K, Luo Z, Huang
S (2022) 2D ultrathin p-type ZnTe with high environmental stability. Adv Electron Mater
8:2101146
70. Li Y, Zhang J, Yang B (2010) Antirereflective surfaces based on biomimetic Nanopillared
arrays. Nano Today 5:117–127
71. Choi HJ, Huh D, Jun J, Lee H (2019) A review on the fabrication and applications of sub-­
wavelength anti-reflective surfaces based on Biomimietics. Appl Spectrosc Rev 54:719–735
72. Sun H, Liu J, Zhou C, Yang W, Liu H, Zhang X, Li Z, Zhang B, Jie W, Xu Y (2021) Enhanced
transmission from visible to terahertz in ZnTe crystals with scalable subwavelength structures.
ACS Appl Mater Interfaces 13:16997–17005
73. Wang Z, Yu R, Wen X, Liu Y, Pan C, Wu W, Wang ZL (2014) Optimizing performance
of silicon-based p-n junction photodetectors by the piezo-phototronic effect. ACS Nano
8:12866–12873
74. Avasthi S, Lee S, Loo YL, Sturm JC (2011) Role of majority and minority carrier barriers
silicon/organic hybrid hetero-junction solar cells. Adv Mater 23:5762–5766
75. Masuko K, Shigematsu M, Hashiguchi T (2014) Achievement of more than 25% conversion
efficiency with crystalline silicon heterojunction solar cell. IEEE J Photovolt 4:1433–1435
76. Lin J, Wei S, Yu Y, Hsu C, Kao W, Chen W, Tseng C, Lai C, Lu J, Ju S, Hsieh J (2013)
Synthesis and characterization of ZnTe thin films on silicon by thermal-furnace evaporation.
Proc SPIE 8913:89130k
77. Maki SA, Hassun HK (2018) Effect of aluminum on characterization of ZnTe/n-Si heterojunc-
tion photodetector. J Phys Conf Series 1003:012085
78. Song Z, Liu Y, Wang Q, Yuan S, Yang Y, Sun X, Xin Y, Liu M, Xia Z (2018) Self-powered
photodetectors based on a ZnTe-TeO2 composite/Si heterojunction with ultra-broadband and
high responsivity. J Mater Sci 53:7562–7570
79. Moger SN, MG, M. (2020) Investigation on ZnTe/CdxZn1-xTe Heterostructure for photodetec-
tor applications. Sens Actuators A 315:112294
80. Averin SV, Kuznetsov PI, Zhtov VA, Zakharov LY, Kotov VM (2018) Multicolour photodetec-
tor based on a ZnSe/ZnTe/GaAs heterostructure. Quant Electron 48:675–678
81. Rai SC, Wang K, Chen J, Marmon JK, Bhatt M, Wozny S, Zhang Y, Zhou W (2015) Enhanced
broad band Photodetection through piezo-Phototronic effect in CdSe/ZnTe Core/Shell nanow-
ire Array. Adv Electron Mater 1:1400050
82. You D, Xu C, Zhang W, Zhao J, Qin F, Shi Z (2019) Photovoltaic-pyroelectric effect coupled
broadband photodetector in self-powered ZnO/ZnTe Core/Shell Nanorod arrays. Nano Energy
62:310–318
83. Pistone A, Arico AS, Antonucci PL, Silvestro D, Antonucci V (1998) Preparation and charac-
terization of thin film ZnCuTe semiconductors. Sol Energy Mater Sol Cell 53:255–267
84. Ernt K, Seiber I, Neumann-Spalart M, Lux-Steiner MC, Könenkamp R (2000) Characterization
of II-VI compounds on porous substrates. Thin Solid Films 361:213–217
85. Winnewiser C, Jepsen PU, Schall M, Schiyja V, Helm H (1997) Electrooptic detection of THz
radiation in LiTaO3, LiNbO3, and ZnTe. Appl Phys Lett 70:3069–3071
86. Tanaka T, Yu KM, Stone PR, Beeman JW, Dubon O, Reichertz LA, Kao VM, Nishio M,
Walukiewicz W (2010) Demonstration of Homojunction ZnTe solar cells. J Appl Phys
108:024502
87. Tang N, Hu Q, Ren A, Li W, Liu C, Zhang J, Wu L, Li B, Zeng G, Hu S (2017) An approach to
ZnTe:O intermediate-band photovoltaic materials. Sol Energy 157:707–712
300 J. Ning

88. Amin N, Yamada A, Konagai M (2002) Effect of ZnTe and CdZnTe alloys at the back contact
of 1-μm-thick CdTe thin film solar cells. Jpn J Appl Phys 41:2834–2841
89. Späth B, Fritsche J, Klein A, Jaegermann W (2007) Nitrogen doping of ZnTe and its influence
on CdTe/ZnTe interfaces. Appl Phys Lett 90:062112
90. Li J, Beach JD, Wolden CA (2014) Rapid thermal processing of ZnTe:cu contacted CdTe solar
cells. In: Proceedings of the 40th IEEE photovoltaic specialist conference, Denver, CO, USA,
8–13 June 2014, pp 2360–2365
91. Uliˆcnâ S, Isherwood PJM, Kaminski PM, Walls JM, Li J, Wolden CA (2016) Development of
ZnTe as back contact material for thin film cadmium telluride solar cells. Vacuum 139:159–163
92. Oklobia O, Kartopu G, Irvine SJC (2019) Properties of arsenic-doped ZnTe thin films as a
back contact for CdTe solar cells. Materials 12:3706
93. Amin N, Yamada A, Konagai M (2000) ZnTe insertion at the back contact of 1 μm-CdTe
thin film solar cells. In: Proceedings of the 28th IEEE photovoltaic specialists conference,
Anchorage, AK, USA, 15–22 September 2000, pp 650–653
94. Makhratchev K, Price KJ, Ma X, Simmons DA, Drayton J, Ludwig K, Gupta A, Bohn RG,
Compaan AD (2000) ZnTe:N back contacts to CdS/CdTe solar cells. In: Proceedings of the
28th IEEE photovoltaic specialists conference, Anchorage, AK, USA, 15–22 September 2000,
pp 475–478
Chapter 13
ZnSe-Based Photodetectors

Ghenadii Korotcenkov

13.1 Introduction

ZnSe, as well as other direct-gap wide-gap semiconductors, is of great interest in the


development of ultraviolet photodetectors (PD). Currently, UV photodiodes are
being developed based on various materials such as silicon, gallium phosphide, sili-
con carbide, gallium nitride, zinc oxide, etc. [44]. The disadvantage of silicon pho-
todiodes is the presence of photosensitivity outside the UV range, namely in the
visible and infrared (IR) region of the spectrum. Photodiodes based on gallium
phosphide have a low sensitivity in the region of 200–250 nm. Photodiodes based
on silicon carbide are quite sensitive to radiation with a wavelength of about 300 nm.
However, their sensitivity in the range of 200–250 nm and 350–400 nm is low.
Gallium nitride photodiodes have good parameters, but thin-film GaN is quite
expensive, which limits its wide application in general-purpose devices. Good
parameters are also demonstrated by photodiodes based on zinc oxide. But, due to
the large band gap (Eg = 3.37 eV), photodetectors have low sensitivity in the near
UV region. The photocurrent begins to decrease after 370 nm, and it continues
decreasing to zero at 400 nm. In this regard, the ZnSe compound has a number of
advantages. The band gap of ZnSe is 2.67–2.82 eV, which makes it possible to
develop sensors sensitive to radiation in a wider spectral range. Photodiodes based
on ZnSe are sensitive in the range of 250–470 nm, which are necessary for the
manufacture of dosimeters in the UV-A (320–400 nm) and UV-B (290–320 nm)
spectral ranges. In addition, photodiodes are not sensitive to optical radiation in the
visible and IR ranges. ZnSe also has a high breakdown electric field strength (1 MV
cm−1) [68], as well as high resistance to degradation under intense UV and X-ray
radiation. In addition, the existing technology for the synthesis and deposition of the

G. Korotcenkov (*)
Department of Physics and Engineering, Moldova State University, Chisinau, Moldova
e-mail: ghkoro@yahoo.com

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 301
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_13
302 G. Korotcenkov

ZnSe compound does not require large capital investments. It should also be taken
into account that ZnSe, unlike many II-VI compounds, can have both n- and p-type
conductivity, which makes it possible to develop all possible types of ZnSe-based
photodetectors. To date, there have been reports of the development of ZnSe based
photodetectors such as photoconductive photodetectors [26], Schottky barrier pho-
todiodes [79], metal-semiconductor-metal (MSM) photodiodes [68], p–n [31],
p–i–n [2] and avalanche [30] photodiodes, phototransistors [15], and photodetectors
based on various heterostructures [60], one-dimensional nanomaterials [36] and
hybrid materials [1]. A theoretical analysis of how these devices work can be found
in [8, 44, 62, 75].

13.2 Photoconductive Photodetectors

The structure of photoconductive detector is usually a semiconductor with two con-


tacts, which represents a simple and low-cost structural design. Ideally, the contacts
should be Ohmic without any energy barrier between the metal and semiconductor,
so that electrons and holes can be injected без каких-либо ограничений into the
semiconductor through the contacts and recombine with the photo-carriers.
Therefore, the forming good Ohmic contacts is crucial in the photoconductive
detector fabrication.
ZnSe photoconductors have been investigated in detail by a number of groups,
especially before the year 2000. These studies have been carried out using both
single crystal samples [11] and polycrystalline materials [21]. The spectral response
of ZnSe photoconductors, measured under continuous excitation, is shown in
Fig. 13.1. Undoped ZnSe single crystals show maximum photosensitivity at about
462 nm (2.68 eV) [11], which corresponds to the absorption edge of ZnSe. In crys-
tals with halide impurity, the maximum is shifted about 10 nm to longer wave-
lengths. In most crystals of ZnSe:Br:Cu, the maximum is shifted even further to
longer wavelengths, by as much as 30 nm. At the same time, doping with selenium
shifts the peak to shorter wavelengths [58]. For polycrystalline ZnSe film, the pho-
toconductivity maximum lies in the same spectral region, λ ~ 450–460 nm [21, 61].
Dima and Vasiliu [21] also showed that heat treatment significantly increases the
photoconductivity of films without a significant change in the position of the photo-
conductivity maximum.
The main feature of photoconductors is that (a) an external bias voltage is
required to operate the device, and (b) they can have an external quantum yield
greater than unity. In general, when photoconductive detectors absorb photons and
generate electron-hole pairs, the majority carriers (electrons) have higher mobility
and shorter transit time between electrodes than carrier lifetime; while the minority
carriers (holes) move slowly and have a transit time longer than the lifetime of the
carriers. In such conditions, the electrons are quickly swept out of the detector,
which leads to the creation of an excess of holes. In this case, to maintain electrical
neutrality, the other electrode must supply electrons. Due to this behavior, the elec-
trons can move back and forth between the electrodes several times during the life
13 ZnSe-Based Photodetectors 303

Fig. 13.1 Spectral response curves for photoconductivity in ZnSe crystals. (1) “Pure” ZnSe, (2)
ZnSe with halide impurity, (3) ZnSe:Br:Cu, and (4) ZnSe:Br:Cu at −183 °C. (Reprinted with per-
mission from Ref. [11]. Copyright 1958: American Physical Society)

of the carrier, and thus the photodetector can obtain a higher gain, which is usually
understood as a persistent photoconductivity. The longer the carrier recombination
time, the greater will be the responsivity of the photodetector. However, this mecha-
nism will inevitably lead to an increase in the recovery time of the device after
switching from on to the off state [33]. It is clear that to increase the carrier life time,
any carrier quenching processes should be minimized. Therefore, a high quality
semiconductor film is desired in the development of semiconductor photoconduc-
tors. Selecting a smaller distance between electrodes in a photoconductor can
shorten the transit time, but shortening the distance also increases the dark current
which affects the performance of device. A balance of these two factors should be
considered in the device design.
The weakness of the photoconductive PD is the exposed optical receiving area to
air, which may suffer from pollutants and surface effect. In addition, the devices
display poor UV/visible contrast, hardly reaching a factor of ten. As a rule, ZnSe
photoconductors exhibit increased sensitivity in the impurity region of the spectrum
as well. For example, in most crystals of ZnSe:Br:Cu a second impurity-associated
shoulder in the spectral response curve occurs at about 510 nm, with a long-­
wavelength tail extending out to about 600 nm [11].
However, the main drawback of photoconductive detectors is the presence of
persistent photoconductivity (PPC), i.e., the photocurrent persists for a long time
(up to hours) after the light is removed (see Fig. 13.2a). As a consequence of persis-
tence, the measured responsivity depends drastically on the time that the sample has
been kept in the dark. However, this effect is most pronounced for polycrystalline
material. The rise and decay of monocryctalline samples are usually quite fast. For
304 G. Korotcenkov

Fig. 13.2 (a) A typical logarithmic plot of the persistent photocurrent decay in MBE-grown
p-ZnSe/GaAs structures at T = 250 K. The arrows indicate the moments at which the excitation
light was turned on and off. Reprinted with permission from [65]. Copyright 2000: Elsevier; (b)
Rise and decay curves of ZnSe single crystals when annealed in the presence of zinc. (Reprinted
with permission from Ref. [9]. Copyright 1987: Springer)

example, Bube and Lind [11] reported that the insulating crystals had a measured
decay time of 2 ms; the more conducting crystals had a decay time of 14 ms, which
indicates the high structural quality of the crystals. If, however, the concentration of
structural defects is increased, which can be done, for example, by annealing ZnSe
in the presence of metal zinc [9], then the rise and decay become quite slow even for
a single crystal material (see Fig. 13.2b).
In fact, the presence of PPC can be explained by the same mechanism responsi-
ble for the high gain in photoconductors [49]. The photoionization of electrons
trapped on extended defects causes a local charge change and, consequently, a nar-
rowing of the space charge region (SCRs) around them. Because of the band bend-
ing, holes are captured by defects, and when the light is off, the electrons have to
overcome the potential barrier before recombination begins. Since the barrier height
increases in proportion to the square of the charge on the defect, the decay of the
photocurrent is non-exponential. This model can be applied if the dominant defects
are located either in dislocations or grain boundaries, or at the ZnSe interfaces with
air or substrate [64]. In a polycrystalline material, PPC manifests itself most clearly,
since, on the one hand, an increased concentration of structural defects is character-
istic of a polycrystalline material, and, on the other hand, grain boundaries act as
trapping centers for photogenerated charge carriers.
If the photoconductors are not encapsulated, then oxygen chemisorbed on the
surface of crystallites begins to play a significant role in PPC. In short, surface traps
tend to chemisorb oxygen molecules from the air, which is accompanied by the
capture of free electrons from an n-type semiconductor [49]. This leads to the
13 ZnSe-Based Photodetectors 305

formation of an electron-depleted region near the surface and band bending, which
significantly reduces the conductivity of the device. The bending of the surface
bands creates an internal electric field that spatially separates the photogenerated
electron-hole pars, which leads to the suppression of photocarrier recombination
and a significant increase in the carrier lifetime. These effects are especially notice-
able in nanocrystalline films, where the surface area is large and depletion regions
can spread throughout the crystallite. When illuminated with a photon energy
greater than or equal to the band gap of the semiconductor, electron-hole pairs are
generated. Holes that migrate to the surface along the potential gradient created by
band bending either discharge the negatively charged chemisorbed oxygen ions,
facilitating desorption of oxygen from the surfaces, or are effectively trapped on the
surface of ZnSe crystallites. This leads to an increase in the concentration of free
carriers and a decrease in the width of the depletion layer, and hence to an increase
in the conductivity of the material. When the UV illumination is turned off, the
residual accumulated holes recombine with unpaired electrons, and oxygen is grad-
ually re-chemosorbed by the surface, which leads to a slow decay of the current.
Considering the dependence of the process of oxygen chemisorption on many fac-
tors, such as pressure, temperature and humidity, it must be stated that in order to
exclude the influence of external factors on the parameters of devices, photoconduc-
tive detectors must be encapsulated.
As follows from our consideration, the slowness of the decay process is mainly
due to the presence of impurities, grain boundaries, and structural defects that play
the role of charge carrier traps. This means that the optimization of the ZnSe synthe-
sis technology is the main way to reduce the effect of PPC on the characteristics of
photodetectors and thereby make photoconductive detectors faster. Research con-
ducted by Rao et al. [61] and Sharma and Tripathi [64] support this claim. Sharma
and Tripathi [66] showed that the photoresponse rate and photocurrent decay depend
on the ZnSe film deposition conditions (see Fig. 13.3). They believe that the
observed improvement in decay time is due to a decrease in the intergrain potential
barrier and trap density. The same results were obtained by Rao et al. [61]. They
found that elevated substrate temperature during ZnSe deposition reduces the den-
sity of traps in the films, and hence the rise and decay processes become much faster.
As for the most efficient photoconductive detector developed on the basis of
ZnSe, such a device is the PD developed by Huang et al. [26]. They reported that a
photoconductor fabricated from a ZnSe epitaxial layer grown on semi-insulating
GaAs by molecular beam epitaxy (MBE) showed a sensitivity of 320 A/W
(U = 10 V) and a response time of 2.3 ms. Ohmic contacts were fabricated by evapo-
rating 0.2 μm Au and annealing at 450 °C for 5 min. In the interdigital electrodes,
the fingers were 2 μm wide and 50 μm long with a distance of 2 μm between them.
The responsivity increased linearly with the applied voltage up to 20 V. This behav-
ior is typical for all photoconductive PCs. It is important to note that these parame-
ters of the photodetector were achieved due to the high quality of the epitaxial layers
and the use of interdigital contacts with a small distance between the electrodes.
However, despite their high sensitivity, photoconductive detectors are unsuitable for
some applications that require speed or certain spectral contrast.
306 G. Korotcenkov

Fig. 13.3 Decay of photocurrent at 298 K for ZnSe thin films deposited at different pressures.
ZnSe thin films have been prepared by inert gas condensation method. Thin films were grown in a
conventional vacuum coating system on well degassed chemically cleaned Corning 7059 glass
substrates. (Reprinted with permission from Ref. [66]. Copyright 2011. Elsevier)

13.3 The p–n Junction Photodiodes

As follows from the previous chapter, photoconductive detectors suffer from a poor
time response, due to the long electron lifetime. In addition, it should be noted that
such PDs are also characterized by significant shot noise from the high level of dark
current. Both of these problems are remedied in the photodiode detector. In this
device, light incident on the p-n junction of a semiconductor creates electron-hole
pairs, which are swept out of the depletion region by the electric field there. Current
flows in the external circuit only while charges are moving through this E field
region, so the time response of the detector can be made quite fast. The p-n junction
also provides a potential barrier for majority charge carriers, greatly reducing the
amount of dark current and associated shot noise.
An example implementation of a ZnSe-based p-n junction photodiode would be
the PD developed by Ishikura et al. [31]. A schematic diode structure and the spec-
tral response of the external quantum efficiencies of these PDs are shown in
Fig. 13.4a. The ZnSe p+–n structure photodiodes were fabricated by conventional
MBE growth on (100) GaAs substrates. This PD operates in the blue-ultraviolet
optical region at room temperature. The diode fabricated has fairly good forward-­
bias characteristics with a built-in voltage Vbi ~ 1.5 V, and an ideal diode factor
(n) ~ 1.3–1.4. The open-circuit photo-voltage (Voc) for the blue incident light with
an intensity of 10 μW is 1.55 V, which is about five times larger in magnitude than
13 ZnSe-Based Photodetectors 307

Fig. 13.4 (a) A schematic structure of a ZnSe p+– n avalanche photodiode grown on GaAs by
MBE. The surface ohmic contact layer consists of ZnTe–ZnSe multi-quantum wells (with five
periods) and a thin p+-ZnTe cap layer (10 nm); (b) Spectral response of an external quantum effi-
ciency in the p+– n ZnSe photodiodes under zero-and reverse-bias conditions. (Reprinted with
permission from Ref. [31]. Copyright 2000: AIP)

the case in conventional Si photodiodes. As it is seen in Fig. 13.4b, in zero-bias


conditions the efficiency with the maximum value in the fundamental absorption-­
edge region is about 38%. The zero-bias efficiency is not so high because the active
region (depletion region of the n-ZnSe layer) is small ~0.2 μm, compared to the
thickness of the ZnSe layer. However, one can see a существенное increase in the
efficiencies with applying reverse bias. In high reverse-bias condition of 15 V, the
efficiencies are found to increase up to 60%. According to Ishikura et al. [31],
reported parameters of the PD were achieved due to optimization of the MBE pro-
cess, perfect lattice matching of the ZnSxSe1-x buffer (x = 5.6–6.0%) on GaAs sub-
strates, high-quality initial preparation of the interface, and correctly chosen doping
conditions. This made it possible to significantly improve the quality of the epitaxial
layers and reduce the dislocation density below 1 × 105 cm−2. Росту
чувствительности способствовало также использование an extremely thin sur-
face contact layer which consisted of ZnTe (well)-ZnSe (barrier) multi-quantum
wells. The high quality of the structure allowed the developed PD to work in the
multiplication mode with an avalanche gain of G = 60 under the electric field
8 × 105 V/cm.
Subsequently, based on the developed technology for growing ZnSe p+–n struc-
tures by MBE, Ishikura et al. [30] developed ZnSSe and ZnSe p+–n avalanche pho-
todiodes (APDs) with back surface photo excitation (hole injection) (see Fig. 13.5).
GaAs substrate removed by wet etching. APDs were characterized by a very stable
308 G. Korotcenkov

Fig. 13.5 Schematic structure of ZnSe (binary) or ZnSSe (ternary) APD with back excitation
grown by MBE on n + -GaAs substrate. (Reprinted with permission from Ref. [30]. Copyright
2002: Wiley)

high field operation with extremely low dark current (~1 pA/mm2). The ZnSSe
APDs on GaAs had a large avalanche gain of G > 60 under the electric field of
~1 × 106 V/cm at RT. ZnSe APDs had an avalanche gain of G ~ 50.

13.4 The p-i-n Junction Photodiodes

As shown in the previous section, p-n junction PDs can have a fairly high an exter-
nal quantum efficiency. However, such PDs suffer from the following disadvan-
tages. First, an external quantum efficiency depends on the bias voltage, and high
values ​​are achieved with high reverse-bias voltages. Second, p-n junction PDs usu-
ally have an extended base (Fig. 13.6a) in which only diffusion, not drift, processes
take place. As is known, diffusion proceeds much more slowly than drift. And this
leads to the fact that PDs with a p-n junction are characterized by a large inertia and
a reduced efficiency of carrier collection with a significant base thickness. The car-
riers generated in the base, upon reaching the space charge region (SCR), contribute
to the photocurrent, but do so with a delay depending on their position relative to the
SCR. Holes formed at different distances from the edge of the depletion region have
different diffusion times, which leads to a diffusion “tail” in the response of the
13 ZnSe-Based Photodetectors 309

Fig. 13.6 Schematic diagram of (a) p-n and (b) p-i-n photodiodes. In a p-n junction photodiode,
charge carriers may be created in a high-field drift region or a low-field diffusion region. In a PIN
photodiode, charge carriers are mostly created in the high-field drift region, which extends almost
to the far electrode. The lightly n-doped “intrinsic” region has a nearly constant E field, which
sweeps charge carriers through the device without diffusion

photocurrent to a rectangular light pulse. This type of signal distortion is generally


undesirable.
The solution of the problem of charge carrier diffusion is to simply eliminate the
diffusion region. This can be achieved by decreasing the donor concentration in the
n-region until the depletion region occupies almost the entire space between the
electrodes. As shown in Fig. 13.6b, the E field extends almost the entire distance to
the far electrode, so charge carriers generated anywhere in the material will drift
rather than diffuse. As for the p+ and n+ layers necessary to create a good ohmic
contact, in order to exclude diffusion processes in them, in the manufacture of pho-
todiodes, they tend to make the p+ and n+ layers as thin as possible. Since the middle
region is very lightly doped (almost intrinsic), it is labeled i and the device is termed
a p-i-n photodiode.
It is important to note that the use of the p-i-n structure not only eliminates car-
rier diffusion, but also has the advantage that the depletion width d is fixed by the
device geometry. The ability to adjust d according to design rather than applied
voltage allows the photodiode performance to be optimized for specific applica-
tions. For example, increasing d increases the length of the light absorption path,
which increases the efficiency of light absorption. This is especially important for
wavelengths close to the semiconductor band gap. However, a larger value of d
worsens the response time by increasing the transit time. Therefore, when develop-
ing p-i-n PD, a compromise is sought.
310 G. Korotcenkov

Fig. 13.7 (a) Conventional p-i-n structure APD on n-GaAs with surface SLE. The thickness of
window layer (W) is 0.57 μm, including surface SLE. (Reprinted with permission from [4].
Copyright 2002: Wiley); (b) n+-i-p structure ZnSSe APD with a simple thin n+ window (W = 30 nm
thickness) grown on p-GaAs substrate. Interface superlattice buffer layer (SLE) of total thickness
of 11.7 nm is inserted between p-ZnSe and p-GaAs. (Reprinted with permission from [43].
Copyright 2006: Wiley)

Implementation examples of ZnSe-based p-i-n PDs are shown in Fig. 13.7. High
gain and high sensitive blue-ultraviolet avalanche photodiodes (APDs) were devel-
oped using high quality ZnSSe n+-i-p heterostructure grown on p- and n-type GaAs
substrates by molecular beam epitaxy (MBE) [2, 4, 32]. The same technology was
used for fabrication of ZnSe-based p-n junction PDs [31]. An essential factor of the
p–i–n diode is a high quality undoped i-layer (active layer) with high resistivity of
over 106 Ω·cm. For n- and p-type conduction control, a chlorine (Cl) and an active
nitrogen (N) by rf-radical doping [56] were used, respectively. The difference in two
device structures is a position of superlattice buffer layer (SLE). In device on
n-GaAs substrates, SLE of ZnTe-ZnSe MQWs is formed to obtain a hole ohmic
contact in p-ZnSSe window layer and top metal (Au) electrode. This surface SLE
has caused severe disadvantage such as incident light absorption loss and device
instability under high electric field operation. The structure n+-i-p on p-GaAs has no
surface SLE, in turn, very thin interface superlattice of 10 layers of ZnTe/ZnSe
MQW (total MQW thickness = 11.7 nm) is formed, which is shown in the inset of
the Fig. 13.7b. The role of this interface superlattice is to overcome energy barrier
(~1 eV) for hole ohmic conduction between p-type GaAs and p-type ZnSe hetero-­
interface. How important the presence of SLE at the interface with the substrate can
be seen from the results shown in Fig. 13.8a. One can see a drastic improvement in
the I–V characteristics in the photodiode with the interface SLE. The Al2O3 coating
also increases the sensitivity of the PD [43].
Ando et al. [4], Abe et al. [2, 3] and Miki et al. [43] showed that the use of a
p-type GaAs substrate and ZnSxSe1-x (x = 5.5–6%) instead of ZnSe gives a notice-
able improvement in the PDs parameters. ZnSxSe1-x has a lattice parameter well
matched to that of GaAs substrate. APDS fabricated on p-GaAs substrate operated
in blue-violet-ultraviolet (450–300 nm) optical region (see Fig. 13.8c). The
avalanche-­gain reached G = 91.6 at 33 V (E = 6.5 × 105 V/cm). Abe et al. [2] also
13 ZnSe-Based Photodetectors 311

Fig. 13.8 (a) Current– voltage characteristics of ZnSSe PIN/p+-GaAs with interface SLE, in com-
parison with ZnSSe PIN/p+-GaAs without interface SLE. A built-in voltage of (a) is drastically
improved below 2 V. (Reprinted with permission from Ref. [43]). Copyright 2006: Wiley; (b)
Improved dark current characteristics of the n+-i-p structure on p+-GaAs in comparison with a
conventional p+-i-n structure on n+-GaAs. For comparison, a GaN-based PIN and a Si-PIN are also
shown. The dark current of ZnSSe p-i-n on p+-GaAs is below sub pA/mm2. (Reprinted from Ref.
[42]). Published 2008 by Korean Physical Society as open access; (c) Spectrum sensitivities of the
n+-i-p structure ZnSSe APD in different reverse bias voltages. The highest sensitivities of 5 A/W
for blue region (450 nm) and 3 A/W for ultraviolet region (300 nm) are obtained at reverse bias
32.8 V. (Reprinted with permission from Ref. [2]). Copyright 2005: Japan Society of Applied
Physics; (d) Photo-voltages of ZnSe and ZnSSe p–i–n photodiodes vs. incident optical power.
Here wavelength of 455 nm (blue region) is used. (Reprinted with permission from Ref. [4].
Copyright 2002: Wiley)

established that PDs had high sensitivities (~3 A/W) even for ultraviolet optical
region (λ = 300 nm). According to Abe et al. [2], such sensitivity was achieved due
to following two effects; (i) thin n+ window layer acting as semi-transparent win-
dow, and (ii) efficient hole impact ionization process in spin-orbit splitting valence
312 G. Korotcenkov

band. It was found that such PDs had small dark current (<10−9 A/ mm2) under
intrinsic avalanche breakdown region, which is one order smaller than in ZnSSe
APDs on n-GaAs, and is smaller in two orders than in GaN based-APDs [2]. A
noticeable decrease in the dark current was achieved by reducing the density of
macro- and microscopic point defects in the developed devices. The same results
were reported by Miki et al. [42]. These results are shown in Fig. 13.8b. Another
important parameter of ZnSe–ZnSSe p–i–n devices is the large open-circuit photo-­
voltages (Voc ~ 1.6–1.7 V) [4]. Voc is about eight times larger in magnitude than that
of practical Si photodiodes (see Fig. 13.8d). Ando et al. [4] also suggested that
ZnSSe p-i-n PDs are stable and can have a long device life in excess of 10,000 h
because they did not observe any change in device parameters and film properties
after 200 h of testing.
Also noteworthy are the developments of p-i-n PDs based on ZnSe-based com-
pounds such as Zn(Mg)BeSe [79, 80] and ZnMgSSe [22]. ZnMgBeSe quaternary
alloys with a lattice parameter matched to that of GaAs can have a band gap in the
range of 2.75 to more than 3.8 eV. These compounds have high crystal quality,
which should provide high sensitivity and detectability. The lattice mismatch
between the ZnMgSSe layers and the GaAs substrate is only 0.03%.

13.5 Schottky Photodiodes

The metal-semiconductor contact, or Schottky barrier, is the equivalent of a p-i-n


photodiode in the short wavelength region of the spectrum. Photons pass through a
partially transparent metal layer (often gold) and are absorbed in the space charge
region (SCR) of an n-type semiconductor. The metal film provides much less series
resistance than a conventional p + −n junction, and the parasitic absorption of short-
wave radiation in the metal film is less than in the p + region due to the large thick-
ness difference. Another advantage of the Schottky photodiode compared to p-n (or
p-i-n) structures is the improved kinetics of the photoresponse. Because it lacks a
p-type layer, there is no residual “diffusion tail” arising from charge carriers gener-
ated in the p-type layer. This becomes especially important at short wavelengths,
when, at a high absorption coefficient, a significant amount of light is absorbed in a
p-n photodiode by a thin p-type layer. As in the case of a p-i-n photodiode, the “dif-
fusion tail” in the response associated with the presence of a diffusion region in the
n-type layer can be minimized by adjusting the concentration of donors in the
n-region in such a way that the depletion region extends to the n+-layer. Additional
advantages of these devices are associated with their relative ease of manufacture
compared to p-n and especially p-i-n photodiodes. Photodiodes with a Schottky bar-
rier are technologically and physically compatible with integrated optics structures.
In addition, Schottky barrier photodiodes can be fabricated on a wide variety of
semiconductors, even those in which p-n junctions cannot be obtained.
13 ZnSe-Based Photodetectors 313

It is important to note that photodiodes with a Schottky barrier are most effective
in the blue-UV region of the spectrum, since at longer wavelengths, due to the
reflection and absorption of light by the metal layer, their efficiency decreases. To
reduce light reflection, it is necessary to apply an anti-reflection coating, which
complicates the manufacture of such devices.
The greatest advances in the development of the Schottky barrier photodiode
have been made at the Center National de la Recherche Scientifique (Valbonne,
France) [10, 45, 79, 81, 82]. To fabricate the PD, they used undoped n-ZnSe layers
grown by MBE on GaAs substrates. Two types of structures have been fabricated.
The first ones are vertical geometry devices which consist of two layers: one nonin-
tentionally doped (NID) layer which is “the absorption layer” and one Cl-doped
n+-type layer which is “the contact layer” (see Fig. 13.9a). The second ones are
planar geometry devices which consist of a unique NID layer (see Fig. 13.9b). In
theory, the vertical structure allows high-quality ohmic contacts (ohmic contact on
a heavily doped layer) and low capacitance (Schottky contact on an undoped layer).
In practice, the etching needed to realize that the mesa structure generates defects
which might lead to bandwidth narrowing and noise increase. A semi-transparent
Ni/Au layer (5 nm/5 nm) was used as a Schottky barrier. I-V characteristics had an

Fig. 13.9 A schematic representation of the (a) vertical and (b) planar geometry Schottky barrier
photodiode. (Reprinted with permission from Ref. [77]. Copyright 2001: Springer)

Table 13.1 Performances of ZnSe-based Schottky barrier PDs


Maximum
Wavelength Bias responsivity Quantum Rejection Detectivity
Device Material cutoff (nm) (V) (A/W) efficiency rate (mHz1/2 W−1)
Schottky ZnSe 460 -2 0.1 27% 3·103 -
vertical −3.5 0.115 31% 3·103 1.4·1010
Schottky ZnSe 460 −5 0.125 33% 103 -
planar −10 0.127 34% 103 1.3·1011
Schottky ZnMgBeSe 375 −2 0.1 31% 5·103 2·1010
planar −3 0.115 38% 5·103 1.5·1010
−10 0.145 57% 5·103 -
Source: Data extracted from Vigué et al. [78, 80]
314 G. Korotcenkov

Fig. 13.10 (a) Spectral response of ZnSe- based Schottky photodiode at −2 V bias. (Reprinted
with permission from Ref. [45]). Copyright 2000: AIP; (b) Spectral response of a ZnMgBeSe
Schottky photodiode under 10-V bias. (Reprinted with permission from Ref. [78]. Copyright 2001:
Springer)

ideality factor of 1.0. The ideality factor, very close to unity, underlines the high
quality of the Schottky junction. The potential barrier height was
1.17–1.19 eV. Table 13.1 displays the results obtained on planar and vertical geom-
etry Schottky devices based on ZnSe. Spectral characteristics are shown in
Fig. 13.10.
A comparison of the parameters of the vertical and planar PDs showed that verti-
cal devices are somewhat less sensitive than planar ones. The detectivity of a planar
PD was ten times higher than that of a vertical one. This effect of Vigué et al. [79]
explained on the basis that a vertical device had a much larger dark current due to
leakage along the surface of the mesa structures. On the other hand, the rather low
quality of ohmic contacts made on a high-resistance NID layer in planar structures
makes it necessary to use a sufficiently high voltage to sufficiently bias the pla-
nar device.
Approximately the same parameters were demonstrated by Ni-ZnSe PDs devel-
oped by Naval et al. [50] and Voronkin [83]. These PDs were fabricated using
single-­crystal ZnSe substrate. Spectral sensitivity was in the range from 200 nm to
460 nm Voronkin [83]. A detector without a UV filter showed a maximum respon-
sivity of about 0.11 A/W at 375 nm wavelength Naval et al. [50] and 0.1 A/W at
400 nm Voronkin [83]. The speed of the unfiltered detector was found to be about
300 kHz primarily limited by the RC time constant determined largely by the detec-
tor area [50].
Given the large surface leakage in ZnSe mesa structures, studies have been car-
ried out to reduce them. In particular, Makhniy et al. [40, 41] proposed to treat the
free ZnSe surface around mesa-structure with hydrogen peroxide to reduce the dark
current in vertical ZnSe-based PDs. The effect of such treatment on the reverse I-V
characteristics of Ni-ZnSe Schottky barriers is shown in Fig. 13.11. It can be seen
that such treatment leads to a significant decrease in the reverse current (Fig. 13.11b)
and an increase in the breakdown voltage (Fig. 13.11a). The most likely reason for
the decrease in reverse current in surface-modified diodes is the formation of a
13 ZnSe-Based Photodetectors 315

Fig. 13.11 (a) Reverse current–voltage characteristics of initial (1) and modified (2) diodes in the
high current region; (b) Comparison of the reverse current–voltage characteristics of initial (1) and
modified (2) diodes. (Reprinted with permission from Ref. [40]. Copyright 2003: Wiley)

high-­resistance zinc oxide layer on the free surface of ZnSe, which acts as a pas-
sivating coating and leads to the suppression of leakage currents. The penetration
depth of this layer in ZnSe is small and practically equal to the thickness of the zinc
oxide layer, but it seems to be quite sufficient to reduce the possibility of surface
breakdown.
Monroy et al. [45] and [79], have also shown that ZnMgBeS is also a very prom-
ising material for the development of Schottky barrier PDs. The spectral response of
such PDs measured under – 2 V bias voltage is shown in Fig. 13.10b. It is seen that
a very flat response is obtained at short wavelengths. A quantum efficiency over
50% is obtained from 380 to 315 nm, that is, over the whole UV-A range. Over the
UV-B range (from 315 to 280 nm), a quantum efficiency over 35% is obtained. It is
to be noted that a quantum efficiency of 57%, as obtained at 360 nm, is close to the
theoretical limit taking into account the fact that this detector is not anti-reflection
coated. Moreover, the cutoff at 380 nm is very sharp with a UV/visible contrast of
more than three orders of magnitude: this is a consequence of the growth of a high-­
quality layer, quasi-lattice matched to the GaAs substrate.

13.6 Metal–Semiconductor–Metal Photodiodes

It should be noted that the metal–semiconductor–metal structure is most often used


in the development of photodetectors based on ZnSe. A metal–semiconductor–
metal photodetector (MSM PD) is a planar device containing two Schottky con-
tacts, i.e., there are two metallic electrodes on a surface of semiconductor material.
During operation, some electric voltage is applied to the electrodes, and therefore
one Schottky barrier is turned on in the forward direction and the other in the reverse
direction. When light impinges on the semiconductor between the electrodes, it
316 G. Korotcenkov

Fig. 13.12 (a) Cross-sectional diagram of the integrated MSM-PD/HBT photoreceiver; (b)
Surface SEM of the fabricated MSM-HBT photoreceiver. (Reprinted with permission from [17].
Copyright 2009: IOP)

generates charge carriers (electrons and holes), which are collected by the electric
field and thus can form a photocurrent.
MSM PD design typically uses an interdigital electrode structure (see Fig. 13.12)
where the finger spacing can be as small as 1 μm. Light entering the device from the
side of the electrodes is partially blocked by the electrodes, which, of course,
reduces the quantum efficiency. For higher quantum efficiency, there are back-illu-
minated devices in which the light comes from the back side so that the electrodes
do not obstruct it. Another possibility to reduce top exposure losses is to use
extremely thin semi-transparent gold contacts. The advantage of overhead illumina-
tion is that the achievable detection bandwidth is usually larger as the charge carri-
ers are generated closer to the contact.
The MSM PD design typically uses an interdigital electrode structure where the
finger spacing can be as small as 1 μm. The light entering the device from the side
of the electrodes is partially blocked by the electrodes, which naturally reduces the
quantum efficiency. For higher quantum efficiency, there are back-illuminated
devices in which the light comes from the back side so that the electrodes do not
obstruct it. Another possibility to reduce top-irradiation losses is to use extremely
thin semi-transparent gold contacts. The advantage of top-illumination is that the
achievable detection bandwidth is usually larger as the charge carriers are generated
closer to the contact.
An important advantage of MFM PD is a rather simple planar structure, which is
convenient for monolithic integration with other components of photonic integrated
circuits. There is also no need to form a low-resistance contact, which is a difficult
task for II-VI compounds, during the MSM PD fabrication. In addition, MSM pho-
todiodes can be made with a faster response than photodiodes. UV photodetectors
based on MSM structures have a very low capacitance in the fF range and, therefore,
can operate at high frequencies. Their detection bandwidth can reach hundreds of
GHz (with an impulse response of less than 1 ps), making them suitable for high-­
speed fiber optic communications. As for ZnSe-based MSM PDs, their monolithic
13 ZnSe-Based Photodetectors 317

integration with InGaP/GaAs heterojunction bipolar transistors (HBTs) has been


demonstrated by Chen and Chang [16–18]. Figure 13.12 shows one such device.
To develop ZnSe-based MSM PDs, undoped high-resistance ZnSe epitaxial lay-
ers grown on a GaAs substrate by molecular beam epitaxy (MBE) are commonly
used. Due to the high quality structure and low concentration of charge carries in
n-ZnSe, MSM photodetectors have a high resistivity, and a high bias voltage can be
applied. This opens up the possibility to enhance the responsivity of ZnSe-based
MSM UV photodetectors using internal gain mechanisms employing impact ioniza-
tion (avalanche breakdown effect) due to the higher internal electric field
strength [69].
As regards the performances of the MFM-PCs developed on the basis of ZnSe,
they are summarized in Table 13.2. The MFM-PCs based on ZnSe have the same
spectral characteristics as the Schottky barrier PDs discussed earlier. To achieve
improved performance of UV photodetectors, a high Schottky barrier height at the
metal-ZnSe interface is required. The high barrier height would result in low leak-
age current and high breakdown voltage, allowing operation at high bias voltage.
This leads to an improvement in the sensitivity and response kinetics of such UV
photodetectors, as well as a reduction in the noise level. Among different metals, the
metals with high work functions like Ni, Cr, Pd, or Au are required to achieve a
large Schottky barrier height on n-ZnSe (read Chap. 17, Vol. 1). In practice, a large
group of metals such as Au [78], Ti/W [13], Pd [24], Cr/Au [69] and Ni/Au
(25/140 nm) [13, 67] was used to fabricate the PD. Due to amphoteric properties
[51], Au atoms may create both donor and acceptor centres during diffusion to
ZnSe, while Cr and Ni create deep acceptor centres in ZnSe [59], which compensate
for uncontrolled background donor impurities in ZnSe. Therefore, Sirkeli et al. [67]
believe that Cr/Au and Ni/Au are attractive for the fabrication of the Schottky bar-
rier structures with good characteristics. However, according to Sirkeli et al. [67], it
is necessary to give preference to the use of Ni/Au contacts with a higher potential
barrier height (Table 13.2). For the deposition of metals forming a Schottky barrier,
either e-beam evaporation for refractory metals and thermal evaporation for gold are
typically used.
Chang et al. [13] and Lin et al. [37] have shown that transparent conductive ITO
(Indium Tin Oxide) layers can also be used for this purpose. In this case, it is the
ITO layer that provides the maximum transparency of the contact (Fig. 13.13a),
while the Ni/Au films have the minimum resistance [13]. However, although trans-
parent ITO contact electrodes can provide detectors with greater photon absorption
and higher photocurrent, the low height of the Schottky barrier between ITO and
ZnSe results in a relatively large dark current (Fig. 13.13c).
During the development of ZnSe-based MSM PDs, the following regularities
were also established:
• The breakdown voltage depends on the distance between the electrodes. For
example, Hong and Anderson [24] reported an increase in breakdown voltage
from 25 to 50 V with increasing distance from 2 to 4 μm.
318

Table 13.2 Performances of ZnSe-based MSM PDs with interdigital contacts


R, S, Φb, Responsivity, Detectivity,
Ω·cm ID width/space, μm μm2 Contact eV Dark current A/W Response time Bandwidth 1010 cm Hz1/2 W−1 Ref.
MBE (2–4)/(2–4) 4·104 Pd 1.3 1 pA (10 V) 1.0 (380 nm, 5 V) 620–800 MHz – [24]
0.6 (450 nm, 5 V)
VP;~ 1012 0.5/1.5 102 Cr/au 1.26 1.64 nA (5 V) 2.23 (325 nm, 15 V) 0.10–0.16 ms 9.9 [67]
Ni/au 1.49 0.82 nA (5 V) 5.4 (325 nm, 15 V) 0.10–0.16 ms 33.7
MBE 6/6 Au 0.13 (460 nm, 20 V) 1–2 [78]
VP: 1010–1012 0.5/1.5 102 Cr/Au 1.26 3.4 nA (20 V) 4.44 (325 nm, 20 V) 0.15 ms 14 [68]
MBE; 0.6 10/10 4·104 ITO 0.78 0.13 (450 nm, 1 V) 0.04 ms [37]
MBE 10/10 5·105 ITO 1.5 nA (1 V) 0.128 (450 nm, 1 V) 93 [38]
MBE 10/10 5·105 ITO 0.66 1.5 nA (1 V) 0.120 (450 nm, 1 V) 87 [13]
Ti/W 0.695 0.35 nA (1 V) 0.051 (450 nm, 1 V) 41
Ni/au 0.715 0.20 nA (1 V) 0.028 (450 nm, 1 V) 77
MBE 20/20 106 ITO 4 nA (5 V) 0.08 (440 nm, 5 V) [18]
ID – interdigital; MBE- molecular beam epitaxy; R – resistivity; S – active area; VP-vapour phase method; Φb,- the height of potential barrier
G. Korotcenkov
13 ZnSe-Based Photodetectors 319

Fig. 13.13 (a) Normalized transmission spectral of ITO, TiW and Ni/Au; (b) Dark I–V character-
istics of the photodetectors with ITO, TiW and Ni/Au contact electrodes. (Reprinted with permis-
sion from [13]. Copyright 2006: Elsevier)

Fig. 13.14 (a) Variation of the quantum efficiency with the applied bias measured at 460 nm for a
(6x6) MSM ZnSe photodetector. (Reprinted with permission from Ref. [78]). Copyright 2001:
Springer; (b) I–V characteristics of the homoepitaxial and heteroepitaxial ZnSe MSM photodetec-
tors with ITO transparent contact electrodes measured both in dark and under illumination.
(Reprinted with permission from Ref. [38]. Copyright 2005: Elsevier)

• Devices with minimal electrode spacing have improved device speed due to the
short carrier transit time between electrodes [24].
• Quantum efficiency increases as the characteristic size of ID contacts (width/
space) in MSM PDs [78] decreases (see Fig. 13.14a).
• ZnSe-based MSM PDs based on homoepitaxial layers, due to better quality, have
better performance compared to devices using heteroepitaxial layers [37, 38].
I-V characteristics of such PDs are shown in Fig. 13.14b. It is known that the
slight lattice mismatch (0.27% at room temperature) between ZnSe and GaAs
will still generate a huge amount of defects when we grow a very thick ZnSe
epitaxial layer on top of the GaAs substrate [85]. The defects generated at the
ZnSe–GaAs interface will significantly reduce the efficiency of the ZnSe-based
photodetectors.
320 G. Korotcenkov

• The RC time of the measurement system is the main limiting factor of frequency
bandwidth of ZnSe-based MSM UV photodetectors [67, 69]. As reported by
Monroy et al. [46], the response time is linearly dependent on load resistance and
can be reduced by 2–3 orders of magnitude by reducing the load resistance from
MΩ to kΩ-range.

13.7 Heterostructure-Based Photodetectors

As a rule, heterostructure-based structures in photodetectors are used to create a


transparency window or to develop photodetectors that are selective in the required
spectral region. To do this, a semiconductor with a larger band gap and low resistiv-
ity is deposited on the surface of the active photosensitive material. Such a wide-gap
window transmits radiation in the active region without significant losses and, at the
same time, is a contact layer with a low series resistance.
Processes in the active region of heterojunctions - absorption of radiation and
generation, separation and accumulation of generated charge carriers - proceed in
the same way as in Schottky barriers or a p-i-n structure. The difference lies in the
fact that by choosing a suitable semiconductor for the photosensitive layer, it is pos-
sible to ensure complete absorption of radiation at a thickness of this layer of the
order of 1 μm. The most important advantage of heterostructure-based photodetec-
tors (HPDs) is also their physical and technological compatibility with integrated
optics devices. The freedom to choose the HPD materials also makes it possible to
achieve higher photo-emf values. However, HPDs are much more difficult to fabri-
cate, require matching of the lattice parameters of the materials forming the hetero-
junction, and they have an increased noise level.
It is important to note that in HPDs containing ZnSe, ZnSe can act both as a
photosensitive material [19, 74, 86] and as a wide-gap window [14, 34, 39], which
cuts off the photoresponse at λ < 450 nm if the ZnSe layer is sufficiently thick.
There are currently reports on the development of HPDs such as ZnSe/Si [14, 60],
ZnSe/InSe and ZnSe/GaSe [34], ZnSe/ZnTe [60], Zn1-xCdxTe /ZnSe [76], ZnO/
ZnSe [74, 86], ZnSTeSe/ZnSe/GaAs [19]. ZnSe/Si HPDs, whose schematic cross
section and schematic band diagram are shown in Fig. 13.15, have a sensitivity
range of 300–800 nm, while ZnSe/InSe and ZnSe/GaSe HPDs are sensitive in the
range of 850–450 and 550–450 nm, respectively (see Fig. 13.16). InSe and GaSe are
indirect band gap semiconductors with Eg equaled 1.25–1.4 eV and 2.0–2.1 eV,
respectively. ZnSe/ZnTe HPDs developed by Rao [60] are sensitive in the range of
300–625 nm with one large maximum at around 455 nm, which corresponds to the
band gap of ZnSe, and a second smaller peak at around 550 nm, which corresponds
to energy bandgap of ZnTe (2.5 eV). At the same time, ZnSe/Si HPDs because of
the narrow bandgap of silicon, exhibit considerably higher amount of photocurrent
in longer wavelength up to 900 nm. However, large maximum lies at around 450 nm
as in ZnSe/ZnTe HPDs. The measured dark current in ZnSe/Si p-i-n HPDs was as
small as 0.2 nA/μm2 at 8 V. The maximum photosensitivity of this device was 2.5
13 ZnSe-Based Photodetectors 321

Fig. 13.15 (a) The schematic cross section of the ZnSe/Si photodiode. The PIN-like visible pho-
todiode was grown on a (111)-oriented p-Si substrate using vapor phase epitaxy technique; (b-d)
The schematic energy-band diagrams of the ZnSe/Si photodiode (b) at equilibrium, under (c)
reverse and (d) forward biased conditions, respectively. (Reprinted with permission from Ref. [39].
Copyright 1996: Elsevier)

Fig. 13.16 (a) Spectral characteristics of the (1) n-ZnSe/p-GaSe heterojunction at 300 K, (2)
n-ZnSe/p-InSe heterojunc tion at 300 K, and (3) n-ZnSe/p-InSe heterojunction at 120 K; (b)
Energy band diagram for the n-ZnSe/p-InSe heterojunction. (Reprinted with permission from Ref.
[34]. Copyright 2014: Springer)
322 G. Korotcenkov

A/W. ZnSe/Si HPDs developed on the basis of porous silicon (PSi) (n-ZnSe/PSi/
p-Si structure) had a significantly lower sensitivity, which did not exceed 0.03
A/W [14].
According to Sunaina et al. [74], a Type II band alignment of ZnO and ZnSe in
the heterostructure leads to slower recombination of electron-hole pairs after illumi-
nation. These ZnO-ZnSe heterostructures were synthesized by a simple hydrother-
mal method. ZnO-ZnSe HPDs gave fast rise and decay times of 23 and 80 ms,
respectively, and relatively high photo responsivity (0.091 A/W). The spectral
region of maximum sensitivity of such HPDs lies in the range of 380–460 nm and
is determined by the band gaps of ZnO (3.2 eV) and ZnSe (2.7 eV).
Chen et al. [19] have shown that using ZnSTeSe/ZnSe/GaAs heterostructure it is
possible to develop high sensitive PDs, which could absorb light with wavelength
between 400 and 900 nm. In other words, the spectral width of this ZnSTeSe photo-
diode could reach 500 nm. In the long wavelength side, a sharp cutoff occurs at
around 870 nm, which corresponds to the band gap of GaAs. In the short wave-
length region, sensitivity is limited by the band gap of ZnSTeSe. The maximum
quantum efficiency of the fabricated ZnSTeSe photodiodes was around 75%.
Responsivity in this spectral range was 0.4–0.5 A/W. High quality epitaxial layers
of ZnSe and ZnSTeSe were grown by MBE. ZnSTeSe epitaxial layers perfectly
matched to the GaAs substrate used.
Another option for using ZnSe-based heterostructures in PDs is the development
of PDs based on superlattices such as ZnSe/ZnTe [6, 7, 35], ZnSe/MgCdS [77],
ZnSe/ZnS [5]. It was shown that with the use of such multilayer heterostructures it
becomes possible to create selective multicolor photodetectors with high sensitivity.
HPDs developed by Averin et al. [6, 7] and Kuznetzov et al. [35] had responsivity
from 0.18 A/W to 0.28 A/W. As for ZnSe/MgCdS superlattice based PDs [77], no
progress has been achieved here. The responsivity of such PDs in the region of
maximum (400–450 nm) was only about 0.001 A/W.

13.8 Phototransistors

Despite the promise of using phototransistors in optoelectronics, there are only a


few reports on the development of film and monolithic phototransistors based on
ZnSe. In particular, Sun et al. [73] developed a phototransistor based on a hybrid
structure: graphene decorated with ZnSe/ZnS core/shell quantum dots (see
Fig. 13.17). Monolayer graphene grown by CVD was decorated with ZnSe/ZnS
core/shell QDs using a simple solution method. The devices were fabricated on sili-
con substrates with a bottom gate. Using this structure, Sun et al. [73] achieved a
high responsivity above 103 AW−1 for UV light. The photodetectors exhibited a
selective photo responsivity for the UV light with the wavelength of 405 nm, con-
firming the main light absorption from QDs. According to Sun et al. [73], the
13 ZnSe-Based Photodetectors 323

Fig. 13.17 (a) Schematic illustration of a graphene FET decorated with ZnSe/ZnS quantum dots
under the illumination; (b) The TEM image of the ZnSe/ZnS quantum dots on the SiO2 substrate
with different scale bars of 10 nm and 5 nm, respectively. (Reprinted from Ref. [73]. Published
2018 by Nature as open access)

sensing mechanism is attributed to the photo-generated charge carriers in the QDs


and the transfer process from QDs to graphene, which may directly modulate the
Fermi Level of graphene and the conductance of graphene channel. This charge
transfer results in a gate-controlled photosensitivity with a maximum value obtained
at VG of about 15 V. It is important to note that the phototransistors showed good
stability during the 21 days test. During this time, responsivity decreased by only
13%. The response time of devices developed was about 0.5 s. Sun et al. [73] believe
that the response time is limited by the ligand covering the surface of QDs. Therefore,
ligand optimization is required to further improve the response time of
phototransistors.
Chang and Wu [15] have proposed a different approach. They developed a pho-
totransistor based on n-ZnSe/p-Si/n-Si heterojunction. The cross-sectional sche-
matic of the studied ZnSe/Si heterojunction phototransistor is shown in Fig. 13.18a.
The 0.45 μm n-ZnSe emitter epilayer was grown on a (111) Si substrate utilizing an
ordinary infrared (IR) furnace CVD system and a two-step process to overcome the
lattice constant mismatch of approximately 4.1%, existing between the ZnSe epi-
layer and the Si substrate. The resulting photo-current responsivity was approxi-
mately 50 A/W at a bias voltage of VG = 15 V, with the highest photo-current
responsivity occurring at a wavelength of 470 nm (Fig. 13.18b). It is seen that due
to большой толщине n-ZnSe слоя, sensitivity is low in the spectral range
λ < 400 nm. As the intensity of the incident light increases, the photosensitivity
decreases. In addition, a strong photo-responsivity was also observed at a longer
wavelength of 700 nm, due to the absorption of impinging photons in the Si sub-
strate region. With a collector current density of 65 mA/ cm2, the I-V current gain
was approximately 28.
324 G. Korotcenkov

Fig. 13.18 (a) Schematic cross section of the n-ZnSc/p-Si/n-Si heterojunction phototransistor (b)
The photocurrent and photo responsivity as a function of the wavelength of the incident light under
the fixed incident intensity of 100 lux. (Reprinted with permission from Ref. [15]. Copyright 2000:
The Institution of Engineering and Technology)

13.9 Nanowire-Based ZnSe Photodetectors

The Chap. 12 (Vol. 1) describes the main methods for synthesizing 1D nanostruc-
tures, and the Chap. 16 (Vol. 2) describes in detail all the approaches that can be
used in the development of photodetectors based on 1D nanostructures of II-VI
compounds. In the same Chap. 16 (Vol. 2) one can find an analysis of the photosen-
sitivity mechanism of nanowires (NWs), as well as the advantages and disadvan-
tages of their use in the development of photodetectors. In principle, all approaches
discussed in the Chap. 16, Vol. 2 can also be used in the development of UV detec-
tors based on ZnSe 1D structures. For example, PDs such as photoconductive PDs
[63], field effect transistors [52], Schottky diodes based PDs [88], PDs with p-n
junctions [52], PD with radial p-n heterojunctions [53], core-shell nanowire based
PD [55], and axial nanowire photodetectors [47] have been developed based on
ZnSe nanowires. In most studies, Cu/Au, Ti/Au or Au films were used as ohmic
contacts.
There are other approaches to the development of 1D-based photodetectors. For
example, Deng et al. [20] developed PDs based on ZnSe NWs decorated with PbSe
nanoparticles (NPs). They showed that the formation of PbSe-ZnSe heterojunctions
makes it possible to expand the spectral range of sensitivity in the visible and IR
regions of the spectrum, and thereby increase the responsivity of the PDs from 5
A/W to 40 A/W (see Fig. 13.19). According to Deng et al. [20], electrons are trans-
ported from PbSe to ZnSe while holes are trapped at PbSe, and electrons are
13 ZnSe-Based Photodetectors 325

Fig. 13.19 (a) Wavelength-dependent responsivity of ZnSe and ZnSe-PbSe NW-based photode-
tectors. (b) Energy band alignment between ZnSe-PbSe heterojunction NWs and Al electrodes.
(Reprinted from Ref. [20]. Published 2022 by Elsevier as open access)

Table 13.3 Performances of ZnSe nanostructure based photodetectors


Responsivity Photoconductive gain Response/recovery
Material (A/W) (EQE) times Ref.
ZnSe:P/Si 1.1× 105 2.9× 105 74/153 μs [71]
ZnSe:N/Si 2.4× 104 6.5× 104 0.4/0.5 s [72]
ZnSe:Bi/Al n.a. >103 0.38/0.4 s [88]
ZnSe:Cl/Si 4.1× 102 1.1× 103 2/2 s [84]
ZnSe:Sb/ZnO 5.2× 105 1.77 × 106 <1 s [52]
ZnSe 4.2 × 105 1.3 × 106 (EQE) 0.55/0.6 s [47]
ZnS0.49Se0.51/ 6.3 × 105 2.08 × 106 (EQE) 23/50 ms [47]
ZnSe

extracted to an external circuit by an Al electrode, resulting in a high gain of the


photoconductor.
However, the most common variants of NW-based photodetectors are photocon-
ductive 1D-based ZnSe photodetectors. In particular, in [36, 57, 88] one can find
reports describing the results of testing such photodetectors. Photodetectors were
fabricated both on the basis of single nanowires [23], and nanowires array [25]. Just
as in the case of film photodetectors, doping can be used to control the parameters
of nanowire-based PDs. Comparative characteristics of photodetectors based on
ZnSe NWs doped with different dopants, is given in the Table 13.3. Functional
principles of photoconductive 1D-based PDs in addition to Chap. 16 (Vol. 2) one
can find in [70].
As can be seen from Table 13.3, photodetectors based on single 1D nanostruc-
tures demonstrate high sensitivity and rate of response. But it should be recognized
that due to technological limitations (see Chap. 16, Vol. 2) it is unlikely that single
NW-based PDs will appear on the market in the near future. Oksenberg et al. [53,
54] developed a technology for the horizontal growth of oriented ZnSe NWs to
326 G. Korotcenkov

Fig. 13.20 Optoelectronic behavior of the guided ZnSe NWs: (a) A schematic illustration of the
device fabrication process. Ti/Au electrodes are deposited over guided ZnSe NWs that grow from
a patterned Au catalyst; (b) SEM images of the fabrication steps illustrated in (a); (c) I–V charac-
teristics of a typical device under different 405 nm illumination intensities and under dark condi-
tions; (d) The response of a typical photodetector set at a 30 V bias to on/off switching of a 405 nm
laser illumination (13 mW cm−2). (Reprinted with permission from Ref. [54]. Copyright
2015: Wiley)

solve this problem and fabricate photodetectors based on them (see Fig. 13.20).
However, this technology is still far from perfect.

13.10 Photodetectors Based on Hybrid Structures

At present, two approaches have been used in the development of ZnSe-based pho-
todetectors based on hybrid structures. Team from Tottori University (Japan) [27–
29] used organic–ZnSe hybrid structures, while Chakrabortya et al. [12] and Xu
et al. [87] developed photodetectors based on the ZnSe-graphene structure.
Chakrabortya et al. [12] used a ZnSe-decorated reduced graphene oxide (RGO)
nanocomposite and Xu et al. [87] used a two-layer graphene-ZnSe structure in
which ZnSe films with a thickness of 60 nm were deposited on graphene by e-beam
evaporation method. ZnSe-RGO nanocomposite for a large area thin film photode-
tector device was synthesized using a simple one step solvothermal reaction. Thin
film of RGO-ZnSe composite was prepared on a pre-cleaned glass substrate by
simple drop casting from dispersed solution in isopropyl alcohol (IPA). Chakrabortia
et al. [12] reported that the photosensitivity of the RGO-ZnSe photodetector was 3
times higher than that of controlled ZnSe PD. Chakrabortia et al. [12] believe that
this improvement is due to the specificity of the hybrid structure, in which ZnSe
nanoparticles are evenly decorated onto the surface of RGO nanosheets. Photocurrent
13 ZnSe-Based Photodetectors 327

Fig. 13.21 Photocurrent generation mechanism in our thin film device. (Reprinted with permis-
sion from Ref. [12]. Copyright 2017: Royal Society of Chemistry)

Fig. 13.22 (a) Schematic structure of organic–inorganic (PEDOT:PSS / ZnSSe) hybrid APD. (b)
Responsivity spectra in the operation bias region of ZnSe-based hybrid APDs fabricated by photo-
lithography. (Reprinted with permission from Ref. [27]. Copyright 2020: Springer)

generation mechanism for such structures is shown in Fig. 13.21. However, the
responsivity of such PDs was small (2 ma/W), and the time constants of photores-
ponse were too large to use such PDs as detectors. Response and recovery times
were about 29 and 41 seconds, respectively.
It should be noted that the photodetectors developed by Xu et al. [87] demon-
strated a significantly higher responsivity, which sharply increased with decreasing
channel width and reached 1.2 × 109 A/W for PD with 70 nm channel. At 5 μm
channel the responsivity was three orders of magnitude lower. Obviously, the shorter
channel allows the faster transport and more times recirculating of the charge carri-
ers and thus produces larger photocurrent and responsivity. As a result, devices with
a 70 nm channel had a response time of about 50 ms.
328 G. Korotcenkov

The approach based on the use of organic–inorganic hybrid structures, proposed


by the team from Tottori University, is also very promising. A view of the structure
of photodetectors developed within the framework of this approach is shown in
Fig. 13.22a. ZnSe-based structure (i-ZnSSe (<1015 cm−3) - active layer/n+-ZnSSe
(4 × 1018 cm−3) - contact layer) was grown by molecular beam epitaxy (MBE) on
n-type GaAs substrates. The organic UV-transparent conducting polymer layer of
poly 3,4-ethylenedioxythiophene:poly-styrenesulfonate (PEDOT:PSS) [48] with a
thickness of 0.15 to 1.0 μm was formed by spin-coating and a photolithography
technique. Smaller thickness provided lower absorption losses in the layer. The
absorption loss of the PEDOT:PSS layer with the thickness of 1 μm is around 60%
[29]. The energy barrier height of interface between organic (PEDOT: PSS) and
inorganic (i-ZnSSe) was about 1.0 eV [28]. It was shown that the developed photo-
detectors are high-gain and highly sensitive ultraviolet avalanche photodiodes
(UV-APDs). For the best samples, the leakage current before the breakdown voltage
was suppressed to <10−10 A/mm2 [27], and the maximum external quantum effi-
ciency reached 90% at 325 nm [28]. APDs also had high-speed photoresponse
(3.25 ns) [1]. The maximum responsivity was about 10 A/W (Fig. 13.22b).
The maximum multiplication factor due to technology optimization was
improved up to M = 3100 [27]. It is assumed that ultraviolet avalanche photodiodes
(UV-APDs) under development are promising devices for application as medical
imaging devices [e.g. positron emission tomography (PET)], astronomical measure-
ment, and next-generation large capacitive optical storage systems. As for the deg-
radation of the properties of the developed photodetectors, studies over 100 days
have shown that polyimide passivation of the developed hybrid APDs can signifi-
cantly improve the stability of their parameters [29]. The devices with polyimide
passivation worked stably throughout this period and retained their dark character-
istics with a steep avalanche breakdown (dark current less than 2 × 10−11 A/mm2).
For comparison, the dark current without polyimide passivation and N2 sealing rap-
idly increased over 4–5 days from 2 × 10−11 to 2 × 10−6 A/mm2 before the avalanche
breakdown.

Acknowledgments This research was funded by the State Program of the Republic of Moldova,
project 20.80009.5007.02.

References

1. Abe T, Inoue R, Fujimoto T, Tanaka K, Uchida S, Kasada H et al (2016) Development of ZnSe-­


based organic–inorganic hybrid UV-APDs array. Phys Status Solidi C 13:677–682
2. Abe T, Ando K, Ikumi K, Maeta H, Naruse J, Miki K, Ehara A, Kasada H (2005) High gain
and high sensitive blue-ultraviolet avalanche photodiodes (APDs) of ZnSSe n+-i-p structure
molecular beam epitaxy (MBE) grown on p-type GaAs substrates. Jpn J Appl Phys 44(17):L
508–L 510
3. Abe T, Maeta H, Naruse J, Ikumi K, Kubota T, Fujiwara T, Kasada H, K. Ando K. (2004)
New blue-ultraviolet PIN photodiodes of II-VI widegap compounds ZnSSe using p-type GaAs
substrates grown by molecular beam epitaxy. phys stat sol a 1(4):1054–1057
13 ZnSe-Based Photodetectors 329

4. Ando K, Ishikura H, Fukunaga Y, Kubota T, Maeta H, Abe T, Kasada H (2002) Highly effi-
cient blue–ultraviolet photodetectors based on II–VI wide-bandgap compound semiconduc-
tors. Phys Stat Sol a 229(2):1065–1071
5. Averin SV, Kuznetzov PI, Zhitov VA, Zakharov LY, Kotov VM (2020) MSM-photodetector
with ZnSe/ZnS/GaAs Bragg reflector. Opt Quant Electron 52:93
6. Averin SV, Kuznetzov PI, Zhitov VA, Zakharov LY, Kotov VM (2019) Electrical, optical and
spectral characteristics of type-II ZnSe/ZnTe/GaAs superlattice and MSM-photodetector on
their base. Opt Quant Electron 50:368
7. Averin SV, Kuznetzov PI, Zhitov VA, Zakharov LY, Kotov VM (2018) Multicolour photode-
tector based on a ZnSe/ZnTe/GaAs heterostructure. Quant Electron 48(7):675 –678
8. Bhattacharya P (1994) Semiconductor optoelectronic devices. Prentice Hall, New Jersey
9. Bhushan S (1987) Photoconductivity of ZnSe crystals. J Mater Sci Lett 6:591–592
10. Bouhdada A, Hanzaz M, Vigue F, Faurie JP (2003) Electrical and optical proprieties of photo-
diodes based on ZnSe material. Appl Phys Lett 83(1):171–173
11. Bube RH, Lind EL (1958) Photoconductivity of Zinc Selenide crystals and a correlation of
donor and acceptor levels in II-VI photoconductors. Phys Rev 110(5):1040–1049
12. Chakraborty K, Chakrabarty S, Pal T, Ghosh S (2017) Synergistic effect of zinc selenide–
reduced graphene oxide towards enhanced solar light-responsive photocurrent generation and
photocatalytic 4-nitrophenol degradation. New J Chem 41:4662–4671
13. Chang SJ, Lin TK, Su YK, Chiou YZ, Wang CK, Chang SP et al (2006) Homoepitaxial ZnSe
MSM photodetectors with various transparent electrodes. Mater Sci Eng B 127:164–168
14. Chang CC, Lee CH (2000) Study and fabrication of PIN photodiode by using ZnSe/PS/Si
structure. IEEE Trans Electron Dev 47(1):50–54
15. Chang CC, Wu KT (2000) Fabrication of n-ZnSe/p-Si/n-Si heterojunction phototransistor
using IR furnace chemical vapour deposition and its optical properties analysis. IEE Proc
Optoelectron 147(2):104–108
16. Chen M-Y, Chang C-C (2009a) Comparison of performance of integrated photodetectors
based on ZnS and ZnSe metal–semiconductor–metal photodiodes. Jpn Appl Phys 48:112201
17. Chen MY, Chang CC (2009b) Monolithic integration of a ZnSe MSM photodiode and an
InGaP/GaAs HBT on a GaAs substrate. Semicond Sci Technol 24:045009
18. Chen M-Y, Chang C-C (2008) Monolithic photoreceiver constructed with a ZnSe MSM photo-
diode and an InGaP/GaAs HBT. IEEE Electron Dev Lett 29(11):1212–1214
19. Chen W-R, Meen T-H, Cheng Y-C, Lin W-J (2006) P-down ZnSTeSe/ZnSe/GaAs heterostruc-
ture photodiodes. IEEE Electron Dev Lett 27(5):347–349
20. Deng J, Lv W, Zhang P, Huang W (2022) Large-scale preparation of ultra-long ZnSe-PbSe het-
erojunction nanowires for flexible broadband photodetectors. J Sci: Adv Mater Dev 7:100396
21. Dima I, Vasiliu G (1967) Photoconductivity of ZnSe thin films. Phys Status Solidi 22:K79–K82
22. Ehinger M, Koch C, Korn M, Albert D, Nurnberger J, Hock V, Faschinger W, G. Landwehr
G. (1998) High quantum efficiency II–VI photodetectors for the blue and blue-violet spectral
range. Appl Phys Lett 73(24):3562–3564
23. Fang X, Xiong S, Zhai T, Bando Y, Liao M, Gautam UK et al (2009) High-performance blue/
ultraviolet-light-sensitive ZnSe-nanobelt photodetectors. Adv Mater 21:5016–5021
24. Hong H, Anderson WA (1999) Cryogenic processed metal-semiconductor-metal (MSM) pho-
todetectors on MBE grown ZnSe. IEEE Trans. Electron Dev. 46:1127–1133
25. Hsiao CH, Chang SJ, Wang SB, Chang SP, Li TC, Lin WJ et al (2009) ZnSe nanowire pho-
todetector prepared on oxidized Silicon substrate by molecular-beam epitaxy. J Electrochem
Soc 156(4):J73–J76
26. Huang ZC, Wie CK, Na I, Luo H, Mott DB, Shu PK (1996) High performance ZnSe photocon-
ductors. Electron Lett 32(16):1507–1505
27. Ichikawa Y, Tanak K, Nakagawa K, Fujii Y, Yoshida K, Nakamura K et al (2020) High-gain
ultraviolet avalanche photodiodes using a ZnSe-based organic-inorganic hybrid structure. J
Electron Mater 49:4589–4593
28. Inagaki Y, Ebisu M, Otsuki M, Ayuni N, Shimizu T, Abe T et al (2012) New ultraviolet ava-
lanche photodiodes (APDs) of organic (PEDOT: PSS)–inorganic (ZnSSe) hybrid structure.
Phys Status Solidi A 209(8):1852–1855
330 G. Korotcenkov

29. Inoue R, Abe T, Fujimoto T, Ikadatsu N, Tanaka K, Uchida S et al (2015) ZnSe-based organic–
inorganic hybrid structure ultraviolet avalanche photodiodes with long lifetime and its device
integration. Appl Phys Express 8:022101
30. Ishikura H, Fukunaga Y, Kubota T, Maeta H, Adachi M, Abe T et al (2002) Blue-violet
avalanche-­photodiode (APD) and its ionization coefficients in II–VI wide bandgap compound
grown by molecular beam epitaxy. Phys Stat Sol (b) 229(2):1085–1088
31. Ishikura H, Abe T, Fukuda N, Kasada H, Ando K (2000a) Stable avalanche-photodiode
operation of ZnSe-based p+– n structure blue-ultraviolet photodetectors. Appl Phys Lett
76(8):1079–1071
32. Ishikura H, Fukuda N, Itoi M, Yasumoto K, Abe T, Kasada H, Ando K (2000b) High quan-
tum efficiency blue-ultraviolet ZnSe pin photodiode grown by MBE. J Crystal Growth
214(215):1130–1133
33. Jia L, Zheng W, Huang F (2020) Vacuum-ultraviolet photodetectors. PhotoniX 1:22
34. Kudrynskyi ZR, Kovalyuk ZD (2014) Photosensitive anisotype n-ZnSe/p-InSe and n-ZnSe/p-­-
GaSe heterojunctions. Technical Phys 59(9):1205–1208
35. Kuznetzov PI, Averin SV, Zhitov VA, Zakharov LY, Kotov VM (2017) MSM optical detector
on the basis of II-type ZnSe/ZnTe superlattice. Semiconductors 51(2):249–253
36. Li S, Su Q, Zhao H (2013) Photoresponse properties of p-type ZnSe nanowire photodetectors.
Micro Nano Lett 8(9):496–499
37. Lin TK, Chang SJ, Su YK, Chiou YZ, Wang CK, Chang CM, Huang BR (2005a) ZnSe homo-
epitaxial MSM photodetectors with transparent ITO contact electrodes. IEEE Trans Electron
Dev 52(1):121–123
38. Lin TK, Chang SJ, Su YK, Chiou YZ, Wang CK, Chang SP, Chang CM, Tang JJ, Huang BR
(2005b) ZnSe MSM photodetectors prepared on GaAs and ZnSe substrates. Mater Sci Eng B
119:202–205
39. Lour W-S, Chang C-C (1996) VPE growth ZnSe/Si PIN-like visible photodiodes. Solid State
Electron 39(9):1295–1298
40. Makhniy VP, Melnik VV, Tkachenko IV, Gorley PN, Horváth ZJ, Horley PP, Vorobiev YV
(2003) Electrical properties of UV detectors based on zinc selenide with modified surface bar-
rier. Phys Status Solidi 0(3):1039–1043
41. Makhniy VP, Mel'nyk VV, Gorley PN, Horley PP, Sletov MM, Zhuo Z (2005) Detectors of
UV and x-ray irradiation on the base of metal-zinc selenide contact. Proc SPIE 6024:60242J
42. Miki K, Oshita Y, Katada D, Nobe K, Nomura M, Abe T, Kasada H, Ando K (2008) High sen-
sitivity ultraviolet PIN photodiodes of ZnSSe n+−i-p structure /p+-GaAs with an extremely thin
n+-window layer grown by using MBE. J Korean Phys Soc 53(5):292–292
43. Miki K, Abe T, Naruse J, Ikumi K, Yamaguchi T, Kasada H, Ando K (2006) Highly sen-
sitive ultraviolet PIN photodiodes of ZnSSe n+–i–p structure/p+-GaAs substrate grown by
MBE. Phys Status Solidi 243(4):950–954
44. Monroy E, Omnes F, Calle F (2003) Wide-bandgap semiconductor ultraviolet photodetectors.
Semicond Sci Technol 18:R33–R51
45. Monroy E, Vigue F, Calle F, Izpura JI, Munoz E, Faurie J-P (2000) Time response analysis of
ZnSe-based Schottky barrier photodetectors. Appl Phys Lett 77(17):2761–2763
46. Monroy E, Calle F, Munoz E, Omnes F (1999) AlGaN metal-semiconductor-metal photodi-
odes. Appl Phys Lett 74:3401
47. Mu Z, Zheng Q, Liu R, Malik MWI, Tang D, Zhou W, Wan Q (2019) 1D ZnSSe-ZnSe axial
heterostructure and its application for photodetectors. Adv Electron Mater 2019:1800770
48. Nakano M, Makino T, Tsukazaki A, Ueno K, Ohtomo A, Fukumura T et al (2008) Transparent
polymer Schottky contact for a high performance visible-blind ultraviolet photodiode based on
ZnO. Appl Phys Lett 93:123309
49. Nasiri N, Tricol A (2019) Nanomaterials-based UV photodetectors. In: Thomas S,
Grohens Y, Pottathara YB (eds) Industrial Applications of Nanomaterials. Elsevier,
New York, pp 123–149. https://doi.org/10.1016/B978-­0-­12-­815749-­7.00005-­0
13 ZnSe-Based Photodetectors 331

50. Naval V, Smith C, Ryzhikov V, Naydenov S, Alves F, Karunasiri G (2010) Zinc selenide-based
Schottky barrier detectors for ultraviolet-A and ultraviolet-B detection. Adv OptoElectron
2010:61957
51. Nedeoglo ND, Nedeoglo DD, Sirkeli VP, Tiginyanu IM, Laiho R, Lähderanta E (2008)
Shallow donor states induced in ZnSe:Au single crystals by lattice deformation. J Appl Phys
104:123717
52. Nie B, Luo L-B, Chen J-J, Hu J-G, Wu C-Y, Wang L et al (2013) Fabrication of p-type ZnSe:Sb
nanowires for high-performance ultraviolet light photodetector application. Nanotechnology
24:095603
53. Oksenberg E, Martí-Sànchez S, Popovitz-Biro R, Arbiol J, Joselevich E (2017) Surface-guided
core-shell ZnSe@ZnTe nanowires as radial p-n heterojunctions with photovoltaic behavior.
ACS Nano 11(6):6155–6166
54. Oksenberg E, Popovitz-Biro R, Rechav K, Joselevich E (2015) Guided growth of horizontal
ZnSe nanowires and their integration into high-performance blue–UV photodetectors. Adv
Mater 27:3999–4005
55. Park S, Kim S, Sun G-J, Byeon DB, Hyun SK, Lee WI, Lee C (2016) ZnO-core/ZnSe-shell
nanowire UV photodetector. J Alloys Comp 658:459–464
56. Park RM, Troffer MB, Rouleau CM, DePuydt JM, Haase MA (1990) p-type ZnSe by nitrogen
atom beam doping during molecular beam epitaxial growth. Appl Phys Lett 57:2127
57. Philipose U, Ruda HE, Shik A, de Souza CF, Sun P (2006) Conductivity and photoconductivity
in undoped ZnSe nanowire array. J Appl Phys 99:066106
58. Rabaco F, Martin JM, Vincent AB, Joshi NV (1991) Photoconductivity and optical absorption
on the high energy side in an se treated ZnSe monocrystal. J Phys Chem Solids 52(4):575–578
59. Radevici I, Sushkevich K, Colibaba G, Sirkeli V, Huhtinen H, Nedeoglo N, Nedeoglo D, Paturi
P (2013) Influence of chromium interaction with native and impurity defects on optical and
luminescence properties of ZnSe:Cr crystals. J Appl Phys 114:203104
60. Rao GK (2017) Electrical and photoresponse properties of vacuum deposited Si/Al:ZnSe and
Bi:ZnTe/Al:ZnSe photodiodes. Appl Phys A Mater Sci Process 123:224
61. Rao KG, Bangera KV, Shivakumar GK (2011) Photoconductivity and photo-detecting proper-
ties of vacuum deposited ZnSe thin films. Sol State Sci 13:1921–1925
62. Razeghi M, Rogalski A (1996) Semiconductor ultraviolet detectors. J Appl Phys 79:7433–7473
63. Salfi J, Philipose U, de Sousa CF, Aouba S, Ruda HE (2006) Electrical properties of Ohmic
contacts to ZnSe nanowires and their application to nanowire-based photodetection. Appl Phys
Lett 89:261112
64. Seghier D, Gislason HP (1999) The observation of persistent photoconductivity in N-doped
p-type ZnSe/GaAs heterojunctions. J. Phys D: Appl Phys 32(4):369
65. Seghier D, Gislason HP (2000) Investigation of persistent photoconductivity in nitrogen-doped
ZnSe/GaAs heterojunctions grown by MBE. J Crystal Growth 214(215):511–515
66. Sharma J, Tripathi SK (2011) Effect of deposition pressure on structural, optical and electrical
properties of zinc selenide thin films. Physica B 406:1757–1762
67. Sirkeli VP, Nedeoglo ND, Nedeoglo DD, Yilmazoglu O, Hajo AS, Preu S et al (2021) ZnSe-­
based solar-blind ultraviolet photodetectors with different Schottky contact metals. Studia
Universitatis Moldaviae 2(142):59–67
68. Sirkeli VP, Yilmazoglu O, Ong DS, Preu S, Küppers F, Hartnagel HL (2017a) Resonant tunnel-
ing and quantum cascading for optimum room-temperature generation of THz signals. IEEE
Trans Electron Dev 64:3482–3488
69. Sirkeli VP, Yilmazoglu O, Hajo AS, Nedeoglo ND, Nedeoglo DD, Preu S et al (2017b)
Enhanced responsivity of ZnSe-based metal–semiconductor–metal near-ultraviolet photode-
tector via impact ionization. Phys Status Sol RRL 12(2):1700418
70. Soci C, Zhang A, Bao X-Y, Kim H, Lo Y, Wang D (2010) Nanowire photodetectors. J Nanosci
Nanotechnol 10(3):1430–1449
71. Su Q, Zhang Y, Li S, Du L, Zhao H, Liu X, Li X (2015) Synthesis of p-type phosphorus doped
ZnSe nanowires and their applications in nanodevices. Mater Lett 139:487–490
332 G. Korotcenkov

72. Su Q, Li L, Li S, Zhao H (2013) Synthesis and optoelectronic properties of p-type nitrogen


doped ZnSe nanobelts. Mater Lett 92:338–341
73. Sun Y-S, Xie D, Sun M-X, Teng C-J, Qian L, Chen R-S et al (2018) Hybrid graphene/cadmium
free ZnSe/ZnS quantum dots phototransistors for UV detection. Sci Rep 8:5107
74. Sunaina, Ganguli AK, Mehta K (2022) High performance ZnSe sensitized ZnO heterostruc-
tures for photo-detection applications. J. Alloys Compounds 894:162263
75. Sze SM, Ng KK (2007) Physics of semiconductor devices. Wiley-Interscience, Hoboken
76. Terui Y, Yoshino M, Ogura M, Nakayama M, Yoneda M, Chikamura T et al (1981) A CCD
imager using ZnSe-Zn1-x CdxTe heterojunction photoconductor, Jpn. J Appl Phys Sup
21–1:237–242
77. Ueno J, Ogura K, Ichiba A, Katsuta S, Kobayashi M, Onomitsu K, Horikoshi Y (2006) MBE
growth of ZnSe/MgCdS and ZnCdS/MgCdS superlattices for UV-A sensors. Phys Stat Sol (c)
3(4):1225–1228
78. Vigue F, Faurie J-P (2001) Zn(MgBe)Se ultraviolet photodetectors. J Electron Mater
30(6):662–666
79. Vigué F, Tournié E, Faurie J-P (2001) Evaluation of the potential of ZnSe and Zn(Mg)BeSe
compounds for ultraviolet photodetection. IEEE J Electron Dev 37(9):1146–1152
80. Vigué F, Tournié E, Faurie J-P (2000a) Zn(Mg)BeSe-based p-i-n photodiodes operating in the
blue-violet and near-ultraviolet spectral range. Appl Phys Lett 76:242–244
81. Vigué F, de Mierry P, Faurie J-P, Monroy E, Calle F, Mufioz E (2000b) High detectivity ZnSe-­
based Schottky barrier photodetectors for blue and near-ultraviolet spectral range. Electron
Lett 36(9):826–827
82. Vigué F, Tournié E, Faurie J-P (2000c) ZnSe-based Schottky barrier photodetectors. Electron
Lett 36:352–354
83. Voronkin E (2013) Schottky diodes based on the Zinc Selenide semiconductor crystals.
Functional Mater 20(4):534–537
84. Wang Z, Jie J, Li F, Wang L, Yan T, Luo L et al (2012) Chlorine-doped ZnSe nanoribbons with
tunable n-type conductivity as high-gain and flexible blue/UV photodetectors. Chem Phys
Chem 77:470–475
85. Yu Y-M, Nam SOB, Lee K-S, Yu PY, Lee J, Choi TD (2002) Strain effect in ZnSe epilayer
grown on the GaAs substrate. J Crystal Growth 243:389–395
86. Xiao C, Wang Y, Yang T, Luo Y, Zhang M (2016) Preparation of ZnO/ZnSe heterostructure
parallel arrays for photodetector application. Appl Phys Lett 109:043106
87. Xu Z, Wu C, Li A, Ma Y, Fei GT, Wang M (2018) Sub-100 nm channel ZnSe film/graphene
hybrid-based photodetectors with an ultrahigh responsivity of 109 A/W. IEEE Electron Dev
Lett 39(2):240–243
88. Zhang X, Jie J, Wang Z, Wu C, Wang L, Peng Q et al (2011) Surface induced negative photo-
conductivity in p-type ZnSe: Bi nanowires and their nano-optoelectronic applications. J Mater
Chem 21:6736–6741
Chapter 14
ZnS-Based UV Detectors

Sema Ebrahimi, Benyamin Yarmand, and Nima Naderi

14.1 Introduction

The technology of photodetectors based on II-VI compound semiconductors has


drawn much research attention over the past decades due to their optical properties,
good transport properties, strong thermal stability, and high electron mobility [1, 2].
Zinc sulfide (ZnS), a prominent II-VI intrinsic semiconductor material, has been
intensely studied because of its broad spectrum of diverse applications such as in
photovoltaic, electroluminescence devices, bio-imaging devices, photodetectors,
sensors, and solar cells for a long time [3–7].
When the size of these structures is comparable to their corresponding Bohr
exciton radius, they show a quantum confinement effect with a considerable blue
shift in absorption spectra, high surface-to-volume ratio, high aspect ratio, and large
Debye length. On the contrary, engineering the composition of the materials is an
alteration option to tuning the bandgap for specific applications [1–3, 8–12], in
which the large bandgap of ZnS in UV regime and high quantum efficiency makes

S. Ebrahimi
Materials and Energy Research Center (MERC), Karaj, Iran
Light, Nanomaterials, Nanotechnologies (L2n) Laboratory, CNRS EMR7004, The University
of Technology of Troyes, Troyes Cedex, France
Department of Physics and Mathematics, University of Hull, Cottingham Road, UK
G.W.Gray Centre for Advanced Materials, University of Hull, Cottingham Road, UK
B. Yarmand
Materials and Energy Research Center (MERC), Karaj, Iran
N. Naderi (*)
Materials and Energy Research Center (MERC), Karaj, Iran
Photonics Research Centre, University of Malaya, Kuala Lumpur, Malaysia
e-mail: n.naderi@merc.ac.ir

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 333
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_14
334 S. Ebrahimi et al.

it desirable for visible-blind UV detectors, revealing remarkable characteristics in


widespread applications such as sensing, astronomy, communication, and medical
instruments.
In the modern scientific area, the motivation for the fabrication of doped semi-
conductor nanostructures has attracted great interest due to their significance in
basic scientific research and potential technological applications. To utilize semi-
conductor nanostructures as building blocks of photodetectors, it is significant to
synthesize nanostructures that have remarkable optoelectrical properties [10, 12,
13]. In the case of doped materials, the addition of impurities into a wide-gap semi-
conductor can often induce dramatic changes in the optical and electrical properties
as well as influence transition probabilities [11, 13, 14]. Since ZnS has a wide band-
gap at room temperature, its nanocrystals are suitable to host material for doping
elements including transition metal and/or rare-earth elements which are optically
active. Therefore, doping a selective element such as Mn, Sn, Ni, W, and Graphene
in ZnS nanostructures has been an important route for raising and controlling the
bandgap and making it suitable to prepare UV-detectors with different cut-off wave-
lengths and high performance [2, 3, 6, 9, 15–22].
In this chapter, we focus on the most recent progress in UV-detectors based on
the ZnS nanostructures. First, ZnS photodetectors including photoconductors,
metal-semiconductor-metal (MSM) photodetectors, Schottky photodiodes, p-n
junction photodiodes are discussed, followed by photodetectors based on the doped
ZnS nanostructures. Finally, we conclude this review chapter with future perspec-
tives in these brand-new fields.

14.2 ZnS-Based Photodetectors

This section discusses ZnS-based photodetectors, including photoconductors,


metal-semiconductor-metal (MSM) photodetectors, and Schottky photodiodes.
Among the photons that strike a semiconductor, those with an energy greater
than the semiconductor bandgap can give enough energy to the valance band elec-
trons and move them into the conduction band. The photo-generated carriers
increase the semiconductor conductivity, and this is the basic operating mechanism
in photoconductors. The photoconductor can be simply placed in series with the
load resistance. The change in the device conductivity due to the illumination
changes the voltage across the load resistance, which can be detected by a high
impedance voltmeter. The typical photoconductor usually contains a semiconductor
layer sandwiched between two ohmic contacts, resulting in a high photoconductive
gain and high responsivity and no need the external amplifying equipment. However,
they show relatively large dark-current, low signal-to-noise ratio, and slow response
speed. Many efforts have been taken to overcome these issues, such as using II–
VI-based nanostructures [12, 20, 23] and, surface treatment [24] to increase the
internal gain and other parameters on-demand.
14 ZnS-Based UV Detectors 335

Fig. 14.1 LPG sensing response (a) at room temperature and (b) Stability response of ZnS
nanoparticles against LPG and inset show the transient response at room temperature. (Reprinted
with permission from [25]. Copyright 2021: Elsevier)

ZnS-based photoconductors have been investigated in detail by several groups.


Different materials such as Au, Ag, Al, Pt, Al/Au, etc., have been used as electrodes.
Nemade et al. [25] demonstrated UVC photoconductors based on the ZnS nanopar-
ticles synthesized using flame-assisted spray pyrolysis method on a SiO2 substrate.
The highly conducting Ag was used as ohmic electrodes. In this report, the ZnS
nanoparticles were applied for the liquefied petroleum gas (LPG) sensing, in which
the response curve featured a decrease in the resistance of the nanoparticles suggest-
ing the intrinsically n-type ZnS in the presence of LPG, as shown in Fig. 14.1a. The
good response stability (Fig. 14.1b) has been recorded for the as-synthesized ZnS
nanoparticles against LPG, with a fast response and recovery time of about 18 s and
11 s, respectively.
Hajimazdarani et al. [26] have fabricated UV detectors based on the ZnS
nanoparticles using a chemical deposition method, followed by an annealing pro-
cess. They focused on the effect of annealing temperature on the optoelectrical
properties of the ZnS nanoparticles. According to this report, at the optimized
annealing temperature of 500 °C, a phase change was applied to the ZnS lattice
system from zinc blende to wurtzite, leading to a significant enhancement in the
absorbance and photoluminescence intensity. The optoelectrical characteristics
indicate the higher generation of electron carriers under the UV exposure for the
annealed samples, resulting in higher photosensitivity and photoresponsivity com-
pared to the as-synthesized sample as demonstrated in Fig. 14.2.
Prasad et al. [27] investigated a UV photoconductor based on the highly crystal-
line ZnS nanoparticles having cubic and wurtzite structure produced by a simple
co-precipitation at room temperature and one-step hydrothermal method at 180 °C,
respectively. They studied the influence of the crystalline structure of ZnS nanopar-
ticles on the photoresponsivity behavior under UV illumination. Both fabricated
devices showed ohmic behavior, which interestingly confirmed a better response of
wurtzite lattice than the cubic one due to the high electron-phonon coupling. The
336 S. Ebrahimi et al.

Fig. 14.2 The sensitivity of UV detectors based on detectors samples with different annealing
temperature. (Reprinted with permission from [26]. Copyright 2021: Elsevier)

Fig. 14.3 (a) I–V characteristic curve of a photodetector with c-ZnS and w-ZnS nanoparticles as
channel and Ag as electrode under dark and illumination of UV light and (b) Logarithmic plot of
I–V characteristics of c-ZnS and w-ZnS nanoparticles-based UV photodetector. (Reprinted with
permission from [27]. Copyright 2021: Elsevier)

I–V characteristic curves exhibited an enhanced photoresponsivity of the wurtzite


ZnS nanoparticles-based UV detectors and high dependence of the performance on
the dynamics of the phonon population illustrated in Fig. 14.3. It was then con-
cluded that the fabricated device shows ohmic behavior and the wurtzite ZnS has a
better optical response towards the UV illumination than the cubic type since a
dominant enhancement in electron-phonon coupling has been achieved for wurtzite
ZnS than for cubic-ZnS.
14 ZnS-Based UV Detectors 337

The Metal-Semiconductor-Metal (MSM) structure is the simplest type of photo-


detector. MSM photodetectors have a simple structure, easy to fabricate, have fast
response, low capacitance per unit area, and large active area. MSM photodetectors
are comprised of two back-to-back Schottky diodes using two interdigitated elec-
trodes on an absorbing layer [28]. The main drawback of MSM photodetectors is
their low photoresponsivity, due to the shadow effect arising from the metallization
of electrodes on the active area. Under the incident optical power, electron-hole
pairs are generated through the absorption of light and flow towards the gap between
the two metal contacts, resulting in a photocurrent. For compound semiconductors,
the light absorption layer is usually deposited on a semi-insulating substrate.
ZnS-based MSM photodiodes have been fabricated by different physical and
chemical techniques such as sputtering [29], sol-gel [30], chemical vapor deposition
[31], molecular beam epitaxy [32], atomic layer epitaxy [33], pulse electrochemical
deposition [34], chemical bath deposition [35], spray pyrolysis [36] and hydrother-
mal/solvothermal method [37].
In the case of MSM photodiodes, Ebrahimi et al. [38] investigated the perfor-
mance of UV detectors based on the different ZnS-based absorbing layers. First,
they studied the optoelectrical properties of ZnS films deposited by a simple and
low-cost spray pyrolytic method. In this report, the effect of the Zn: S molar ratio in
the precursor solution on the structural, optical, and electrical properties of the ZnS
layers has been investigated. For all the samples having a different molar ratio
(Zn:S) of precursor, (1:1), (1:2), and (1:3), at the fixed substrate temperature of
about 400 °C, a single-phase ZnS cubic with good crystallinity was obtained using
a sulfur-rich solution. The band gap energy was found to be increased by the incre-
ment of sulfur concentration up to three times of zinc content in the ZnS lattice due
to the improvement of crystallinity.
To continue, Ebrahimi et al. [36] reported the UV detectors based on the chalco-
genide Mn-doped ZnS films deposited by the spray pyrolysis method. They studied
the modification of the ZnS-based films by evaluating their optical properties
including the optical density, optical band gap energy, Urbach energy, steepness
parameter, electron-phonon interactions, and the relationship between them, to
explore the photodetection and photoresponsivity of the UV detectors. From the
optical properties in Fig. 14.4, by increasing the Mn2+ ions concentration from 0.05
to 0.15 mol, a slight redshift was observed in the absorption band-edge, leading to a
decrease in the optical band gap energy of the samples from 3.98 eV to
3.86 eV. Otherwise, the Urbach energy corresponds to the extended defect levels
and disorders introduced by Mn ions increased from 354 meV to 420 meV. Considering
the linear relationship between the optical band gap and Urbach energies, two
important phenomena were found. First, the value of the optical bandgap energy in
the absence of defects and band tailing of about 4.61 eV. Second, is the inverse rel-
evance between the steepness parameter and electron-phonon interactions. The PL
spectra demonstrated two emission peaks located in UV and orange regions, corre-
sponding to the band-to-band transition of host ZnS and 4T1-6A1 inter-band transi-
tion of Mn2+ ions, respectively.
338 S. Ebrahimi et al.

Fig. 14.4 (a) the variation of (αhυ)2 versus hυ, and (b) the schematic illustration of the energy
diagram of pure ZnS and Mn-doped ZnS films. (Reprinted with permission from [36]. Copyright
2021: Elsevier).

Fig. 14.5 The current-voltage curves of (a) ZnS: 0.15 Mn films-based UV detectors and (b) The
photocurrent gain of Pure ZnS and Mn-doped ZnS films-based UV detectors. (Reprinted with
permission from [36]. Copyright 2021: Elsevier)

The MSM UVA detectors based on the Mn-doped ZnS films were fabricated
using two Schottky contacts of Au, as sketched in Fig. 14.5. Compared to the pure
ZnS-based UV detectors, the as-fabricated devices revealed a high photocurrent
gain of about 235.0 (Fig. 14.5b). The response and recovery times recorded the low-
est values of about 9.9 ms and 13.2 ms, respectively, by which the results confirmed
the performance of the fabricated visible-blind UV detectors based on Mn-doped
ZnS films can be enhanced by introducing Mn2+ ions as dopants into the ZnS lattice.
As the results featured a strong dependence of the photodetection and sensing
behavior on the Mn-content, they investigated structural, optical, and electrical
properties of the composition-tunable ternary MnxZn1-xS thin films to use as UV
detectors. At a high concentration of Mn2+ions (>0.2 mol), a gradual decrease of the
bandgap energy was found followed by an increase up to 4.05 eV. As Ebrahimi et al.
[39] explained, when the Mn2+ions are introduced into the ZnS lattice, they can
occupy either substitutional or interstitial sites of Zn2+ ions, which caused the
emerging of the Mn2+ impurity bands formed by the overlapped impurity states,
14 ZnS-Based UV Detectors 339

Fig. 14.6 (a) The variations of the optical absorption coefficient, and (b) (αhυ)2 versus hυ for the
pure ZnS and MnxZn1-xS thin films. (Reprinted with permission from [39]. Copyright 2021:
Elsevier)

leading to a slight decrease of the ZnS bandgap energy. On the other hand, widening
the bandgap energy at the higher content of Mn2+ ions is attributed to another phe-
nomenological event named the Burstein-Moss shift theory, by which the blue shift
of the bandgap energy is aroused by lying at the Fermi level (EF) in the valence band
for the ZnS structure, as depicted in Fig. 14.6. By increasing the Mn2+ content, the
PL emission intensity located in the UV region was found to be enhanced which
modifies the rate of electron-hole pair generation.
The performance of the UV detectors based on the MnxZn1-xS thin films showed
an enhanced photoresponsivity and photoswitching behavior in the UVB region of
the spectrum [39]. The current gain was improved over 4.5 times and the photo-
switching behavior became faster over 5 times compared to the pure ZnS-based UV
detectors (Fig. 14.7). The working mechanism of the Au/ MnxZn1-xS thin film/Au
UV detectors can be described based on the energy band theory. When the MnxZn1-­x
S alloys as a photo-sensing semiconductor connect to Au electrodes, due to a rela-
tively lower work function of the Au contacts of about 5.31 eV than the MnxZn1-xS
layer of about 5.43 eV, the electrons drift from the Au electrodes toward the MnxZn1-­x
S thin films until their fermi levels are aligned. Under equilibrium conditions, a
Helmholtz double-layer will be formed at the Au/MnxZn1-xS junction, where the Au
is positively charged and MnxZn1-xS is negatively charged near their interface attrib-
uted to the electrostatic induction. Due to the electric field induced by the Helmholtz
double-layer, a depletion layer is established by decreasing the free charge carrier
concentration near the surface. Therefore, energy bands of MnxZn1-xS bend down-
ward at the inter-face due to the electric field, which forms a Schottky barrier at the
interface of Au and MnxZn1-xS thin film. When devices are illuminated by UV radia-
tion with photon energy more than the optical bandgap of the samples, electron-hole
pairs are generated. The generated carriers can be separated by a built-in potential
or external electric field and then a photocurrent is produced. As well as the states
caused by Mn impurities will function as effective trapping states. By trapping the
minority charge carriers, the lifetime of the carriers is remarkably prolonged, which
leads to enhancing the generated photocurrent which in turn decreases the SBHs
340 S. Ebrahimi et al.

Fig. 14.7 The schematic energy band diagram of Au/MnxZn1-xS/Au based UV detector under (a)
dark and (b) illumination conditions. The inset shows the circuitry of the back-to-back Schottky
device. Time response of (c) pure ZnS, (d) Mn0.2Zn0.8S, (e) Mn0.3Zn0.7S and (f) Mn0.4Zn0.6S thin
films-based UV detectors. (Reprinted with permission from [39]. Copyright 2021: Elsevier)

and depletion width across the junction. It can be concluded that the UV detector
based on the MnxZn1-xS thin film is a high-speed performance device, which is much
faster than many oxide and chalcogenide nanomaterial UV detectors. It can be
attributed to the high surface-to-volume ratio, enhanced density of mobile charge
carriers, optically active trap and defect states at the surface of the photo-detecting
MnxZn1-xS layers, rearrangement and modification of the energy band structure as
well as an efficient excitation transfer between the Mn2+ and ZnS energy levels,
14 ZnS-Based UV Detectors 341

which drastically affects the photosensing and photoresponsivity of the devices. The
results show that the visible-blind UV detector based on ternary MnxZn1-xS thin
films is a promising candidate for future optoelectronic integration devices because
it can offer remarkable characteristics for the high detection of the dangerous UV-B
radiations.
Using different methods and ZnS-based compounds, designing simple and cost-­
effective self-powered UV detectors with high photoresponsivity, and photo-­
switching speed based on the ZnS nanostructures is still a noticeable challenge. An
efficient and high-performance self-powered UV detector must satisfy 5S require-
ments, including high sensitivity, high spectral detectivity, high signal-to-dark cur-
rent ratio, high stability, and high speed. Generally, the reported self-powered UV
detectors suffer from complicated fabrication processes and high dark current. In
this case, Ebrahimi et al. [37] aimed to study the self-powered UV detectors based
on the ZnS thin films and their compounds through engineering the bandgap energy
by doping and/or alloying processes. They developed a modified low-temperature
solvothermal method using a capping agent of Ethylenediamine under short deposi-
tion time to synthesize the well-aligned SnxZn1-xS nanostructured thin films on the
glass substrate for MSM UV detector applications. As depicted in Fig. 14.8, by
incorporation of the Sn2+ ions into the ZnS lattice, a growth mechanism was

Fig. 14.8 The schematic illustration of the growth mechanism of the SnxZn1-xS thin films by the
low-temperature solvothermal method; (a) the formation of the SnxZn1-xS thin film on the glass
substrate in the presence of binary solvent of EN/water and (b-d) the formation of cubic-like,
pencil-like and rod-like nanostructured SnxZn1-xS thin films developed by the variation of the Sn2+
doping content. (Reprinted with permission from [37]. Copyright 2021: Elsevier)
342 S. Ebrahimi et al.

Fig. 14.9 (a) the relationship between the bandgap energy and Urbach energy, and (b) the varia-
tion of the steepness parameter and electron-phonon interaction of the pure ZnS and the SnxZn1-xS
thin films. (Reprinted with permission from [37]. Copyright 2021: Elsevier).

proposed based on morphological evolution from aligned nanoflakes to cubic-like,


pencil-like, and finally, rod-like nanostructures [37, 40, 41].
Considering the effect of the morphology on the optical properties, a linear rela-
tionship between the optical band gap energy and Urbach tails was found, detail
shown in Fig. 14.9.
The fabricated UV detectors based on the SnxZn1-xS thin film junctions showed
excellent photoresponse in the UVB region and high visible rejection. Additionally,
they featured an excellent rectification behavior and attractive photovoltaic effect
under UVB exposure, confirming the self-powered characteristic of the devices
based on SnxZn1-xS thin films. The photodetection characteristics of the self-­powered
UV detectors demonstrated a high Ion/Ioff ratio over 103 and a fast-photoswitching
speed with superior stability and reproducibility at zero voltage. Notable photosen-
sitivity and detectivity of more than ~20 times were recorded in the self-powered
mode of these UV detectors (Fig. 14.10). The simplified circuitry of the UV detec-
tors demonstrates that in the dark condition, the dark current emanates from the
inbuilt potential of n-type ZnS semiconductor thin films by the cause of the free
charge carriers. Upon UV exposure, the separation of the electron-hole pairs towards
the conduction (CB) and valence (VB) bands leads to the generation of the photo-
current response due to the absorption of the energy of the incident light within the
active region of the device. Ultimately, the more the charge carrier concentration
increases, the more the photocurrent produces in the UV detector. The responsivity
and photodetectivity of the aligned SnxZn1-xS nanostructured thin films-based pho-
todetectors are the record-highest values obtained in the UV region, as compared in
Table 14.1, which even considerably have the highest sensitivity and photoswitch-
ing performance at the self-power mode. The higher Ion/Ioff ratio can be attributed to
not only the larger surface-to-volume ratios and one−/two-dimension nanostruc-
tured SnxZn1-xS thin films, with which more absorbing UV and higher efficiency of
electron-hole pair generation/separation are beneficial to produce larger photocur-
rent; but also Schottky junction formed between the thin films and Au contacts,
which effectively hinders the transport of free charge careers and further gives rise
to a lower dark current.
14 ZnS-Based UV Detectors 343

Fig. 14.10 (a) The schematic view of the n-type band energy diagram of the Schottky-type con-
figuration, (b) photoresponsivity and (c) photosensitivity behavior vs. Sn2+ doping content for the
as-grown and annealed SnxZn1-xS thin films-based MSM UV detectors at the self-power mode with
zero bias voltage. (Reprinted with permission from [37]. Copyright 2021: Elsevier)

Table 14.1 Comparison of the important figures-of-merit of recently fabricated ZnS-based


photodetectors
Bias Wavelength Responsivity Dark Photo
Photodetector (V) (nm) (AW−1) current (A) current (A) Ion/Ioff References
ZnS 10 325/442 1.86 3 × 10−12 24 × 10−12 – [42]
nanowires
Sb-doped 1 254 – ~10−9 ~10−7 >102 [43]
ZnS
N-doped ZnS 5 254 1.4 × 105 – – 5 [44]
Mn-doped 5 325 – ~10−8 ~10−4 235 [36]
ZnS
Sn-doped 0 280 >10−2 <9× 10−7 >1 × 10−3 >103 [37]
ZnS
MgxZn1-xS 5 400 1.523 × 10−6 0.89 × 10−6 1.52 × 10−6 96.74 [45]
CdxZn1-xS 0 325 130 ~10−4 ~10−2 – [46]
MnxZn1-xS 5 280 – ~ 10−6 ~10−2 579.9 [39]
SnxZn1-xS 0 280 212 × 10−3 4.75 × 10−6 6.98 × 10−3 2406.1 [41]
344 S. Ebrahimi et al.

In the other similar research carried out by the same group, Ebrahimi et al. [41]
proposed that the photoresponse performance of the UV detectors can be tuned by
varying the content of Sn2+ ions in the SnxZn1-xS absorbing layer, attributing to the
advantages of the microstructural and optical properties of the SnxZn1-xS thin films.
The Schottky barrier heights of the UV detectors were decreased by increasing the
Sn2+ incorporation content, revealing good rectifying characteristics of the n-type
UV detectors. Further, for each UV detector, superior self-powered characteristics
were found in the difference and relation, which can be adjustable by the variation
of the Sn2+ concentration in the SnxZn1-xS thin films. To understand the mechanism
of the enhancement of UV photocurrent gain by introducing the Sn2+ ions into the
ZnS network, the energy-band diagram in both junctions is illustrated in the dark
and under UV illumination. From Fig. 14.11a, the orange arrows exhibit the photo-
current in the SnxZn1-xS thin films generated by the motion of the electron-hole pairs
under UV illumination. In the dark condition, the electrons flowing into the n-type
SnxZn1-xS barrier layer accelerate due to the built-in electric field and/or applied
external bias, resulting in a high photocurrent. Thus, the Au/SnxZn1-xS thin film/
Au devices exhibit a good rectification behavior in the dark current. On the other

Fig. 14.11 (a) The schematic energy-band alignments of the self-powered MSM UV detectors
based on ternary SnxZn1-xS alloy thin films in the dark and under UV illumination conditions. I–V
curve of the self-powered UV detector based on the ternary (b) Sn0.40Zn0.60S alloy thin films mea-
sured under the dark and UV illumination, and (c) the plot of VOC and ISC vs. Sn2+ content.
(Reprinted with permission from [41]. Copyright 2021: Elsevier)
14 ZnS-Based UV Detectors 345

hand, upon UVB exposure, the electron-hole pairs are produced in the SnxZn1-­xS
nanostructures, in which the photogenerated holes migrate to the surface and dis-
charge the negatively charged ions adsorbed at the surface of the photoactive layer.
Thus, the high photoresponsivity and sensitivity are strongly dependent on effective
photocurrent transport through the active SnxZn1-xS layer, and the introduction of
the Sn2+ as donor states, resulting in the tunable self-powered SnxZn1-xS thin films-
based UV detectors with tremendous photoresponsivity and photodetection perfor-
mance. These UV detectors featured the highest photosensitivity of about 2406.1
(Ion/Ioff ratio ≈ 8.47 × 103) and ultrafast photoresponse rise and decay times of 1.9
and 2.6 ms, respectively at the self-powered mode, which are the best results
reported in the ZnS-based self-powered UV detectors until now. A clear comparison
from Table 14.1 represents the important figures-of-merit of recently produced self-­
powered ZnS-based photodetectors. Generally, the data reveals the superiority of
the self-powered UV detector based on the SnxZn1-xS nanostructured thin films with
different Sn2+ content devices used in photodetection applications. It can be claimed
that both the optical and electrical properties of the samples suggest the high photo-
responsivity and excellent stability of the self-powered ultraviolet detectors.

References

1. Jiang D, Cao L, Su G, Liu W, Qu H, Sun Y et al (2009) Synthesis and luminescence properties


of ZnS: Mn/ZnS core/shell nanorod structures. J Mater Sci 44(11):2792–2795
2. Mall M, Kumar L (2010) Optical studies of Cd2+ and Mn2+ Co-doped ZnS nanocrystals. J
Lumin 130(4):660–665
3. Al-Rasoul KT, Abbas NK, Shanan ZJ (2013) Structural and optical characterization of Cu and
Ni doped ZnS nanoparticles. Int J Electrochem Sci 8(4):5594–5604
4. Cao J, Yang J, Zhang Y, Wang Y, Yang L, Wang D et al (2010) XAFS analysis and luminescent
properties of ZnS: Mn2+ nanoparticles and nanorods with cubic and hexagonal structure. Opt
Mater 32(5):643–647
5. Salem JK, Hammad TM, Kuhn S, Draaz MA, Hejazy NK, Hempelmann R (2014) Structural
and optical properties of Co-doped ZnS nanoparticles synthesized by a capping agent. J Mater
Sci Mater Electron 25(5):2177–2182
6. Salem JK, Hammad TM, Kuhn S, Nahal I, Draaz MA, Hejazy NK et al (2014) Luminescence
properties of Mn and Ni doped ZnS nanoparticles synthesized by capping agent. J Mater Sci
Mater Electron 25(12):5188–5194
7. Zhang L, Qin D, Yang G, Zhang Q (2012) The investigation on synthesis and optical properties
of ZnS: Co nanocrystals by using hydrothermal method. Chalcogenide Lett 9(3):93–98
8. Chen Y-C, Wang C-H, Lin H-Y, Li B-H, Chen W-T, Liu C-P (2010) Growth of Ga-doped ZnS
nanowires constructed by self-assembled hexagonal platelets with excellent photocatalytic
properties. Nanotechnology 21(45):455604
9. Dong L, Liu Y, Zhuo Y, Chu Y (2010) General route to the fabrication of ZnS and M-doped
(M= Cd2+, Mn2+, Co2+, Ni2+, and Eu3+) ZnS Nanoclews and a study of their properties. Wiley
Online Library
10. Kang T, Sung J, Shim W, Moon H, Cho J, Jo Y et al (2009) Synthesis and magnetic properties
of single-crystalline Mn/Fe-doped and Co-doped ZnS nanowires and nanobelts. J Phys Chem
C 113(14):5352–5357
346 S. Ebrahimi et al.

11. Zhu G, Zhang S, Xu Z, Ma J, Shen X (2011) Ultrathin ZnS single crystal nanowires: con-
trolled synthesis and room-temperature ferromagnetism properties. J Am Chem Soc
133(39):15605–15612
12. Tian Y, Zhao Y, Tang H, Zhou W, Wang L, Zhang J (2015) Synthesis of ZnS ultrathin nanow-
ires and photoluminescence with Mn2+ doping. Mater Lett 148:151–154
13. Shanmugam N, Cholan S, Kannadasan N, Sathishkumar K, Viruthagiri G (2014) Effect of
polyvinylpyrrolidone as capping agent on Ce3+ doped flowerlike ZnS nanostructure. Solid
State Sci 28:55–60
14. Cao J, Han D, Wang B, Fan L, Fu H, Wei M et al (2013) Low temperature synthesis, photolu-
minescence, magnetic properties of the transition metal doped wurtzite ZnS nanowires. J Solid
State Chem 200:317–322
15. Karan NS, Sarkar S, Sarma D, Kundu P, Ravishankar N, Pradhan N (2011) Thermally con-
trolled cyclic insertion/ejection of dopant ions and reversible zinc blende/wurtzite phase
changes in ZnS nanostructures. J Am Chem Soc 133(6):1666–1669
16. Zhai T, Li L, Ma Y, Liao M, Wang X, Fang X et al (2011) One-dimensional inorganic nano-
structures: synthesis, field-emission and photodetection. Chem Soc Rev 40(5):2986–3004
17. Azimi H, Ghoranneviss M, Elahi S, Yousefi R (2016) Photovoltaic and UV detector applications
of ZnS/rGO nanocomposites synthesized by a green method. Ceram Int 42(12):14094–14099
18. Kumar V, Rawal I, Kumar V, Goyal PK (2019) Efficient UV photodetectors based on ni-doped
ZnS nanoparticles prepared by facial chemical reduction method. Phys B Condens Matter
575:411690
19. Prasad N, Balasubramanian K (2018) Effect of morphology on optical and efficiently enhanced
electrical properties of W-ZnS for UV sensor applications. J Appl Phys 124(4):045702
20. Sun Y-L, Xie D, Sun M-X, Teng C-J, Qian L, Chen R-S et al (2018) Hybrid graphene/
cadmium-­free ZnSe/ZnS quantum dots phototransistors for UV detection. Sci Rep 8(1):1–8
21. Song K-K, Lee S (2001) Highly luminescent (ZnSe) ZnS core-shell quantum dots for blue to
UV emission: synthesis and characterization. Curr Appl Phys 1(2–3):169–173
22. Kim Y, Kim SJ, Cho S-P, Hong BH, Jang D-J (2015) High-performance ultraviolet photode-
tectors based on solution-grown ZnS nanobelts sandwiched between graphene layers. Sci Rep
5(1):1–8
23. Fang X, Zhai T, Gautam UK, Li L, Wu L, Bando Y et al (2011) ZnS nanostructures: from
synthesis to applications. Prog Mater Sci 56(2):175–287
24. Tiwari A, Dhoble S (2017) Critical analysis of phase evolution, morphological control, growth
mechanism and photophysical applications of ZnS nanostructures (zero-dimensional to three-­
dimensional): a review. Cryst Growth Des 17(1):381–407
25. Nemade K, Waghuley S (2016) Ultra-violet C absorption and LPG sensing study of zinc sul-
phide nanoparticles deposited by a flame-assisted spray pyrolysis method. J Taibah Univ Sci
10(3):437–441
26. Hajimazdarani M, Naderi N, Yarmand B (2019) Effect of temperature-dependent phase
transformation on UV detection properties of zinc sulfide nanocrystals. Mater Res Express
6(8):085096
27. Prasad N, Karthikeyan B (2019) Phase-dependent structural, optical, phonon and UV sensing
properties of ZnS nanoparticles. Nanotechnology 30(48):485702
28. Chatterjee A (2016) Performance analysis and optimization of MSM photodetectors using
group-IV semiconductors. Project report, Kalyani Government Engineering College, Kalyani,
India. https://www.researchgate.net/publication/304347238
29. Ghosh P, Ahmed SF, Jana S, Chattopadhyay K (2007) Photoluminescence and field emission
properties of ZnS: Mn nanoparticles synthesized by rf-magnetron sputtering technique. Opt
Mater 29(12):1584–1590
30. Tang W, Cameron D (1996) Electroluminescent zinc sulphide devices produced by sol-gel
processing. Thin Solid Films 280(1–2):221–226
31. Lee EY, Tran NH, Russell JJ, Lamb RN (2003) Structure evolution in chemical vapor-­deposited
ZnS films. J Phys Chem B 107(22):5208–5211
14 ZnS-Based UV Detectors 347

32. Schön S, Chaichimansour M, Park W, Yang T, Wagner B, Summers C (1997) Homogeneous


and δ-doped ZnS: Mn grown by MBE. J Cryst Growth 175:598–602
33. Torimoto T, Obayashi A, Kuwabata S, Yasuda H, Mori H, Yoneyama H (2000) Preparation of
size-quantized ZnS thin films using electrochemical atomic layer epitaxy and their photoelec-
trochemical properties. Langmuir 16(13):5820–5824
34. Hennayaka H, Lee HS (2013) Structural and optical properties of ZnS thin film grown by
pulsed electrodeposition. Thin Solid Films 548:86–90
35. Goudarzi A, Aval GM, Park SS, Choi M-C, Sahraei R, Ullah MH et al (2009) Low-temperature
growth of nanocrystalline Mn-doped ZnS thin films prepared by chemical bath deposition and
optical properties. Chem Mater 21(12):2375–2385
36. Ebrahimi S, Yarmand B, Naderi N (2019) Enhanced optoelectrical properties of Mn-doped
ZnS films deposited by spray pyrolysis for ultraviolet detection applications. Thin Solid Films
676:31–41
37. Ebrahimi S, Yarmand B (2020) Solvothermal growth of aligned SnxZn1-xS thin films for tun-
able and highly response self-powered UV detectors. J Alloys Compd 827:154246
38. Ebrahimi S, Yarmand B, Naderi N (2017) Effect of the sulfur concentration on the optical
band gap energy and Urbach Tail of spray-deposited ZnS films. Advanced Ceramics Progress
3(4):6–12
39. Ebrahimi S, Yarmand B, Naderi N (2020) High-performance UV-B detectors based on
MnxZn1-xS thin films modified by bandgap engineering. Sensors Actuators A Phys 303:111832
40. Ebrahimi S, Yarmand B (2019) Morphology engineering and growth mechanism of ZnS nano-
structures synthesized by solvothermal process. J Nanopart Res 21(12):1–12
41. Ebrahimi S, Yarmand B (2021) Tunable and high-performance self-powered ultraviolet detec-
tors using leaf-like nanostructural arrays in ternary tin zinc sulfide system. Microelectron J
116:105237
42. Liang Y, Liang H, Xiao X, Hark S (2012) The epitaxial growth of ZnS nanowire arrays and
their applications in UV-light detection. J Mater Chem 22(3):1199–1205
43. Peng Q, Jie J, Xie C, Wang L, Zhang X, Wu D et al (2011) Nano-Schottky barrier diodes
based on Sb-doped ZnS nanoribbons with controlled p-type conductivity. Appl Phys Lett
98(12):123117
44. Zhou Y, Chen G, Yu Y, Feng Y, Zheng Y, He F et al (2015) An efficient method to enhance the
stability of sulphide semiconductor photocatalysts: a case study of N-doped ZnS. Phys Chem
Chem Phys 17(3):1870–1876
45. Dive AS, Huse NP, Gattu KP, Sharma R (2017) Soft chemical growth of Zn0.8Mg0.2S one
dimensional nanorod thin films for efficient visible light photosensor. Sensors Actuators A
Phys 266:36–45
46. El-Shazly O, Farag A, Rafea MA, Roushdy N, El-Wahidy E (2016) Light scattering and pho-
tosensitivity characteristics of nanocrystalline Zn1− xCdxS (0≤ x≤ 0.9) films for photosensor
diode application. Sensors Actuators A Phys 239:220–227
Chapter 15
Photodetectors Based on II-VI
Multicomponent Alloys

Ghenadii Korotcenkov and Tetyana Semikina

15.1 Introduction

As was shown earlier in the (Chap. 2, Vol. 1), compounds II-VI have a specific set of
parameters, which allows them to be widely used in the development of various devices
(read Chaps. 1 and 2, Vol. 1). In particular, the band gap of these compounds can vary
from −0.3 eV (HgTe) to 3.54 eV (ZnS). This means that photodetectors based on these
compounds can cover the entire spectral region from infrared (IR) to ultraviolet (UV).
However, in some cases it turns out that the band gap of binary compounds does not
correspond to the optimal value required to achieve maximum efficiency.
The same situation arises when trying to develop monolithic optoelectronic
devices in which it would be possible to combine the existing achievements of the
well-established technology of III-V compounds and the properties of II-VI com-
pounds, in other words, to develop technologies that allow fabrication on the same
substrate both the III-V-based and II-VI-based high-quality optoelectronic devices.
When using epitaxial methods for growing semiconductor compounds, the main
obstacle to this is the lack of matching of the crystal lattice parameters of III-V and
II-VI binary compounds.
However, it turned out that the above problems can be successfully solved. II-VI
semiconductors can form a continuous series of solid solutions, which creates con-
ditions for a smooth change in both the band gap and the crystal lattice parameter in
a certain range of values determined by the materials used. It is these prospects that

G. Korotcenkov (*)
Department of Physics and Engineering, Moldova State University,
Chisinau, Republic of Moldova
e-mail: ghkoro@yahoo.com
T. Semikina
Lashkarev Institute of Semiconductor Physics, National Academy of Science of Ukraine,
Kiev, Ukraine

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 349
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_15
350 G. Korotcenkov and T. Semikina

are the main stimulus for the development of technology for the synthesis and study
of multicomponent semiconductors based on II-VI compounds. Optimization of the
electrophysical and physical properties of II-VI compounds, or giving them new
properties not characteristic of binary compounds through the introduction of addi-
tional components, is another direction in the study of multicomponent II-VI
compounds.

15.2 Photodetectors with Controlled Spectral Response

15.2.1 Solar Cells

The efficiency of any photodetector is determined by the consistency of the region


of maximum sensitivity with the radiation wavelength. If this agreement does not
exist, then high efficiency cannot be expected. This feature of photodetectors is
most clearly manifested in solar cells. If the spectral distribution of the intensity of
solar radiation is not taken into account, then the efficiency of using solar energy
incident on the Earth will be insignificant. For example, the band gap of silicon
(Eg ~ 1.1 eV, λ = 1127 nm), the material most widely used in solar cells, is not opti-
mal for these applications. Most of the sunlight is not involved in the generation of
photocarriers. In addition, due to the Si indirect band gap and the related low absorp-
tion coefficient, a thickness on the order of 100 μm are required for adequately
absorb sunlight. Therefore, numerous studies are currently being carried out aimed
at finding materials and configurations of solar cells that contribute to the most
complete use of the energy of solar radiation (Fig. 15.1).
In particular, it has been shown that a II-VI compound such as CdTe with a band
gap of 1.45–1.5 eV (λ ~ 840 nm) has more optimal properties for these applications

Fig. 15.1 Spectrum of solar radiation


15 Photodetectors Based on II-VI Multicomponent Alloys 351

[26]. CdTe has become a strong candidate for photovoltaic applications due to its
optimum bandgap, high absorption coefficient, and ease of deposition. Currently
the development of CdTe-based solar cells is one of the most promising thin film
technologies that have already reached the commercial stage. At the research and
development scale, solar submodules above 18% have been reported for this tech-
nology, while commercially available modules show overall area efficiencies in the
range of 14–16%, depending on the manufacturer and the technology. But even the
use of this material cannot solve the problem of the full use of solar energy.
One of the approaches being developed to solve these problems is the develop-
ment of tandem solar cells, consisting of two cells of different bandgap semiconduc-
tors on the top of each other (1.7–1.8 eV (λ ~ 708 nm) on 1.1 eV). The Fig. 15.2
shows how effective the use of tandem solar cells can be to improve the efficiency
of solar radiation conversion. Semiconducting alloys such as Cd1-xZnxTe and Cd1-­x
MnxTe are good candidates for the top cell since their bandgaps can be tailored
between 1.45 eV (CdTe) and 2.26 eV (ZnTe) or 2.85 eV (MnTe) by adjusting the
film composition. Studies have shown that these solid solutions have the required
band gap at x = 0.42 and x = 0.16 for Cd1-xZnxTe and Cd1-xMnxTe, respectively.
However, it turned out that Cd1-xZnxTe films have a more uniform structure and a
sharper interface, which ensured their predominant use in the development of vari-
ous types of solar cells [6, 58]. Cd1-xZnxTe thin films have been deposited by several
different techniques, and depending on the growing parameters used, whether epi-
taxial [54] or polycrystalline thin films with different grain sizes, they lead to varia-
tions of their optical and electrical properties [5]. This ternary system has been
shown to be one of the semiconductor materials allowing tandem solar cell to absorb
most of the energy in the solar spectrum.
Research has also shown that the CZT film can be more heavily doped with Cu
than CdTe. This indicates that the Fermi level is closer to the valence band edge in
CZT than in CdTe. The use of CZT:Cu/Te as the back buffer layers (BBLs) between
CdTe and the back electrode, results in improved device performance by reducing
recombination at the back of the device. Temperature dependent current-voltage

Fig. 15.2 The theoretical efficiency of solar cells can be improved by sequentially harvesting the
Sun’s constituent spectral components in tandem, using multi-junction cells that reduce losses
associated with intraband relaxation. (Reprinted with permission from Ref. [18]. Copyright 2012:
Springer Nature)
352 G. Korotcenkov and T. Semikina

analysis shows that the valence band edge of Cd0.5Zn0.5Te is close to the CdTe edge.
Taken together, these results indicate that it should be possible to develop a high
performance back contact with the desired Fermi level alignment at the back of the
device [26] and upward band bending at the back of the device.
However, some developers believe that ternary Cd1-xMnxTe and Cd1-xMgxTe
alloys still have some prospects for use in CdTe-based tandem solar cells [30]. To
reach Eg = 1.7–1.8 eV, the Mn and Mg alloys require only about 15% alloying in
contrast to more than 40% for Zn in Cd1-xZnxTe. These two alloys also have smaller
lattice contraction with x-value than does Zn. These two advantages are balanced
somewhat by the greater sensitivity of the alloys of Mn and Mg to oxygen and water
vapor. Therefore, the realization of the prospects of these alloys will depend on
whether it is possible to reduce the change in their properties under the influence of
the environment.
Another II-VI alloy promising for use in solar cells is Cd1-xZnxS. Ternary
cadmium-­zinc sulfide combined with CdS is a good candidate for a wide bandgap
window material for conventional CdTe/CdS cell. Its band gap can be tuned from
2.42 eV (CdS) to 3.6 eV (ZnS). Of course, ZnS can play the role of a wide band gap
window, but the required efficiency of a solar cell can only be achieved if ZnS is
heavily doped, which is difficult. At the same time, Cd1-xZnxS, unlike ZnS, can have
a much lower resistance. That is why the use of Cd1-xZnxS improves the spectral
response of solar cells to wavelength without compromising the transport properties
and shunt resistance of the cell. However, the use of a CdZnS film in a CdTe-based
device is associated with some problems. First, as the band gap changes, the lattice
constant also changes linearly. Second, as the Zn content in the compound increases,
a significant increase in electrical resistivity is observed. Electrical resistivity
increases from <1 Ω·cm to >1010 Ω·cm as x increases from 0 to 1.0 [57]. This means
that we have restrictions on the band gap engineering during forming transparency
window of the solar cell.
Previously, we considered II-VI solid solutions as a material for wide-gap trans-
parency windows or top cells in tandem solar cells. However, it turned out that some
ternary alloys can also be used in bottom cells. Such material is Hg1-xCdxTe alloy
[34, 51]. These alloys at x ~ 80% have a band gap Eg ≈ 1.1 eV, which is optimal for
bottom cells in tandem devices. Moreover, this compound appeared to have a slight
tendency to lose initial stoichiometry during the optimizing annealing process.
Another promising direction in the development of solar cells based on ternary
and quaternary II-VI alloys is the development of flexible solar cells [21] and solar
cells using as an absorbing layer a material with a bandgap that varies in thickness
[31], which can be easily implemented using a multicomponent alloy (Fig. 15.3).
The development of cheap thin film solar cells using solution processing is also
an important area for application of multicomponent chalcogenides. Solution based
approaches are especially interesting for their potential low production costs and
their easy scalability. It was found that for these purposes it is very promising
Cu2ZnSnS4, a quaternary p-type semiconductor which presents great absorption
coefficient (>104 cm−1) and have tunable band gap (1.0–1.5 eV) (see Fig. 15.4) by
varying the S/Se ratio [48]. The main advantage of this alloy is that it is formed by
15 Photodetectors Based on II-VI Multicomponent Alloys 353

Fig. 15.3 Proposed structure for new solar cells based on CdZnTe with variable Zn content for
which higher efficiencies than conventional CdTe solar cells are expected. (Adapted with permis-
sion from Ref. [31]. Copyright 2011: Elsevier)

Fig. 15.4 The calculated band gap of Cu2ZnSn(S1 − xSex)4 at different composition (x). (Adapted
with permission from Ref. [8]. Copyright 2011: American Physical Society)

abundant and nontoxic elements. Record solar cell efficiency for this material is
12.6% at the research scale denoting a high improvement potential [52].

15.2.2 Detectors for Visible Range

As we can see, solar energy converters require devices that are sensitive to the entire
spectrum of solar radiation, which means that devices must respond to radiation in
the entire spectral range from IR to UV (see Fig. 15.1). In the case of photodetec-
tors, the requirements for devices may change radically, since in many cases it is not
broadband photodetectors that are required, but devices characterized by spectral
354 G. Korotcenkov and T. Semikina

selectivity. It is important to note that this problem, as well as in solar cells, can be
solved using multicomponent II-VI alloys. For example, if we need a photodetector
for the visible region of the spectrum that is not sensitive to radiation in the near-IR
region, then according to Ren et al. [36], photodetectors based on a single crystal of
Cd0.96Zn0.04Te can be useful in this case. It has been shown that metal-­semiconductor-­
metal (MSM) photodetectors based on Cd0.96Zn0.04Te have very sharp edges in the
region of 900–800 nm (see Fig. 15.5a). The discrimination ratio between the near
infrared region of 800 nm and 900 nm is almost 102, which is enough for the accu-
rate spectra selectivity. Benefitting from the high-quality single crystallization, an
ultra-low dark current of ~10−10 A was obtained at a high applied voltage of 10 V,
leading to a photo-to-dark-current ratio of more than 103 at 700 nm light illumina-
tion. The highest responsivity was estimated to be 1.43 A/W with a specific detec-
tivity of 3.3 × 1012 Jones at −10 V at a relatively lower injection power density. In
addition, the MSM photodetector also exhibited a fast response speed of ~800 μs
(Fig. 15.5b), and extremely low persistent photoconductivity (PPC), while the PPC
is inhibited at high temperatures.
If, however, solid solutions are used in the development of heterojunction, n-n+,
or p-n junction-based photodetectors, then in this case it is possible to manufacture
selective photodetectors for the visible region of the spectrum with a tunable sensi-
tivity region [14, 27]. For the visible region of the spectrum, the most optimal and
most studied material is CdSSe [30, 38] and CdSTe alloys [47]. For example,
Fig. 15.6 shows the spectral characteristics of injection photodetectors developed on
the basis of CdSYSe1–Y alloy [27]. The photodetectors had an n–n + structure and
operated in the forward bias mode, when the injection amplification of the photo-
current is realized. Such photodiodes are designed to work with weak light fluxes.
The mechanism of the internal injection amplification of the photocurrent under
forward bias is based on the redistribution of the bias voltage between the high-­
resistance and low-resistance regions of the structure. The reason for this redistribu-
tion is the change in the conductivity of the high-resistance region when it is

Fig. 15.5 (a) Spectral photoresponse/specific detectivity of fabricated Cd0.96Zn0.04Te photodetec-


tor; (b) transient response time measurement at a bias voltage of 0.2 V with a chopping frequency
of 100 Hz. (Reprinted from Ref. [36]. Published 2019 by Optica Publishing Group as open access)
15 Photodetectors Based on II-VI Multicomponent Alloys 355

Fig. 15.6 Spectral characteristics of injection photodiodes based on CdSYSe1–Y: 1– y = 0.96,


Ub = 2 V; 2 – y = 0.95, Ub = 1 V; 3 – y = 0.94, Ub = 0.5 V; 4 – y = 0.9, Ub = 4 V; 5 – y = 0.85,
Ub = 0.5 V; 6 – у = 0.55, Ub = 4 V; 7 – у = 0.3, Ub = 4 V; 8 – у = 0.06, Ub = 6 V. (Data extracted
from Ref. [27])

Fig. 15.7 (a) UV − visible absorption spectra of CdSeS NWs with varied Se:S ratio. Reprinted
with permission from [22]. Copyright 2014: ACS; (b) Responsivity of CdS and CdS-CdSxTe1-x-­
CdTe-based photodetectors relative to modulation frequency under incident light density of
15.4 mW/cm2 with a bias voltage of 5 V. (Reprinted with permission from Ref. [47]. Copyright
2018: Elsevier)

illuminated with light with a wavelength λ close to the absorption edge of the single
crystal used, and the appearance of a positive current feedback in connection with
this. This mechanism is well described in [50]. As can be seen, the positions of the
photosensitivity maxima of the structures were in the wavelength range of
520–850 nm. The spectral sensitivity band at a level of 0.5 for most detectors was
15–30 nm, which indicates the high selectivity of the developed photodetectors.
Besides traditional photodetectors [38], CdSSe and CdSTe alloys are being used
to develop nanowire- [22], and core-shell nanobelt-based photodetectors [47]. The
steady-state absorption spectra of the CdSeS NWs with the absorption of the NWs
normalized at 400 nm are shown in Fig. 15.7a. By varying the Se:S precursor ratio
we were able to tune the optical band gap of the CdSeS NWs from 2.36 to 1.79 eV. At
the same time, the measured response spectrum of CdS-CdSxTe1-x-CdTe core-shell
nanobelt photodetector covers a wide range from 355 nm (3.4 eV) to 785 nm
356 G. Korotcenkov and T. Semikina

(1.58 eV) [47]. At that, Tang et al. [47] reported that CdS-CdSxTe1-x-CdTe photode-
tector had a rise time of 11 μs, which is the fastest among CdS based photodetectors,
reported previously. As a result, CdS-CdSxTe1-x-CdTe photodetector demonstrated
the best frequency characteristics (Fig. 15.7b). Tang et al. [47] believe that the core-
shell structure plays an influential role in improving the photoresponse rate. For
pure CdS NB-based photodetectors, the photoresponse is mainly governed by
desorption and adsorption of oxygen [16]. Since the equilibrium of gas molecules
interacting with carriers takes longer time, the response speed of pure nanostructure
based photodetectors is limited. In contrast, in core-shell NB photodetectors, the
intrinsic positive charge in CdTe shell will avoid the absorption of oxygen and thus
operation of this kind NB photodetectors mainly depends on the built-in potential at
the interface, which has a much faster carriers-separating speed. Because of the
abovementioned mechanism, the responsivity and response speed of the CdS-­
CdSxTe1-­x-CdTe core-shell NB photodetectors have been extraordinarily optimized.

15.2.3 UV Detectors

The need to optimize the absorption spectrum also exists in UV detectors. However,
unlike solar cells or visible radiation sensors, in UV sensors, on the contrary, it is
necessary to suppress the sensitivity of devices in the visible region of the spectrum.
Research has shown that this problem can be successfully solved by using ternary
Zn1-xMnxS samples [13, 45]. Manganese (Mn2+) is one of the most well studied dop-
ant compounds II-VI, used in the development of luminescent-based optical sensors
[23]. Due to the relatively close ionic radii of Mn and Zn, as well as the same ionic
charge, Mn can fundamentally be introduced in the form of Mn2+ ions into various
sites of the ZnS host lattice and change its structural and optoelectric properties.
Ebrahimi et al. [13] found that all the deposited Zn1-xMnxS showed a stable Zinc-
Blende structure. The nanocrystalline, uniform and smooth surface of the alloys led
to increasing the optical transmittance over 95% at the high-content of Mn2+ ions
(see Fig. 15.8a). At a high concentration of Mn2+ ions (>0.2 mol), the values of
bandgap energy increased dramatically from 3.91 eV to 4.05 eV, while the Urbach
energy decreased considerably from 361 meV to 225 meV. Therefore, the bandgap
energy in the absence of defects and band tailing and at the rearrangement of elec-
tronic localized states was obtained of about 4.28 eV. For comparison the Eg of ZnS
equals 3.54 eV.
The performances of UV detectors based on Zn1-xMnxS thin films also showed a
huge improvement in photosensitivity and photoswitching compared to ZnS devices.
For example, the current gain of the detectors was improved from ~120 for pure
ZnS to ~580 for the Mn0.4Zn0.6S sample. The photoresponse rate of UV detectors has
also increased significantly. For Mn0.4Zn0.6S detectors, rise time (τr) and fall time
(τd) decreased from 28.1 and 34.3 ms observed for ZnS devices to 5.3 and 6.1 ms,
respectively. An important advantage of Zn1-xMnxS-based detectors is also the
expansion of the spectral region where these devices can be effectively used as
15 Photodetectors Based on II-VI Multicomponent Alloys 357

Fig. 15.8 (a) Optical transmittance of the pure ZnS and MnxZn1-xS thin films. Adapted with per-
mission from [13]. Copyright 2020: Elsevier. (b) Photoresponse of four Zn1-xMgxS Schottky bar-
rier photodiodes with long-wavelength cutoffs at 325, 305, 295, and 270 nm, respectively. (Adapted
with permission from Ref. [45]. Copyright 2001: AIP Publishing)

selective UV detectors [45]. For example, with Zn0.25Mg0.75S compositions, it is pos-


sible to manufacture a detector that will not only be insensitive to visible radiation,
but also to UV radiation from the UV-A (320–400 nm) and UV-B (280–320 nm)
regions (see Fig. 15.8b). For the device with the highest Mg composition of 75%
and with a thickness of the active layer as thin as 30 nm, the external quantum effi-
ciency still can be maintained around 15% in the plateau region. In particular, this
device has a cutoff wavelength at 270 nm and offers a fall of almost three orders in
responsivity at 300 nm relative to its peak, indicating very good solar–blind charac-
teristics. The ability to manufacture sensors for different spectral ranges can be
useful for pollution detection, medical applications, and as special purpose detec-
tors. For example, such sensors can be used for development of flame sensing sys-
tem. This system requires UV detectors that are insensitive to radiation above
280 nm [25].
Averin et al. [3] have shown that ZnCdS and ZnMgS can also be used in the
development of metal-semiconductor-metal (MSM) photodetectors based on peri-
odic heterostructures with ZnCdS quantum wells separated by ZnMgS and ZnS
barrier layers. Heterostructures were grown on semi-insulating GaP substrates by
MOVPE. It was established that these detectors exhibited very low dark currents
and electrically tunable spectral response. At low bias detectors provide narrowband
UV-response determined by a composition of ZnCdS quantum well. A shift of the
peak position of the narrowband detector response to longer wavelengths is observed
with increasing Cd content. A higher operating bias shifts the maximum sensitivity
of the detector in the visible part of spectrum due to the penetration of external
electric field down to the semi-insulating GaP substrate while a narrowband
UV-response remains. Averin et al. [3] believe that developed highly selective two-­
color photodetector allows discriminating the optical channels and increasing
dynamic range and noise immunity of optical informational and measuring systems.
Other ternary compounds, such as ZnSSe [44], ZnSeTe, ZnSTe [28, 44], CdMgTe
[30], and CdSxTe1-x [47] have also been tried to control the spectral response region
358 G. Korotcenkov and T. Semikina

of UV detectors. However, the use of ternary compounds instead of binary ones did
not always give the expected results. In particular, Sou et al. [44] studied ZnSSe and
ZnSTe-based Schottky barrier UV detectors and concluded that of these compounds,
only ZnSSe is a good candidate for an active material for UV detection applications
that require excellent visible rejection and tunable turn-on wavelength capabilities.
As seen in Fig. 15.9b, ZnSSe diodes have a sharper long wavelength limit compared
to ZnSTe diodes. This is due to the fact that, unlike ZnSTe, ZnSSe does not contain
isoelectronic traps responsible for the appearance of a photoresponse in the long-­
wavelength region of the spectrum in ZnSTe-based detectors (see Fig. 15.9a). In
addition, good activation of dopant donor impurities up to 50% Se in the alloy is
achieved for ZnSSe. In ZnSTe, this cannot be achieved, since high Te containing
ZnSTe alloy cannot be n-doped. If we compare ZnSSe with CdS and ZnS, which are
traditional materials for UV detectors, we will see that ZnSSe, unlike these semi-
conductors, can be used to detect UV radiation in the UV-A range (320–400 nm),
where CdS and ZnS have very low sensitivity. This indicates that ZnSSe detectors
can be used to measure UV radiation in all three spectral regions of UV radiation –
UV-A (320–400 nm), UV-B (280–320 nm), and UV-C (100–280 nm).
The same results were obtained when using quadruple solid solutions, such as
ZnMgBeSe [49]. Figure 15.10 shows the spectral response of a typical ZnMgBeSe-­
based detector measured at −10 V bias, with front-side illumination through the
semitransparent Schottky contact.
It can be seen that the response is very flat above the band gap, due to the position
of the depleted region on top of the structure: this is an advantage of Schottky bar-
rier devices in comparison with p – n photodiodes. The cutoff at 380 nm is very
sharp with a visible rejection rate of more than three orders of magnitude (5х103),
which is a consequence of the high crystalline quality of the structure. A maximum
responsivity of 0.17 A/W is obtained at 375 nm corresponding to a quantum effi-
ciency of 54%. It is important to note that a quantum efficiency remains over 50%
above the whole UV range from 380 to 315 nm, and drops off slightly as the wave-
length decreases further. This is close to the theoretical limit. Moreover, due to the

Fig. 15.9 Photoresponse curves of (a) ZnS and ZnS0.94Te0.06, and (b) ZnS0.9Se0.1 and ZnS0.58Se0.42
based Schottky barrier UV detectors. (Adapted with permission from Ref. [44]. Copyright 2000:
Springer)
15 Photodetectors Based on II-VI Multicomponent Alloys 359

Fig. 15.10 Spectral response of a ZnMgBeSe Schottky photodiode at −10 V bias. Theoretical
curves corresponding to different quantum efficiencies (η) are indicated. (Adapted with permission
from Ref. [49]. Copyright 2001: AIP Publishing)

use of epitaxial films, no internal gain mechanism has been observed. And this
means that a change in the state of the outer atmosphere will not affect the param-
eters of these detectors. A detectivity of 2 × 1010 mHz1/2 W−1 has also been mea-
sured, highlighting the fact that these detectors are both sensitive and little noisy.
In addition, Vigué et al. [49] compared the fabricated detectors with detectors
based on other materials and showed that the parameters of ZnMgBeSe detectors
are superior to those of many analogues. ZnS- and ZnSSe-based structures have
responsivities of 0.08 A/W at 335 nm and 0.09 A/W at 370 nm, respectively [43].
ZnSTe devices exhibit a high response of 0.13 A/W at 320 nm, but the tellurium
incorporation leads to a significant decrease of the rejection rate and generates a
large cutoff in the spectral response [28]. ZnMgBeSe photodetectors also compare
advantageously with 6H-SiC (0.15 A/W) [2], GaN (0.10–0.18 A/W) [7, 35], and
AlGaN (0.07 A/W) [33].

15.3 Solid Solutions Providing Lattice Matching


of Contacting Semiconductor Materials

The most obvious example of a problem that has been solved through the use of
II-VI ternary compounds is the growth of CdZnTe single crystals with lattice param-
eters matched to the lattice parameter of CdHgTe, the main material of IR technol-
ogy. It turned out that for the manufacture of high-performance photodetectors array
by epitaxial methods, large-diameter substrates (> 50 cm) with the same lattice
parameter are needed. Unfortunately, at the moment there is no CdHgTe technology
that makes it possible to grow single crystals of this diameter. At the same time,
360 G. Korotcenkov and T. Semikina

CdZnTe technology allows solving this problem (read the Chap. 9, Vol. 1). With the
composition Cd1-xZnxTe (0.02 < x < 0.06), the lattice parameter of this compound
coincides with the lattice parameter of CdxHg1-xTe (0.19 < x < 0.23 and
0.27 < x < 0.32), optimized under the required IR range, which makes it possible to
grow high-quality CdHgTe epitaxial layers on their basis, which are required for the
manufacture of high-performance matrix photodetectors (Chap. 15, Vol. 1).
Moreover, the use of a wide-gap material as a substrate made it possible to develop
IR photodetectors with illumination through the substrate (Fig. 15.11), which sig-
nificantly facilitated the technology for manufacturing IR photodetector arrays and
improved their performances.
The problem of integration of II-VI and III-V compounds is solved in a similar
way. Depending on the tasks set, optoelectronic devices based on III-V compounds
are manufactured on the basis of single-crystal substrates of these compounds, but
mainly on GaAs. Therefore, when developing optoelectronic devices based on II-VI
compounds integrated with devices based on III-V compounds, it is necessary to
match the lattice parameters of GaAs and II-VI compounds in order to achieve the
required device parameters. Matching the lattice parameters of the substrate and the
epitaxial layers grown on their surface significantly reduces the mechanical stresses
at the interface and the imperfection of the grown layer itself, which naturally
affects the improvement of the operational parameters of the manufactured devices.
The experiment showed that II-VI multicomponent solid solutions used as transi-
tion or buffer layers make it possible to solve this problem. In this case, the struc-
tures of II-VI-based photodiodes fabricated on a GaAs substrate may look as shown
in Fig. 15.12.
Using this approach, it was possible to match the lattice parameters of AIIBVI
compounds with the lattice parameters of other semiconductors, including Si [42],

Fig. 15.11 Schematic diagram: (a) back side illuminated n+/ν/p + HgCdTe e-APD (b) bandgap
grading in the device structure. (Reprinted with permission from Ref. [40]. Copyright 2018:
Elsevier)
15 Photodetectors Based on II-VI Multicomponent Alloys 361

Fig. 15.12 (a) A schematic structure of a ZnSe p+ − n avalanche photodiode grown on GaAs by
MBE. The surface ohmic contact layer consists of ZnTe–ZnSe multiquantum wells (with five
periods) and a thin ZnTe cap layer. Reprinted with permission from [20]. Copyright 2000: AIP
Publishing. (b) Schematic representation of a typical ZnBeSe–ZnMgBeSe p-i-n photodiode.
(Adapted with permission from Ref. [49]. Copyright 2001: AIP Publishing)

GaP [42], InSb [56], and on single-crystal substrates of these semiconductors to


produce high-performance optoelectronic devices based on II-VI compounds,
including lasers and photodetectors. In particular, using ZnS1-xTex solid solutions as
buffer layers, it was possible to fabricate highly efficient UV detectors on GaP sub-
strates [42]. The structure of these photodetectors and some parameters are shown
in Fig. 15.13. Sou et al. [42] believe that a direct-gap semiconductor ZnS1-­xTex alloy
with a band gap covering 2.1–3.7 eV is a promising material for visible and ultra-
violet optoelectronic applications. An important advantage of ZnSTe alloys is that
their lattice can be matched with a number of commercially available substrates
such as Si, GaAs, and GaP, which was confirmed experimentally. In particular, the
3%-Te alloy has a lattice-matched silicon substrate, and thus ZnSTe-based devices
are compatible with advanced silicon integration technology. At the same time,
Zhang and co-workers [24, 55] believe that for CdTe-based double heterostructures
grown on InSb substrate, the Mg0.46Cd0.54Te layer is an ideal barrier layer, which can
effectively confine both electrons and holes and reduce the surface recombination.
As we can see, ternary alloys are used as a buffer layer in the given examples.
However, it has been found that in addition to ternary alloys, more complex com-
pounds can be used. For example, Vigué et al. [49] for such matching used the
ZnMgBeSe solid solution, which can also have a lattice parameter equal to that of
GaAs (Fig. 15.14). It is important to note that the use of a quadruple alloy instead of
a ternary one has significant advantages, as it provides more opportunities for band
gap engineering. If a ternary alloy has one composition with a fixed band gap, which
ensures the equality of the lattice parameters with the substrate, then in a quaternary
362 G. Korotcenkov and T. Semikina

Fig. 15.13 (a) Photovoltage for Schottky barrier Au-ZnSTe structure grown on a GaP substrate;
and (b) external quantum efficiency vs photon wavelength for Schottky barrier Au-ZnSTe struc-
tures grown on GaP and Si substrates. (Reprinted with permission from Ref. [42]. Copyright 1997:
AIP Publishing)

Fig. 15.14 Energy of wide-band gap II-VI semiconductors as a function of their lattice parameter.
(Adapted with permission from Ref. [49]. Copyright 2001: AIP Publishing)

alloy there are much more such possibilities. As shown in Fig. 15.13, ZnMgBeSe
quaternary alloys can be grown on lattice-matched GaAs substrates with a direct
band gap ranging from 2.75 to greater than 4.2 eV. These compounds present high
crystalline quality, which should lead to high sensitivity and detectivity.
15 Photodetectors Based on II-VI Multicomponent Alloys 363

15.4 Optimization of Electrophysical and Physical Properties


of II-VI Compounds

Optimizing the parameters of optoelectronic devices is a multifactorial task that can


be solved in various ways. Improving the electrical and physical properties of II-VI
semiconductors through the complication of the composition is one of such
approaches. Earlier, we noted that by using Cd0.96Zn0.04Te alloys, it is possible to
optimize the spectral sensitivity of photodetectors [36]. However, it turned out that
the transition to solid solutions also makes it possible to get rid of some of the dis-
advantages inherent in CdTe [36]. Compared to the commercially used Si semicon-
ductor, CdTe is much more sensitive to the light from visible to infrared region
because it has a direct band gap and a relatively higher optical absorption coeffi-
cient. The higher carrier mobility also makes it promising for photovoltaic applica-
tions. However, polycrystalline CdTe photodetectors have been found to have a
rather slow response rate due to the strong persistent photoconductivity (PPC),
which was induced from abundant grain boundaries and surface state traps [53].
Single crystal CdTe has better parameters. However, even in this case, due to the
specifics of the technology, the CdTe single crystal has a significant concentration
of intrinsic point defects and residual impurities, which cause the appearance of
localized levels in the band gap. This resulted in high leakage current, severe carrier
trapping, and low responsivity of X-ray detectors. Another important issue in single
crystal CdTe is polarization charges under bias, which can act as trapping centers
affecting the space charge distribution and electric field profile [29]. Studies have
shown that the use of single crystal Cd1-xZnxTe instead of CdTe can be a solution to
the above problems. By increasing the Zn concentration, it is possible to adjust the
band gap of alloys from 1.44 eV (for CdTe) to 2.2 eV (for ZnTe) and synthesize a
material with high electrical resistivity (ρ = (1–4) × 1010 Ohm·cm) close to the theo-
retical maximum [46]. In addition, it was reported that the polarization effect can
also be eliminated in Cd1-xZnxTe crystals [41].
A promising direction is also the doping of II-VI compounds with Mn, Cr, Co,
and Fe [1]. Wide-gap II-VI compounds, which do not have magnetic properties,
after doping with these elements, begin to exhibit magnetic and feromagnetic prop-
erties, which significantly expands the possible areas of their application. Such
semiconductors belong to the class of so-called diluted magnetic semiconductors
(DMS) [11, 12]. An important feature of dilute magnetic semiconductors is the abil-
ity to control the magnetic order through the concentration of carriers, which is
easily changed by the gate voltage, and the ability to remagnetize layers using spin-­
polarized currents with a much lower density compared to traditional ferromagnetic
metals. Due to their uniqueness, these materials are of extreme interest for both
fundamental science and applied science. Diluted magnetic semiconductors are cur-
rently being considered as materials for spin polarized charge carrier injectors in
semiconductor spin electronics devices [1], and for the development of a new gen-
eration of magnetic memory elements. Since ferromagnetic spin-spin interactions
are mediated by holes in the valence band, changing the Fermi level using
364 G. Korotcenkov and T. Semikina

co-doping, electric fields, or light can directly affect magnetic ordering. Moreover,
engineering the Fermi level position by co-doping makes it possible to modify solu-
bility and self-compensation limits, affecting magnetic characteristics in a number
of surprising ways. The Fermi energy can even control the aggregation of magnetic
ions, providing a new route to self-organization of magnetic nanostructures in a
semiconductor host [11, 12].
The most studied II-VI-based diluted magnetic semiconductors are Zn1-xMnxTe,
Cd1-xMnxTe, Zn1-xCrxTe, Cd1-xCrxTe, Cd1-x-yHgxMnyTe, and ZnMnSe [1, 11, 12, 19,
32]. An exchange interaction between the sp-band electrons of the alloy and the
localized d electrons associated with Mn2+ results in a strongly enhanced g-value for
Zeeman splitting of electronic levels, which in turn causes a strong Faraday effect
near the band gap of the DMS material. The large magneto-optical effects of single
crystalline Cd1-xMnxTe and Cd1-x-yHgxMnyTe are already find practical application as
a Faraday rotator in optical isolators for the wavelength region in which magnetic
garnet crystals are not applicable due to presence of strong absorption [19, 32].
However, the magnetic properties of most II-VI-based semiconductors are either
paramagnetic or spin-glass state, which is not very attractive for practical applica-
tions. Recently, the development of a new II-VI semiconductor doping technology
has made it possible to obtain p-type DMS materials based on II-VI allays. Theory
predicts ferromagnetism in some of these DMS materials if they are heavily-doped
p-type. Experiment confirmed that prediction [15]. Heavy p-doped Zn1-xMnxTe and
Cd1-xMnxTe epilayers were synthesized [15].
II-VI compounds doped with indium, aluminum and copper can also be classi-
fied as multicomponent semiconductors, since the concentration of dopants reaches
5%. Moreover, doping is accompanied by a change not only in electrophysical
parameters, but also in the optical properties of semiconductors, which affect the
performances of optoelectronic devices. For example, Cui et al. [9] studied the pho-
toelectron characteristics of ZnSe/ZnS/L-cys and Cu-doped ZnSe/ZnS/L-cys core
shells and found that the surface photovoltaic property of Cu-doped ZnSe QDs is
significantly improved when compared with ZnSe QDs. More specifically, the
intensity of the surface photovoltaic (SPV) response of Cu-doped ZnSe QDs is
approx. Twenty-one times stronger than that of un-doped ZnSe QDs. Cui et al. [9]
believe that the excellent characteristics of Cu-doped ZnSe QDs may result from
their unique electron structure and photogenerated charge-transfer (CT) transition
behaviors. The new SPV response peaks of Cu-doped QDs can be ascribed to the
defect-state levels that are closely related to the doped Cu2+ ions.
Doping with copper gives a positive result in the development of solar cells as
well. It was established that the use of p-type ZnTe:Cu (III-V%) as a back contact
in CdTe solar cells can significantly increase its efficiency [39]. A major challenge
for CdTe solar cell technology is the formation of high quality ohmic back contacts
[10]. p-CdTe has a high work function (~5.7 eV) which is higher than most common
metals [4]. A direct metal contact with a low work function will form a negative
Schottky barrier band bend in the back contact region [10], which is an important
limiting factor for cell performance. Theory and experiment have shown that p-type
ZnTe:Cu is an ideal back contact buffer material with CdTe absorber in both lattice
15 Photodetectors Based on II-VI Multicomponent Alloys 365

matching and energy band matching. ZnTe has a high work function of 5.3–5.8 eV
and a band gap of 2.2 eV, which can not only reduce the contact barrier with a small
valence band gap, but also act as an electron reflector to suppress recombination at
the interface [17, 37].

15.5 Summary

This chapter, which considers various types of photodetectors, has shown that ter-
nary and quaternary solid solutions based on II-VI compounds are indeed promising
materials that allow engineering of both band gap and lattice constant, and thereby
develop devices with the required parameters.

Acknowledgments G. Korotcenkov is grateful to the State Program of the Republic of Moldova,


project 20.80009.5007.02, for supporting his research.

References

1. Akai H, Ogura M (2006) Half-metallic diluted antiferromagnetic semiconductors. Phys Rev


Lett 97:026401
2. Anikin M, Andreev AN, Pyatko SN, Rastegaeva MG, Savkina NS, Strelchuk AM et al (1992)
UV photodetectors in 6H-SiC. Sens Actuators A 33:91–93
3. Averin SV, Kuznetzov PI, Zhitov VA, Zakharov LY, Kotov VM, Alkeev NV (2016) Wavelength
selective UV/visible metal-semiconductormetal photodetectors. Opt Quant Electron 48:303
4. Bätzner DL, Romeo A, Zogg H, Wendt R, Tiwari AN (2001) Development of efficient and
stable back contacts on CdTe/CdS solar cells. Thin Solid Films 387(1–2):151–154
5. Becerril M, Zelaya-Angel O, Fragoso-Soriano R, TiradoMejia L (2004) Band gap energy in
Zn-rich Zn1−xCdxTe thin films grown by r.f. sputtering. Rev Mex Fis 50(6):588–593
6. Chander S, De AK, Dhaka MS (2018) Towards CdZnTe solar cells: an evolution to post-­
treatment annealing atmosphere. Solar Cells 174:757–761
7. Chen Q, Yang JW, Osinsky A, Gangopadhyay S, Lim B, Anwar MZ, Khan MA (1997) Schottky
barrier detectors on GaN for visible–blind ultraviolet detection. Appl Phys Lett 70:2277
8. Chen S, Walsh A, Yang JH, Gong XG, Sun L, Yang P-X et al (2011) Compositional depen-
dence of structural and electronic properties of Cu2ZnSn(S,Se)4 alloys for thin film solar cells.
Phys Rev B 83:125201
9. Cui JY, Li KY, Ren L, Zhao J, Shen TD (2016) Photogenerated carriers enhancement in
Cu-doped ZnSe/ZnS/L-cys self-assembled core-shell quantum dots. J Appl Phys 120:184302
10. Demtsu SH, Sites JR (2006) Effect of back-contact barrier on thin-film CdTe solar cells. Thin
Solid Films 510(1–2):320–324
11. Dietl T, Ohno H (2003) Ferromagnetic III-V and II-VI Semiconductors. MRS Bull
28(11):714–719
12. Dietl T, Ohno H (2006) Engineering magnetism in semiconductors. Mater Today 9(11):18–26
13. Ebrahimi S, Yarmand B, Naderi N (2020) High-performance UV-B detectors based on MnxZn1-­
xS thin films modified by bandgap engineering. Sens Actuators A 303:111832
14. Faschinger W, Ehinger M, Schallenberg T (2000) High-sensitivity p-i-n-detectors for the vis-
ible spectral range based on wide-gap II-VI materials. J Crystal Growth 214(215):1138–1141
366 G. Korotcenkov and T. Semikina

15. Ferrand D, Cibert J, Bourgognon C, Tatarenko S, Wasiela A, Fishman G et al (2000) Carrier


induced ferromagnetic interactions in p-doped Zn(1-x)MnxTe epilayers. J Appl Phys 87:6451
16. Gao T, Li QH, Wang TH (2005) CdS nanobelts as photoconductors. Appl Phys Lett 86:173105
17. Gessert T, Mason A, Sheldon P, Swartzlander A, Niles D, Coutts T (1996) Development of
Cu doped ZnTe as a back contact interface layer for thin film CdS/CdTe solar cells. J Vac Sci
Technol A 14:806–812
18. Graetzel M, Janssen RAJ, Mitzi DB, Sargent EH (2012) Materials interface engineering for
solution-processed photovoltaics. Nature 488:304–312
19. Imamura M, Tashima D, Kitagawa J, Asada H (2020) Magneto-optical properties of wider
gap semiconductors ZnMnTe and ZnMnSe films prepared by MBE. J Electron Sci Technol
18(3):201–211
20. Ishikura H, Abe T, Fukuda N, Kasada H, Ando K (2000) Stable avalanche-photodiode operation
of ZnSe-based p+ –n structure blue-ultraviolet photodetectors. Appl Phys Lett 76:1069–1071
21. Khalil MI, Bernasconi R, Lucotti A, Le Donne A, Mereu RA, Binetti S et al (2021) CZTS
thin film solar cells on flexible molybdenum foil by electrodeposition-annealing route. J Appl
Elecrochem 51:209–218
22. Kim J-P, Christians JA, Choi H, Krishnamurthy S, Kamat PV (2014) CdSeS nanowires: com-
positionally controlled band gap and Exciton dynamics. J Phys Chem Lett 5:1103–1109
23. Li D, Qin J, Xu Q, Yan G (2018) A room-temperature phosphorescence sensor for the detection
of alkaline phosphatase activity based on Mn-doped ZnS quantum dots. Sensors Actuators B
Chem 274:78–84
24. Liu S, Zhao X-H, Campbell CM, Lassise MB, Zhao Y, Zhang Y-H (2015) Carrier lifetimes and
interface recombination velocities in CdTe/MgxCd1−xTe double heterostructures with different
Mg compositions grown by molecular beam epitaxy. Appl Phys Lett 107:041120
25. Liu Y, Pang L-X, Liang J, Cheng M-K, Liang JJ, Chen JS, Lai Y-H, So I-K (2017) A compact
solid-state UV flame sensing system based on wide-gap II-VI thin film materials. IEEE Trans
Ind Elektron 65(3):2737–2744
26. Liyanage GK, Phillips AB, Alfadhili FK, Ellingson RJ, Heben MJ (2019) The role of Back
buffer layers and absorber properties for >25% efficient CdTe solar cells. ACS Appl Energy
Mater 2(8):5419–5426
27. Lubegin G, Onishchenko D, Guslyannikov V (2011) Injection photodiodes based on AIIBVI
single crystals. Photonics (Russia) 1:34–37. (in Russian)
28. Ma ZH, Sou IK, Wong KS, Yang Z, Wong GKL (1998) ZnSTe-based Schottky barrier ultravio-
let detectors with nanosecond response time. Appl Phys Lett 73:2251
29. Malm HL, Martini M (1974) Polarization phenomena in CdTe nuclear radiation detectors.
IEEE Trans Nucl Sci 21(1):322–330
30. Mathew X, Drayton J, Parikh V, Compaan AD (2006) Sputtered Cd1-xMgxTe films for top cells
in tandem devices. In: Proceedings of IEEE 4th world conference on PVEC, pp 327–331
31. Morales-Acevedo A (2011) Analytical model for the photocurrent of solar cells based on
graded band-gap CdZnTe thin films. Solar Energy Mater Solar Cells 95:2837–2841
32. Onodera K, Matsumoto T, Kimura M (1994) 980 nm compact optical isolators using Cd1-x-­
yHgxMnyTe single crystals for high power pumping laser diodes. Electron Lett 39:1954–1955
33. Osinsky A, Gangopadhyay S, Lim BW, Anwar MZ, Khan MA (1998) Schottky barrier photo-
detectors based on AlGaN. Appl Phys Lett 72:742
34. Parikh VY, Marsillac S, Collins RW, Chen J, Compaan AD (2007) Hg1-xCdxTe as the bottom
cell material in tandem II-VI solar cells. MRS Proc 1012(Y12):37
35. Ravikiran L, Radhakrishnan K, Dharmarasu N, Agrawal M, Wang Z, Bruno A et al (2017)
GaN Schottky metal–semiconductor–metal UV photodetectors on Si(111) grown by ammonia-­
MBE. IEEE Sensors J 17(1):72–77
36. Ren B, Zhang J, Liao M, Huang J, Sang L, Koide Y, Wang L (2019) High-performance visible
to near-infrared photodetectors by using (Cd,Zn)Te single crystal. Opt Express 27(6):8935
37. Rioux D, Niles DW, Hochst H (1993) ZnTe: a potential interlayer to form low resistance back
contacts in CdS/CdTe solar cells. J Appl Phys 73:8381–8385
15 Photodetectors Based on II-VI Multicomponent Alloys 367

38. Sahana NM, Mahesh MG (2020) Colour tunable co-evaporated CdSxSe1-x (0 <x < 1) ternary
chalcogenide thin films for photodetector applications thin films for photodetector applica-
tions. Mater Sci Semicond Process 120:105288
39. Shen K, Wang X, Zhang Y, Zhu H, Chen Z, Huang C, Ma Y (2020) Insights into the role of
interface modification in performance enhancement of ZnTe:Cu contacted CdTe thin film solar
cells. Sol Energy 201:55–62
40. Singh A, Pal R (2018) Impulse response measurement in the HgCdTe avalanche photodiode.
Solid State Electron 142:41–46
41. Sordo SD, Abbene L, Caroli E, Mancini AM, Zappettini A, Ubertini P (2009) Progress in the
development of CdTe and CdZnTe semiconductor radiation detectors for astrophysical and
medical applications. Sensors (Basel) 9(5):3491–3526
42. Sou IK, Man CL, Ma ZH, Yang Z, Wong GKL (1997) High performance ZnSTe photovoltaic
visible-blind ultraviolet detectors. Appl Phys Lett 71:3847
43. Sou IK, Ma ZH, Zhang ZQ, Wong GKL (2000a) Temperature dependence of the responsivity
of II-VI ultraviolet photodiodes. Appl Phys Lett 76:1098
44. Sou IK, Ma ZH, Wong GKL (2000b) ZnS-based visible-blind UV detectors: effects of isoelec-
tronic traps. J Electron Mater 29(6):723–726
45. Sou IK, Wu MCW, Sun T, Wong KS, Wong GKL (2001) Molecular-beam-epitaxy-grown
ZnMgS ultraviolet photodetectors. Appl Phys Lett 78:1811
46. Takahashi T, Watanabe S (2001) Recent progress in CdTe and CdZnTe detectors. IEEE Trans
Nucl Sci 48(4):950–959
47. Tang M, Xu P, Wen Z, Chen X, Pang C, Xu X et al (2018) Fast response CdS-CdSxTe1-x-CdTe
core-shell nanobelt photodetector. Sci Bull 63(17):1118–1124
48. Todorov TK, Reuter KB, Mitzi DB (2010) High-efficiency solar cell with earth-abundant
liquid-­processed absorber. Adv Mater 22:E156–E159
49. Vigué F, Tournié E, Faurie J-P, Monroy E, Calle F, Muñoz E (2001) Visible-blind ultraviolet
photodetectors based on ZnMgBeSe Schottky barrier diodes. Appl Phys Lett 78:4190
50. Vikulin IM, Kurmashev SD, Stafeev VI (2008) Injection photodetectors. Phys Technol
Semicond 42(1):113–127. (in Russian)
51. Wang SL, Lee SH, Gupta A, Compaan AD (2004) (2004) RF sputtered HgCdTe films for tan-
dem cell applications. Phys Status Solidi C 1(4):1046–1049
52. Wang W, Winkler MT, Gunawan O, Gokmen T, Todorov TK, Zhu Y, Mitzi DB (2014) Device
characteristics of CZTSSe thin-film solar cells with 12.6% efficiency. Adv Energy Mater
4(7):1301465
53. Yang G, Kim D, Kim J (2016) Photosensitive cadmium telluride thin-film field-effect transis-
tors. Opt Express 24(4):3607–3612
54. Zhang ZZ, Shan DZ, Shan CX, Zhang JY, Lu YM, Liu YC, Fan XW (2003) The growth of
the CdxZn1−xTe epilayers by low-pressure metalorganic vapor-phase epitaxy. Thin Solid Films
429:211–215
55. Zhao X-H, DiNezza J, Liu S, Campbell CM, Zhao Y, Zhao Y, Zhang Y-H (2014) Determination
of CdTe bulk carrier lifetime and interface recombination velocity of CdTe/MgCdTe double
heterostructures grown by molecular beam epitaxy. Appl Phys Lett 105:252101
56. Zhao X-H, Liu S, Zhao Y, Campbell CM, Lassise MB, Kuo Y-S, Zhang Y-H (2016) Electrical
and optical properties of n-Type indium-doped CdTe/Mg0.46Cd0.54Te double Heterostructures.
IEEE J Photovoltaics 6(2):552–556
57. Zhou J, Wu X, Teeter G, To B, Yan Y, Dhere RG, Gessert TA (2004) CBD-Cd1−xZnxS thin films
and their application in CdTe solar cells. Phys Status Solidi B 241:775–778
58. Zhou C, Chung H, Wang X, Bermel P (2016) Design of CdZnTe & crystalline silicon tandem
junction solar cells. IEEE J Photovoltaics 6(1):301–308
Part III
New Trends in Development of II–VI
Semiconductors–Based Photodetectors
Chapter 16
Nanowire-Based Photodetectors
for Visible-UV Spectral Region

Ghenadii Korotcenkov and Victor V. Sysoev

16.1 Introduction

In recent decades, there has been a tremendous interest in the development of vari-
ous devices based on (quasi-)1D nanostructures, such as nanowires (NWs), nano-
tubes (NTs), nanoribbons (NRs) and nanobelts (NBs) [17, 36, 54]. Semiconducting
1D nanostructures are quasi-one-dimensional nanocrystals, which normally exhibit
a transverse width at the 10–100 nm range and length to be substantially longer, up
to micrometers or even millimeters. NWs and nanobelts, which are most employed
in the development of photodetectors, are so far the smallest structures, which could
transport charge carriers efficiently. It has been shown that these 1D structures are
promising nanostructures for the next generation of nanoscaled electrical and opti-
cal devices such as field-effect transistors, solar cells, diodes, LEDs, and lasers. The
development of photodetectors for visual and UV range is no exception [8, 45, 65,
66]. This interest is explained by special shapes and specific fundamental chemical,
physical and electro-optical properties of 1D structures, mostly grown as single
crystals. In particular, it was found that the unique photonic features of 1D struc-
tures make them an excellent candidate for photodetectors. Due to the nanoscale
size and matching the thickness or diameter of NWs and NBs to the Debye length,
the electronic and electrophysical properties of 1D nanostructures strongly depend
on the state of the surface and the processes occurring on it, that ensures appearing
conditions to observe a superior sensitivity when compared to one of conventional
thin-film photodetectors [64, 65]. The mechanism of this influence can be under-
stood from the diagram presented in Fig. 16.1.

G. Korotcenkov (*)
Department of Physics and Engineering, Moldova State University, Chisinau, Moldova
e-mail: ghkoro@yahoo.com
V. V. Sysoev
Yuri Gagarin State Technical University of Saratov, Saratov, Russia

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 371
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_16
372 G. Korotcenkov and V. V. Sysoev

Fig. 16.1 (a) Schematic of a NW photodetector. (b, c) the energy band diagrams of a NW in dark
and upon illumination, indicating band-bending and surface trap states. The bottom drawing shows
oxygen molecules adsorbed at the NW surface. VB and CB are the valence and conduction band,
respectively. (Reproduced with permission from Ref. [66]. Copyright 2007: American Chemical
Society)

As illustrated in Fig. 16.1, oxygen molecules adsorbed on the NW surface under


dark conditions, capture free electrons via O2(g) + e− → O2−(ad), and a low-­
conductivity depletion layer appears/extends near the surface (Fig. 16.1b). Once
illuminated with the light at a photon energy above the energy gap (hv > Eg),
electron-­hole pairs are photogenerated [hv → e− + h+] (Fig. 16.1a). The holes
migrate to the surface, where they can be easily trapped at the surface state, leaving
behind unpaired electrons, which results in the increase of the conductivity under an
applied dc electric field. At the same time, the illumination stimulates the oxygen to
desorb from the NW surface [h+ + O2−(ad) → O2(g)] that causes a reducing the
height of the potential barrier at the surface and the return of electrons captured by
chemisorbed oxygen to the conduction band of the semiconductor, thereby enhanc-
ing the photoresponse (Fig. 16.1c). It is thought that the maximum effect is achieved
at a certain ratio of the NW diameter (d) and Debye length (DL). For example, if
d >> DL, then no signal enhancement is expected. Therefore, it is very important to
choose NWs of a suitable diameter while fabricating the photodetectors. A more
detailed description of the mechanism of photosensitivity in NWs can be found
somewhere else [61, 65, 66].
The low electrical cross-section of NWs implies a low electrical capacitance, but
this comes without degradation of total light absorption due to antenna effects. With
a certain configuration of highly ordered arrays of 1D structures, it is also possible
to create conditions under which high absorption is achieved with a minimum vol-
ume of photosensitive material, required to achieve a high signal-to-noise ratio.
Therefore, NW arrays can exhibit higher absorption than a thin film of the equiva-
lent thickness [38]. In addition, the small footprint of NWs leads to an efficient
relaxation of strain in lattice-mismatched heterostructures and facilitates
16 Nanowire-Based Photodetectors for Visible-UV Spectral Region 373

heterogeneous integration of II-VI semiconducting NW devices with mass-scaled


Si technologies. As a result, we get new degrees of design freedom to optimize
electrical and photoelectrical performance of photodetectors. The advantages of
nanostructures include also the possibility of growing one-dimensional nanostruc-
tures or converting them into flexible materials [46], which can be used to develop
of wearable devices. In the following, we will mostly use the designation NWs, as
the most common while discussing the properties of 1D nanostructures, for
convenience.

16.2 Synthesis of NWs

Currently, high-quality II-VI semiconducting 1D nanostructures to be employed in


photodetectors aimed mostly to visible and UV spectral range, can be rationally
synthesized in a single-crystal phase whose key parameters such as chemical com-
position, shape, doping state, diameter, length, etc., are well controlled. For these
purposes, a number of techniques have been developed. These methods can gener-
ally be classified into one of the following approaches: bottom-up or top-down. A
more detailed discussion of the methods used to grow II-VI semiconducting NWs
can be found in Chap. 12 (Vol. 1). Currently, the most common method to grow
II-VI semiconducting 1D nanostructures is a thermal evaporation of II-VI com-
pounds powders via a vapor–liquid–solid (VLS) process.

16.3 Fabrication Features of NW-Based Photodetectors

It should be noted that photodetectors based on individual NWs have a specific


configuration and manufacturing technology, that is fundamentally different from
the traditional one used to manufacture conventional photodetectors. The NW pho-
todetectors can be fabricated under various designs (see Fig. 16.2). The most popu-
lar is the in-plane architecture where NWs are placed to lie on a substrate. This
architecture is employed to fabricate photodetectors based on both single wires and
their arrays. In this design, the NWs are either randomly dispersed or prepositioned
using some alignment techniques, as dielectrophoresis, contact printing, and others.
Contacts to the NWs are usually made at the last step with the help of metal
evaporation or sputtering by various methods and lithographic techniques. One can
find the description of this approach in [36]. A typical example of photodetectors
based on individual 1D structures and manufactured using this approach is shown in
Fig. 16.2a. Another, a more progressive, approach to manufacture NW array-­
matured photodetectors employs a vertical architecture where NWs are stand gener-
ally on their native growth substrate (see Fig. 16.2c). The vertical configuration was
mainly designed to demonstrate NW array devices using a large number of parallel-­
connected NWs as the active medium [8].
374 G. Korotcenkov and V. V. Sysoev

Fig. 16.2 Schematic of some of the approaches used to fabricate semiconducting NW-based pho-
todetectors. (a) NWs can be directly grown between two electrodes using both top down or bottom
up approaches. (b) NW bridges can form on isolated electrode contacts as a result of a serendipi-
tous fusing phenomenon during their growth. (c) Vertically oriented NWs atop mother wafer can
be fashioned into a PD by a lateral 3D-1D contact printing onto receiver substrates followed by
deposition of electrical contacts or by a 3D-3D print transfer mechanism onto thermally pliable
matrices followed by top contact deposition. Or whilst still atop the mother wafer, isolation and top
contact layers can be deposited onto the array to create a device. (Reprinted with permission from
Ref. [59]. Copyright 2016: Elsevier)
16 Nanowire-Based Photodetectors for Visible-UV Spectral Region 375

16.3.1 Direct NW Integration

If the deposition of the catalyst on the side walls of the platform used is acceptable,
the semiconducting NWs can be most effectively integrated into the required sites
of the device via direct growing to form a mechanically and electrically stable
monolithic assembly. This process is shown schematically in Fig. 16.3. The positive
output of this method depends on the proper making catalytic sites in the desired
location. For this purpose, one can use technologies such as nanoimprint, electron
beam lithography or evaporation followed by annealing. Many semiconducting
NWs of Group IV [30], Group III-V [87], and Group II-VI [39] were integrated in
frames of this technology. Unfortunately, due to the nucleation and migration of
nanoparticles of the catalytic material upon annealing, it is difficult to control the
catalyst diameter and its position. This leads to uncertainty in the position and diam-
eter of the NWs [35].
It should be noted that the described integration method opens up new techno-
logical possibilities for fabricating NW-based photodetectors with advanced param-
eters. For large scale integrated circuits or embedded sensor applications, the
“bridging” process can meet industry requirements for relatively low cost and high
throughput. The in situ fabricated NW devices offer a massively parallel, self-­
assembling technique that provides a controlled connection of NWs between elec-
trodes by means of just a coarse lithography [59].

16.4 Single NW-Based Photodetectors

16.4.1 Photoconductive Detectors Employing Single NWs

Single NWs are the simplest platform to develop photodetectors. Currently, such
NW-based photoconductors are produced taking all wide-gap II-VI semiconducting
compounds such as CdS, CdSe, ZnSe, CdTe, ZnTe and ZnS. They detect electro-
magnetic waves in a broad spectrum range, from infrared to ultraviolet region, due
to their broad coverage of bandgap energies. CdS, ZnSe and ZnS NWs are the most

Fig. 16.3 (a) A schematic illustrating a lithography aided deposit of a gold catalyst onto the side
of side wall of a trench. (b) Catalyst grown NW produces oriented NW directional growth. (c)
Continued NW growth causes the NW to grow onto the opposing sidewall. (Reprinted with per-
mission from Ref. [59]. Copyright 2016: Elsevier)
376 G. Korotcenkov and V. V. Sysoev

studied photoconductors in the group II–VI aimed to a visible and ultraviolet light
detection [10]. As in the case of metal oxide NWs, the photoresponse of these low-­
dimensional structures is strongly affected by the photochemistry of surface oxy-
gen, which can significantly change the relaxation dynamics of photocarriers [32,
33]. The characteristics of photodetectors developed on the basis of II-VI semicon-
ducting NWs are shown in Table 16.1. The photoconductive detectors are highly
sensitive ordinarily in the visible-UV spectral range (Fig. 16.4b, d) and have stable,
fully reversible and periodic characteristics (Fig. 16.4c). Ideally, photoconductive
detectors should have ohmic contacts (Fig. 16.4a). However, this condition is not
met in many cases.
As shown in numerous works, the performance of NW-based photodetectors can
be significantly improved through their doping. For instance, to optimize the param-
eters of CdS-based photodetectors, this material is doped with In [53], Ga [79], and
Cl [83]. The observed effect to improve the parameters of photodetectors is
explained by enhancing the concentration of charge carriers after doping. Typically,
as-synthesized NWs exhibit a high resistance. Following doping, the resistance of
NWs might drop by 3–4 orders of magnitude. For example, the photoconductive
properties of CdSe nanobelts have been investigated in both intrinsic and n-doped
structures [34]. Fast speed, response/recover times to be ~15/31 μs, and high photo-
sensitivity, ca. 100, were observed in i-CdSe, while a high-gain is preserved in
n-CdSe. These differences seem to mature from the traps created by impurities
within the nanobelt [81]. In the visible-NIR range, of 400–800 nm, CdTe

Table 16.1 Performance of II-VI semiconducting NW-based photoconductive detectors


Min. Pλ,
size, mW/ λ, Res. Decay
Mater. Type nm cm2 nm Rλ, AW−1 Ilight/Idark EQE, % Idark, pA time time Ref.
CdS:Sn NW >300 16 405 14.6 ~30 4.5 104 9 × 102 0.26 s 0.26 s [105]
CdS NB – 405 4.1 1.3 103 ~1 0.27 s 0.28 s [105]
CdS:Er NR 30– 0.03 457 3.46 × 104 ~103 9.3 104 – – [26]
80
CdS NW 30 0.1 450 ~6 × 102 – – [3]
CdS NB 40 – – 2 × 103 20 1s 3s [16]
CdS NW >500 3.75 365 3.6 × 102 104 8.6 × 105 72 10 ms 10 ms [18]
CdS NR ~400 0.5 450 1.2 × 104 – 3.5 × 106 0.82 s 0.84 s [104]
CdS NB 0.21 484 0.036 1.5 106 30 ms 40 ms [57]
CdSe NB 60– 3.4 650 3 103 1.7 × 104 20 μs 27 μs [31]
80
ZnSe NW 4.2 × 105 1.3 × 106 0.55 s 0.6 s [56]
ZnSe NB 40 – 400 0.12 >102 37.2 <10−2 <0.3 s <0.3 s [14]
ZnSe:Mn NW 3–4 0.02– 365 0.13 <0.5 s <0.4 s [40]
9
ZnS0.44Se0.56 NW – 1.5 × 106 4.5 × 106 0.52 s 0.93 s [78]
ZnS NB 20 – 300 0.18 <0.3 s <0.3 [15]
ZnS NT, 40 350 50 1.6 105 ~8 0.12 0.14 [2]
array
NW nanowires, NB nanobelts, NR nanoribbons, NT nanotubes
16 Nanowire-Based Photodetectors for Visible-UV Spectral Region 377

Fig. 16.4 (a) I–V curves ofs the ZnS:Ga nanoribbon-based PDs illuminated with varied wave-
lengths at a constant light intensity of 200 mW cm−2. Inset: schematic diagram of the PCD for
measurement. Reprinted with permission from [90]. Copyright 2014: RSC. (b) Spectral photores-
ponse measured from an individual ZnS nanobelt-based UV-light sensor for a bias of 10 V. (Reprinted
with permission from Ref. [14]). Copyright 2009: Wiley. (c) Time response of CdS:Cl NW photo-
detector. (Reprinted with permission from Ref. [83]). Copyright 2010: IOP. (d) Spectral response
of the CdSe nanoribbon. The curves represent the response spectra plotted at a constant light
intensity (1) and a constant photon density (2), respectively. (Reprinted with permission from Ref.
[31]. Copyright 2007: Wiley)

nanoribbons with p-type conductivity present a significant photoresponse to irradia-


tion with high responsivity and gain of 7.8 × 102 AW−1 and 2.4 × 105%, respectively
(Fig. 16.5).
An interesting effect was observed in CdS NWs doped with Sn [105]. It was
found that a photodetector based on this material cannot only respond to light with
energies exceeding the band gap, but can also respond to light with energies below
the band gap, while maintaining excellent photoconductivity parameters. This phe-
nomenon is distinctly different from that observed in pure CdS nanobelt photode-
tector, which can only response to light with energy exceeding the band gap. Zhou
et al. [105] believed that both the trapped state induced by tin impurity and the
optical whispering gallery mode microcavity effect in the Sn-doped CdS NWs con-
tribute to improved conductivity and broad spectral response consistent with photo-
luminescence measurements.
378 G. Korotcenkov and V. V. Sysoev

Fig. 16.5 (a) and (b) Dependence of gain and responsivity on light intensity of a CdTe nanoribbon
(λ = 400 nm). (Reproduced with permission from Ref. [93]. Copyright 2012: Royal Society of
Chemistry)

Changing the structure of NWs is also a powerful tool to modify the parameters
of photodetectors. For example, taking hierarchical CdS NWs allowed Li et al. [41]
to advance the sensitivity of the detectors by almost two orders of magnitude when
compared with detectors based on individual CdS NWs without branches. When the
device was exposed upon a visible light of 470 nm with a light intensity of 0.934 mW/
cm2 at a bias voltage of 5 V, the sensitivity of the hierarchical CdS NW photodetec-
tor, Ilight/Idark, was ~2 104. The hierarchical CdS NW has much larger specific surface
area, which can adsorb more oxygen molecules and lead to a higher resistance and
smaller dark current compared to conventional CdS NWs.
Another technique to optimize the parameters of NW-based photodetectors is
based on a decoration of the crystal surface with plasmonic gold nanoparticles
(PGNs). In particular, Luo et al. [48] showed that both sensitivity and detecting abil-
ity of CdSe NWs has been improved significantly after decorating their surface
with PGNs.
It is important to note that nanostructures of II-VI ternary alloys are also utilized
in the development of photodetectors for the visible-UV spectral range [49]. The
most studied in this area are 1D structures of CdSxSe1-x. The photodetectors based
on CdSxSe1-x NWs were primarily reported by Choi et al. [7] who grown them by
pulsed laser deposition on conductive electrodes. These devices have exhibited a
high sensitivity to UV light with a reasonably fast response. An important advantage
of photodetectors based on the specified II-VI ternary alloy, as shown by Choi et al.
[7] and, further, by Lu et al. [50, 51], is the ability to control their spectral range of
sensitivity by changing the stoichiometry of the materials. The spectral responsivity
of CdSxSe1-x NWs with different x value is shown in Fig. 16.6b. Lu et al. [50] also
clarified that employing a modification of 1D structures via a focused laser could
significantly improve the performance of CdSxSe1-x nanoribbon photodetectors. In
another study conducted by Guo et al. [21] it was found that the CdSxSe1-x nanorib-
bons with lateral heterostructures could serve also as photodetectors with high pho-
toresponsivity and quantum efficiency.
16 Nanowire-Based Photodetectors for Visible-UV Spectral Region 379

Fig. 16.6 (a) Schematic illustrating a network structure of CdSxSe1-x NWs floating above the SiO2
/Si substrate. (b) Normalized spectral responsivity curves of CdSxSe1-x NWs with a different x
value. The applied bias to the device was 5 V. (Reproduced with permission from Ref. [7].
Copyright 2010: IOP Publishing Ltd)

Fig. 16.7 (a) Schematic showing the contact printing method and the final device construction for
CdSSe-based photodetectors. (b) The spectral response from different NW-array devices corre-
sponding to S-rich, middle, and Se-rich regions. (Reprinted with permission from Ref. [68].
Copyright 2012: IOP Publishing Ltd)

Takahashi et al. [68] used another approach to fabricate photodetectors with


II-VI ternary NWs. They designed wavelength-tunable detectors taking spatially
composition graded CdSxSe1-x NW arrays. A schematic illustration of their experi-
ments is shown in Fig. 16.7a. CdSxSe1-x NWs at varied compositions can be grown
on the SiO2/Si substrate taking an advantage to vary a distance to the vapour source.
The as-synthesized NWs covered a wide range of visible light from 525 nm, S-rich
area, to 650 nm, Se-rich area. The Ti/Au, of 5 nm/40 nm, thin film contact elec-
trodes were finally deposited by electron beam evaporation to form an array of
wavelength tunable photodetectors. As it is seen in Fig. 16.7b, NWs in different
places on the substrate had different spectral responses, which contributed to an
380 G. Korotcenkov and V. V. Sysoev

increase in the response range and demonstrated the successful integration of detect-
ing different wavelengths on a single chip with NWs with spatially varying
composition.
As for other II-VI ternary alloys, the photoresponse of ZnSxSe1-x and ZnxCd1-xSe
NWs was also reported [78, 88] in addition to photodetectors based on CdSxSe1-x.
ZnxCd1-xSe (0.31 < x < 0.72)-based variable-wavelength photodetectors, as well as
the ones described in work by [7], were fabricated by a selective growth of NWs on
patterned Au catalysts, thus forming NW air-bridges (see Fig. 16.6a) between two
Pt pillar electrodes [88].

16.4.2 Phototransistors

Typically, NW field-effect phototransistors (FEPTs) have been fabricated via dis-


persing NWs on a dielectric-semiconductor substrate [22, 25, 85, 103] or by pat-
terning NWs through conventional lithographic methods. A gate bias was applied
through a lithographically patterned top gate, or a back gate. An example of NW
FEPT made on the basis of II-VI compounds is given in Fig. 16.8a, b. A high per-
formance CdS NW- based FEPT was developed by Ye et al. [85]. This FEPT with
Au Schottky contact exhibited an ultrahigh photoresponse ratio with Ilight/Idark of
2.7 × 106 at Pλ,=5.3 mW/cm2, the high current responsivity of 2.0 × 102 AW−1, high
external quantum efficiency of 5.2 × 102, and fast rise and decay time of 137 and
379 μs (Fig. 16.8c, d). Such a good performance was achieved by applying a gate
shift near the threshold potential under illumination, which was more negative than
the threshold potential in the dark. This facilitated the depletion of the channel
charge carriers and led to a fast photoresponse [85]. Dai et al. [9] fabricated CdSe
NBs-based metal-semiconductor field-effect phototransistors. The gate electrode
was Au, which formed a good Schottky contact with the CdSe NB. Typical photo-
transistors exhibited a high responsivity, of ca. 1.4 × 103 A/W, high gain, of ca.
2.7 × 103, and fast response speed, of 35–60 μs, under a gate voltage of −1 V, which
was attributed to the unique advantages of the high-performance MESFET structure.
In order to develop FEPTs based on ZnS:Ga nanoribbons, Yu et al. [90] used an
ITO gate electrode, 80 nm thick, deposited by pulsed laser deposition. This study
reported that the ZnS doping with Ga, combined with improved ohmic contact and
the use of ITO as the contact, enabled the photodetectors to yield a high photocon-
ductive gain under extremely low intensity of the light. Upon an ultralow operating
voltage, of ~0.01 V, the device exhibited excellent photoresponse properties to
1·10−14 W UV light such as a high gain of 106, and fast response speed. Response
and decay times were equaled to 3.2 ms and 10 ms, respectively. These authors
believe that doped ZnS NRs have a great potential as the building blocks for low-­
intensity UV phototransistors. The detectivity of the ZnS:Ga FEPT at 0.01 V was
estimated to be 1.3 × 1019 cmHz 1/2 W−1 under a light intensity of 1.0 μW cm−2,
which is higher than the detectivity of traditional and most nanophotodetectors.
Phototransistors based on II-VI ternary alloys have also been developed. In par-
ticular, the results of testing FEPTs based on CdSxSe1-x are given in [43].
16 Nanowire-Based Photodetectors for Visible-UV Spectral Region 381

Fig. 16.8 (a) Schematic illustration of a single CdS nanobelt MESFET-based photodetector. (b)
Typical FESEM image of a single CdS NB MESFET-based photodetector. (c) Transient photocur-
rent response of (A) at VG = 3.8 V and VDS = 0.5 V. (d) On/off photocurrent response of the CdS
NB MESFET-based photodetector with VG = −3.8 V as a function of time on an exponential scale.
(Reprinted with permission from Ref. [85]. Copyright 2010: American Chemical Society)

As follows from the principle of FEPT operation, the transistor effect occurs due
to the presence of a lateral electric field across the NW, which modulates the con-
ductance in the channel. However, a similar effect is also present in NW photocon-
ductors where a radial electric field is appeared, too, due to trapping charge carriers
in deep trap states or surface defect states. The appearance of a charge on the surface
is accompanied by band bending. This causes a separation of photogenerated carri-
ers in the NW channel, which greatly extends the carrier recombination lifetime
with enabling a much higher sensitivity. Thus, NW photoconductors can be viewed
as phototransistors where the internal electric field in combination with light illumi-
nation acts as a photogate [8].

16.5 NWs-Based Heterostructures

Currently, heterojunctions are considered to be fundamental elements for modern


electronic and optoelectronic devices such as light emitting diodes, laser diodes,
photodetectors, and solar cells, which provide them with high performance
382 G. Korotcenkov and V. V. Sysoev

parameters. Experiment has shown that heterojunctions based on II-VI semicon-


ducting 1D nanostructures can also contribute to the development of photodetectors
suitable for the market [95]. Many developers believe that the use of heterojunctions
based on 1D nanostructures opens up new possibilities for improving the efficiency
of photoconversion in photodetectors. For example, efficient and fast separation of
photogenerated charge carriers in the space charge region of a heterojunction can
lead to a fast photoresponse. This response differs from the slow- relaxed one to be
associated with the behavior of adsorbates on the surface of nanostructuresas
observed in 1D-structure-based photoconductive detectors without a heterojunc-
tion [37].

16.5.1 Core-Shell Heterojunctions

Core-shell heterojunctions or radial heterostructures based on 1D semiconducting


nanostructures versus bulk materials exhibit unique advantages in terms of increased
surface-to-bulk ratio, shortened carrier collection path, and reduced reflection,
which are vitally important for high-performance photodetectors [96]. One of the
examples of such photodetectors is shown in Fig. 16.9. Radial heterostructures are
also interesting in that they provide a passivating envelope that reduces the sensitiv-
ity of the NW to the chemical composition of the environment. Additional advan-
tages of such structures are associated with the ability to control their properties
through changing the diameter and chemical composition of both the core and
the shell.
Currently, a number of strategies have been developed to construct core-shell
heterojunctions. These include chemical vapor deposition (CVD), atomic
layer deposition (ALD), pulsed laser deposition (PLD), solution-based cation
exchange reaction, sputtering, and molecular beam epitaxy (MBE). Core-shell

Fig. 16.9 (a) Schematics of a photovoltaic device based on an individual CdSe/ZnTe core-shell
NW. The electrode materials for the core and shell are indium and nickel respectively. (b) A repre-
sentative SEM image of a photodetector developed based on CdSe/ZnTe core-shell NW. The ZnTe
shell was partially etched by a saturated, aqueous FeCl3 solution. (Reprinted with permission from
Ref. [75]. Copyright 2014: RSC)
16 Nanowire-Based Photodetectors for Visible-UV Spectral Region 383

Table 16.2 II-VI semiconducting based core-shell structure designed for photodetector application
Core-shell Core Method of Shell Method of Type of
structure material synthesis material synthesis PD Ref.
Ge/CdS, NWs Ge PVD CdS ALD Ind. [67]
(NGrs)
Si/CdS, NWs p-Si CE CdS PLD Array [55]
(NGrs)
CdSe/ZnTe, NWs CdSe TE ZnTe PLD Ind. [75]
CVD Array [62]
CdS/ZnTe, NWs n-CdS:Ga TE p-ZnTe TE Ind. [74]
(NGr)
CdS/Cu2S, NWs CdS CSM Cu2S CSM Ind. [70]
CdS/CdTe, NBs CdS CVD CdTe CVD Ind. [69]
ZnSe/ZnO, NWs p-ZnSe Sputt. n-ZnO Sputt. Ind. [97]
(NPs)
ZnS/CdS, NBs ZnS TE CdS TE Ind. [47]
ZnS/InP, NWs ZnS CVD InP CVD Ind. [95]
ZnS/ZnO, biaxial ZnS TE ZnO TE Ind. [27]
NBs
ZnO/ZnSe, NWs ZnO TE ZnSe TE Array [60]
ZnO/ZnS, NRs ZnO CSM ZnS CSM Array [44]
ZnO/ZnS, NWs ZnO CVD ZnS (NPs) PLD Array [63]
ALD atomic layer deposition, CE chemical etching, CSM chemical solution methods, CVD chem-
ical vapour deposition, Ind. individual, NB nanobelts, NGr nanograins, NPs nanoparticles, NRs
nanorods, NW nanowire, PD photodetector, PLD pulsed laser deposition, PVD physical vapor
deposition, SC solar cell, Sputt. spattering, TE thermal evaporation

heterojunctions employed in photodetector design are listed in the Table 16.2. For
core-shell heterojunctions based on II-VI compound semiconductors, two-steps
methods are usually applied as construction approaches. In this case, the NWs are
first grown to build a core, and then shells are formed on the surface of these NWs
in various ways. A schematic representation of a typical manufacturing process for
a photodetector based on core-shell heterojunctions is shown in Fig. 16.10.

16.5.2 1D Axial Heterojunctions

The axial geometry of 1D-based heterojunctions is another option for II-VI semi-
conducting heterostructures to be promised for use in photodetectors. Numerous
works have suggested that 1D axial heterojunction should demonstrate lower leak-
age currents and, therefore, they should possess a superior rectifying behavior com-
pared to radial structures [52]. Currently, two approaches have been proposed for
the implementation of such structures based on II-VI semiconductors. The first
approach deals with modifying the composition of NWs during their growth by
changing the precursors served as a source (Fig. 16.11a). In particular, Mu et al. [56]
384 G. Korotcenkov and V. V. Sysoev

Fig. 16.10 Schematic of the fabrication process of the photodetector based on a Ge/CdS core–
shell heterojunction NW. (Reprinted with permission from Ref. [67]. Copyright 2016: RSC)

Fig. 16.11 (a) Schematic diagram of the growth process for ZnS0.49Se0.51/ZnSe axial heterojunc-
tion. (b) SEM image of as-prepared HNW nanodevice. Inset is schematic model of photodetector
based on single ZnS0.49Se0.51/ZnSe HNW. (c) Wavelength-dependent photocurrent response of
ZnS0.49Se0.51/ZnSe heterojunction and pure ZnSe NW photodetectors. Inset is corresponding semi-
logarithmic plot. (d) Experimental and fitted photocurrent versus power plot under 375 nm illumi-
nation and Vds = 3 V. (e) Time-resolved photocurrent of the ZnS0.49Se0.51/ZnSe heterojunction
photodetector in response to 375 nm laser on/off at Vds = 3 V. (Reprinted with permission from Ref.
[56]. Copyright 2019: Wiley)
16 Nanowire-Based Photodetectors for Visible-UV Spectral Region 385

employed this approach to grow 1D ZnSSe-ZnSe axial heterostructures and to fab-


ricate the related photodetectors. SEM image and characteristics of these photode-
tectors are presented in Fig. 16.11b, c, d. Zhang et al. [101] also applied this method
with a variation. They constructed a kind of p-n heterojunction arrays by direct
growing the p-type ZnSe NRs on highly-aligned n-type Si NWs arrays. Cross-­
section SEM images showed that these quasi-aligned nanoribbons were just grown
as an array on the top of Si NWs array.
Another approach to employ the local oxidation of NWs of II-VI compounds
allows one preparing single 1D-axial heterostructures. Thus, Zhang et al. [99], fab-
ricated a single ZnSe-ZnO NW axial p-n junction following such a technique. For
this, as-synthesized p-type ZnSe NWs were transferred onto the SiO2/Si substrate
by a sliding transfer process. Then, a Si3N4 protection layer was applied to one half
of the NW with further oxidizing the exposed part via placing the sample into a fast
heat treatment system. After the thermal oxidation, the Si3N4 protective layer was
removed by reactive ion etching. As a result, the authors could get a NW whose one
half consists of unoxidized ZnSe, and another one consists of post-oxidized
ZnO. This process is illustrated in Fig. 16.12.

16.5.3 Crossed NW Heterojunctions

As previously demonstrated, fabricating core-shell heterojunctions and 1D axial


heterojunctions is always a complicated multi-step preparation process [95].
Therefore, the search for simpler technological solutions is constantly being carried
out. In this regard, using the crossed NWs to make heterojunctions is such a solu-
tion. The crossed NW heterojunction is a kind of post-assembly structure. According
to Huang et al. [29], such a configuration of heterojunctions presents many advan-
tages, including the ability: (1) to flexibly choose component materials; (2) to inde-
pendently tune the dopant concentration of the component materials; (3) to define
abrupt, nanoscale junctions that are ideal for high spatial resolution; and (4) to
assemble arrays for integrated nano-optoelectronic devices. The manufacturing pro-
cess of crossed NW heterojunctions is shown schematically in Fig. 16.13.

16.5.4 1D Nanostructure/Thin Film or Si


Substrate Heterojunctions

Coupling of II-VI group semiconducting 1D nanostructures and Si substrates is


another approach to design heterojunctions. Moreover, this method to create hetero-
junctions based on 1D nanostructures is the simplest one. This approach has the
following advantages [95]: (1) this could have avoided the complex multi-step pro-
cess of preparing the device for fabrication; (2) it is compatible with conventional
386 G. Korotcenkov and V. V. Sysoev

Fig. 16.12 (a) Schematic illustration of the construction process of the ZnSe-ZnO axial p-n junc-
tion. (b) The SEM image of a NW whose bare part has already been oxidized. (c) The SEM image
of a prepared ZnSe-ZnO NW axial p-n junction. (Reproduced with permission from Ref. [99],
Copyright 2015: Elsevier B.V)

microelectronic technology; and (3) it is suitable for developing aligned plane array
integrated devices.
Ordinarily, the manufacturing process of such heterojunctions composes of the
following operations [98] (see Fig. 16.14): (i) the formation of SiO2 insulating pads
on a n-Si substrate; (ii) the transfer of NWs from the growth substrate onto the pat-
terned SiO2/Si substrate; (iii) after dispersing, some NWs would cross on the edges
of the SiO2 pads and partially contact with the underlying n-Si substrate, where the
heterojunction is formed; and (iv) finally, a metal film (usually, Au) is applied on the
SiO2 pads to provide the ohmic contacts to the NWs. Using this approach, such
16 Nanowire-Based Photodetectors for Visible-UV Spectral Region 387

Fig. 16.13 (a) Schematic illustration of the step-wise process (1,2,3,4) for the fabrication of a
ZnSe NW/ZnO NW p-n junction. (b) SEM image of ZnSe/ZnO crossed NW heterostructure.
(Reproduced with permission from Ref. [58]. Copyright 2013: Institute of Physics)

Fig. 16.14 Schematic illustration of the fabrication process of ZnSeNW/Si p-n junction.
(Reproduced with permission from Ref. [98]. Copyright 2016, Elsevier B.V)
388 G. Korotcenkov and V. V. Sysoev

heterojunctions as p-ZnSe NW / n-Si [98], ZnSe nanoribbon/Si [73], CdS:Ga


nanoribbon/Si [82], n-CdSe NWs/p+-Si [28], p-ZnS nanoribbon/n-Si [91], p-CdS
nanoribbon/n-Si [92], n-CdS NWs/p+-Si [4], ZnSe NWs/Si [76], ZnSe NWs/Si
[100], CdTe nanoribbons/ n-Si NWs array [94], and so on, were fabricated.

16.5.5 Photodetector Performance

It is important to note that heterojunctions prepared by different methods were


tested and showed acceptable photoelectrical characteristics. So, CdS-ZnTe core-­
shell heterojunction photodetectors tested under the light illumination with a wave-
length of 638 nm and light intensity of 2 mWcm−2 at 1 V, exhibited responsivity of
1.55 × 103 AW−1, conductive gain of 3.3 × 103, and detectivity of
8.7 × 1012 cm·Hz1/2 W−1 [74]. Ge-CdS core-shell heterojunction NW photodetec-
tors, developed by Sun et al. [67], showed photodetection sensitivity of 18,000%,
which is remarkably much higher than sensitivity of arrays of CdS NWs and CdS
quantum dots [1, 24]. CdSe/ZnTe core/shell NW array, fabricated by Wang et al.
[75] exhibited even greater photodetection sensitivity. Rai et al. [63] and Lin et al.
[44] developed an efficient and highly sensitive broad band UV/VIS photodetector
based on wide band gap ZnO/ZnS heterojunction 3D core/shell NW array. It is
worth noting that the photoresponse of heterojunction was always greater and faster
than the response of individual NWs which are employed to form such a hetero-
junction (see Fig. 16.15).
Photodetectors based on a crossed ZnSe/ZnO p-n heterojunction and ZnSe-ZnO
axial p-n junction, designed by Zhang et al. [95, 99], exhibited excellent diode

Fig. 16.15 (a) The photoresponse of the p-ZnSe:Sb/n-ZnO:Ga core shell NW junction diode
under illumination with 365 nm UV light. (b) The photoresponse of the single ZnO:Ga NW and
the ZnSe:Sb NW under illumination with 365 nm UV light. During the photodetection measure-
ment, the light intensity and the bias voltage were kept at 100 μW cm−2 and − 5 V, respectively.
The UV light was switched on and off manually. (Reprinted with permission from Ref. [58].
Copyright 2013: IOP)
16 Nanowire-Based Photodetectors for Visible-UV Spectral Region 389

behaviors and high sensitivity to ultraviolet light illumination with a good reproduc-
ibility and quick photoresponse. Both photodetectors showed much higher sensitiv-
ity than the detectors based on ZnO, ZnTe, and CdSe NWs [5, 6, 80]. CdS/CdSxSe1 − x
axial heterostructure NWs, fabricated via a temperature-controlled chemical vapor
deposition method [20] also demonstrated an appropriate photodetector perfor-
mance. The photodetector based on the single CdS/CdSxSe1 − x NWs showed higher
responsivity (1.2 × 102 A / W), faster response speed (rise ∼68 μs, decay ∼137 μs),
higher Ilight/Idark ratio (105), higher EQE (3.1 × 104%), and broader detection range
(350–650 nm) at room temperature versus the photodetector based on single nano-
structures. The performance of several photodetectors based on II-VI semiconduct-
ing NW heterojuntions is shown in Fig. 16.16.
Based on crossed CdS/p-Si NW heterojunction, Hayden et al. [23] designed ava-
lanche photodiodes (APDs), which exhibited a photocurrent response (IPC/Idark) to be
approx. 104 times higher than that in individual n-CdS or p-Si NW photoconductors.
This effect was observed due to avalanche multiplication at the p-n crossed NW

Fig. 16.16 (a) I-V curves of the ZnS/CdS photodetector under various wavelength light with dif-
ferent intensities and in the dark; (b) Spectroscopic photoresponses of the ZnS/CdS detector mea-
sured at different wavelengths ranging from 300 to 900 nm at a bias of 2 V. (Reprinted with
permission from Ref. [47]. Copyright 2016: RSC). (c) Spectral photoresponse of CdS/CdS:SnS2
superlattice NW. (Reprinted with permission from Ref. [18]. Copyright 2016: RSC). (d) Spectral
photoresponse of ZnS/ZnO biaxial nanobelt with Cr/Au electrodes measured at a bias of
5.0 V. (Reprinted with permission from Ref. [27]. Copyright 2012: Wiley)
390 G. Korotcenkov and V. V. Sysoev

Table 16.3 Comparison of the spectral responsivity, external quantum efficiency, and response
time of II-VI semiconducting NW heterostructure-based photodetectors
Materials Responsivity, A·W−1 Iligh/Idark EQE, % Rise time Decay time Ref.
CdS/CdSSe 1.18 × 102 3.1 × 102 68 μs 137 μs [20]
ZnS/ZnO – – 0.77 s 0.73 s [71]
ZnO/ZnSe 8.6 – – – [84]
CdS/CdSSe 1.2 × 103 106 280 ms 550 ms [21]
ZnS0.49Se0.51/ZnSe 6.3 × 105 2.1 × 106 23 ms 90 ms [56]
CdS/ZnTe 1.6 × 103 3.3 × 103 – – [74]
CdS/CdS:SnS2 2.5 × 103 ~105 8.6 × 105 10 ms 10 ms [18]
ZnSe:Sb/ZnO 5.2 × 105 1.8 × 106 – – – [58]
ZnS/InP – 5 103 1.1 × 103 0.75 s 0.5 s [96]
EQE External quantum efficiency

junction. A detection limit of about 75 photons was estimated for these devices.
Polarization dependence of the photoresponse has also been observed in the
“crossed” structure due to the predominant optical absorption in the CdS NW.
Xie et al. [94] have developed a photodetector, based on a p-CdTeNR/n-SiNWs
array heterojunction, which can operate in the light, of visible to near-infrared
range, with a good stability, high sensitivity, and fast response speed. The photode-
tector had small rise time, ca. 1.2 ms, and fall time, ca. 1.58 ms, which are much
faster than the parameters reported for CdTe NWs and CdTe NRs-based photodetec-
tors [1, 86]. Even better performance was demonstrated by photodiodes based on
CdS NR/Si heterojunctions which were produced by Xie et al. [92]. Detectors had
the rise time of ca. 300 μs and the fall time of ca. 740 μs under white light illumina-
tion with an intensity of ca. 5.3 mW cm−2. Xie et al. [92] believed that such a per-
formance could attribute to the high-quality of p-n junction appeared between the
CdS NR and the Si substrate. Due to a low concentration of interface defects, the
photogenerated carriers could be quickly separated by the space charge region and
then transferred to the electrodes. Photodetectors made by Yu et al. [91], who
employed heterojunctions of p-ZnS nanoribbon/n-Si substrate, also had a high sen-
sitivity. The measurements showed that the photodetectors had stable optoelectrical
properties with high sensitivity to UV-VIS-NIR light and an enhancement of respon-
sivities of ca. 1.1 × 103 AW−1 for λ = 254 nm under a reverse bias of 0.5 V. Parameters
of other heterostucture-based photodetectors are listed in the Table 16.3

16.6 Schottky Barrier-Based Photodetectors

Schottky photodiodes based on metal/semiconductor junction exploit such an inter-


face to separate and collect the photogenerated carriers. Electron-holes pairs (EHP)
generated in the vicinity of the depletion region are swept out by the built-in electric
field of the Schottky junction. It is believed [8, 81, 106] that one of the major advan-
tages of Schottky photodiodes relative to photoconductive structures with Ohmic
16 Nanowire-Based Photodetectors for Visible-UV Spectral Region 391

Fig. 16.17 (a) I-V curves of the Schottky barrier-based detectors (ITO-p-ZnS) measured in dark
and under UV light intensity of ~1 μWcm−2. (b) Time response of the ITO-p-ZnS detectors mea-
sured under UV light intensity of 1 μWcm−2 and 2.5 nWcm−2, respectively. The external bias volt-
age is fixed at 0.01 V. (Reprinted with permission from Ref. [89]. Copyright 2015: Springer)

contacts in addition to the simplicity of the manufacturing technology are the fast
response rate due to the high electric field (hence, the short transit time of charge
carriers) through the junction under reverse bias, lower dark currents, and higher
photoresponse. For example, Wu et al. [79] fabricated CdS:Ga NRs/Au Schottky
barrier diodes, which demonstrated a high light sensitivity and fast response speed,
response/decay times of ca. 95/290 μs, at a zero bias. A CdS single-NB Schottky
contact photodetectors fabricated by Li et al. [42] were also characterized by a high
sensitivity and short decay time. The photosensitivity, dark current and the decay
time of the sensor were ca. 4 × 104, 0.2 pA, and 31 ms, respectively. Yu et al. [89]
suggested using the ITO-p-ZnS contact as a Schottky barrier-based photodetectors.
The device exhibited a pseudo-photovoltaic behavior which can allow one detecting
UV light irradiation with incident power of ca. 6 × 10−17 W, or ~ 85 photons/s on the
NR, at room temperature with excellent reproducibility and stability. The corre-
sponding detectivity and photoconductive gain were calculated to be ca.
3.1 × 1020 cm·Hz 1/2·W−1 and 6.6 × 105, respectively (see Fig. 16.17).
More results from the Schottky photodiodes study based on CdS, and CdSe NWs
can be found elsewhere [11, 12, 19, 77]. During these studies, it was shown that the
photocurrent – voltage characteristics of such structures were typically asymmetric
and photocurrent could be strongly localized near the metal electrode-NW contact
in dependence on the biasing conditions.

16.7 Summary

In this chapter, we have described the various approaches utilized to fabricate quasi-
­1D NW-based photodetectors. Based on this consideration and analysis of the
results presented in numerous articles, it can be concluded that photodetectors based
on II-VI semiconductor-based 1D structures really have advantages over bulk and
392 G. Korotcenkov and V. V. Sysoev

thin-film photodetectors, which can lead to cost-effective, superior and novel func-
tionalities. The major advantages, according to [8], are the following:
• Large photosensitivity due to internal photoconductive or phototransistive
gain [102];
• Ability to integrate NWs and substrates of different materials, including silicon;
• Enhanced light absorption with a small amount of active material in NW arrays.
A dense network of NWs enables multiple reflections and scatterings, which
enhance incident photon trapping, a feature, which does not exist in conventional
PDs [102];
• Ability to build devices, as photoresistor, photodiode, phototransistor, on the
basis of single NWs with different functionality;
• Opportunity for dense-packed device integration.
At the same time, along with advantages, the technologies dealing with NWs also
has a number of significant disadvantages [34, 36, 59]. Primarily, in order to imple-
ment the most progressive ideas based on NW integration with silicon technology,
a technology of direct growth and integration of II-VI semiconducting NW-based
devices onto large Si wafers is required. Unfortunately, there are some challenges
such as lattice constant and thermal mismatch, lack of good control over atomic and
crystallographic structures, troubles with NWs assembly into functional devices, a
complicated forming of ternary and quaternary NW alloys, high contact resistance
and increased noise due to this fact [45], the lack of a well-controlled doping tech-
nology to develop sharp homo- and heterojunctions, catalyst-induced pollution, etc.
makes it difficult to grow and design NW devices on Si or any other substrates
which is a failure in the industrial application of NW PDs. Therefore, the technol-
ogy for integrating II-VI NWs with Si-integrated circuits is still at an infancy stage
despite on the significant research progress ongoing in this direction for recent
years. There still exists a a large gap to overcome towards the practical integration
and application of photodetectors based on 1D nanomaterials [34, 36]. Two issues
should be resolved. First, we do not have an inexpensive, simple and reproducible
method to synthesize one-dimensional nanomaterials with precisely managed geo-
metric and physical properties, such as diameter, length, orientation, crystal quality,
and doping concentration. Second, the controlled and scalable post-growth transfer
and assembly of one-dimensional nanomaterials over target substrates is rather dif-
ficult in most cases to manufacture such photodetectors. Some methods have been
proposed so far, including the use of fluid-assisted alignment, the Langmuir-­
Blodgett technique, the dielectrophoretic method, contact printing, etc. [36].
However, their ability to precisely control the position and orientation of each one-­
dimensional nanomaterial is still insufficient. Therefore, one cannot expect that
there will be monolithic devices integrating II-VI semiconducting NWs and Si inte-
grated circuits in the coming years. Most likely, hybrid approaches will be used at
the initial stage, which would not require drastic changes in the current manufactur-
ing process of Si-integrated circuits.
Another limitation to use the NWs-based photodetectors is their rather low quan-
tum efficiency. While the most conventional semiconductor detectors operate at
16 Nanowire-Based Photodetectors for Visible-UV Spectral Region 393

almost 100% quantum efficiency, then NW-based ones’ efficiencies are much lower
[59]. Although II-VI semiconducting 1D nanomaterials have a high absorption
coefficient near the band edge, they absorb a little light due to thin diameters. With
purpose to enhance the light absorption in 1D nanostructures, Wang et al. [72] and
Fang and Hu [13] analyzed six routes to realize localized fields enhancement in
two-dimensional material-based photodetectors. They considered ferroelectric
field, photogating electric field, floating gate-induced electric field, interlayer built-
­in field, localized optical field, and photo-induced temperature gradient field. It is
proved that localized fields substantially enhance the optical absorption, suppress
background noise, and advance the efficiency of separation of electrons and holes.
Wang et al. [72] and Fang and Hu [13] believe that combining 1D II-VI nanomaterial-­
based photodetectors with the idea of localized fields enhancement might be an
efficient method to further improve the performance of photodetectors in the future.
There are also challenges associated with the scatter of parameters and the nega-
tive impact of surface traps on the photoelectrical properties of semiconducting
NWs. Semiconducting NWs, especially those obtained by the bottom-up method,
have stochastic deviations from each other due to their defects, which affect the
charge capture and carrier mobility. Although tighter control of NWs growth and the
use of high purity chemical precursors eliminate some of these defects, it is virtually
impossible to prepare identical NWs with bottom-up approaches due to other physi-
cal and thermodynamic factors that arise during NWs growth. These problems were
discussed in sufficient detail in [36, 45].
As we indicated earlier, involving surface effects in the photoresponse allows
one to achieve very high sensitivityin the corresponding detectors. However, the
strong dependence of the NW resistance on the state of the surface also has negative
consequences, since the photoresponse becomes dependent on the state of the sur-
rounding atmosphere, namely, on air humidity, temperature, and the presence of
toxic and reducing gases. This means that the parameters of the photodetectors will
alter when these factors vary, and this change can be significant. These effects have
been detailed in Chap. 2 (Vol. 1). The encapsulation of NWs eliminates the influ-
ence of the surrounding atmosphere on the response of photoconductive detectors.
In this case, the internal photoconductive or phototransistor gain will also disappear.
In this regard, photodetectors based on Schottky barriers and core-shells structures
as well as phototransistors may be more preferable when employed in an uncon-
trolled atmosphere.
There are also problems related to the mechanical stability of photodetectors
based on 1D structures. A vibration might cause displacing NWs relative to each
other or to the substrate, which should undoubtedly be accompanied by a change in
the properties of homo- and heterojunctions having such contacts. The operational
reliability of NW-based photodetectors against to Joule self-heating is also an issue
for single nanostructures that needs to be addressed in high-power NW photodetec-
tors [36, 45].

Acknowledgments G. Korotcenkov is grateful to the State Program of the Republic of Moldova,


project 20.80009.5007.02, for supporting his research. V. Sysoev thanks the Russian Science
Foundation, project No. 22-29-00793 (https://rscf.ru/project/22-­29-­00793/), for a partial support.
394 G. Korotcenkov and V. V. Sysoev

References

1. Amos FF, Morin SA, Streifer JA, Hamers RJ, Jin S (2007) Photodetector arrays directly
assembled onto polymer substrates from aqueous solution. J Am Chem Soc 129:14296–14302
2. An Q, Meng X, Xiong K, Qiu Y, Lin W (2017) One-step fabrication of single-crystalline ZnS
nanotubes with a novel hollow structure and large surface area for photodetector devices.
Nanotechnology 28:105502
3. An B-G, Chang YW, Kim H-R, Lee G, Kang M-J, Park J-K, Pyun J-C (2015) Highly sen-
sitive photosensor based on insitu synthesized CdS nanowires. Sensors Actuators B Chem
221:884–890
4. Cai J, Jie J, Jiang P, Wu D, Xie C, Wu C et al (2011) Tuning the electrical transport proper-
ties of n-type CdS nanowires via Ga doping and their nano-optoelectronic applications. Phys
Chem Chem Phys 13:14663–14667
5. Cao YL, Liu ZT, Chen LM, Tang YB, Luo LB, Jie JS et al (2011) Single-crystalline ZnTe
nanowires for application as high-performance green/ultraviolet photodetector. Opt Express
19:6100–6108
6. Chang SP, Lu CY, Chang SJ, Chiou YZ, Hsueh TJ, Hsu CL (2011) Electrical and optical
characteristics of UV photodetector with interlaced ZnO nanowires. IEEE J Sel Top Quantum
17:990–995
7. Choi Y-J, Park K-S, Park J-G (2010) Network-bridge structure of CdSxSe1−x nanowire-based
optical sensors. Nanotechnology 21:505605
8. Dai X, Tchernycheva M, Soci C (2015) Compound semiconductor nanowire photodetectors.
In: Semiconductors and semimetals, vol 94. Elsevier, Cambridge, MA, pp 75–107
9. Dai Y, Yu B, Ye Y, Wu P, Meng H, Dai L, Qin G (2012) High-performance CdSe nanobelt
based MESFETs and their application in photodetection. J Mater Chem 22:18442–18446
10. Deng K, Li L (2014) CdS nanoscale photodetectors. Adv Mater 26(17):2619–2635
11. Doh Y-J, Maher KN, Ouyang L, Yu CL, Park H, Park J (2008) Electrically driven light emis-
sion from individual CdSe nanowires. Nano Lett 8:4552–4556
12. Dufaux T, Burghard M, Kern K (2012) Efficient charge extraction out of nanoscale Schottky
contacts to CdS nanowires. Nano Lett 12:2705–2709
13. Fang H, Hu W (2017) Photogating in low dimensional photodetectors. Adv Sci 4:1700323
14. Fang X, Xiong S, Zhai T, Bando Y, Liao M, Ujjal K, Gautam UK et al (2009) High-performance
blue/ultraviolet-light-sensitive ZnSe-nanobelt photodetectors. Adv Mater 21:5016–5021
15. Fang X, Bando Y, Liao M, Gautam UK, Zhi C, Dierre B et al (2009b) Single-crystalline ZnS
nanobelts as ultraviolet-light sensors. Adv Mater 21:2034–2039
16. Gao T, Li QH, Wang TH (2005) CdS nanobelts as photoconductors. Appl Phys Lett
86:173105–173107
17. Garnett E, Mai L, Yang P (2019) Introduction: 1D nanomaterials/nanowires. Chem Rev
119(15):8955–8957
18. Gou G, Dai G, Qian C, Liu Y, Fu Y, Tian Z et al (2016) High-performance ultraviolet photo-
detectors based on CdS/CdS:SnS2 superlattice nanowires. Nanoscale 8:14580–14586
19. Gu Y, Romankiewicz JP, David JK, Lensch JL, Lauhon LJ, Kwak ES, Odom TW (2006)
Local photocarrent mapping as a probe of contact effects and charge carrier transport in
semiconductor nanowire devices. J Vac Sci Technol B 24:2172–2177
20. Guo P, Xu J, Gong K, Shen X, Lu Y, Qiu Y et al (2016) On-nanowire axial heterojunction
design for high-performance photodetectors. ACS Nano 10:8474–8481
21. Guo P, Hu W, Zhang Q, Zhuang X, Zhu X, Zhou H et al (2014) Semiconductor alloy nanorib-
bon lateral heterostructures for high-performance photodetectors. Adv Mater 26:2844–2849
22. Han S, Jin W, Zhang DH, Tang T, Li C, Liu XL et al (2004) Photoconduction studies on GaN
nanowire transistors under UV and polarized UV illumination. Chem Phys Lett 389:176–180
23. Hayden O, Agarwal R, Lieber CM (2006) Nanoscale avalanche photodiodes for highly sensi-
tive and spatially resolved photon detection. Nat Mater 5:352–356
16 Nanowire-Based Photodetectors for Visible-UV Spectral Region 395

24. He J, Chen J, Yu Y, Zhang L, Zhang G, Jiang S et al (2014) Effect of ligand passivation on


morphology, optical and photoresponse properties of CdS colloidal quantum dots thin film. J
Mater Sci Mater Electron 25:1499–1504
25. Heo YW, Tien LC, Kwon Y, Norton DP, Pearton SJ, Kang BS, Ren F (2004) Depletion-mode
ZnO nanowire field-effect transistor. Appl Phys Lett 85:2274–2276
26. Hou D, Liu Y-K, Yu D-P (2015) Multicolor photodetector of a single Er3+-doped CdS nanorib-
bon. Nanoscale Res Lett 10:285
27. Hu L, Yan J, Liao M, Xiang H, Gong X, Zhang L, Fang X (2012) An optimized ultraviolet-a
light photodetector with wide-range photoresponse based on ZnS/ZnO biaxial nanobelt. Adv
Mater 24:2305–2309
28. Hu Z, Zhang X, Xie C, Wu C, Zhang X, Bian L et al (2011) Doping dependent crystal struc-
tures and optoelectronic properties of n-type CdSe:Ga nanowries. Nanoscale 3:4798–4803
29. Huang Y, Duan X, Wei Q, Lieber CM (2001) Directed assembly of one-dimensional nano-
structures into functional networks. Science 291:630–633
30. Islam MS, Sharma S, Kamins TI, Williams RS (2004) Ultrahigh-density silicon nano-bridges
formed between two vertical silicon surfaces. Nanotechnology 15:L5–L8
31. Jiang Y, Zhang WJ, Jie JS, Meng XM, Fan X, Lee ST (2007) Photoresponse properties of
CdSe single-nanoribbon photodetectors. Adv Funct Mater 17:1795–1800
32. Jie JS, Zhang WJ, Jiang Y, Meng XM, Li YQ, Lee ST (2006a) Photoconductive characteris-
tics of single-crystal CdS nanoribbons. Nano Lett 6:1887–1892
33. Jie JS, Zhang WJ, Jiang Y, Lee ST (2006b) Single-crystal CdSe nanoribbon field-effect tran-
sistors and photoelectric applications. Appl Phys Lett 89:133118
34. Jin W, Hu L (2019) Review on quasi one-dimensional CdSe nanomaterials: synthesis and
application in photodetectors. Nano 9:1359
35. Kayes BM, Filler MA, Putnam MC, Kelzenberg MD, Lewis NS, Atwater HA (2007) Growth
of vertical aligned Si wire arrays over large areas (> 1 cm2) with au and cu catalysis. Appl
Phys Lett 91:10310–103113
36. Korotcenkov G (2020) Current trends in nanomaterials for metal oxide-based conductometric
gas sensors: advantages and limitations. Part 1: 1D and 2D nanostructures. Nano 10:1392
37. Kum MC, Jung H, Chartuprayoon N, Chen W, Mulchandani A, Myung NV (2012) Tuning
electrical and optoelectronic properties of single cadmium telluride nanoribbon. J Phys Chem
C 116:9202–9208
38. Lähnemann J, Browne DA, Ajay A, Jeannin M, Vasanelli A, Thomassin J-L, BelletAmalric
E, Monroy E (2019) Near- and mid-infrared intersubband absorption in topdown GaN/AlN
nano- and micro-pillars. Nanotechnology 30:054002
39. Lee JS, Islam MS, Kim S (2006) Direct formation of catalyst-free ZnO nanobridge devices
on an etched Si substrate using a thermal evaporation method. Nano Lett 6:1487–1490
40. Li D, Xing G, Tang S, Li X, Fan L, Li Y (2017) Ultrathin ZnSe nanowires: one-pot synthesis
via a heat-triggered precursor slow releasing route, controllable Mn doping and application
in UV and near-visible light detection. Nanoscale 9:15044–15055
41. Li L, Lou Z, Shen G (2015) Hierarchical CdS nanowires based rigid and flexible photodetec-
tors with ultrahigh sensitivity. ACS Appl Mater Interfaces 7(42):23507–23514
42. Li L, Yang S, Han F, Wang L, Zhang X, Jiang Z, Pan A (2014) Optical sensor based on a
single CdS nanobelt. Sensors 14:7332–7341
43. Li L, Lu H, Yang ZY, Tong LM, Bando Y, Golberg D (2013) Bandgap-graded CdSxSe1–x nanow-
ires for high-performance field-effect transistors and solar cells. Adv Mater 25:1109–1113
44. Lin H, Wei L, Wu C, Chen Y, Yan S, Mei L, Jiao J (2016) High-performance self-powered
photodetectors based on ZnO/ZnS core-shell nanorod arrays. Nanoscale Res Lett 11:420
45. Logeeswaran VJ, Oh J, Nayak AP, Katzenmeyer AM, Gilchrist KH, Grego S et al (2011) A
perspective on nanowire photodetectors: current status, future challenges, and opportunities.
IEEE J Sel Topics Quant Electron 17(4):1002–1032
46. Lou Z, Shen G (2016) Flexible photodetectors based on 1D inorganic nanostructures. AdvSci
3:1500287
396 G. Korotcenkov and V. V. Sysoev

47. Lou Z, Li L, Shen G (2016) Ultraviolet/visible photodetectors with ultrafast, high photosen-
sitivity based on 1D ZnS/CdS heterostructures. Nanoscale 8:5219–5225
48. Luo L-B, Xie W-J, Zou Y-F, Yu Y-Q, Liang F-X, Huang Z-J, Zhou K-Y (2015) Surface plasmon
propelled high-performance CdSe nanoribbons photodetector. Opt Express 23(10):12979
49. Lu J, Liu H, Zhang X, Sow CH (2018) One-dimensional nanostructures of II-VI ternary
alloys: synthesis, optical properties, and applications. Nanoscale 10:17456–17476
50. Lu J, Lim X, Zheng M, Mhaisalkar SG, Sow C-H (2012) Direct laser pruning of CdSxSe1–x
nanobelts en route to a multicolored pattern with controlled functionalities. ACS Nano
6:8298–8307
51. Lu J, Sun C, Zheng M, Nripan M, Liu H, Chen GS et al (2011) Facile one step synthe-
sis of CdSxSe1–x nanobelts with uniform and controllable stoichiometry. J Phys Chem C
115:19538–19545
52. Lysov A, Vinaji S, Offer M, Gutsche C, Regolin I, Mertin W et al (2011) Spatially resolved
photoelectric performance of axial GaAs nanowire pn-diodes. Nano Res 4:987–995
53. Ma RM, Dai L, Huo HB, Yang WQ, Qin GG, Tan PH et al (2006) Synthesis of high quality
n-type CdS nanobelts and their applications in nanodevices. Appl Phys Lett 89:203120
54. Machín A, Fontánez K, Arango JC, Ortiz D, De León J, Pinilla S et al (2021) One-dimensional
(1D) nanostructured materials for energy applications. Materials 14:2609
55. Manna S, Das S, Mondal SP, Singha R, Ray SK (2012) High efficiency Si/CdS radial
nanowire heterojunction photodetectors using etched Si nanowire templates. J Phys Chem
C 116:7126–7133
56. Mu Z, Zheng Q, Liu R, Iqbal Malik MW, Tang D, Zhou W, Wan Q (2019) 1D ZnSSe-ZnSe
axial heterostructure and its application for photodetectors. Adv Electron Mater 2019:1800770
57. Nawaz MZ, Xu L, Zhou X, Shah KH, Wang J, Wu B, Wang C (2021) CdS nanobelt-based
self-powered flexible photodetectors with high photosensitivity. Mater Adv 2:6031–6038
58. Nie B, Luo LB, Chen JJ, Hu JG, Wu CY, Wang L et al (2013) Fabrication of p-type ZnSe:Sb
nanowires for high-performance ultraviolet light photodetector application. Nanotechnology
24:095603
59. Ombaba MM, Karaagac H, Polat KG, Islam MS (2016) Nanowire enabled photodetection.
In: Nabet B (ed) Photodetectors: Materials, Devices and Applications. Elsevier, pp 87–120
60. Park S, Kim S, Sun G-J, Byeon DB, Hyun SK, Lee WI, Lee C (2016) ZnO-core/ZnSe-shell
nanowire UV photodetector. J Alloys Comp 658:459–464
61. Prades JD, Jimenez-Diaz R, Hernandez-Ramirez F et al (2008) Toward a systematic under-
standing of photodetectors based on individual metal oxide nanowires. J Phys Chem C
112:14639–14644
62. Rai SC, Wang K, Chen J, Marmon JK, Bhatt M, Wozny S et al (2015a) Enhanced broad band
photodetection through piezo-phototronic effect in CdSe/ZnTe core/shell nanowire array.
Adv Electron Mater 1:1400050
63. Rai SC, Wang K, Ding Y, Marmon JK, Bhatt M, Zhang Y et al (2015b) Piezo-phototronic
effect enhanced UV/visible photodetector based on fully wide band gap type-II ZnO/ZnS
core/shell nanowire array. ACS Nano 9:6419–6427
64. Shen G, Chen D (2010) One-dimensional nanostructures for photodetectors. Recent Patents
Nanotechnol 4:20–31
65. Soci C, Zhang A, Bao X-Y, Kim H, Lo Y, Wang D (2010) Nanowire photodetectors. J Nanosci
Nanotechnol 10:1430–1449
66. Soci C, Zhang A, Xiang B, Dayeh SA, Aplin DPR, Park J et al (2007) ZnO nanowire UV
photodetectors with high internal gain. Nano Lett 7:1003–1009
67. Sun Z, Shao Z, Wu X, Jiang T, Zheng N, Jie J (2016) High-sensitivity and self-driven photo-
detectors based on Ge-CdS core–shell heterojunction nanowires via atomic layer deposition.
Cryst Eng Comm 18:3919–3924
68. Takahashi T, Nichols P, Takei K, Ford AC, Jamshidi A, Wu MC et al (2012) Contact printing
of compositionally graded CdSxSe1-x nanowire parallel arrays for tunable photodetectors.
Nanotechnology 23:045201
16 Nanowire-Based Photodetectors for Visible-UV Spectral Region 397

69. Tang M, Xu P, Wen Z, Chen X, Pang C, Xu X et al (2018) Fast response CdS-CdSxTe1-x-CdTe


core-shell nanobelt photodetector. Sci Bull 63(17):1118–1124
70. Tang J, Huo Z, Brittman S, Cao H, Yang P (2011) Solution-processed core–shell nanowires
for efficient photovoltaic cells. Nat Nanotechnol 6:568–572
71. Tian W, Zhang C, Zhai T, Li SL, Wang X, Liu J et al (2014) Flexible ultraviolet photodetec-
tors with broad photoresponse based on branched ZnS-ZnO heterostructure nanofilms. Adv
Mater 26:3088–3093
72. Wang J, Fang H, Wang X, Chen X, Lu W, Hu W (2017) Recent progress on localized field
enhanced two-dimensional material photodetectors from ultraviolet—visible to infrared.
Small 13:1700894
73. Wang L, Chen R, Ren ZF, Ge CW, Liu ZX, He SJ et al (2016) Plasmonic silver nanosphere
enhanced ZnSe nanoribbon/Si heterojunction optoelectronic devices. Nanotechnology
27:215202
74. Wang L, Song HW, Liu ZX, Ma X, Chen R, Yu YQ et al (2015) Core–shell CdS:Ga-ZnTe:Sb
p-n nano-heterojunctions: fabrication and optoelectronic characteristics. J Mater Chem C
3:2933–2939
75. Wang K, Rai SC, Marmon J, Chen J, Yao K, Wozny S et al (2014) Nearly lattice matched all
wurtzite CdSe/ZnTe type II core–shell nanowires with epitaxial interfaces for photovoltaics.
Nanoscale 6:3679–3685
76. Wang L, Lu M, Wang X, Wu D, Xie C, Wu C et al (2013) Tuning the p-type conductivity of
ZnSe nanowires via silver doping for rectifying and photovoltaic device applications. J Mater
Chem A 1:1148–1154
77. Wei T-Y, Huang C-T, Hansen BJ, Lin Y-F, Chen L-J, Lu S-Y, Wang ZL (2010) Large enhance-
ment in photon detection sensitivity via Schottky-gated CdS nanowire nanosensors. Appl
Phys Lett 96:013508
78. Wu D, Chang Y, Lou Z, Xu T, Xu J, Shi Z, Tian Y, Li X (2017) Controllable synthesis of ter-
nary ZnSxSe1-x nanowires with tunable band-gaps for optoelectronic applications. J Alloys
Compd 708:623–627
79. Wu D, Jiang Y, Zhang Y, Yu Y, Zhu Z, Lan X et al (2012) Self-powered and fast-speed photo-
detectors based on CdS:Ga nanoribbon/Au Schottky diodes. J Mater Chem 22:23272
80. Wu D, Jiang Y, Zhang Y, Li J, Yu Y, Zhang Y et al (2012b) Device structure-dependent field-­
effect and photoresponse performances of p-type ZnTe: Sb nanoribbons. J Mater Chem
22:6206–6212
81. Wu P, Dai Y, Sun T, Meng H, Fang X, Yu B, Dai L (2011a) Impurity-dependent photoresponse
properties in single CdSe nanobelt photodetectors. ACS Appl Mater Interfaces 3:1859–1864
82. Wu D, Jiang Y, Li S, Li F, Li J, Lan X et al (2011b) Construction of high-quality
CdS:Ga nanoribbon/silicon heterojunctions and their nano-optoelectronic applications.
Nanotechnology 22:405201
83. Wu CY, Jie JS, Wang L, Yu YQ, Peng Q, Zhang XW et al (2010) Chlorine-doped n-type CdS
nanowires with enhanced photoconductivity. Nanotechnology 21:505203
84. Yan S, Rai SC, Zheng Z, Alqarni F, Bhatt M, Retana MA, Zhou W (2016) Piezophototronic
effect enhanced UV/visible photodetector based on ZnO/ZnSe heterostructure core/shell
nanowire array and its self-powered performance. Adv Electron Mater 2:1600242
85. Ye Y, Dai L, Wen X, Wu P, Pen R, Qin G (2010) High–performance single CdS nanobelt
metal-semiconductor field-effect transistor-based photodetector. ACS Appl Mater Interfaces
2:2724–2727
86. Ye Y, Dai L, Sun T, You LP, Zhu R, Gao JY et al (2010b) High-quality CdTe nanowires: syn-
thesis, characterization, and application in photoresponse devices. J Appl Phys 108:044301
87. Yi SS, Girolami G, Amano J, Islam MS, Sharma S, Kamins TI et al (2006) InP nanobridges
epitaxially formed between two vertical Si surface by metal-catalyzed chemical vapor depo-
sition. Appl Phys Lett 89:133121
88. Yoon Y-J, Park K-S, Heo J-H, Park J-G, Nahm S, Choi KJ (2010) Synthesis of ZnxCd1-­
xSe (0 < x <1) alloyed nanowires for variable-wavelength photodetectors. J Mater Chem
20:2386–2390
398 G. Korotcenkov and V. V. Sysoev

89. Yu Y, Luo L, Wang M, Wang B, Zeng L, Chunyan WC et al (2015) Interfacial states induced
ultrasensitive ultraviolet light photodetector with resolved flux down to 85 photons per sec-
ond. Nano Res 8(4):1098–1107
90. Yu Y, Jiang Y, Zheng K, Zhu Z, Lan XZ, Yan ZY et al (2014) Ultralow-voltage and high gain
photoconductor based on ZnS:Ga nanoribbons for the detection of low-intensity ultraviolet
light. J Mater Chem C 2:3583–3588
91. Yu YQ, Luo LB, Zhu ZF, Nie B, Zhang YG, Zeng LH et al (2013) High-speed ultraviolet-­
visible-­near infrared photodiodes based on p-ZnS nanoribbon-n-silicon heterojunction. Cryst
Eng Comm 15:1635–1642
92. Xie C, Li F, Zeng L, Luo L, Wang L, Wu C, Jie J (2015) Surface charge transfer induced
p-CdS nanoribbon/n-Si heterojunctions as fast-speed self-driven photodetectors. J Mater
Chem C 3:6307–6313
93. Xie X, Kwok SY, Lu Z, Liu Y, Cao Y, Luo L et al (2012a) Visible–NIR photodetectors based
on CdTe nanoribbons. Nanoscale 4:2914–2919
94. Xie C, Luo LB, Zeng LH, Zhu L, Chen JJ, Nie B et al (2012b) p-CdTe nanoribbon/n-silicon
nanowires array heterojunctions: photovoltaic devices and zero-power photodetectors. Cryst
Eng Comm 14:7222–7228
95. Zhang X, Wu D, Geng H (2017) Heterojunctions based on II-VI compound semiconductor
one-dimensional nanostructures and their optoelectronic applications. Crystals 7:307
96. Zhang K, Ding J, Lou Z, Chai R, Zhong M, Shen G (2017b) Heterostructured ZnS/InP
nanowires for rigid/flexible ultraviolet photodetectors with enhanced performance. Nanoscale
9:15416–15422
97. Zhang X, Meng D, Hu D, Tang Z, Niu X, Yu F, Ju L (2016) Construction of coaxial ZnSe/
ZnO p-n junctions and their photovoltaic applications. Appl Phys Express 9:025201
98. Zhang X, Tang Z, Hu D, Meng D, Jia S (2016b) Nanoscale p-n junctions based on p-type
ZnSe nanowires and their optoelectronic applications. Mater Lett 168:121–124
99. Zhang X, Hu D, Tang Z, Ma D (2015) Construction of ZnSe-ZnO axial p-n junctions via
regioselective oxidation process and their photo-detection applications. Appl Surf Sci
357:1939–1943
100. Zhang X, Zhang X, Wang L, Wu Y, Wang Y, Gao P et al (2013) ZnSe nanowire/Si p-n hetero-
junctions: device construction and optoelectronic applications. Nanotechnology 24:395201
101. Zhang X, Zhang X, Zhang X, Zhang Y, Bian L, Wu Y et al (2012) ZnSe nanoribbon/Si
nanowire p-n heterojunction arrays and their photovoltaic application with graphene trans-
parent electrodes. J Mater Chem 22:22873–22880
102. Zhang A, You SF, Soci C, Liu YS, Wang DL, Lo YH (2008) Silicon nanowire detectors show-
ing phototransistive gain. Appl Phys Lett 93:121110
103. Zhang D, Li C, Han S, Liu X, Tang T, Jin W, Zhou C (2003) Ultraviolet photo-detection prop-
erties of indium oxide nanowires. Appl Phys A Mater Sci Process 77:163–166
104. Zhao W, Liu L, Xu M, Wang X, Zhang T, Wang Y et al (2017) Single CdS nanorod for high
responsivity UV–visible photodetector. Adv Optical Mater 5(12):1700159
105. Zhou W, Peng Y, Yin Y, Zhou Y, Zhang Y, Tan D (2014) Broad spectral response photodetec-
tor based on individual tin-doped CdS nanowire. AIP Adv 4:123005
106. Zhou J, Gu Y, Hu Y, Mai W, Yeh PH, Bao G et al (2009) Gigantic enhancement in response
and reset time of ZnO UV nanosensor by utilizing Schottky contact and surface functional-
ization. Appl Phys Lett 94:191103
Chapter 17
QDs of Wide Band Gap II–VI
Semiconductors Luminescent Properties
and Photodetector Applications

M. Abdullah, Baqer O. Al-Nashy, Ghenadii Korotcenkov,


and Amin H. Al-Khursan

17.1 Introduction

The optoelectronic device that converts the incident light into an electric signal is
the photodetector (PD). The principle of the photosensitive semiconductor devices
depends mainly on the electron transition under high energy to the conduction band
(CB), leaving a hole behind. Ideally, each photon creates a single electron-hole pair.
This pair can recombine at a characteristic time (lifetime), releasing additional
energy in the form of heat. In an applied electric field, the drift of electrons and
holes creates an electric current, and the way the field is applied determines the type
of detector. The basics of PDs are explained well in the textbooks; see, for example
[18, 75], and the reviewed paper [45, 89]. In semiconductor PDs, three types of
mechanisms convert light into current: photoconductive effect, photogate effect,
and photovoltaic effect [45]. In addition, PDs can be divided into photoconductive
PDs and photodiode PDs. The photodiode accumulates a charge at the junction of
n- and p-type semiconductors or between a semiconductor and a metal contact
(Schottky junction). At the same time, the photoconductors have a semiconductor

M. Abdullah
Nasiriya Nanotechnology Research Laboratory (NNRL), College of Science, University of
Thi-Qar, Nasiriya, Iraq
Department of Physics, College of Science, University of Thi-Qar, Nasiriya, Iraq
B. O. Al-Nashy
Department of Physics, College of Science, University of Misan, Omarah, Iraq
G. Korotcenkov
Department of Physics and Engineering, Moldova State University, Chisinau, Moldova
e-mail: ghkoro@yahoo.com
A. H. Al-Khursan (*)
Nasiriya Nanotechnology Research Laboratory (NNRL), College of Science, University of
Thi-Qar, Nasiriya, Iraq

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 399
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_17
400 M. Abdullah et al.

crystal with ohmic contacts. The phototransistor is a photoconductor-type with an


additional gate insulated by a dielectric material. The applied gate voltage modifies
the conductivity and reduces noise through the dark current suppression.
The required PD characteristics are the high sensitivity, the spectral selectivity,
the high signal-to-noise ratio, the fast photoresponse, the remarkable stability, and
the simple manufacturing process [88]. PDs with high sensitivity in various spectral
ranges are in demand in a wide range of applications, such as defense, industry,
healthcare, imaging, biomedicine, object discrimination, advanced optical wireless
communication systems, chemical analysis, and scientific research in various fields
[2, 10]. PDs can be fabricated using single crystalline materials, epitaxial layers and
polycrystalline films deposited by various methods. In recent decades, photodetec-
tors have also been developed on the basis of nanomaterials such as 1D, 2D and
quantum dots (QDs). Studies have shown that the use of these materials, especially
QDs, can improve the performance of photodetectors [56, 85, 88].
Quantum dots (QDs), artificial atoms, particles of 2–20 nm have discrete atomic-­
like energy levels due to the size quantization. Such QDs mimic the atomic solids.
One of the main characteristics of QDs is the tuning of their energy states by chang-
ing the QD size.The energy configuration in QDs is different for the same QD semi-
conductor crystal, depending on the QD size. This tuning in electronic properties is
accompanied by a change in their optical characteristics [4]. Thus, the advantages
of QDs lie in their tunable bandgap, as well as their superior electronic, surface and
optical properties. Their surface-to-volume ratio, in addition to a short diffusion
lengths, provides excellent photogeneration of charged carriers and then best used
for PD operation. Therefore, QDs are used in various sciences including physics,
material science, chemistry, energy technology, chemical and biosensing, medicine
and biology [5, 26, 46, 48, 58, 87, 93]. In bio-detection, QDs are used in bio-­
imaging, bio-labeling, and bio-molecular detection [86]. In particular, in chemical
sensors, QDs are used as probes for ions, such as Cu2+, K+, Na+, Hg2+, and small
molecules, like glucose and aromatic hydrocarbons. The QDs as molecular detec-
tors exhibit simplicity, selectivity, and cost-effectiveness [20, 62].
Experiment has shown that colloidal QD-based PDs fabricated by solution pro-
cessing technology are cheaper than thin-film devices fabricated by vacuum deposi-
tion technology [49, 60, 95]. QD-based photodetectors are now being manufactured
as self-powered devices, that generate potential when illuminated to operate in
short-circuit or open-circuit mode [10, 50, 57]. However, a more promising direc-
tion is the development of flexible and UV photodetectors based on QDs. However,
a more promising direction is the development of flexible and UV photodetectors
based on QDs. Currently, there is a great need for flexible PDs, since wearable elec-
tronics, in contrast to traditional “hard” PDs, require both high sensitivity and flex-
ibility. Wearable electronics, the milestone in future development of optoelecttronic
devices intended for applications in the fields of biomedicine, sports and disease
diagnostics. As far as UV photodetectors are concerned, with zero background
detection of the deep ultraviolet (DUV, 200-280 nm) signal under sunlight or room
lighting, the DUV (SBDUV) PD solar curtains are an excellent tool for applications
such as ozone detection, space communications, missile plume detection, flame
17 QDs of Wide Band Gap II–VI Semiconductors Luminescent Properties… 401

monitoring, biological medical analysis, and offshore oil inspection [2, 89].
Ultraviolet (UV) nanostructure PDs are reviewed in many articles [2, 55, 85, 89].
Except for ref. [2], which reviews ZnSe, all these reviews do not take into account
II-VI PDs. Wang et al. [87] reviewed the zinc-containing nanocrystals for energy
conversion applications covering the ZnS and ZnSe nanocrystals. Type I and type II
core-shell quantum dot (QD) synthesis is reviewed well in the chapter of Drofs
et al. [25].
Charge carriers in QD-based PDs are transferred in three dimensions resulting in
great trapping states on the surface. Therefore, there are some difficulties in the
process of electron transfer between QDs. While colloidal QDs can effectively cre-
ate electron-hole pairs, all the photogenerated charge carriers are not transferred to
the corresponding electrodes, as is required for excellent device performance. They
travel through the adjacent QDs and are trapped before reaching the electrodes. This
problem is solved by bringing the QDs closer enough using short-chain ligands
(such as trioctylphosphine oxide (TOPO), TOP, Oleic acid, Oleylamine) to boost the
performance through the best inter-dot charge transport. In this case the electron
wave function is strongly delocalized between the neighboring QDs, enhancing the
photocurrent.

17.2 II-VI Semiconductor Quantum Dots-Based PDs


and Their Fabrication Methods

The widest band gap II-VI semiconductor, ZnS, has a large bandgap of 3.72 eV
(2.5 nm Bohr radius) for cubic zinc-blende and 3.77 eV for hexagonal wurtzite
crystals, making it suitable for visible-blende UV devices [93]. The quantum size
effect can tune the bandgap beyond 4.43 eV by reducing the QD size below the
Bohr radius.
A large-scale, cost-effective method is vital for QDs production (read the Chaps.
12 and 13, Vol. 1). Liquid phase epitaxy, hot injection, sputtering, thermal evapora-
tion, chemical vapour deposition (CVD) are expensive due to their equipment, lim-
iting large scale production, or require high temperature and pressure [54]. Wet
chemical methods are simpler and cheaper methods for the synthesis of II-VI com-
pounds. It is a low temperature, inexpensive method suitable for mass production.
This liquid-phase synthesis of QDs results in a good dispersion and stability in a
solvent, easy preparation, pot reaction, low-cost, and controlling particle size [54].
The most important factor in this method is the choice of the appropriate composi-
tion of the solution for synthesis, which, in addition to the precursors of the compo-
nents of II-VI compounds, may include surfactants and polyelectrolytes. For
example, for the synthesis of ZnS QDs by this method, Li et al. [54] used sodium
dodecyl sulfate as a passivating agent, controlling their nucleation growth rate and
confirming its stability. The resulted PD is highly suitable for UV detection by its
wide bandgap and high electron mobility (600 cm2/V.s). He et al. [36] introduced
402 M. Abdullah et al.

the 3-mercaptopropionic acid as a stabilizing reagent and a source of sulfur in the


synthesis of Cu:ZnS QDs assisted by microwave irradiation. Compared with the
conventional method, the microwave-assisted wet chemical method accelerates the
synthesis of Cu:ZnS QDs, while no phase transformation occurs, and the crystal
quality is improved. The QD tunable range is possible by controlling the reac-
tion time.
Large bandgap ZnS QDs have high environmental stability and can be used for
photocatalysts, solar energy conversion, phosphor, and photodetectots. Xia et al.
[95] fabricated a solar-blind, deep-UV (SBDUV) ZnS QDs PD by adopting ZnS
QDs via solution process with a bandgap tuned beyond 4.68 eV and emitting a sharp
exciton peak at 256 nm with a fast and stable response. Premkumar et al. [72] dem-
onstrated visible blind UV detection using ZnS QDs grown by a simple reflux tech-
nique. The ZnS QDs UV sensor exhibits the best photocurrent response and high
responsitivity. However, their long recombination lifetime slows the device
response, limiting their applications.
ZnSe QDs also have a wide-direct bandgap (the bulk bandgap is 2.67 eV),
enabling UV detection. Their spectrum cover 300–450 nm wavelength (A and B UV
regions), broader than the UV spectrum covered by GaN. A slight mismatch between
ZnSe and GaAs can be corrected by growing ternary or quaternary ZnSe structures
on the available GaAs substrates [2]. ZnSe is the least toxic and high photolumines-
cence QD material. However, the ZnSe microfluidic platform still remains a chal-
lenge due to the low solubility of most Zn precursors and the low reactivity of Zn-Se
[33, 39, 79]. Bratskaya et al. [11] reported on the green synthesis of ZnSe QDs on
an aqueous solution of polyampholyte chitosan derivatives-N-(2-carboxyethyl) chi-
tosan, producing a high intensity of photoluminescence. ZnSe devices are environ-
mentally friendly and used for biological imaging, DNA analysis, and therapy
diagnosis. Their high optical nonlinearity makes them adequate for optical limiting
and ultrafast switch devices [15, 65]. Fang et al. [27] demonstrated the first ZnSe
nanobelts sensor by chemical vapor deposition (CVD) method that had an ultralow
dark current (below 10−14 A, the detection limit of the current meter) and response
time (τ < 0.3 s) less than conventional ZnSe PD sensors.
Mishra et al. [64] synthesized QDs of CdS, another wide band gap II-VI semi-
conductor, by a simple co-precipitation method. Absorption in the UV-visible region
has an absorption peak at 427 nm where it is blue-shifted by 0.48 eV compared to
bulk CdS. In [32] CdS QDs were synthesized by liquid phase technique at room
temperature and atmospheric pressure using the sodium alkyl sulfonate as capping
agent. Magic size CdS QDs were synthesized by Dickon and Hu [24] by a no-­
injection reaction using 1-octadecene as a solvent, a cadmium salt as a source of
cadmium, and the organic sulfur 1-dodecanethiol as a source of sulfur. The acid
acted as a capping ligand on the surface and gently controlled the solubility. As a
result, single-sized CdS QDs with a fixed position of the emission wavelength were
formed due to homogeneous nucleation and size uniformity. Husham et al. [38]
propose a simple microwave-assisted chemical bath deposition (CBD) to synthesize
a self-powered PDs from nanocrystalline CdS.
17 QDs of Wide Band Gap II–VI Semiconductors Luminescent Properties… 403

Jiang et al. [40] also developed CdS QDs-based photodetector. They modified
the CdS QDs by Au nanoparticles to increase the performance via surface plasmon
resonance (SPR) and then used MXene as an electrode. Ti3C2Tx MXene, a 2D struc-
ture, exhibits Ohmic contacting with CdS at a non considerable Schottky barrier of
0.07 eV height. First, the prepared CdS core-Au nanoparticles were dispersed in
ethanol. Then the uniform solution above was dropped onto fluorophlogopite sub-
strate (1 cm × 1 cm) and natural air-drying at room temperature to fabricate the
device. The MXene colloidal solution was dropped on both sides of CdS core-Au
nanoparticles as electrodes. Under illumination, hot electrons are injected from to
CdS QDs from two sides: due to SPR from Au nanoparticles and due to its photo-
thermal property from MXene. The photoresponse polarity depends on the higher
electrons transfer to CdS. Jiang et al. [40] have found that a broad-spectrum, begin-
ning from deep UV to near IR can be detected at high response by developed
PD. The specific detectivity at 405 nm was equaled ~1.3·1011 Jones [40].
However, the use of CdS QDs is prohibited in Europe from an environmental
point of view. Therefore, when using CdS QDs, it is necessary to form protective
coatings on its surface. In this case, we get a core-shell structure, where CdS per-
forms the functions of the core, and the less toxic material forms the shell. One such
example is the quantum dots of the CdTeSe/ZnS core with a surface modified with
trioctylphosphine oxide (TOPO) [12].

17.3 QD Core/Shell Structures

The nanoscale QDs have a high surface-to-volume ratio. A significant fraction of


atoms are located on the surface, and some atoms are not fully coordinated. As a
result, we have a surface with increased reactivity, prone to chemical interaction
with the surrounding atmosphere, primarily with oxygen and water vapor. Dangling
bonds in QD surface defects also create nonradiative recombination centers, which
worsen their optical and transport properties. To eliminate unwanted interactions
that are the main cause of instability of QDs parameters, the QDs surface needs to
be passivated by surface ligands. The loss of ligands that occurs during film forma-
tion, leads to the formation of dangling bonds and defects on the surface, which can
be evidenced by low PL QY. Coating the QD with an inorganic shell material with
a larger band gap can reduce the reactivity of the QD surface with oxygen and water.
Encapsulation of QD cores with shells effectively passivates anionic and cationic
surface areas, eliminates harmful effects, prevents heavy metals from entering the
environment, improves bothe the photostability against photooxidation, and the
photoluminescence (PL) quantum efficiency. It is important to note that the match
of the crystal structure and the lattice constant between the core and the shell are
important to avoid lattice strain at the interface. With a big mismatch and when the
shell becomes thick, the strain-induced defects increase, resulting in the decrease of
PL QY. It was found that alloying and graded shells relax strain with increasing
shell thickness [29, 33, 52].
404 M. Abdullah et al.

Fig. 17.1 (a) Schematic illustration of shell growth. Route I: layer-by-layer growth or CBD of
CdSe–CdS core–shell QDs. Route II: The cation-exchange reaction used to convert core-only PbS
QDs into core–shell PbS–CdS QDs. (Reprinted from Ref. [80]. Published 2017 by Oxford
University Press as open access)

Fig. 17.2 Types of core-shell QDs based on band alignment. (Reprinted with npermission from
Ref. [86]. Copyright 2015: Elsevier)

In most cases, core-shell structures are formed using the two paths shown in
Fig. 17.1. As a result, depending on the properties of the materials used, two types
of core/shell structures can be formed: type-I and type-II. Figure 17.2 shows the
band gap alignments for each of these types. Table 17.1 compares QDs classifica-
tion based on their band alignment, the location of carriers, quantum yields, stokes
shift, and the average absorption and emission ranges [1]). It also provides examples
of core/shell QDs based on II-VI semiconductors and lists some of their
limitations.
17 QDs of Wide Band Gap II–VI Semiconductors Luminescent Properties… 405

Table 17.1 Types of core/shell QDs based on band alignment


Parameter Type I Inverse Type I Type II Inverse Type II
Band gap The band gap of The band gap of Valence band Conduction band
the core is smaller the core is greater edge of the core edge of the core is
than the band gap than the band gap is within the band within the band gap
of the shell, as well of the shell, as well
gap of the shell of the shell or
as the band gap of as the band gap of or conduction valence band edge
the core falls the shell falls band edge of the of the shell is within
within the band within band gap of shell is within the the band gap of the
gap of the shell the core band gap of the core
core
Excited Excited electrons The excited One charge One of the excited
electrons/ and holes are electrons and holes carrier either electrons or the
holes completely are completely or excited electron holes are
positions confined in the partially confined or hole is delocalized in the
core region in the shell based confined to the core/shell structure,
on the thickness of core, while the and the other one is
the shell. other is mostly confined within the
confined to the core.
shell
Quantum Higher QY and Lower QY and Lower QY and Relatively higher
yield (QY) long-term stability poor stability poor stability QY and fair
stability
Stokes shift Small Significantly large Large Large and tunable
via controlling the
size of the core and
thickness of the
shell
Average (400–500) nm (400–500) nm (600–800) nm (300–1600) nm
absorption
range
Average (430–600) nm (400–700) nm (700–1000) nm (700–1000) nm
emission
range
Limitations The shell can trap Both the excited One of the The excited electron
charge carriers electrons and holes excited electrons or hole can be
which leads to may leak to the or hole leak to absorbed leading to
reduced surface the surface reduced excited
fluorescence QY decay time one
carrier is mostly
confined to the core,
while the other is
mostly confined to
the shell
Construct/ CdSe/ZnS CdS/HgS CdTe/CdSe InP/CdS
materials CdSe/CdS CdS/CdSe CdSe/ZnTe PbS/CdS
CdS/ZnS ZnSe/CdSe CdSe/ZnSe
Source: Reprinted from Ref. [1]. Published 2019 by MDPI as open access. Table 1
406 M. Abdullah et al.

Table 17.1 shows that when going from Type-I to Type-II, the difference in the
bandgap between the core and shell decreases, and the excited electrons are spa-
tially depart from each other. The stokes shift increased while the quantum yield and
stability decreased with increasing absorption wavelength. As a result, Type-I is
more suitable as concentrators compared to Type-II because they show no leakage
of excited electrons, show better stability and also offer higher QY [1]. It is neces-
sary to note that in comparison with organic passivation, core-shell passivation pro-
vides higher quantum yield, extraordinary photoluminescence, enhanced optical
properties, increased half life time, improved stability towards photo-oxidation and
finally better structural properties. As a result, the core/shell confinement results in
a higher power conversion efficiency than the only core QD structures [41].
CdS/ZnS core/shell nanoparticles were successfully synthesized using a two-­
step chemical synthesis method. Compared to CdS and ZnS nanoparticles, CdS/
ZnS core/shell QDs exhibit a photocatalytic enhancement due to the reduced recom-
bination of photogenerated electron-hole pairs and increasing the CdS/ZnS core/
shell photocatalytic properties [74]. Increasing cadmium concentration in the shell
for ZnS/CdxZn1-xS core/shell QDs decreases the interface optical phonon frequen-
cies due to the reduced mass of the CdZnS shell [37].
CdSe/ZnS, CdSe/CdS, and CdSe/ZnS/CdS nanocrystals are of great importance
because their luminescence wavelength overlaps the full vision wavelength and
high photoluminescence [6, 96]. Coating CdSe core with CdS shell (or CdS/ZnS
multishell) exhibits redshift and boost absorption and PL spectra. Therefore, the
semiconductor CdSe/CdS/ZnS core-shell QDs are an alternative to fluorescent
nanoparticles for data storage equipment, sensors, optoelectronic devices, and solar
cells. CdSe/ZnS core-shell can also be synthesized via the successive ion layer
adsorption and reaction (SILAR) method. This conventional hot injection synthesis
method injects Zn and S precursors into a solution containing CdSe QDs. The few
structural differences and the similarity of internal energies result in zinc blende and
wurtzite structures for CdSe QD morphology. CdSe/ZnS QDs of zinc blend crystal
structure has the best optoelectronic properties compared to the wurtzite type. The
element of high average atomic number (Z) has high intensity. The average Z value
of the CdSe core (ZCd = 48 and ZSe = 34) is twice that of the ZnS shell (ZZn = 30 and
ZS = 16) [29].
ZnSe has numerous advantages such as high stability and better doping effects.
However, it takes some time before this becomes decisive for its application due to
problems in realizing high-quality ZnSe QDs. ZnSe QDs adhere to any nanoparticle
surface, creating a nano-hybrid system with enhanced lifetime and high perfor-
mance. Passivating the ZnSe QD core with a broader bandgap (like ZnS) shell
reduces the recombination and allows better stability [2, 61, 76, 83]. It is proved that
ZnS-based PDs have a lower leakage current. A colloidal ZnTe/ZnSe core/shell
QDs of the type-II structure extends the PL wavelength from blue to amber [9]. This
structure has an energy conversion efficiency 11 times higher than ZnSe QDs for
photovoltaic devices. This improvement is a result of the properties of the Type-II
core shell structure that increase carrier extraction and extend wavelength [9]. When
the shell thickness increases, the ZnSe/CdSe core/shell QDs are tuned from Type-I
17 QDs of Wide Band Gap II–VI Semiconductors Luminescent Properties… 407

to Type-II and then back to Type-I, see Fig. 17.2, when the shell thickness is
increased [13]. ZnSe is used as a shell for CdSe QD optical amplifier with a high
emission coefficient and low noise [7, 13]. ZnS is reliable as a shell for type-I align-
ment with ZnSe QDs for high fluorescence. The growing mode of the ZnS shell is
modified by tuning the precursor from kinetic (fast) to thermodynamic (slow)
regimes. In the thermodynamic regime, the quantum yield increases, and on-off
blinking reduces, which prevents the growth of traps at the core/shell interface [43].
A ZnO/ZnS core-shell nanorod array as a photoanode is synthesized as a self-­
powered photoelectrochemical cell, using facile chemical solution method such as
SILAR [57]. This PD showed an improvement in both photosensitivity and sensitivity.
This improvement is a result of excess light collection and rapid separation of photo-
generated electron-hole pairs due to ZnS coating. The highest photoresponsivity was
obtained at seven SILAR cycles for ZnS deposition. Further increment ZnS SILAR
cycles reduces the photoresponsivity of ZnO/ZnS array UV PDs, where a thick layer
of ZnS nanoparticles may hinder the hole transport and increase the recombination of
the photon-generated electron-hole pairs. Lin et al. [57] also found that this PD had
the best wavelength selectivity making it suitable for the UV-A range [57].

17.4 Doping Influence on QDs Properties

In addition to the controllability of the size and shape of QDs, doping with impuri-
ties is vital for the developing electric and optical properties. The energy levels of
the dopant lie in the bandgap of the host crystal and change the electronic and opti-
cal properties of the host crystal. In particular, the emission and absorption spectra
vary depending on the dopant and its concentration (see Fig. 17.3). II-VI semicon-
ductor QDs can be doped with transition metals and lanthanide ions such as Ni+,
Mn2+, Ag2+, Cr3+, Eu+3, Sm3+, and Tb3+ [42, 60]. For example, Mn2+ doped ZnS QD
shows luminescence at 413 and 585 nm wavelengths. ZnS:Mn QDs were synthe-
sized by employing a simple aqueous method using starting materials: zinc sulfate,
sodium sulfide, manganese acetate tetrahydrate, a complex agent, and a solvent
[16]. On the other hand, defects generated at a high dopant concentration quench the
PL. Precisely this effect was observed at a high Mn2+ concentration; in addition to
PL quenching, a redshift of the orange (impurity) luminescence wavelength and a
blueshift of the edge luminescence of the initial ZnS QDs were also observed [16].
This means that this effect must be taken into account when doping QDs.
It should be noted that Mn2+ ions are the common dopant used for ZnS QDs due
to their low cost and ability to change the electronic and optical properties of ZnS
QDs. The most important advantages of Mn-doped ZnS QDs are low toxicity, lon-
ger excited-state lifetimes (on millisecond), significant Stokes shift between excita-
tion and emission wavelengths which avoids self-absorption, the absorption range
of doped QDs is extended, insignificant interference from associated compounds,
no light scattering, increased thermal and environmental stability. Mn2+ dopant is
also exhibiting a hyperfine splitting in the paramagnetic resonance. The excitation
408 M. Abdullah et al.

Fig. 17.3 Emission spectra of undoped and doped ZnS, ZnSe, and CdS QDs

occurs first through the host material; then, energy is transferred to the impurity. The
Mn2+ doped ZnS QDs range is limited from yellow to orange [16, 67]. Lee et al. [53]
also observed that QDs from an Mn2+ doped ZnS core with a ZnS shell show
improved PL intensity, but electric current decreases due to the core–shell electron
confinement effect. Al-Haddad et al. [3] prepared Mn2+ and Ce3+ doped ZnS nano-
crystals by simple microwave irradiation and found that responsivity, detectivity,
and quantum efficiency for Ce3+ doped ZnS are higher than that for Mn2+ doped ZnS
nanocrystals.
Conversely, Cu dopants existing as either Cu+ or Cu2+ state, offer more possibili-
ties to develop optical properties and extend the visible range of ZnS QDs. The
Cu:ZnS QDs have a tuned range between 500–595 nm wavelength [19, 36]. The
samples of Cu- and Fe-codoped ZnS QDs exhibit a UV peak at 315 nm wavelength
in addition to a visible one at 500 nm for the co-doped sample. These samples were
prepared using a hot injection method [31]. According to Grandhi et al. [31] mid-­
gap energy states occur due to doping with transition metals and a charge transfer
between the energy band of the host material and the level of doping in the band
gap. Cui et al. [19] suggest that doped copper atoms locate at the vacancy of the Zn
atom of the QDs in the Cu2+ ion form. The defect levels are the shallow accepter
levels reducing the n-type photovoltaic characteristic in the Cu-doped QDs.
Cr-doped ZnS QDs, prepared by co-precipitation method, have two absorption
peaks in the visible range at 425 nm and 595 nm, related to Cr3+, in addition to the
UV peak below 370 nm, characterized the intrinsic ZnS bandgap absorption [97].
17 QDs of Wide Band Gap II–VI Semiconductors Luminescent Properties… 409

Study has also found that due to the dopant intrinsic property, Cr-doped ZnS QDs
exhibit weak ferromagnetism or super-paramagnetism at room temperature. The
Cr-doped ZnS QDs demonstrate a sharp electron paramagnetic resonance, which
indicates an increase in the number of unpaired electrons upon doping. The fourfold
spin degeneracy of the ground state of the Cr3+ is removed by a subsequent low sym-
metric field and splits into Kramers doublets, Ms. = ±3/2 and ± 1/2. Hence, Cr:ZnS
QDs are promising for spintronic device applications [71].
Ni-doped ZnS QDs have also recognized ferromagnetic properties compared to
paramagnetism with bare ZnS QDs [78]. It was found that Ni-doped ZnS QDs
exhibit size-reduction attributed to the small Ni-ion radius than Zn-ion. The ionic
radii of Ni2+ and Zn2+ are 0.07 nm and 0.074 nm, respectively. Structural character-
ization of ZnS:Ni has shown that Ni during doping incorporates in the ZnS matrix
at interstitial or substitutional sites. This difference between ion radius diminishes
the bond length that connects S-anion and Ni-cation, vice versa to Zn-S bond length,
causes a lattice strain and leads to stress in the ZnS crystal. This stress shifts the PL
peak position of the Ni-doped ZnS QDs [49]. The sample color is also changed from
light green to dark green at high Ni-ion concentrations.
Ag:ZnS QDs were prepared by wet chemical co-precipitation. These samples
were found to show improved PL [77]. A blue PL emission peak with a short radia-
tive lifetime was found. Sahai et al. [77] believe that Ag:ZnS QDs are the best
choice for display devices due to undelayed image transformation.
As for the nature of the emerging photoluminescence peaks, the studies per-
formed in [35] for ZnS QDs synthesized via the vapor-liquid-solid technique and
dopped with metal ions emited in the visible band, have shown that blue emission is
ascribed to S vacancies, green emission is due to Zn vacancies, and yellow-orange
and red colors are responsible for impurity defects in the host ZnS. Defect states
alert the balance of participating states in the recombination processes and substan-
tially modify the radiative recombination kinetics in the host matrix [35].
Similar studies were carried out for ZnSe QDs. It was found that ZnSe QDs
doped with transition metals, such as Cu:ZnSe, Mn:ZnSe, and Fe:ZnSe QDs, still
have the same cubic zinc blende crystal type after doping. However, the photolumi-
nescence intensity of Mn:ZnSe QDs and Fe:ZnSe QDs is lower than that of Cd:ZnSe
and Cu:ZnSe QDs, which indicates a greater number of structural defects in these
QDs compared to ZnSe QDs doped with Cd and Cu [99]. In addition, for these
samples, due to the increase in size after doping, a red shift in photoluminescence is
observed for ZnSe QDs doped with transition metal ions.
An analysis of the optical properties of doped ZnS and ZnSe quantum dots shows
that doping with rare earth elements is promising in applications such as high-­
performance multicolor photodetectors and nonlinear optics. A multicolor light-­
detecting PD was also fabricated using Er3+-doped CdS nanoribbons (Er-CdS NRs)
[22]. Dedong et al. [22] reported that this PD can detect blue, red, and near-infrared
light. This wide range of detection results from the energy band structure of Er-CdS
NRs and especially the transitions of Er3+ ions energy levels.
Some parameters of photodetectors developed on the basis of quantum dots of
wide band gap II-VI compounds are listed in Table 17.2.
410 M. Abdullah et al.

Table 17.2 Photoresponsivity and the on-off ratio of photodetectors based on ZnS, ZnSe, and CdS
nanostructures
Photodetectors Responsivity (R) (A/W) λ (nm) On-off ratio Ref.
ZnS nanobelts 0.12 320 [27]
ZnS QDs 0.31 390 413 [72]
ZnS QDs 5.83 365 [54]
CdS QDs 0.3 μA/W 454 [32]
CdS nanorods 1.23 × 104 450 [100]
Er:CdS nanorods 9.19 × 103 – 3.46 × 104 457–955 [22]
Pd/Ce doped ZnS/Pd 5.59 300 [3]
ZnS QDs 0.1 mA/W 254 [95]
ZnO/ZnS core-shell nanorod 0.056 340 [57]
ZnO/ZnS MC 177 320–380 [10]
ZnSe QD-CdTe nanoneedle – [76]
ZnSe-PbSe nanowire 40 350–510 [23]
1D ZnS/CdS – 300–550 1.2 × 105 [59]
CdS/CdSSe axial nanowire 1.18 × 102 500 105 [34]
CdS/CdS:SnS2 superlattice nanowire 2.5 × 103 350–650 1.5 × 105 [30]
CdS QD-Au/MXene 86 mA/W 405 [40]

17.5 1D Structures and QDs


in Heterostructure-Based Photodetectors

The fabrication of heterostructures based on 1D structures and QDs is another


approach to fabrication of photodetectors. For the hybrid or heterostructure-based
PDs, the enhanced photoresponse is due to large nanostructure surface-to-volume
ratio allowing more oxygen molecules to adsorb on the surface, giving a thicker
charge depletion layer and reducing the dark current. The multiple scattering light
among the nanostructure branches can further increase the light absorption, increas-
ing the photocurrent. The energy diagram of heterostructures based on QDs of vari-
ous materials can be represented as shown in Fig. 17.2. Electrons generated under
illumination are transferred into the lower bandgap side while holes are into the
higher side. The separation of the photogenerated electron-hole pairs reduces the
rate of charge carriers recombination and significantly increases the photoresponse
[30, 59].
Currently, there are reports on the formation of heterostructures using a wide
variety of materials. For example, Deng et al. [23] fabricated PbSe-ZnSe nanowire
PDs as flexible PDs. The PbSe-ZnSe nanowire has type II alignment where elec-
trons and holes are separated at the interface. This PD covers UV-visible-NIR spec-
tral regions. Lou et al. [59] proposed to use one-dimensional ZnS/CdS
heterostructures synthesized by thermal evaporation for the development of such
photodetectors. For these PDs, an ultrahigh on-off ratio (105) was obtained, which
is several orders of magnitude higher than for devices with pure ZnS or CdS nano-
structures. At a wavelength of 300–550 nm, the sensitivity was high due to the
17 QDs of Wide Band Gap II–VI Semiconductors Luminescent Properties… 411

coupling between parts of the ZnS/CdS heterostructure, and the response time was
short. It was also found that the integration of these ZnS and CdS II-VI semiconduc-
tors into the ZnS/CdS heterostructure is beneficial for broadband photodetection in
the UV-visible range (Lou et al. [59]. ZnS works in the UV range, while CdS works
in the visible range. Such photodetectors are required for light image processing,
memory storage or switches with high on-off ratio, significant stability and high
response speed. Lou et al. [59] also studied PDs based on one-dimensional ZnS/
CdS heterostructures fabricated on flexible PET substrates. Under bending condi-
tions, stability, photoresponse to visible light, and photocurrent were also proved to
be excellent for this flexible PDs, promising portable and flexible devices [59].
Yadav et al. [89] also fabricated flexible PDs, using ZnO/CdS core/shell micro/
nanowire synthesized by the facile hydrothermal method. They have found that the
created piezoelectric potential inside this structure enhances charge carrier genera-
tion and separation in the metal-semiconductor contact or p-n junction. The broad
polarization creates a piezoelectric potential under stress in the wurtzite crystal of
the ZnO nanowire. Yadav et al. [89] determined that this PD is sensitive both to
green light (548 nm), and UV light (372 nm). The sensitivity of ZnO/CdS core/shell
micro/nanowire was three orders higher than CdS nanoribbon and six times higher
than CdS nanowire. In addition, the responsivity was significantly increasing by
more than ten times compared to the unstrained nanowire. Yadav et al. [89] believe
that the improvement in the characteristics of a photonic sensor was due to coupling
of piezoelectric, electrical, and optical properties in the developed semiconductor
device. The internal piezoelectric field in ZnO controls the charge transfer through
carrier separation, optimizing the photoresponse, and the piezoelectric potential
itself acts as a “gate” voltage [99].
Benyahia et al. [10] elaborated an eco-friendly ZnO/ZnS composite-based
UV-visible PDs with high photoresponse. PDs exhibited a very few dark-current
(5 nA) and high on-off ratio (78 dB). In addition, prepared ZnO/ZnS PDs had low
optical losses and high absorption at UV-visible, guaranteeing 270% enhancement
than the conventional ZnS thin-film PDs. The thermal treatment evaporates the ZnS
layer and creates cavities. It was found that the forming such ZnO/ZnS structures
with voided regions increases the light-trapping due to multiple reflections acting as
optical confinement regions, making this structure a promising high photoresponse
multispectral photodetector [10].
CdS/CdS:SnS2 superlattice nanowires synthesized using co-evaporation technol-
ogy with local environmental control have also been used to develop UV photode-
tectors [30]. Its photo-responsivity was about seven times higher than pure CdS
PD. However, it was found that the photocurrent of superlattice devices increased
linearly rapidly at low light intensity and then slowly increased at high light inten-
sity. This means that the photosensitivity decreases at high light intensity. According
to Gou et al. [30], this effect is associated with a slow increase in the photocurrent
upon saturation of photogenerated carriers.
Guo et al. [34] have shown that high-quality CdS/CdSxSe1 − x axial heterostruc-
ture nanowires, grown using a moving-source improved CVD method, allowing a
large-scale synthesis, are also promising for development of UV-visible PDs. For
412 M. Abdullah et al.

these photodetectors, the spectral range of 350–650 nm, response time (∼68 μs),
and on-off ratio of 105 were determined. This performance was higher compared to
PD developed on CdS and CdSxSe1 − x nanostructures separately (Guo et al. [34].
It should be noted that 1D structures and quantum dots can also form hetero-
structures. It is known that the fast capture of photogenerated electrons at the inter-
face is still a challenge for efficient light-harvesting from photonic devices using
ZnSe QDs. Given this problem, it has been suggested that hybridizing QDs with
other nanostructures can contribute to more efficient separation of the photogene-
rated electron-hole pairs and enhance light harvesting. It is this approach that Saeed
et al. [76] used when developing a photodetector based on ZnSe QDs. ZnSe QDs,
synthesized by the wet chemical method, were adhered to one-dimensional (1D)
nanostructures. Said et al. (2020) showed that the attachment of QDs to one-­
dimensional (1D) nanostructures indeed promotes efficient electron transfer from
the QD surface to the 1D nanostructure. It was found that for the ZnSe QD-CdTe
nanoneedle heterostructure, the PL peak is redshifted, which is related to nanohy-
brid interfaces. In addition, the PL peak of the nanohybrid structure becomes more
intense, which is associated with the rapid transfer of excitation energy from the QD
as it is the higher energy side (ZnSe band gap ~2.7 eV) to the nanoneedle with a
small band gap (CdTe band gap ~1.2 eV). Estimates have shown that the charge-­
energy transfer from the QD to the nanoneedle occurs in <800 ps. All these results
promise nanohybrid photovoltaic devices with better performance [76].
Quantum dot-quantum dot is another form of heterostructure. In particular, UV
PDs with such configuration were developed by Peng et al. [69]. ZnO and ZnS
spherical nanoshells connected as hollow-sphere bilayer nanofilm were constructed
by the self-assembly method. The developed bilayer QDs-based UV PDs demon-
strated higher sensitivity, excellent stability, and short response times compared to
monolayer devices. The ZnO/ZnS QDs and ZnS/ZnO QDs bilayer nanofilms have
type II staggered heterostructure. The electron affinity of ZnO is higher than that of
ZnS. According to Peng et al. [69], holes are transported to the ZnS layer and elec-
trons to the ZnO layer. The photocurrent enhancement is correlated to the electron-­
hole separation by the internal electric field in the nanofilm bilayer, reducing their
recombination and increasing photocurrent compared to the monolayer devices.
The experiment showed that heterostructures for the manufacture of UV photo-
detectors can be formed not only between nanostructured chalcogenides. In particu-
lar, Zhao et al. [100] fabricated UV PDs based on CdS–Si heterojunctions. CdS NRs
were grown on SiO2/Si substrate over a large scale using CVD. The PD developeted
had a high responsivity due to best crystallinity and a large surface of CdS NR
[100]. While indirect bandgap of Si is an obstacle for its UV-visible detection,
porous silicon (PSi) as a nano-Si structure with direct bandgap is a good counterpart
with limitations on UV detection. Electrochemical etching is usually used to form a
porous structure in silicon [47]. It has been found that wider PSi bandgap, con-
trolled as a nanostructure, is favored for good band alignment between QDs and
PSi. It facilitates the transfer of photo-induced electrons from the QDs to PSi. In
particular, good band alignment between QDs (1.91 eV) and PSi (1.77 eV) was
17 QDs of Wide Band Gap II–VI Semiconductors Luminescent Properties… 413

achieved by Chou et al. [17], which contributed to the improvement of optoelec-


tronic device performance. CdSe/CdS/ZnS QDs in (Chou et al. [17] were synthe-
sized by chemical reaction and incorporated into PSi layer, prepared by anodic
etching of a p-type silicon substrate. Das et al. [21] used the same approach to fab-
ricate UV-visible PDs based on a ZnS-PSi:p-Si hybrid heterostructure.

17.6 QDs-Polymer Hybrid Strcutures

QDs-polymer hybrid strcutures are a kind of analogue of inorganic heterojunctions,


in which one of the components is organic. At present, in the development of pho-
todetectors based on QDs-polymer hybrid strcutures, and as an organic semicon-
ductor, such polymers as poly(3,4-ethylenedioxythiophene): poly(styrene sulfonate)
(PEDOT: PSS) [51, 90], cellulose [73], PQT-12 polymer [50], 6,13-Bis(triisopropy
lsilylethynyl)- pentacene (TIPS-pentacene) [91], and poly-(3-hexylthiophene-­2,5-
diyl) (P3HT) [51] have been tested. Quantum dots of CdS, ZnS, CdSe and CdSe/
CdS, CdS/ZnS and PbS/CdS core-shell QDs were used. When developing solar
cells with different polymers and quantum dots, even more II-VI compounds
were tested.
The formation of hybrid heterostructures was usually carried out by layer-by-­
layer deposition of polymers and QDs using a low-cost spin coating method. It is
important to keep in mind that the manufacture of a multilayer structure in this way
requires the use of orthogonal solvents to ensure the integrity of the underlying
layer when applying the upper layers. A typical structure of photodetectors based on
QDs-polymer hybrid strcutures is shown in Fig. 17.4.

Fig. 17.4 Device structure and TEM images of the all solution-processed SWIR photodetectors. a
Schematic device structure, b cross-sectional (scale bar, 50 nm). (Reprinted from Ref. [51].
Published 2020 by Springer as open access)
414 M. Abdullah et al.

In such structures, metal oxides such as an indium tin oxide (ITO), TiO2 or ZnO
are used as transparent conductive contacts and electron transport layers, QDs of
II-VI semiconductors act as an absorbing photoactive layer, and the polymer pro-
vides separation of photogenerated carriers and acts as the hole transport layer.
Metal oxides also serve as a kind of protective coating that improves the stability of
photodetectors.
Using the considered approach, photodetectors were developed both for SWIR
[51] and for the visible spectral region [50, 90, 91]. At the same time, the developed
photodetectors have acceptable characteristics. For example, the solution-­
processable PbS/CdS core-shell QD-based SWIR photodetector exhibited a high
on/of (light/dark) ratio of 11, high detectivity of 4.0 × 1012 Jones, fast response
(110 ms) and fall (133 ms) times [51]. CdSe QDs-PQT-12 hybrid structure-based
PD developed by [50] showed band-pass response over the visible spectrum with a
sharp cutoff for higher wavelengths at ∼610 nm. The maximum responsivity and the
detectivity of the self-powered photodetector was ∼3.3 mA/W and 5.4 × 109
cmHz1/2 W−1, respectively, at a wavelength of ∼420 nm under the optical power
density of ∼130 μW/cm2. The rise time and fall time of the device were found to be
∼12 and ∼15 ms, respectively. Yang et al. [91] reported that the pentacene/CdSe@
ZnS QDs composite device exhibited excellent electrical and optical properties with
current switch ratio Ion/Ioff of 104 (λ = 513 nm), photosensitivity (P) of 105, respon-
sivity R of 0.33 mA/W and detectivity (D) of 1.5 × 1011 Jones at drain voltage of
~35 V and light intensity of 1.6 mW/cm2. Yang et al. [91] believe that the pentacene/
CdSe@ZnS QDs hybrid strusture could be one of the most promising candidates for
channel transport layer materials for photodetectors.

17.7 Photodetectors Based on QDs-Polymer Composites

QDs-polymer composites is another interesting combination used for improvement


performance of photodetectors. The use of QDs-polymer composites allows us to
solve two problems. The first is to improve the stability of QDs. As is known, QDs
do not have stable parameters and their properties change during contact with the
environment. For example, Xiao et al. [101] established that the photocurrent den-
sity of the pure CdS QD-based photodetector exhibited a big drop of more than 80%
in 24 hours.
QDs in solution are quite stable formations. QDs are usually synthesized in solu-
tion containing surface ligands with a headgroup tethered to the QD surface and a
hydrocarbon tail directed away. The surface ligands act as stoppers to control the
QD growth and surfacepassivating ligands prohibit the formation of dangling bonds.
Commonly used ligand head groups are carboxyl, amino, thiol, phosphate, etc. [80].
During the formation of films, QDs usually lose ligands, which is accompanied by
the formation of defects that promote oxygen adsorption, oxidation of QDs, and
affect the transport of carriers between them. Shell growth, i.e., the formation of
core-shell structures, which was discussed in the previous sections, is an effective
17 QDs of Wide Band Gap II–VI Semiconductors Luminescent Properties… 415

strategy to mitigate the effects of ligand loss. Polymers can also have the same func-
tion. To improve film stability and conductivity, QDs are typically dispersed in a
conductive organic matrix that acts as a medium for charge carrier transport. In
addition, organic molecules grafted onto the surface of QDs can protect them from
aggregation and enhance solubility in non-polar solvents, facilitating the fabrication
of uniform and stable QD-based films.
The second task is to improve the photoelectric characteristics of polymer opto-
electronic devices [82]. In recent years, flexible, low-power, and low-cost polymer-­
based devices have been increasingly used [84]. But, polymers do not belong to
materials with high absorption coefficient and high quantum efficiency of radiation.
At the same time, the QDs of II-VI compounds have these parameters. Therefore,
the combination of these materials in one device is quite logical.
QDs-polymer composites can be prepared using various approaches. To incorpo-
rate CdS QDs into the polypyrrole (PPy) polymer network, the synthesized CdS
QDs were added during the polymerization of PPy. The electrospray method was
used to deposit the PPy/CdS QD composite onto the substrate [8]. ZnS quantum
dots were incorporated in polyacrylamide by adding an aqueous suspension of ZnS
sol in anacrylamide:bisacrylamide copolymer with following polymerasation [63].
To prepare the CdTe/CdS core-shell QD-polymethyl methacrylate (PMMA) com-
posite, PMMA powder was dissolved in chloroform, then QDs were added to the
PMMA solution and sonicated for 1 hour. The mixture was then heated to evaporate
the solvents [28]. Films were prepared using spin-coating method. The CdSe/PVA
nanocomposite thin films were deposited on chemically clean glass substrate by
reacting Cd2+ dispersed PVA with sodium selenosulphite [70]. The final solution
containing Cd2+ Cd2+ and Se2− ions in the polymeric matrix was coated onto chemi-
cally clean glass substrate by dip coating technique and then subjected to thermoly-
sis at 300∘ C. For incorporation of CdSe/ZnS QDs in PMMA, the PMMA was
stirred with QDs in a chloroform solution to form polymer. Then the polymer was
paved on a quartz substrate, and kept drying for 72 hours to obtain a uniform film
with hundred micrometers thickness [94]. To form a P3HT-CdTe QDs composite, a
mixed solution of P3HT and CdTe QDs was spin-coated onto a hole transport layer
coated with indium tin oxide (ITO) glass [88]. Subsequently, the films were annealed
in a solvent in an atmosphere of 1,2-dichlorobenzene (DCB) for 8 hours. However,
Stiff-Roberts et al. [82] noted that one of the biggest problems with these methods
of forming QDs-polymer composites is the inability of many solution-based meth-
ods to control the internal QDs morphology in the polymer bulk film. For example,
they observed that drop-cast and spin-cast CdSe/polymer hybrid nanocomposite
films demonstrated the aggregation of CdSe CQDs into micrometre scale clusters
inside the polymer bulk. Furthermore, the CdSe CQD concentration demonstrated a
strong spatial dependence on the distance from the film centre. Stiff-Roberts et al.
[82] believe that resonant IR matrix-assisted pulsed laser evaporation (RIR-MAPLE)
has more potential to control the internal and surface morphologies of conjugated
polymers, as well as the distribution of QDs in hybrid nanocomposites. However,
this method does not belong to the simple and cheap methods of film formation,
which drastically limits the possibilities of its use.
416 M. Abdullah et al.

Experiment has shown that polymer-QDs composites can find the widest appli-
cation in the development of various detectors. Phukan and Saikia [70] found that a
thin film of CdSe/PVA nanocomposite is suitable for use as a window layer in the
fabrication of a CdSe/CdTe solar cell. NIR PDs with good responsivity at 850 nm
based on PPy doped with CdS QDs were fabricated by Amiri and Alizadeh [8]. The
PDs were prepared with a simple and cost-effective method and had planner struc-
ture. The PPy/CdS QDs (22 ppm) PD exhibited a high overall performance with a
photoresponsivity of 3.8 mA/W, on/off ratio of 120, external quantum effi ciency
(EQE) of 560% and specific detectivity of up to 2.1x1016 Jones. According to Amiri
and Alizadeh [8], the increase mobility of charge carriers in PPy/CdS QDs was
indicated as one of the most important factors for enhance sensitivity of PDs. Wei
et al. [88] have shown that by selective passivating deep traps on CdTe QDs surfaces
with P3HT, the photodetectors maintain a high gain of 50 but have a 25,000 times
shorter response time of 2 μs compared to photodetectors based on CdTe QDs. The
high specific detectivity of approaching 1013 Jones in the UV–vis spectral range
makes it possible to directly detect weak light with an intensity of less that
1 pW cm−2. The Fig. 17.5 shows the PD structure and the EQE spectra of UV-visible
photodetectors based on P3HT-CdTe QDs composite developed by Wei et al. [88].
Feizi and Zare [28] developed a new gamma sensor based on CdTe/CdS-PMMA
composite. The results showed that the dose rate response of the prepared sensor
was linear in the dose rate range of 50–145 mGy/min (see Fig. 17.6a). Yang et al.
[92] suggested using CdSe/ZnS core–shell QDs-polyaniline nanocomposites to
develop a broadband photodetector based on a single nanowire (NW) (see
Fig. 17.6b). The photodetector showed excellent light response in the spectral range
of 350 nm to 700 nm. Moreover, the spectral range could be tuned via the size
change of QDs. The external quantum efficiency value reaches up to 106, the respon-
sivity value reaches up to 105 A/W, and the response time value is down to 8 ms.

Fig. 17.5 (a) The device structure of P3HT:CdTe nanocomposite photodetectors; (b) The EQE
spectra at different reverse bias for the P3HT:CdTe device. (Reprinted with permission form Ref.
[88]. Copyright 2015: Wiley)
17 QDs of Wide Band Gap II–VI Semiconductors Luminescent Properties… 417

Fig. 17.6 (a) Dose rate–response curve of CdTe/CdS-PMMA nanocomposite-based gamma


detector. Reprinted with permission from [28]. Copyright 2018: Springer. Fig. 5; (b) Schematic
diagram of the CdTe/CdS-PMMA NW-based photodetecror. (Reprinted with permission from Ref.
[92]. Copyright 2016: RSC)

The presented results indicate that QDs-polymer composites are indeed of inter-
est for the development of photodetectors. However, it must be recognized that the
problem of long-term stability of the parameters of QDs-polymer composites detec-
tors during operation, thermal exposure and UV irradiation has not been completely
resolved.

17.8 PDs Based on 0D-2D Hybrid Structures

In the last decade, great hopes in the development of electronics and optoelectronics
are associated with the use of 2D materials. The term 2D materials refers to crystal-
line solids consisting of a single layer of atoms. Of all known 2D materials, gra-
phene is the most studied. While the low-cost, low toxicity, biocompatibility,
chemical inertness, stretching, wearable flexibility, high conductivity, high carrier
transport characteristics, and broad absorption bandwidth make graphene PDs at the
top of materials used for optoelectronics and photodetection applications, their
absorption rate is only 2.3%, originates from its gapless property limiting their
response and then reducing graphene applications [56, 83].
The experiment and calculations showed that the most effective way to improve
the parameters of graphene-based photodetectors is the hybridization of graphene
with quantum dots such as QDs of II-VI semiconductors. In these graphene-QDs
hybrid devices: graphene offers high carrier transport conductive channels, while
the QDs of II-VI semiconductors work as a light-harvesting part, where photoelec-
tric conversion occurs by providing high absorption. When a QD absorbs a photon,
the created electron-hole pair separates at the graphene-QD interface under the
built-in field resulting from their different work functions. This contributes to a
decrease in the resistance of graphene and the appearance of a photoresponse, which
418 M. Abdullah et al.

is much higher than the photoresponse caused by the absorption of radiation by


graphene itself [6, 56, 83].
Graphene-QDs hybrid structures can be fabricated using various approaches. For
example, the hybrid structure of graphene-CdSe/CdS/ZnS QDs (core/multishell)
was fabricated by the Langmuir-Blodgett method [6]. It has been shown that the
transfer of photogenerated carriers from QDs to graphene in hybrid graphene-QDs
structure indeed improves the optical properties of graphene and provides superior
light absorption, helping to improve photocurrent performance and achieve good
detection performance [6]. Kan et al. [44] also used hybrid heterostructure for fab-
rication of deep ultraviolet (DUV) photodetectors. They proposed a novel method-
ology to construct a hybrid zero−/two-dimensional DUV photodetector (p-type
graphene/ZnS QDs/4H-SiC) with photovoltaic characteristics. The schematic illus-
tration of the preparation process is shown in Fig. 17.7. Here, the single-layer p-type
graphene (p-Gr), ZnS QDs film and 4H-SiC single crystal substrate are constructed
into a PIN junction. The device exhibited excellent selectivity for the DUV light and
has an ultrafast response speed (rise time: 28 μs and decay time: 0.75 ms), which are
much better than those reported for conventional photoconductive photodetectors.
The same effect was achieved after the fabrication of the hybrid graphene-­CuInS2/
ZnS core/shell QDs structure [68]. In hybrid graphene-CuInS2/ZnS QDs PD gra-
phene was fabricated by CVD [55, 56], and CuInS2/ZnS QDs were synthesized via
hot injection. The direct semiconductor CuInS2 QDs emits at 680 nm but is chemi-
cally unstable. Introducing ZnS shell solves the stability problem of CuInS2 QDs and
improves the PL. The emission wavelength for this core/shell QD structure is extended
to 610–760 nm. In addition, this ternary structure has a long PL lifetime. These merits
make CuInS2/ZnS QDs adequate for application in PDs. According to [56, 68], the
high response of hybrid graphene-CuInS2/ZnS QDs is achieved due to the long life-
time of carriers of CuInS2/ZnS QDs and high mobility due to single-­layer graphene.
There are also reports of a flexible, friendly environment, large area, and cost-­
effective UV PDs with a high response based on hybrid structures such as graphene-­
ZnSe/ZnS core/shell QDs [83] and hybrid graphene-CdS QDs [14, 81]. It is believed
that in the hybrid graphene-nanostructure PDs, the high photoconductive gain
results from the ratio of the nanocrystal (long) carrier lifetime to the (short) gra-
phene transit time. Nevertheless, the slow recombination limits devices-speed below

Fig. 17.7 Fabrication procedure of the DUV photodetector (p-type graphene/ZnSQDs/4H-SiC.


(Reprinted with permission from Ref. [44]. Copyright 2019: ACS)
17 QDs of Wide Band Gap II–VI Semiconductors Luminescent Properties… 419

a 100 Hz. However, Spirito et al. [81] showed that it is possible to synthesize hybrid
graphene-CdS QDs UV-PDs with a fast photocurrent decay time, which make it
possible to detect repetition rates above the required limit, for example, during
video recording.
It is important to note that the same technique can be used to improve the param-
eters of photodetectors based on other 2D nanomaterials such as ReS2 [66] and
antimonene []. For example, Qin et al. [66] suggested using this approach when
developing photodetectors based on 2D transition metal dichalcogenide ReS2. ReS2
is considered as a promising candidate for optoelectronic applications due to its
direct bandgap character and optical/electrical anisotropy. However, the narrow
spectrum and the low absorption of ReS2 limits its optoelectronic applications. At
the same time it is known that CdSe/CdS/ZnS multi-core/shell QDs are stable and
have high absorption in UV-visible spectral range. Qin et al. [66] fabricated the
hybrid CdSe/CdS/ZnS multi-core/shell QD-ReS2 film heterostructure using process
shown in Fig. 17.7. The spin-coated process fabricates the QDs while the thick ReS2
films are prepared by CVD and then exfoliated into a monolayer using a novel
ultrasonic-assisted exfoliation approach. It was established that under 589 nm laser
irradiation, the responsivity of ReS2 phototransistor decorated with II-VI quantum
dots could be enhanced by more than 25 times (up to ~654 A/W), and the rising and
recovery time can be also reduced to 3.2 and 2.8 s, respectively. According to Qin
et al. [66], the excellent optoelectronic performance of developed PD is originated
from the coupling effect of quantum dots light absorber and cross-linker ligands
1,2-ethanedithiol. Photoexited electron-hole pairs in quantum dots can separate and
transfer efficiently due to the type-II band alignment and charge exchange process
at the interface. Thus, the built-in field at the interface prevents the photogenerated
carriers in QDs from recombination.
The responsivity and the on-off ratio of some hybrid-based PDs are listed in the
Table 17.3.

Table 17.3 Photoresponsivity and the on-off ratio of hybrid-based ZnS, ZnSe, and CdS
photodetectors
Photodetectors Responsivity (R) (A/W) λ (nm) Ref.
Hybrid ZnS-PSi:p-Si 0.9–1.1 365, 400 [21]
Hybrid CdSe/CdS/ZnS QDs-PSi MSM 0.24 ~900 [17]
Graphene-CdS QDs 40 450 [14]
Graphene-CdS nanocrystals 3.4 × 104 349 [81]
Graphene-CdSe/CdS/ZnS QDs 46 365 [6]
Graphene-ZnSe/ZnS core/shell QD 2 × 103 405 [83]
Graphene-CuInS2/ZnS core/shell QDs 2.5 × 105 650 [68]
Graphene-CuInS2/ZnS QDs 35 660 [56]
p-type graphene/ZnS QDs/4H-SiC 0.29 mA/W 250 [44]
ReS2-CdSe/CdS/ZnS multicore/shell QDs 654 589 [66]
Antimonene-CdS QDs 10 μA/W 700 [97]
420 M. Abdullah et al.

Fig. 17.8 Lifetime characteristics of SWIR PD1 (PbS QDs-polymer) and SWIR PD3 (PbS/CdS
QDs-polymer) without encapsulation, under constant voltage operation −1 V at room temperature.
(Reprinted from Ref. [51]. Published 2020 by Springer as open access)

17.9 Outlook

This chapter discusses various photodetectors based on QDs of wide band gap II-VI
compounds. It is shown that these QDs are indeed a promising materials for creating
efficient photodetectors for various purposes. No doubts, there are problems that
require additional research and hinder the widespread use of QD-based PDs. The
main problem is the low parameter stability of QDs-based photodetectors. For
example, Kwon et al. [51] studied unencapsulated SWIR PDs, developed using PbS
QDs (PD1) and PbS/CdS core-shell QDs-based hybrid structures (see Fig. 17.1),
and established that the lifetime of PbS QDs-based photodetectors (PD1) is only
32 h (Fig. 17.8). The use of PbS/CdS core-shell structures (PD3) improves the sta-
bility of photodetectors. The increase in device stability was attributed to the pas-
sivating characteristics of the thick CdS shell, which served as a physical barrier to
oxygen and moisture penetration. But even in this case lifetime was only 182 h. Of
course, this is a very short lifetime of detectors intended for the market. But there is
hope that this problem will be solved.

Acknowledgments G. Korotcenkov is grateful to the State Program of the Republic of Moldova,


project 20.80009.5007.02, for supporting his research.

References

1. Abou Elhamd AR, Al-Sallal KA, Hassan A (2019) Review of core/shell quantum dots tech-
nology integrated into building’s glazing ew of core/shell quantum dots technology integrated
into building’s glazing. Energies 12:1058
17 QDs of Wide Band Gap II–VI Semiconductors Luminescent Properties… 421

2. Alaie Z, Nejad SM, Yousefi MH (2015) Recent advances in ultraviolet photodetectors. Mater
Sci Semicond Proces 29:16–55
3. Al-Haddad RM, Ali IM, Ibrahim IM, Ahmed NM (2015) Photoconductivity and performance
of Mn2+ and Ce3+ doped ZnS quantum dot detectors. Eng Tech J B 33:1608–1618
4. Al-Khursan AH (2006) Gain of excited states in the quantum-dots. Phys E 35:6–8
5. Al-Khursan AH (2005) Intensity noise characteristics in quantum-dot lasers: four-level rate
equations analysis. J Luminescen 113:129–136
6. Al-Alwani AJK, Chumakov AS, Shinkarenko OA, Gorbachev IA, Pozharov MV, Venig S,
Glukhovskoy EG (2017) Formation and optoelectronic properties of graphene sheets with
CdSe/CdS/ZnS quantum dots monolayer formed by langmuir-schaefer hybrid method. Appl
Surf Sci 424:222–227
7. Al-Mossawi MA, Al-Shatravi AG, Al-Khursan AH (2012) CdSe/ZnSe quantum-dot semicon-
ductor optical amplifiers. Insciences J 2:52–62
8. Amiri M, Alizadeh N (2020) Highly photosensitive near infrared photodetector based on
polypyrrole nanoparticle incorporated with CdS quantum dots. Mater Sci Semicond Proces
111:104964
9. Bang J, Park J, Lee JH, Won N, Nam J, Lim J et al (2010) ZnTe/ZnSe (Core/Shell) type-II
quantum dots: their optical and photovoltaic properties. Chem Mater 22:233–240
10. Benyahia K, Djeffal F, Ferhati H, Bendjerad A, Benhaya A, Saidi A (2021) Self-powered
photodetector with improved and broadband multispectral photoresponsivity based on ZnO-­
ZnS composite. J Alloys Comp 859:158242
11. Bratskaya S, Sergeeva K, Konovalova M, Modind E, Svirshchevskaya E, Sergeev A et al
(2019) Ligand-assisted synthesis and cytotoxicity of ZnSe quantum dots stabilized by N-(2-­
carboxyethyl) chitosans. Colloids Surf B: Biointerfaces 182:10342
12. Bursa B, Rytel K, Skrzypiec M, Prochaska K, Wróbel D (2018) Thin film of CdTeSe/ZnS
quantum dots on water subphase: thermodynamics and morphology studies. Dyes Pigments
155:36–41
13. Cao J, Jiang Z-J (2020) Thickness dependent shell homogeneity of ZnSe/CdSe core/shell
nanocrystals and their spectroscopic and electron and hole transfer dynamics properties. J
Phys Chem C 124:12049–12064
14. Chan Y, Dahua Z, Jun Y (2020) Fabrication of hybrid graphene/CdS quantum dots film with
the flexible photo-detecting performance. Phys E 124:114216
15. Chen C, Zhu B, Zhang X, Gao Y, Wang G, Gu Y (2018) Synthesis and enhanced third-order
nonlinear optical effect of ZnSe/graphene composites. AIP Adv 8:065025
16. Chen X, Liu W, Zhang G, Wu N, Shi L, Pan S (2015) Efficient photoluminescence of Mn2+-
doped ZnS quantum dots sensitized by hypocrellin A. Adv Mater Sci Eng 2015:412476
17. Chou C, Cho H, Hsiao VKS, Yong KT, Law WC (2012) Quantum dot-doped porous silicon
metal–semiconductor metal photodetector. Nanoscale Res Lett 7:291
18. Chuang SL (2009) Physics of photonic devices. Wiley, New York
19. Cui JY, Li KY, Ren L, Zhao J, Shen TD (2016) Photogenerated carriers enhancement in
Cu-doped ZnSe/ZnS/L-cys self-assembled core-shell quantum dots. J Appl Phys 120:184302
20. Dakhil T, Abdulalmuhsin SM, Al-Khursan AH (2018) Quantum efficiency of CdS quantum
dot photodetectors. Micro Nano Lett 13:1185–1187
21. Das M, Sarmah S, Sarkar D (2021) Distinct and enhanced ultraviolet to visible ZnS-porous
silicon (PS):p-Si hybrid metal-semiconductor-metal (MSM) photodetector. Mater Today
Proc 46:247–6252
22. Dedong H, Ying-Kai L, Yu DP (2015) Multicolor photodetector of a single Er3+-doped CdS
nanoribbon. Nanoscale Res Lett 10:285
23. Deng J, Lv W, Zhang P, Huang W (2022) Large-scale preparation of ultra-long ZnSe-PbSe
heterojunction nanowires for flexible broadband photodetectors. J Sci Adv Mater Dev
7:100396
24. Dickson RE, Hu MZ (2015) Chemical synthesis and optical characterization of regular and
magic-sized CdS quantum dot nanocrystals using 1-dodecanethiol. J Mater Res 30:890–895
422 M. Abdullah et al.

25. Dorfs D, Hickey SG, Eychmüller A (2010) Type-I and type-II core-shell quantum dots:
synthesis and characterization. In: Kumar CSSR (ed) Semiconductor nanomaterials. Wiley-­
VCH, Weinheim, pp 331–366
26. El-Toni AM, Habila MA, Labis JP, ALOthman ZA, Alhoshan M, Elzatahry AA, Zhang F
(2016) Design, synthesis and applications of core–shell, hollow core, and nanorattle multi-
functional nanostructures. Nanoscale 8:2510–2531
27. Fang X, Bando Y, Liao M, Gautam UK, Zhi C, Dierre B et al (2009) Single-crystalline ZnS
nanobelts as ultraviolet-light sensors. Adv Mater 21:2034–2039
28. Feizi S, Zare H (2018) Masoumeh hoseinpour investigation of dosimetric characteristics of a
core–shell quantum dots nano composite (CdTe/CdS/PMMA): fabrication of a new gamma
sensor. Appl Phys A Mater Sci Process 124:420
29. Fernández-Delgado N, Herrera M, Tavabi AH, Luysberg M, Borkowski RED, Cantóc PJR
et al (2018) Structural and chemical characterization of CdSe-ZnS core-shell quantum dots.
Appl Surf Sci 457:93–97
30. Gou G, Dai G, Qian C, Liuc Y, Fu Y, Tianb Z et al (2016) High-performance ultraviolet pho-
todetectors based on CdS/CdS:SnS2 superlattice nanowires. Nanoscale 8(30):14580–14586
31. Grandhi GK, Krishna M, Mondal P, Viswanatha R (2020) Cation co-doping into ZnS quan-
tum dots: towards visible light sensing applications. Bull Mater Sci 43:301
32. Gu Y, Tang L, Guo X (2019) Preparation and photoelectric properties of cadmium sulfide
quantum dots. Chinese Phys B 28:047803
33. Guidelli EJ, Lignos I, Yoo JJ, Lusardi M, Bawendi MG, Baffa O, Jensen KF (2018)
Mechanistic insights and controlled synthesis of radioluminescent ZnSe quantum dots using
a microfluidic reactor. Chem Mater 30:8562–8570
34. Guo P, Xu J, Gong K (2016) On-nanowire axial heterojunction design for high-performance
photodetectors. ACS Nano 10:8474–8848
35. Hafeez M, Al-Asbahi BA, Hj Jumali MH, Yahaya M, Inam F, Bhopal MF, Bhatti AS (2020)
Critical role of defect states on visible luminescence from ZnS nanostructures doped with Au,
Mn and Ga. Mater Sci Semicond Proces 117:105193
36. He L, Yang L, Liu B, Zhang J, Zhang C, Liu S et al (2019) One-pot synthesis of color-­
tunable copper doped zinc sulfide quantum dots for solid-state lighting devices. J Alloys
Comp 787:537–542
37. Huang W, Yuan Z, Ren Y, Zhou K, Cai Z, Yang C (2019) Interface optical phonons and its
ternary mixed effects in ZnS/CdxZn1-xS quantum dots. Phys E 108:60–67
38. Husham M, Hassan Z, Selman AM (2016) Synthesis and characterization of nanocrystalline
CdS thin films for highly photosensitive self-powered photodetector. Eur Phys J Appl Phys
74:10101
39. Ippen C, Greco T, Kim Y, Kim J, Oh MS, Han CJ et al (2014) ZnSe/ZnS quantum dots as
emitting material in blue QD-LEDs with narrow emission peak and wavelength tenability.
Organic Electron 15:126–131
40. Jiang T, Huang Y, Meng X (2020) CdS core-Au/MXene-based photodetectors: positive deep-
­UV photoresponse and negative UV–Vis-NIR photoresponse. Appl Surf Sci 513:145813
41. Jamshidi A, Yuan C, Chmyrov V, Widengren J, Sun L, Ågren H (2015) Efficiency enhanced
colloidal Mn-doped type II Core/Shell ZnSe/CdS quantum dot sensitized hybrid solar cells.
J Nanomater 2015:921903
42. Jbara AS, Abood HI, Al-Khursan AH (2012) Effect of doping and in-composition on gain of
long wavelength III-nitride QDs. J Opt 41:214–223
43. Ji B, Koley S, Slobodkin I (2020) ZnSe/ZnS core/shell quantum dots with superior optical
properties through thermodynamic shell growth. Nano Lett 20:2387–2395
44. Kan H, Zheng W, Lin R, Li M, Fu C, Sun H et al (2019) Ultra-fast photovoltaic-type deep-­
ultraviolet photodetector using hybrid zero−/two dimensional hetero junctions. ACS Appl
Mater Interfaces 11(8):8412–8418
45. Kaushik S, Singh R (2021) 2D layered materials for ultraviolet photodetection: a review. Adv
Optic Mater 9(11):2002214
17 QDs of Wide Band Gap II–VI Semiconductors Luminescent Properties… 423

46. Khani O, Rajabi HR, Yousefi MH, Khosravi AA, Jannesari M, Shamsipur M (2011) Synthesis
and characterizations of ultra-small ZnS and Zn(1−x)FexS quantum dots in aqueous media and
spectroscopic study of their interactions with bovine serum albumin. Spectrochim Acta A
79:361–369
47. Korotcenkov G (ed) (2015) Porous silicon: from formation to application. Vol. 1: formation
and properties. Taylor and Fracis Group/CRC Press, Boca Raton
48. Korotcenkov G (2014) Handbook of gas sensor materials properties, advantages and short-
comings for applications, New trends and technologies, vol 2. Springer, New York
49. Kumar V, Rawal I, Kumar V, Goyal PK (2019) Efficient UV photodetectors based on Ni-doped
ZnS nanoparticles prepared by facial chemical reduction method. Physica B 575:411690
50. Kumar H, Kumar Y, Rawat G, Kumar C, Mukherjee B, Pal BN, Jit S (2017) Colloidal CdSe
quantum dots and PQT-12-based low-temperature self-powered hybrid photodetector. IEEE
Photon Technol Lett 29(20):1715–1718
51. Kwon J-B, Kim S-W, Kang B-H, Yeom S-H, Lee W-H, Kwon D-H et al (2020) Air-stable
and ultrasensitive solution-cast SWIR photodetectors utilizing modifed core/shell colloidal
quantum dots. Nano Convergence 7:28
52. Lad AD, Rajesh C, Khan M, Ali N, Gopalakrishnan IK, Kulshreshtha SK, Mahamuni S
(2007) Magnetic behavior of manganese-doped ZnSe quantum dots. J Appl Phys 101:103906
53. Lee Y-J, Cha J-M, Yoon C-B, Lee S-E (2018) Study on UV opto-electric properties of
ZnS:Mn/ZnS core-shell QD. J Korean Ceram Soc 55:55–60
54. Li R, Tang L, Zhao Q, Teng KS, Lau SP (2020) Facile synthesis of ZnS quantum dots at room
temperature for ultra-violet photodetector applications. Chem Phys Lett 742:137127
55. Li Z, Xu K, Wei F (2018a) Recent progress in photodetectors based on low-dimensional
nanomaterials. Nanotechnol Rev 7:393–411
56. Li F, Guo C, Pan R, Zhu Y, You L, Wang J et al (2018b) Integration of green CuInS2/ZnS
quantum dots for high-efficiency light-emitting diodes and high-responsivity photodetectors.
Opt Mater Express 8:314–323
57. Lin H, Wei L, Wu C, Chen Y, Yan S, Mei L, Jiao J (2016) High-performance self-powered
photodetectors based on ZnO/ZnS core-shell nanorod arrays. Nanoscale Res Lett 11:420
58. Liu H, Zhong H, Zheng F, Xie Y, Li D, Wu D, Zhou Z, Sun XW, Wang K (2019) Near-infrared
lead chalcogenide quantum dots: synthesis and applications in light emitting diodes. Chinese
Phys B 28:128504
59. Lou Z, Li L, Shen G (2016) Ultraviolet/visible photodetectors with ultrafast, high photosen-
sitivity based on 1D ZnS/CdS heterostructures. Nanoscale 8:5219–5225
60. Maity P, Singh SV, Biring S, Pal BN, Ghosh AK (2019) Selective near-infrared (NIR) photo-
detectors fabricated with colloidal CdS: Co quantum dots. J Mater Chem C 7:7725
61. Memon UB, Chatterjee U, Gandhi MN, Tiwari S, Duttagupta SP (2014) Synthesis of ZnSe
quantum dots with stoichiometric ratio difference and study of its optoelectronic property.
Procedia Mater Sci 5:1027–1033
62. Miao S, Cho Y (2021) Toward green optoelectronics: environmental-friendly colloidal quan-
tum dots photodetectors. Front Energy Res 9:666534
63. Mir FA (2010) Structural and optical properties of ZnS nanocrystals embedded in polyacryl-
amide. J Optoelectron Biomed Mater 2(2):79–84
64. Mishra SK, Srivastava RK, Prakash SG, Prakash SG, Yadav RS, Panday AC (2011) Structural,
photoconductivity and photoluminescence characterization of cadmium sulfide quantum dots
prepared by a co-precipitation method. Electron Mater Lett 7:31–38
65. Nikesh VV, Lad AD, Kimura S, Nozak S, Mahamuni S (2006) Electron energy levels in ZnSe
quantum dots. J Appl Phys 100:113520
66. Qin JK, Ren DD, Shao WZ, Li Y, Miao P, Sun ZY et al (2017) Photoresponse enhancement
in monolayer ReS2 phototransistor decorated with CdSe-CdS-ZnS quantum dots. ACS Appl
Mater Interfaces 9:45
67. Pacheco ME, Castells CB, Bruzzone L (2017) Mn-doped ZnS phosphorescent quantum dots:
coumarins optical sensors. Sensors Actuators B Chem 238:660–666
424 M. Abdullah et al.

68. Pan R, Wang J (2019) Ultra-high responsivity graphene-CIS/ZnS QDs hybrid photodetector.
Proc SPIE 10843:108431B
69. Peng L, Han S, Hu X (2014) Photocurrent enhancement mechanisms in bilayer nanofilm-­
based ultraviolet photodetectors made from ZnO and ZnS spherical nanoshells. Nanoscale
Res Lett 9:388
70. Phukan P, Saikia D (2013) Optical and structural investigation of CdSe quantum dots dis-
persed in PVA matrix and photovoltaic applications. Intern J Photoenergy 2013:728280
71. Poornaprakash B, Kumar KN, Chalapathi U, Reddeppa M, Poojitha PT, Park SH (2016)
Chromium doped ZnS nanoparticles: chemical, structural, luminescence and magnetic stud-
ies. J Mater Sci Mater Electron 27:6474–6479
72. Premkumar S, Nataraj D, Bharathi G, Ramya S, Thangadurai TD (2019) Highly respon-
sive ultraviolet sensor based on ZnS quantum dot solid with enhanced photocurrent. Sci
Rep 9:18704
73. Ranjan PS, Bhuyan RK (2021) Organic polymer and perovskite CdSe–CdS QDs hybrid thin
film: a new model for the direct detection of light elements. J Mater Sci Mater Electron
32:5538–5547
74. Reddy CV, Shim J, Cho M (2017) Synthesis, structural, optical and photocatalytic properties
of CdS/ZnS core/shell nanoparticles. J Phys Chem Solids 103:209–217
75. Rosencher E, Vinter B (2002) Optoelectronics. Cambridge University Press, Cambridge
76. Saeed S, Iqbal A, Iqbal A (2020) Photo-induced charge carrier dynamics in a ZnSe quantum
dot-attached CdTe system. Proc Royal Soc A 476:20190616
77. Sahai S, Husain M, Shanker V, Singh N, Haranath D (2011) Facile synthesis and step by
step enhancement of blue photoluminescence from Ag-doped ZnS quantum dots. J Colloid
Interface Sci 357:379–383
78. Sahin O, Horoz S (2018) Synthesis of Ni: ZnS quantum dots and investigation of their prop-
erties. J Mater Sci Mater Electron 29:16775–16781
79. Senthilkumar K, Kalaivani T, Kanagesan S, Balasubramanian V (2012) Synthesis and charac-
terization studies of ZnSe quantum dots. J Mater Sci Mater Electron 23:2048–2052
80. Shang Y, Ning Z (2017) Colloidal quantum-dots surface and device structure engineering for
high-performance light-emitting diodes. National Sci Rev 4:170–183
81. Spirito D, Kudera S, Miseikis V, Giansante C, Coletti C, Krahne R (2015) UV light detection
from CdS nanocrystal sensitized graphene photodetectors at kHz frequencies. J Phys Chem
C 119:23859–23864
82. Stiff-Roberts AD, Lantz KR, Pate R (2009) Room-temperature, mid-infrared photodetection
in colloidal quantum dot/conjugated polymer hybrid nanocomposites: a new approach to
quantum dot infrared photodetectors. J Phys D Appl Phys 42:234004
83. Sun YL, Xie D, Sun MX, Teng CJ, Qian L, Chen RS et al (2018) Hybrid graphene/cadmium
free ZnSe/ZnS quantum dots phototransistors for UV detection. Sci Rep 8:5107
84. Tavasli A, Gurunlu B, Gunturkun D, Isci R, Faraji S (2022) A review on solution-processed
organic phototransistors and their recent developments. Electronics 11:316
85. Tian W, Li L, Lu H (2015) Nanoscale ultraviolet photodetectors based on one dimensional
metal oxide nanostructures. Nano Res 8:382–405
86. Vasudevan D, Gaddam RR, Trinchi A, Cole I (2015) Core-shell quantum dots: properties and
applications. J Alloys Comp 636:395–404
87. Wang A, Buntine MA, Jia G (2019) Recent advances in zinc-containing colloidal semicon-
ductor nanocrystals for optoelectronic and energy conversion applications. Chem Electro
Chem 6:1–17
88. Wei H, Fang Y, Yuan Y, Shen L, Huang J (2015) Trap engineering of CdTe nanoparticle
for high gain, fast response, and low noise P3HT:CdTe nanocomposite photodetectors. Adv
Mater 27(34):4975–4981
89. Yadav PVK, Ajitha B, Reddy YAK, Sreedhar A (2021) Recent advances in development of
nanostructured photodetectors from ultraviolet to infrared region: a review. Chemosphere
279:130473
17 QDs of Wide Band Gap II–VI Semiconductors Luminescent Properties… 425

90. Yan Y, Wu X, Chen Q, Liu Y, Chen H, Guo T (2019) High-performance low-voltage flexible
photodetector arrays based on all-solid-state organic electrochemical transistors for photo-
sensing and imaging. ACS Appl Mater Interf 11:20214–20224
91. Yang Z, Lin S, Liu J, Zheng K, Lu G, Ye B et al (2020) High performance phototransistors
with organic/quantum dot composite materials channels. Org Electron 78:105565
92. Yang X, Liu Y, Lei H, Li B (2016) An organic–inorganic broadband photodetector based on a
single polyaniline nanowire doped with quantum dots. Nanoscale 8:15529–15537
93. Yaraki MT, Tayebi M, Ahmadieh M, Tahriri M, Vashaee D, Tayebi L (2017) Synthesis and
optical properties of cysteamine-capped ZnS quantum dots for aflatoxin quantification. J
Alloys Comp 690:749–758
94. Yingming S, Pan H, Chu H, Mamat M, Abudurexiti A, Li D (2021) Core-shell CdSe/ZnS
quantum dots polymer composite as the saturable absorber at 1.3 μm: influence of the doping
concentration. Phys Lett A 400:127307
95. Xia Y, Zhai G, Zheng Z, Lian L, Liu H, Zhang D, Gao J, Zhai T, Zhang J (2018) Solution-­
processed solar-blind deep ultraviolet photodetectors based on strongly quantum confined
ZnS quantum dots. J Mater Chem C 6:11266–11271
96. Xia X, Liu Z, Du G, Li Y, Ma M (2010) Wurtzite and zinc-blende CdSe based core/shell
semiconductor nanocrystals: structure, morphology and photoluminescence. J Luminescence
130:1285–1291
97. Zeng X, Zhang J, Huang F (2012) Optical and magnetic properties of Cr-doped ZnS nano-
crystallites. J Appl Phys 111:123525
98. Zhang F, Ding Y, Zhang Y, Zhang X, Wang ZL (2012a) Piezo-phototronic effect enhanced
visible and ultraviolet photodetection using a ZnO-CdS core-shell micro/nanowire. ACSNano
6:9229–9236
99. Zhang Y, Shen Y, Wang X, Zhu L, Han B, Ge L, Tao Y, Xie A (2012b) Enhancement of
blue fluorescence on the ZnSe quantum dots doped with transition metal ions. Mater Lett
78:35–38
100. Zhao W, Liu L, Xu M (2017) Single CdS nanorod for high responsivity UV–visible photode-
tector. Adv Optic Mater 5:1700159
101. Xiao Q, Hu C-X, Wu H-R , Ren Y-Y, Li X-Y, Yang Q-Q, et al. (2020) Antimonene-based flex-
ible photodetector. Nanoscale Horizons 5(1):124–130
Chapter 18
Solution-Processed Photodetectors

Shaikh Khaled Mostaque, Abdul Kuddus, Md. Ferdous Rahman,


Ghenadii Korotcenkov, and Jaker Hossain

18.1 Introduction

The current technological revolution has been spurred by design of novel materials
that can be systematically adjusted to gain distinctive features for usage in a number
of applications. Till now, variety of researches have been carried out with II-VI
semiconductor materials as they generally encompass the entire spectrum from
ultraviolet (UV) to far infrared (IR). Besides, their electron energy band gap corre-
sponds to visible region and can be tuned with adjusting the ternary or quaternary
compounds providing the facility to fabricate optoelectronic devices for various
wavelengths [20]. Additional information one can find in the Chaps. 1, 2, 3, 4, 5, and
6 (Vol. 1).
II-VI semiconductors have been extensively investigated over a long period,
since the observation of phosphorescence from ZnS crystals in 1866, especially in

S. K. Mostaque · J. Hossain (*)


Solar Energy Laboratory, Department of Electrical and Electronic Engineering, University of
Rajshahi, Rajshahi, Bangladesh
e-mail: jak_apee@ru.ac.bd
A. Kuddus
Solar Energy Laboratory, Department of Electrical and Electronic Engineering, University of
Rajshahi, Rajshahi, Bangladesh
Graduate School of Science and Engineering, Saitama University, Saitama, Japan
M. F. Rahman
Department of Electrical and Electronic Engineering, Begum Rokeya University,
Rangpur, Bangladesh
G. Korotcenkov
Department of Physics and Engineering, Moldova State University, Chisinau, Moldova

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 427
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_18
428 S. K. Mostaque et al.

the areas of light emitters (e.g. phosphor for television CRT, electro-luminescent
cells for flat display, light emitting diodes (LEDs), Schottky contact diodes, lasers),
photo−/radiation detectors (e.g. photoconductive detectors, solar cells, thermal/far-­
infrared detectors, radiation and particle detectors), optics (e.g electro-optics,
magneto-­optics, optical switching and bistability), electronics (e.g. thin film transis-
tors, integrated optoelectronics, charge couple devices, varistors, passivation and
blocking layer), sensors, data storage and so on [20]. Thus far, research involving
II-VI band gap semiconductors have obtained significant achievements in various
applications via formation of different p-n junctions and patterns while controlling
the doping, applying suitable contacts, and performing various etching and passiv-
ation methods. Despite the fact that current formation techniques such as ion beam
sputtering, magnetron sputtering, ion beam assisted deposition, plasma source
assisted deposition, and electron beam deposition enable high performance applica-
tions, neither single coating technique or model is better compared across all cir-
cumstances [34].
Currently, there are two main directions in the development of electronic and
optoelectronic devices. The first direction is microminiaturization, aimed at the
development of monolithic integrated circuits, and the second is the search for
technological approaches that make it possible to manufacture cheap devices for
wide application. Solution-processed technology refers to the second direction in
the development of semiconductor optoelectronic devices such as solar cells and
photodetectors. Low temperature solution processed semiconductors are an
emerging class of photoactive materials that can be processed using wet chemical
methods. They are attractive from a technological point of view for several rea-
sons [10, 8, 12, 41, 42]: they can be deposited from a solution over large areas
using readily available cheap manufacturing techniques. This simple manufactur-
ing technology is processed at low temperatures and under ambient conditions
with minimal material consumption. In addition, these low cost deposition tech-
niques are compatible with (i) a wide range of versatile and uncommon substrates,
including flexible substrates, (ii) various types of auxiliary materials, which can
be used for contacts and transport layers, and (iii) contemporary sensing systems.
Modern photodetectors mainly use photodiodes based on crystalline and single-­
crystal inorganic semiconductors, such as silicon or III–V compounds. As a rule, for
the manufacture of such devices, epitaxial methods are used, which require expen-
sive equipment. Therefore, photodetectors made from solution-processed semicon-
ductors have in recent years been considered as candidates for the next generation
of light detectors [10, 8, 12].
II-VI compounds fully meet the requirements for materials that can be used in
solution-processed technology. First, nanocrystals and colloidal quantum dots of
II-VI compounds used in this technology can be synthesized by various chemical
methods. Secondly, II-VI compounds have specific optical and electrical properties.
In addition to the manufacturing benefits offered by solution-processed semicon-
ductors, II-VI semiconductors have the advantage that their optoelectronic
18 Solution-Processed Photodetectors 429

properties can be tailored. Especially appealing for photodetection is the ability to


control the semiconducting optical band gap. Compositional tuning of the II-VI
semiconductor alloys allows the modification of the semiconductor bandgap from
the visible through to the near-infrared.
In this chapter, we intend to discuss about the state of the art performance of
solution processed techniques for II-VI semiconductor compounds, especially in
the photodetector applications. Here, the focus has been given on non-oxide II-VI
wide band compounds.

18.2 Solution Processed II-VI Semiconductor-Based Solar


Cells and Photodetectors

When developing solution processed photodetectors, two approaches are used. The
first approach involves the formation of photosensitive layers of II-VI compounds
on substrates directly during their synthesis [4, 37]. These methods include chemi-
cal bath deposition (CBD), electrochemical deposition, spray pyrolysis, successive
ionic layer adsorption and reaction (SILAR) or successive ionic layer deposition
(SILD) [24, 38, 40, 49, 54]. A description of these methods for II-VI compounds
can be found in the Chap. 10 (Vol. 1). This is a fairly simple method for forming
films of II-VI compounds. But if it is necessary to deposit semiconductor films over
large areas, problems arise with their uniformity and reproducibility of their param-
eters. In addition, the properties of the films strongly depend on the deposition
parameters.
The second approach is based on the synthesis of nanocrystallites (NCs) or
colloidal particles, their separation from the solution, and subsequent use for
deposition on the substrate surface by various methods [12, 41, 42]. Typically,
nanocrystals and colloidal particles are synthesized using methods such as sol-
gel, hydrothermal, solvothermal, hot injection, and so on [16]. These methods are
described in sufficient detail in the Chap. 11 (Vol. 1). From the synthesized
nanoparticles, a paste or solution is formed, which are used to form layers of
II-VI compounds. The advantage of this approach is the ability to control the
parameters of nanoparticles at the stage of synthesis. The deposition of films of
II-VI compounds in this case is carried out using methods such as spin coating,
roll-to-roll printing, spraying, ink jet printing, drop coating, dip coating, and doc-
tor blading (Fig. 18.1). A brief description of these methods is presented in
Table 18.1.
430 S. K. Mostaque et al.

Fig. 18.1 Solution-processed materials are synthesized in the form of colloidal semiconductor
inks. These can be deposited and assembled in solid films using spray coating or spin coating, or
manufacturing techniques such as inkjet printing, doctor blading or roll-to-roll printing. (Reprinted
with permission from Ref. [12]. Copyright 2017: Springer Nature)

Table 18.1 Methods usually used for preparing films in the frame of solution-processed
technology
Method Description
Casting In this process the material to be deposited is dissolved in liquid form in a solvent.
The material can be applied to the substrate by spraying or spinning. Once the solvent
is evaporated, a thin film of the material remains on the substrate. The thicknesses
that can be cast on a substrate range all the way from a tens of nanometers to tens of
micrometers. The control on film thickness depends on exact conditions, but can be
sustained within +/−10% in a wide range. Delamination and cracking can occur if the
film is too thick. This method gives a more uniform and a more reproducible
membrane than dip coating.
Spin In the spin coating process, the substrate spins around an axis which should be
coating perpendicular to the coating area. The quality of the coating depends on the rotation
velocity, rheological parameters of the coating liquid and surrounding atmosphere.
The coating thickness varies between several hundreds of nanometers and up to 10
micrometers. Desired thickness obtained by precursor dilution, spin speed and
number of layers. Equipment similar to that of spin-coat tracks used for photoresist
deposition.
Spray Precursor is atomized to form fine aerosol which then is deposited on a wafer.
coating Deposition enhanced by electrostatic charging of aerosol. Desired thickness is
controlled by adjusting deposition time and number of layers. The coating step is
suitable for establishing an in-line process.
Slip Slip casting is a technique in which a suspension (slip) is poured into a porous mold.
casting The mold’s pores absorb the liquid, and particles are compacted on the mold walls by
capillary forces, i.e. solidify, producing parts of uniform thickness. Once dried to the
leather-hard stage, the molds are opened and the cast object removed to dry
completely before firing.
(continued)
18 Solution-Processed Photodetectors 431

Table 18.1 (continued)


Method Description
Tape The tape casting is a technique for continuous production of the films according to
casting the “Doctor-Blade-principle”. In this process a suspension of solid state particles in
an organic solvent or water, mixed together with strengthening plasticizers and/or
binders can be used. The slip is cast onto a precisely machined stone plate, on which
the carrier film is moved smoothly and without perturbations. By the doctor blade the
slurry is spread homogeneously on the surface of the tape. Drying and firing are final
stages of the actual tape forming.
Dip Dip coating techniques can be described as a process where the substrate to be coated
coating is immersed in a liquid and then withdrawn with a well-defined withdrawal speed
under controlled temperature and atmospheric conditions. The coating thickness is
mainly defined by the withdrawal speed, by the solid content and the viscosity of the
liquid. The applied coating may remain wet for several minutes until the solvent
evaporates. This process can be accelerated by heated drying. In addition, the coating
be exposed to various thermal, UV, or IR treatments for stabilization.
Screen Screen printing is a printing technique that uses a woven mesh to support an
printing ink-blocking stencil. The paste (ink) used is a mixture of the material of interest, an
organic binder, and a solvent. The attached stencil forms open areas of mesh that
transfer ink or other printable materials which can be pressed through the mesh as a
sharp-edged image onto a substrate. A roller or squeegee is moved across the screen
stencil, forcing or pumping ink past the threads of the woven mesh in the open areas.
After printing, the wet films are allowed to settle for 15–30 min to flatten the surface
and are dried. This removes the solvents from the paste. Subsequent firing burns off
the organic binder and metallic particles are reduced or oxidized and glass particles
are sintered. It can be used to print on a wide variety of substrates, including paper,
paperboard, plastics, glass, metals, fabrics, and many other materials.
Inkjet Inkjet technologies, which are based on the 2D printer technique of using a jet to
printing deposit tiny drops of ink onto substrate, are perfectly suited to controllably dispense
small and precise amounts of “liquid” to precise locations. The available inkjet
technologies include: (1) Continuous inkjet; (2) Drop on demand inkjet; (3) Thermal
inkjet; and (4) Piezo inkjet.
The “liquid” materials can encompass low to high viscosity fluids, colloidal
suspensions, frits, metallic suspensions, and almost any other material that can be
dispersed in a liquid carrier material. The carrier material can be aqueous or
non-aqueous based solvent material. When printed, liquid drops of these materials
instantly cool and solidify to form a layer of the part. Usually inkjet printing is
accompanied by thermal treatment.
Source: Data extracted from Ref. [23]

18.3 Solution-Processed Photodetectors with Direct Wet


Chemical Synthesis of Photosensitive Layers

As an example, let us consider two methods for the synthesis of II-VI semiconduc-
tor films, which are the most commonly used in the manufacture of solution-treated
photodetectors.
432 S. K. Mostaque et al.

18.3.1 Chemical Bath Deposition (CBD)

Chemical bath deposition is an old method for fabrication of photo detectors and a
still a popular method for thin film deposition because of simplicity, scalability, low
temperature, and less energy consumption [38, 40, 54]. Hou et al. [16] believe that
thin film of II-VI compounds deposited by soft chemical solution method can be use-
ful in production of large area thin films with low cost. In CBD method, the substrate
is dipped into the solution consisting chalcogen source material, a metal ion, an extra
base and a complexing agent. The method is based on the delayed release of chalco-
genide ions into an alkaline mixture, with occurrence of film formation during the
surpassing of ionic product (IP) to the solubility product (SP).
Chemical bath deposition method is acceptable for all II-VI compounds. In this
process ZnS can be synthesized using zinc sulfate heptahydrate (ZnSO4⋅7H2O) as
the targeted ion source for Zn, and thiourea (CS(NH2)2) served as an ion source of S
[38, 40]. 0.03 M ZnSO4⋅7H2O and 0.5 M (CS(NH2)2) with 2.5 M of ammonia were
continuously being stirred at room temperature and kept in a water bath. The sub-
strates were dipped into solution and the temperature was then raised to 60 °C with
continuous stirring to ensure pH at around 9.8. Keeping the substrates for 45 min
resulted deposition of films, which were finally washed with distilled water to get
rid of porous ZnS. Consecutive drying resulted uniform ZnS thin film which is
ready to be annealed. The annealed films were more homogeneous and came out
with less cracks.
To deposit thin films of ZnTe, solid tellurium oxide, TeO2, and zinc chloride,
ZnCl2, can be used as a source of Te and Zn, respectively [54]. While ZnCl2 was
melted with distilled water, both HCl and distilled water took part in melting TeO2.
The precipitation solution was prepared with mixing all of these compounds with a
magnetic stirrer at 70 °C. With mixing the following reactions occurred:

TeO2  4 HCl  TeCl4  2 H 2O (18.1)

and

ZnCl2  TeCl4  ZnTe  Cl2   H 2O  (18.2)

After the deposition of films, the precipitation was done for different times and at
different temperatures. Though the amount of time did not affect the film perfor-
mance, the film prepared with 85 °C exhibited the lowest energy gap.
Likewise, CdSe film, specially studied for optoelectronic and photodetecting
applications, can be prepared with chemical bath deposition technique [14, 17].
Example of such process is shown in Fig. 18.2.
During deposition time, temperature, pH of solution and concentration of the
metal ions played the key role in above method. The proper optimization was pos-
sible by controlling the emancipation of Cd2+ and Se2− ions. For regulation of
18 Solution-Processed Photodetectors 433

Sodium Final pH: 10


Water bath
Selenosulphate temperature SEM
(dropwise) (°C)
d
p.
m
Glass slide te
m 40
Mag. Stirrer oo
R
50
60
Complexing pH stabilizer
agent (TEA) 70
(NH4OH)

80
Substrate
catcher As-grown
Cadmium on glass slides
acetate
Water bath Annealed
(400 °C)
Mag. Stirrer

Fig. 18.2 (a) The procedure for making precipitation solution (b) Samples after deposition on
glass slides at different constant precipitation temperature (c) CdSe film after annealing at 400 °C
and (d) SEM image. (Adapted from Ref. [17]. Published 2021 by Frontiers Media as open access)

cadmium hydrolysis, triethanolamine (TEA) worked as the complexing agent here.


The following reactions took place for the preparation of precipitation solution.

2
Cd 2   TEA  Cd  TEA   (18.3)

NH3  H 2O  NH 4   OH  (18.4)

2
Cd  TEA    Na2 SeSO3  2OH   CdSe  Na2 SO4  TEA  H 2 O (18.5)

In this process, smaller peaks at XRD tends to get larger with deposition tempera-
ture and the best result arrived between 75 and 80 °C, exhibited in Fig. 18.3a.
Though XRD small peaks started exhibiting at 25 °C annealing, an elevated (111)
peak for the particular case arrived at 450 °C because of reduced crystallite size with
annealing temperature. The transmittance data in Fig. 18.3b reveals that with higher
deposition and annealing temperatures, the transmittance falls. Accordingly, the
band gap values get decreased and it has been found that the band gap values for
50 °C, 70 °C and 80 °C falls in the region of visible spectrum. Thus, the film can be
suitably applied for photodetection in the visible region of solar spectrum.
Chemical bath deposition for the formation of high photosensitive CdSe films in
Ag/CdSe/ITO-based photodetectors was also used in [17]. Figure 18.4a represents
the setup for chemical bath deposition where ITO coated glass substrate has been
emerged. The 50 mL bath has been filled with 10 mL 0.1 M CdCl2⋅2H2O, 3.5 mL of
30% NH3 aqueous, 10 mL solution of Na2SeSO3 and distilled water for laboratory
purpose [39, 40]. Figure 18.4b represents the experimental setup for measuring
434 S. K. Mostaque et al.

Fig. 18.3 (a) Characterization of CdTe thin films deposited at (a) 50 °C, (b) 70 °C, (c) 80 °C and
(d) 50 °C (annealed at 400 °C), (b) Transmittance plot. (Adapted from Ref. [17]. Published 2021
by Frontiers Media as open access)

Fig. 18.4 (a) Deposition of CdSe films in chemical bath; (b) Experimental setup to measure pho-
toresponse. (Reprinted from Ref. [39]. Published by SAMRIDDHI as open access)

metal-semiconductor-metal, i.e. Ag/CdSe/Ag structure photodetector. Owing to


suitable direct bandgap, high absorption coefficient and high photosensitivity, CdSe
can take part in optical activity in infrared and visible part of the spectrum [17]. The
I-V characteristics of Ag/CdSe/Ag metal-semiconductor-metal structure confirm
the ohmic nature of the fabricated device and it exhibited a good photosensitivity
and photo-responsitivity under the illumination of visible-light.
18 Solution-Processed Photodetectors 435

Another interesting option for manufacturing a solution processed photodetector


was proposed by Chen et al. [3]. In this process, chemical bath deposition of ZnSe
nanoparticles on Si wire (SiW) arrays makes fabricating well-aligned 3D hetero-
structured ZnSe NP/SiW arrays with a high on/off ratio photocurrent and photocata-
lytic activity. ZnSe/SiW 3D heterostructures demonstrated excellent performance
UV photocurrent response due to a large surface-to-volume ratio and active area, as
well as a fast conductive pathway [3]. The coating of ZnSe nanowires on Si wafer
using CBD method provides excellent photodetection with immediate decay greater
than 99.85%, on/off ratio greater than 700 and fast photoresponse speed less than
0.4 μs. Under the exposure of UV radiation, the formation of electron-hole pair
occurs in ZnSe. The electron goes to conduction band and hole stays in the valance
band. The recombination was possible to avoid because of the ready formation of
3D heterostructure that has the advantage for migration of photogenerated electrons
to the surface.

18.3.2 Successive Ionic Layer Adsorption and Reaction


Technique (SILAR)

Adsorption, which is the base for SILAR method, is a surface activity that takes
place between ions and the substrate’s surface and is caused by the cohesive force,
Van der Walls force, chemical attractive force etc. between the ions in the solution
and the substrate’s surface. Here, the substrate particles are held in place by the
imbalanced or residual force of the atoms or molecules on the substrate surface. The
approach entails two steps (i) adsorption and rinsing, and (ii) reaction and rinsing.
The cations in the precursor solution are adsorbed on the surface of the substrate in
this initial stage, forming the Helmholtz electric double layer [39]. For instance, a
self-powered photoelectrochemical cell-type UV detector with a ZnO/ZnS core-­
shell nanorod array as the active photoanode and deionized water as the electrolyte
demonstrates a quick photoresponse time of around 0.04 s and outstanding spec-
trum selectivity [3]. Here, the shell consist of ZnS nanoparticles was grown through
the steps of SILAR method. First, a ZnO nanorod array substrate was immersed into
0.1 M zinc nitrate aqueous solution and rinsed with deionized water. On the second
step it was dipped into 0.1 M Na2S aqueous solution and again rinsed with deion-
ized water. This process was repeated for 2, 7, 10, and 15 cycles to generate differ-
ent thicknesses of ZnS nanoparticles.
Figure 18.5c shows that ZnS nanoparticles have been uniformly deposited on
ZnO nanorod in SILAR process and have come out with a rough surface which is
favorable for light scattering. The light scattering causes photo-anodes to exhibit
lower transmittance than bare FTO. The transmittance gets lowered with increase of
SILAR cycles as visible in Fig. 18.5a. From Fig. 18.5b, it is clear that the photosen-
sitivity gets improved with more than 0.3 A/W just after adding ZnS with only two
SILAR cycles. The sample with ZnS seven SILAR cycles at 340 nm has the best
436 S. K. Mostaque et al.

Fig. 18.5 (a) UV-visible transmission spectrum, (b) spectral sensitivity characteristics of UV pho-
todetectors at different SILAR cycles, (c) FESEM view of ZnO/ZnS core-shell nanorod array with
high magnification of nanorod at inset. (Adapted from Ref. [31]. Published 2016 by Springer as
open access)

photoresponsivity of 0.056 A/W which is 180% enhancement in comparison with


bare ZnO. The new band energy alignment of the ZnO/ZnS contact is responsible
for accelerating the separation of e-h pairs with this regard. The research further
suggests that the improved photo responsivity has been evolved from the applica-
tion of ZnS with SILAR method [31].

18.4 Solution Processed Photodetector Fabricated Using


Methods of Thick Film Technology

As we indicated earlier, the second approach used in the manufacture of solution


processed photodetectors involves the separation of the processes of nanoparticle
synthesis and their deposition on a substrate to form a photosensitive layer. As an
example of this approach to the development of photodetectors, consider several
technological processes used in the manufacture of solution processed photodetec-
tors based on II-VI semiconductors. For example, Saeed et al. [44] fabricated
ZnS:Mn-based metal-semiconductor-metal junction UV photodetector based on the
nanorod networks using this approach. Mn2+-doped ZnS nanorods were synthesized
by a facile hydrothermal method. The ZnS:Mn NRs were dispersed into DI water to
make a suspension. The typical concentration was 0.2 mg/ml. A droplet of the
ZnS:Mn NRs suspension was drop-cast on Ag electrodes and allowed to dry at room
temperature. After soaking the substrate in a 7:3 mixture of butylamine and aceto-
nitrile, the substrate was annealed for 3 min at 80 °C to prepare the films. The device
exhibited visible blindness, superior ultraviolet photodetection with a responsivity
of 1.62 A/W, and significantly fast photodetection response with the rise and decay
times of 12 and 25 ms, respectively. The shift in sensitivity to the UV region is due
to the fact that when the particle size of ZnS is minimized far below the Bohr
18 Solution-Processed Photodetectors 437

diameter (5 nm) depending on strong quantum confinement, it becomes conceivable


to adjust the band gap of ZnS quantum dot above 4.43 eV for photodetection in the
ultraviolet region.
The same approach was used by Kuang et al. [25] and Xia et al. [53] in develop-
ing solar-blind deep ultraviolet (DUV) photodetectors (PDs). Xia et al. [53] fabri-
cated PDs using colloidal ZnS quantum dots (QDs) via spin coating and ligand
exchange. The ZnS QD solar-blind DUV PDs showed a fast response (tr = 0.35 s,
td = 0.07 s) and good responsivity (∼0.1 mA/W). Colloidal ZnS QDs with an exci-
ton peak at 265 nm, corresponding to a band gap of 4.68 eV, were synthesized by a
hot injection method. Kuang et al. [25] used cubic ZnS QDs with a particle size of
~2.29 nm synthesized by wet chemical method to fabricate UV photodetectors. The
layer of ZnS QDs on the top of interdigital Au electrodes was formed using the drop
coating method. The photodetector showed a cutoff at 300 nm, a photosensitivity of
8, responsivity of 1.60 mA/W, and detectivity of 5.51 × 109 cm Hz1/2 W−1 at 254 nm,
when operated at a bias voltage of 15 V.
Mei et al. [35] showed that the spin-casting method can also be used to form
multilayer structures, such as TiO2/CdS/CdTe-based solar cells. A simple solution
process under ambient conditions developed by Mei et al. [35] is shown in Fig. 18.6.
A TiO2 film with a thickness of 40 nm was prepared by depositing a Ti2+ precursor
onto the FTO substrate and spin-casted at 2500 rpm for 15 s, then the substrate was
annealed at 500 °C for 1 h to eliminate any organic solvent and form a compact TiO2
thin film. Several drops of the CdS NC solution with different concentrations (5 mg/
mL, 10 mg/mL, 15 mg/mL, and 20 mg/mL) were then deposited onto the FTO/TiO2
and spin-casted at 3000 rpm for 20 s. Following this, the substrate was transferred
to a hot plate and annealed at 150 °C for 10 min, then transferred to another hot plate
and annealed at 380 °C for 30 min. One wash with isopropanol was used to remove
any impurities. The CdTe NCs were then deposited layer by layer onto the FTO/

Fig. 18.6 A schematic of the fabrication process of the NC solar cells. (Reprinted from Ref. [35].
Published 2018 by MDPI as open access)
438 S. K. Mostaque et al.

TiO2/CdS substrate with a process described previously in [40]. Finally, several


drops of saturated CdCl2/methanol were put onto the FTO/TiO2/CdS/CdTe substrate
and spin-casted at 1100 rpm for 20 s, then transferred onto a hot plate at 330–420 °C
for 15 min. Sixty nanometers of Au was deposited via thermal evaporation through
a shadow mask with an active area of 0.16 cm2 to make the electrode contact. The
introduction of a thin layer of CdS NC film (~5 nm) between the CdTe and TiO2
resulted in optimized band alignments and reduces the interface defects. As a result,
solar cells manufactured using the described technology showed a power conver-
sion efficiency (PCE) of 5.16% [35].
For the manufacture of photodetectors by this method, commercially available
powders of II-VI semiconductors as well can be used. This approach was used by
Rahman et al. [41, 42]. They showed that the deposition of CdS thin films by spin
coating method using thiol-amine co-solvents has good potential in the fabrication
of low cost films for high efficiency applications. The process involves commer-
cially available CdS powder (99.9999%), ethylene-di-amine, 1,2 ethanedi-thiol, and
Triton X-100 for the preparation of CdS precursor solution. The simple process first
takes a mixture of ethylene-di-amine and 1,2 ethane-di-thiol in 10:1 ratio. CdS pow-
der of 0.3 wt% is then dissolved with the mixture. This can be done with a magnetic
stirrer with keeping the solution more than the room temperature. The process may
take around 15 h at 50 °C to completely dissolve the CdS particles. The process is
illustrated in Fig. 18.7a. Addition of small amount of surfactant (e.g. Triton X-100
with 0.1 wt %) has been found to produce consistent and high quality films [41, 42].

Fig. 18.7 (a) Processing steps for preparation of CdS precursor solution; (b) Processing steps for
deposition of CdS film in simple spin coating method. (Reprinted with permission from Ref. [41].
Copyright 2020: Springer)
18 Solution-Processed Photodetectors 439

The deposition can be completed on glass substrate with simple spin coating
technique. Rahman et al. [41] coated one time on perfectly cleaned glass substrate
at speed more than 1000 rpm for 45 s. To remove the remaining solvents, the coated
films were pre-annealed at 90 °C. The high temperature annealing requires a glass
protector to get rid from oxidation. The process is described in Fig. 18.7b. In this
process, the method offers good cyrstallinity at annealing temperature of 300 °C.
Photodetectors based on II-VI semiconductor-polymer composites can also be
fabricated using the principles of solution processed technology. Huynh et al. [18]
fabricated a CdSe/P3HT-based photodetector using this approach. CdSe/P3HT
200 nm layer between an aluminum electrode and a transparent conducting elec-
trode of PEDOT:PSS was spin-cast from a solution containing 90% wt % CdSe
nanorods in P3HT and pyridinechloroform as solvent. The active area of the device
was 1.5 mm by 2.0 mm. Adjusting the band gap by changing the radius of the
nanorod allowed Huynh et al. [18] to match the absorption spectrum of the photo-
detector with the spectrum of solar radiation. As a result, a photovoltaic device
consisting of CdSe nanorods 7 nm in diameter and 60 nm long and a conjugated
poly-3(hexylthiophene) polymer demonstrated an external quantum efficiency of
more than 54%, and a monochromatic power conversion efficiency of 6.9% under
illumination of 0.1 mW/cm2 at 515 nm. It is important that with a decrease in the
length of nanorods to 7 nm, the external quantum efficiency decreased to 20%.

18.5 Combined Approach to the Fabrication


of Solution-Processed Photodetectors

Solution-processed photodetectors can also be produced by combining the above


methods. This approach is mainly used in the formation of multilayer structures. For
example, Sekhar Reddy et al. [45] in the manufacture of NiO/CdS heterojunction
based photodetectors (see Fig. 18.8), the CdS layers upon ITO coated PET substrate
were deposited by photochemical deposition technology at ambient condition, and
the NiO layers were deposited from NiO nanoparticles by the spin-coating method.
The solution for CdS deposition was prepared by mixing 7 ml of 0.05 M CdCl2 with
6.8 ml of 0.25 M sodium citrate at maintained pH 5, then 2.5 ml of KOH of a pH ten
followed by 5 ml of buffer to reach the pH 12, and finally 27.5 ml of deionized
water. The total solution was mixed 5 min; then the ITO-PET substrate was
immersed in this reaction solution and put in a sealed box illumined by UV source
(313 nm) in a room temperature setting. After 1.5 hrs, the CdS-coated substrate was
taken out, washed with deionized water, and dried up with nitrogen. NiO nanopar-
ticle powders were dispersed in ethanol and subjected to spinning at 3000 rpm for
30 s. After deposition, the p-n heterojunction was heated at 60 °C for 30 mins. The
final device was completed by depositing the top aluminium metal contacts (120 nm)
on the heterojunction using a shadow mask.
440 S. K. Mostaque et al.

Fig. 18.8 Scheme of the solution manufacturing process for CdS, NiO and the photodetector final
device construction. (Reprinted with permission from Ref. [45]. Copyright 2021: Elsevier)

The same approach was taken by Li et al. [30] in the fabrication of organic-­
inorganic metal halide perovkite solar cells (PKSCs) (see Fig. 18.8). ZnSe layer was
synthesizes using chemical bath deposition method, while the TiO2 and metal halide
perovkite were spin coated. Li et al. [30] have found that low-temperature solution-­
processed ZnSe can be used as a potential electron transportation layer (ETL) for
PKSCs. Optimized device with ZnSe ETL has achieved a high power conversion
efficiency (PCE) of 17.78% with negligible hysteresis, compared with the TiO2
based cell (13.76%). Li et al. [30] believe that this enhanced photovoltaic perfor-
mance is attributed to the suitable band alignment, high electron mobility, and
reduced charge accumulation at the interface of ETL/perovskite. Encouraging
results were obtained when the thin layer of ZnSe cooperated with TiO2. It shows
that the device based on the TiO2/ZnSe ETL with cascade conduction band level can
effectively reduce the interfacial charge recombination and promote carrier transfer
with the champion PCE of 18.57%. In addition, the ZnSe-based device exhibits a
better photostability than the control device due to the greater ultraviolet (UV) light
harvesting of the ZnSe layer, which can efficiently prevent the perovskite film from
intense UV-light exposure to avoid associated degradation (see Fig. 18.9b).
Rose et al. [43] for the fabrication of solution-processed СdTe/CdS photodetec-
tors (see Fig. 18.10) proposed another way. If the CdS layer was grown by chemical-­
bath deposition (CBD), then the CdTe layer was formed by close-spaced sublimation.
The I-V results (average and standard deviation) with AM1.5 illuminations are
shown in the Table 18.2. It is seen that photodetectors made in this way showed an
average efficiency of ~12.6%.
18 Solution-Processed Photodetectors 441

Fig. 18.9 (a) Schematic view of the typical cell architecture of organic−inorganic metal halide
perovkite solar cells (PKSCs). (b) The normalized power conversion efficiency (PCE) decay of
devices based on ZnSe and TiO2 ETLs as a function of storage time upon 1 sun irradiation.
(Reprinted with permission from Ref. [30]. Copyright 2018: ACS)

Fig. 18.10 The structure of CdTe/CdS-based solar cell. The CdS layer was grown by chemical-­
bath deposition (CBD). (Reprinted with permission from Ref. [43]. Copyright 1999: Wiley)

Table 18.2 AM1.5 I-V results for 38 solar cell baseline set
Parameter Efficiency Voc (mV) Jsc (mA/cm2) FF (%) Area (cm2)
Average 12.6% 820 21.8 70.6 0.86
Standard deviation 0.5% 8 0.4 2.0 0.39
Source: Reprinted with permission from Ref. [43]. Copyright 1999: Wiley

A combined approach to the manufacture of solution-processed photodetectors


can also offer such a way when wet chemical methods are used to synthesize core
shell structures, and the photosensitive layer of the solution-processed photodetec-
tor is formed by the spin coating method. In particular, Kwon et al. [26] proposed
such an approach for the development of a PbS/CdS core-shell-based photodetector.
While PbS can be a cost-effective option due to its tunable bandgap, photosensitiv-
ity, and, most crucially, solution-processability, findings suggest that depositing
inorganic CdS onto the PbS core can significantly increase photo and thermal
442 S. K. Mostaque et al.

stabilities, and therefore device performance. The charge transfer is improved by the
gradient interfacial layer between the PbS core and the CdS shell, which permits
excitons to partially leak into the shell [26]. The device operates when incoming
photons are absorbed in the active material and photoexcited electron–hole pairs
(EHPs) are driven to the electrodes by an applied electric field. The performances
under dark and IP illumination with 0.1 mW/cm2 power density and reverse bias of
−1 V are shown in Table 18.3 for three different combinations. It is seen that the
detectors with PbS/CdS core shell structure show a significant improvement in on/
off ratio, detectivity and photo response. In addition, the IR photo response indicate
that the devices with thicker shell exhibit a faster response with -1 V bias and fall
time performance which is displayed in Fig. 18.11a, b respectively [26].
Zhou et al. [56] have shown that solution-processed technology makes it possible
to implement more complex structures. For example, Zhou et al. [56], using the
principles of solution-processed technology, fabricated upconversion photodetec-
tors, the structure of which is shown in Fig. 18.12a. The infrared photodetector
converted incident light of 1600 nm into visible emitted light (see Fig. 18.12b).
Luminescent CdSe/ZnS QDs with emission at 525 nm were used as the active mate-
rial for the LED, and narrow-bandgap PbS QDs (300–1600 nm), synthesized via the
hot injection method, were used as the active sensitizing material in the photodetec-
tor. The PbS nanoparticles were uniformly dispersed with a diameter of

Table 18.3 Performance of SWIR photodetector with IR illumination


Device Jdark (mA/cm2) Jlight (mA/cm2) Light/Dark ratio D* (Jones)
PbS-QD 9.32 12.514 1.34 6.16 × 1011
Thin shell PbS/CdS QD 5.884 35.3 5.99 1.35 × 1012
Thick shell PbS/CdS QD 5.0526 56.856 11.25 7.14 × 1012
Source: Reprinted from Ref. [26]. Copyright 2020: Springer Open. Open access

Fig. 18.11 (a) IR photo response for the thicker PbS/CdS QD under −1 V bias and (b) Transient
photo response of the shell under -1 V bias. The light intensity of 0.1 mW/cm2. (Reprinted from
Ref. [26]. Published 2020 by Springer Open as open access)
18 Solution-Processed Photodetectors 443

Fig. 18.12 (a) Structure and composition of upconversion devices herein. (b) Photograph of a
sample clamped in the measurement box (left), and images of a device (area of 0.05 cm2) with
(bottom) and without (top) infrared light (940 nm, 10 mW cm−2). (Reprinted with permission from
Ref. [56]. Copyright 2020: Springer Nature)

approximately 6 nm. Zinc oxide (ZnO) with embedded Ag nanoparticles was used
as the ETL and poly-(N,N-bis(4-butylphenyl)-N,N-bis(phenyl) benzidine) (poly-­
TPD) as the hole transport layer. The PbS layer was deposited via layer-by-layer
spin coating. For each layer, two drops of PbS solution were spin cast on to the ZnO
substrate at 2500 rpm for 10s. CdSe/ZnS core/shell QDs dissolved in octane (10 mg/
ml) were spin coated at 2000 rpm for 30 s, followed by heat treatment at 80 °C for
10 min. The ZnO film was spin cast at 2000 rpm for 30 s, followed by heat treatment
at 80 °C for 10 min. Finally, 100 nm Ag or 140 nm ITO were deposited using an
Angstrom Engineering deposition system.
The sensitivity of the resulting photodetectors corresponds to the absorption of
QDs and exhibits a definite exciton peak at 1500 nm. When reverse biased, the
devices show gain with a peak sensitivity of >20 AW−1 in the shortwave infrared and
>60 AW−1 in the visible range. However, devices without Ag nanoparticles show no
gain and provide sensitivity below 0.3 AW−1. The devices show typical diode char-
acteristics in the dark. The value of the dark current at a bias of −1 V does not
exceed 13 μAcm−2. The resulting noise equivalent power (NEP) reached
3 × 10–14 W/Hz1/2, resulting in a D* of 6 × 1012 Jones at 180 Hz. Zhou et al. [56]
consider this to be in line with the best previously published PbS-based infrared
photodetectors with a similar bandgap. According to Zhou et al. [56], this top-­
emitting upconversion device can be used for bioimaging.
444 S. K. Mostaque et al.

18.6 Outlook and Perspectives


of Solution-Processed Technology

There is no doubt that solution-processed technology, using colloidal particles and


nanocrystals, is a promising technology for wide application. Currently, this tech-
nology is widely used for applying organic materials. However, colloidal inorganic
semiconductor nanocrystals share the important advantages of organic materials
such as low-temperature solution-processing and controllable synthesis. Meanwhile,
compared to organics, nanocrystals-based devices with proper particle surface pas-
sivation exhibit broader absorption spectrum and more efficient charge transport.
The QD wavelength selectivity can be tuned by adjusting the QD materials and
dimensions, which eliminates the need for color filters for image detectors.
Optoelectronic applications based on these nanomaterials include solar cells, photo-
detectors, phosphors, and light-emitting diodes (LEDs). But it must be recognized
that the traditional methods of colloidal nanoparticles deposition, which are widely
used at the moment, are good for research purposes or for forming layers over large
areas used in solar cells (see Fig. 18.13). In the manufacture of photodetectors
intended for the market, the use of such methods as spin coating, dip coating or roll-­
to-­roll technique may not be effective. In this regard, inkjet-printing techniques [2,
15, 28, 47, 57] as well as a number of other printing technologies [11, 21], which
have been intensively developed over the past two decades, seem very promising.
Currently, inkjet printing has begun to be used in the development of ZnS:Mn,
CdS, CdSe, and CdSe/ZnS light emitted diodes, phosphors [19, 22] and color con-
version element for full-color LED displays [50]. An example of the

Fig. 18.13 (a) Schematic diagram of roll-to-roll technique. Using a roll-to-roll technique to man-
ufacture solution-processed solar cells reduces the manufacturing, deployment and energy costs
involved. (Reprinted with permission from Ref. [13]. Copyright 2012: Springer Nature). (b)
Schematic diagram of inkjet printing system. The inkjet printer consists of a XY moving stage, a
print head with cartridge, and a control board. (Reprinted from Ref. [7]. Published 2020 by
Springer as open access)
18 Solution-Processed Photodetectors 445

Fig. 18.14 Fluorescence microscope images of (a–d) single-colour and (e) multi-colour (1x1 cm2)
arrays of CdS (a) and CdSe@ZnS NCs (b–e) in 5 wt% PS. In (f–g) optical profiler 3D images of
pixels formed of (f) CdS and (g) CdSe@ZnS NCs (3.4 nm) in 5 wt% PS, respectively. In (h) 2D
view of the AFM image of CdS NCs in 5 wt% PS. In the graph RMS roughness values of pixels
with CdS and CdSe@ZnS NCs in 3 wt% and 5 wt% PS. The value corresponding to 0 nm size is
referred to bare PS pixels. (Reprinted with permission from Ref. [19]. Copyright 2009: Elsevier)

implementation of luminescent structures based on colloidal CdS nanoparticles and


differently sized CdSe@ZnS nanocrystals (NCs) using the inkjet printing technique
is shown in Fig. 18.14. Figure 18.14 shows that regular shape disk-like pixels, lumi-
nescent from blue to red, were dispensed with a diameter of 400–600 μm and with
no satellites. The microstructures based on CdSe@ZnS NCs present a slight con-
cavity as evidenced by the slight change in shading, with no pronounced “coffee-
staining effect” (Fig. 18.14f), while multiple rings are in the pixels formed of CdS
NCs (Fig. 18.14e). Ingrosso et al. [19] believe that the manufactured highly lumi-
nescent and non-bleachable microstructures can be integrated in polymer displays
and coloured wall papers. The reported approach can be extended to functionalize a
variety of polymers with different functional colloidal NCs to fabricate polymer
based micro- and nano-electronic components.
But experiment shows that inkjet printing technology can be easily transferred to
the production of photodetectors [1, 9, 13, 46]. For example, there are already
446 S. K. Mostaque et al.

reports of fabrication using this method of HgTe NCs based photodetectots operat-
ing up to 3 μm wavelengths [1]. Detectors operating in this spectral region are of
particular importance for biological applications, remote sensing and night-vision
imaging. The hydrophobic HgTe nanocrystals (NC) were initially synthesized in
aqueous solution at room temperature via a reaction between Hg(ClO4)2 and H2Te
gas in the presence of short-chain hydrophilic thiols such as thioglycerol or mercap-
toethanolamine as stabilizer. For inkjet-printing a 2 wt% HgTe NC/chlorobenzen
solution was found to have a suitable viscosity and surface tension for ejection
from the used piezo-driven print head. Boberl et al. [1] reported that fabricated pho-
todetectors demonstrated room temperature detectivities up to D* = 3.2 × 1010 cm
Hz1/2 W−1 at a wavelength of 1.4 μm close to the important telecommunication
wavelength region. The long-wavelength cut-off of the photoresponse can be tuned
by the size of the used HgTe nanocrystals. In particular, for a shift in the cut-off
wavelength from 1–2 μm to 3 μm the size of the HgTe nanocrystals was increased
from 3–4 nm up to 6 nm.
Wu et al. [51] showed that the inkjet-printing method can be combined with
other methods of solution-processed technology. To optimize the parameters of UV
ZnO-based detectors fabricated by inkjet printing, Wu et al. [51] suggested to use
the dip coating of as-prepared ZnO films by CdS nanoparticles. They reported that
using this approach it was achieved a 3 orders of magnitude enhancement of the
ultraviolet photoresponse of ZnO thin film. In addition, the decay time of the pho-
toresponse was reduced to about 4 ms. Thus, capping with CdS not only suppressed
the detrimental passivation layer of ZnO thin films, but also generated an interfacial
carrier transport layer to reduce the probability of carrier recombination.
But we are sure that the main advantage of the inkjet-printing technique is that it
allows to combine microminiaturization and solution processed technology. And
this means that a cheap technology for manufacturing pixelated photodetectors is
emerging. Currently, the minimum nozzle diameter is ~100 nm (https://www.fujif-
ilm.com). This means that if a drop with a diameter of 100 nm is formed, it can form
a round film with a diameter of ~200 nm [48]. However, the best devices for inkjet
printing actually used provide a resolution of ~2 μm [27], while the resolution of
conventional devices is in the range of 20–50 μm [21]. Therefore, in its current
incarnation, inkjet printing is not particularly well-suited to the production of
extremely high-resolution electronics on the order of those currently produced by
conventional means. However, by delivering droplets that are a handful of microme-
tres in diameter, reasonably well-resolved ordinary optoelectronic and photoelec-
tronic devices may be produced.
In addition, the use of the inkjet printing technique makes it possible to produce
multicolor photodetectors without much difficulty. In particular, the possibility of
manufacturing such photodetectors were demonstrated by Cook et al. [5]. Schematic
diagram of printing of ZnO QDs (black), PbS QDs (blue), and FeS2 (red) NCs on
graphene channels between two nearest-neighbor Au electrodes on SiO2 (500 nm)/
Si substrates is shown in Fig. 18.15. Three different nozzles containing ZnO QDs,
PbS QDs, and FeS2 NCs were employed for printing each of them on the different
places. The above example and the presence of a large number of methods for the
18 Solution-Processed Photodetectors 447

Fig. 18.15 (a) Schematic diagram of printing of ZnO QDs (black), PbS QDs (blue), and FeS2
(red) NCs on graphene channels defined between two nearest-neighbor Au electrodes on SiO2
(500 nm)/Si substrates. (b) Normalized spectral photosensitivity of elements forming a multicolor
photodetector. (Reprinted with permission from Ref. [5]. Copyright 2019: ACS)

Table 18.4 Commercially available conductive inks for inject printing


Name Material Manufacturer
NanoGold Au nanoparticle Sigma https://www.sigmaaldrich.
Aldrich com
NanoSilver Ag nanoparticle Sun https://www.sunchemical.
Chemical com
Clevios ™PH 1000 PEDOT:PSS Heraeus https://www.heraeus.com
Polyaniline (Emeraldine PANi Sigma https://www.sigmaaldrich.
salt) nanoparticles Aldrich com
PEDOT:PSS poly(3,4-ethylenedioxythiophene) polystyrene sulfonate

synthesis of colloidal particles and nanocrystals of II-VI compounds (Chaps. 11, 12,
and 13, Vol. 1) indicates that the use of inkjet-printing technology in the manufac-
ture of photodetectors based on II-VI compounds should not cause any particular
difficulties. Ink-jet printing of colloidal nanocrystals is material effective and highly
reproducible, and can be applied not only for the photosensitive materials but also
for fabrication of the electrodes [52], opening up prospects for low cost, all ink-jet
printed photodetector devices. Commercially available conductive inks for inject
printing are listed in Table 18.4. With an appropriate selection of QDs of II-VI com-
pounds, it is possible to manufacture multicolor selective photodetectors simultane-
ously sensitive in the IR, visible and UV spectral regions.
The inkjet-printing technique can also be used to form the heterostructures [33]
needed to fabricate tandem photodetectors. It is also possible to simultaneously
deposit nanoparticles of different II-VI semiconductors. In particular, Miethe et al.
[36] thus formed a CdSe/CdS gel-network, which may be a prototype of the quasi-­
type-­II superlattice structure. The use of inkjet-printing techniques also facilitates
the integration of II-VI semiconductor-based photodetectors with signal processing
silicon integrated circuits. Inkjet printing offers unique advantages in mass
448 S. K. Mostaque et al.

Table 18.5 Typical inkjet ink composition


Loading
Component Function (w/w%)
II-VI semiconductor colloidal Key component 0.1–10
NPs or NCs
Solvent Dispersion/dissolution medium 50–90
Co-solvent(s) Controls drying (“coffee-ring”) 0–50
Viscosity modification Surface tension modification 0–50
Surfactant Modifies surface tension improves wetting 0–5
Viscosity modifier (dissolved) Generally, increases viscosity <1
Humectant Low volatility, prevents ink drying in nozzles 0–20
Other pH buffer; biocide; fungicide; dispersant; <1
defoamer; binder (polymer)
Source: Data extracted from Refs. [2, 6, 28, 32, 48, 57]

scalability, cost reduction, low waste, and direct deposition on targeted regions. This
means that inkjet printing makes it possible to embed photonic components based
on II-VI semiconductor functional nanomaterials and quantum nanostructures into
highly crystalline structures of desired morphology with CMOS read-out circuits
without the need for any additional chemical treatments that affect the properties of
these circuits.
It is understood that the ink preparation process can be very complex depending
on the material used, its compatibility with carrier solvents and other ink compo-
nents, and its ease of dissolution or dispersion. Therefore, ink formulation becomes
an absolutely critical aspect of functional material application and device fabrica-
tion. Although inks may be formulated in many different ways, the composition of
the ink is generally similar to that given in Table 18.5. Ink preparation requires
careful control of fluid properties, wetting behavior, drying behavior, interaction
with a given substrate, maintaining dispersion and, above all, maintaining function-
ality. We have to admit that at the moment, there is a big gap between a simple
suspension or solution and real inks suitable for creating optoelectronic structures
[32]. But there is no doubt that these problems in relation to II-VI compounds will
be successfully solved, which will contribute to the rapid development of inkjet-­
printing technology for manufacturing II-VI semiconductor-based photodetectors.

18.7 Conclusion

This chapter has reviewed some modern methods of II-VI semiconductor synthesis
as well as approaches to photodetectors fabrication. The analysis showed that the
development of photodetectors based on solution-processed II-VI semiconductors
has advanced significantly over the past decade due to advances in materials science
and device development. This has led to the development of photodetectors that
combine desired performance with manufacturing advantages based on
18 Solution-Processed Photodetectors 449

state-of-­the-art technology. The achieved results indicate that solution-processed


devices can indeed be considered as promising candidates for the commercializa-
tion of low-­cost photodetectors.
Since II-VI semiconductors and their quantum dots have a wide range of control-
lable band gaps, the solution- processed technology allows to tune this value to
detect radiation in any part of the solar spectrum. However, in all cases, the solution
process often requires proper optimization of its parameters in order to get the best
possible result from the intended device. Solution- processed photovoltaic materials
typically contain a significant degree of disorder, often due to the boundaries that
define randomly oriented nanocrystals. Therefore, transport within these domains
as well as at their boundaries (including transport-limiting mechanisms such as
charge carrier capture) requires attention and optimization. For example, studies
carried out by Zhang et al. [55] have shown that manipulating the lifetime and trans-
port domains of majority and minority carriers, one can significantly increase the
sensitivity of CdTe photoconductors and achieve a detectivity of up to 5 × 1017 Jones
in the visible and near infrared ranges. Likewise, the best reported gain (G) × band-
widths (BW) products (above 6 × 106 Hz) was achieved for visible CdS photocon-
ductors after improvement of carrier mobility [29].

Acknowledgments G. Korotcenkov is grateful to the State Program of the Republic of Moldova,


project 20.80009.5007.02, for supporting his research.

References

1. Boberl M, Kovalenko MV, Gamerith S, List EJW, Heiss W (2007) Inkjet-printed nanocrystal
photodetectors operating up to 3 μm wavelengths. Adv Mater 19:3574–3578
2. Calvert P (2001) Inkjet printing for materials and devices. Chem Mater 13:329
3. Chen Y-H, Li W-S, Liu C-Y, Wang C-Y, Chang Y-C, Chen L-J (2013) Three-dimensional het-
erostructured ZnSe nanoparticles/Si wire arrays with enhanced photodetection and photocata-
lytic performances. J Mater Chem C 1:1345–1351
4. Chesman ASR, Duffy NW, Martucci A, De Oliveira Tozi L, Singh TB, Jasieniak JJ (2014)
Solution-processed CdS thin films from a single source precursor. J Mater Chem C
2(2014):3247–3253
5. Cook B, Gong M, Ewing D, Casper M, Stramel A, Elliot A, Wu J (2019) Inkjet printing mul-
ticolor pixelated quantum dots on graphene for broadband photodetection. ACS Appl Nano
Mater 2:3246–3252
6. Croucher M, Hair M (1989) Design criteria and future directions in inkjet ink technology. Ind
Eng Chem Res 28(11):1712
7. Da Costa TH, Choi J-W (2020) Low-cost and customizable inkjet printing for microelectrodes
fabrication. Micro Nano Syst Lett 8:2
8. Das S, Dhara S (eds) (2021) Chemical solution synthesis for materials design and thin film
device applications. Elsevier, Amsterdam
9. Dong Y, Zou Y, Song J, Li J, Han B, Shan Q et al (2017) All-inkjet-printed flexible UV photo-
detector. Nanoscale 9:8580
10. Eslamian M (2017) Inorganic and organic solution-processed thin film devices. Nano-Micro
Lett 9:3
450 S. K. Mostaque et al.

11. Fukuda K, Someya T (2017) Recent progress in the development of printed thin-film transis-
tors and circuits with high-resolution printing technology. Adv Mater 29(25):1602736
12. García de Arquer FP, Armin A, Meredith P, Sargent EH (2017) Solution-processed semicon-
ductors for next-generation photodetectors. Nat Rev Mater 2:16100
13. Graetzel M, Janssen RAJ, Mitzi DB, Sargent EH (2012) Materials interface engineering for
solution-processed photovoltaics. Nature 488:304–312
14. Hankare PP, Bhuse VM, Garadkar KM, Delekar SD, Mulla IS (2004) Chemical deposition of
cubic CdSe and HgSe thin films and their characterization. Semicond. Sci. Technol. 19:70.
15. Haverinen H, Myllyla R, Jabbour G (2009) Inkjet printing of light-emitting quantum dots.
Appl Phys Lett 94:073108
16. Hou X, Aitola K, Lund PD (2021) TiO2 nanotubes for dye-sensitized solar cells—A review.
Energy Sci Eng 9:921–937
17. Hussain S, Iqbal M, Khan AA, Khan MN, Mehboob G, Ajmal S et al (2021) Fabrication
of nanostructured cadmium selenide thin films for optoelectronics applications. Front Chem
9:661723
18. Huynh WU, Dittmer JJ, Alivisatos AP (2002) Hybrid nanorod-polymer solar cells. Science
295:2425–2427
19. Ingrosso C, Kim JY, Binetti E, Fakhfouri V, Striccoli M, Agostiano A et al (2009) Drop-on-­
demand inkjet printing of highly luminescent CdS and CdSe@ZnS nanocrystal based nano-
composites. Microelectron Eng 86:1124–1126
20. Jain M (1993) II-VI semiconductor compounds. World Scientific, Singapore
21. Khan S, Lorenzelli L, Dahiya RS (2015) Technologies for printing sensors and electronics
over large flexible substrates: a review. IEEE Sensors J 15:3164–3185
22. Kim JY, Ingrosso C, Fakhfouri V, Striccoli M, Agostiano A, Curri ML, Brugger J (2009)
Inkjet-printed multicolor arrays of highly luminescent nanocrystal-based nanocomposits.
Small 5(9):1051–1057
23. Korotcenkov G (2014) Handbook of gas sensor materials, Vol. 2: new trends and technologies.
Springer, New York
24. Korotcenkov G, Tolstoy V, Schwank J (2006) Successive ionic layer deposition (SILD) as
a new sensor technology: synthesis and modification of metal oxides. Meas Sci Technol
17:1861–1869
25. Kuang W-J, Liu X, Li Q, Liu Y-Z, Su J, Harm TH (2018) Solution-processed solar-blind ultravi-
olet photodetectors based on ZnS quantum-dots. IEEE Photon Technol Lett 30(15):1384–1387
26. Kwon J-B, Kim S-W, Kang B-H, Yeom S-H, Lee W-H, Dae-Hyuk Kwon D-H et al (2020)
Air-stable and ultrasensitive solution-cast SWIR photodetectors utilizing modified core/shell
colloidal quantum dots. Nano Converg 7:28
27. Kwon HJ, Chung S, Jang J, Grigoropoulos CP (2016) Laser direct writing and inkjet printing
for a sub-2 μm channel length MoS2 transistor with high-resolution electrodes. Nanotechnology
27:405301
28. Le H (1998) Progress and trends in ink-jet printing technology. J Imaging Sci Technol 42(1):49
29. Lee J-S, Kovalenko MV, Huang J, Chung DS, Talapin DV (2011) Band-like transport,
high electron mobility and high photoconductivity in all-inorganic nanocrystal arrays. Nat
Nanotechnol 6:348–352
30. Li X, Yang J, Jiang Q, Lai H, Li S, Xin J, Chu W, Hou J (2018) Low-temperature solution-­
processed ZnSe electron transport layer for efficient planar perovskite solar cells with negli-
gible hysteresis and improved photostability. ACS Nano 12:5605–5614
31. Lin H, Wei L, Wu C, Chen Y, Yan S, Mei L, Jiao J (2016) High-performance self-powered
photodetectors based on ZnO/ZnS core-shell nanorod arrays. Nanoscale Res Lett 11(1):420
32. Magdassi S (ed) (2010) The chemistry of inkjet inks. World Scientific Publishing, Singapore
33. Marjanovic N, Hammerschmidt J, Perelaer J, Farnsworth S, Rawson I, Kus M et al (2011)
Inkjet printing and low temperature sintering of CuO and CdS as functional electronic layers
and Schottky diodes. J Mater Chem 21:13634
18 Solution-Processed Photodetectors 451

34. Mbam SO, Nwonu SE, Orelaja OA, Nwigwe US, Gou XF (2019) Thin-film coating; historical
evolution, conventional deposition technologies, stress-state micro/nano-level measurement/
models and prospects projection: a critical review. Mater Res Express 6(12):122001
35. Mei X, Wu B, Guo X, Liu X, Rong Z, Liu S et al (2018) Efficient CdTe nanocrystal/TiO2
hetero-junction solar cells with open circuit voltage breaking 0.8 V by incorporating a thin
layer of CdS nanocrystal. Nano 8:614
36. Miethe JF, Luebkemann F, Schlosser A, Dorfs D, Bigall NC (2020) Revealing the correlation
of the electrochemical properties and the hydration of inkjet-printed CdSe/CdS semiconductor
gels. Langmuir 36:4757–4765
37. Miskin CK, Dubois-Camacho A, Reese MO, Agrawal R (2016) A direct solution deposition
approach to CdTe thin films. J Mater Chem C 4(2016):9167–9171
38. Mohammed RY (2021) Annealing effect on the structure and optical properties of CBD-ZnS
thin films for windscreen coating. Materials 14:6748
39. Nikam CP, Gosavi NM, Gosavi SR (2020) Low-cost visible-light photodetector based on Ag/
CdSe Schottky diode fabricated using soft chemical solution method. SAMRIDDHI: J Phys
Sci Eng Technol 12(2):62–67
40. Pawar SM, Pawar BS, Kim JH, Joo O-S, Lokhande CD (2011) Recent status of chemical bath
deposited metal chalcogenide and metal oxide thin films. Curr Appl Phys 11:117–161
41. Rahman MF, Hossain J, Kuddus A, Tabassum S, Rubel MHK, Shirai H, Ismail ABM (2020a)
A novel synthesis and characterization of transparent CdS thin films for CdTe/CdS solar cells.
Appl Phys A Mater Sci Process 126:145
42. Rahman MF, Hossain J, Kuddus A, Tabassum S, Rubel MHK, Rahman MM et al (2020b) A
novel CdTe ink-assisted direct synthesis of CdTe thin films for the solution-processed CdTe
solar cells. J Mater Sci 55:7715–7730
43. Rose DH, Hasoon FS, Dhere RG, Albin DS, Ribelin RM, Li XS, Mahathongdy Y (1999)
Fabrication procedures and process sensitivities for CdS/CdTe solar cells. Prog Photovolt Res
Appl 7:331–340
44. Saeed S, Dai R, Janjua RA, Huang D, Wang H, Wang Z, Ding Z, Zhang Z (2021) Fast-response
metal−semiconductor−metal junction ultraviolet photodetector based on ZnS:Mn nanorod
networks via a cost effective method. ACS Omega 6(48):32930–32937
45. Sekhar Reddy KC, Selamneni V, Syamala Rao MG, Meza-Arroyo J, Sahatiya P, Ramirez-­
Bon R (2021) All solution processed flexible p-NiO/n-CdS rectifying junction: applications
towards broadband photodetector and human breath monitoring. Appl Surf Sci 568:150944
46. Sliz R, Lejay M, Fan JZ, Choi M-J, Kinge S, Hoogland S et al (2019) Stable colloidal
quantum dot inks enable inkjet-printed high-sensitivity infrared photodetectors. ACS Nano
13(10):11988–11995
47. Taylor R, Church K, Sluch M (2007) Red light emission from hybrid organic/inorganic quan-
tum dot AC light emitting displays. Displays 28:92
48. Tekin E, Smith P, Schubert U (2008) Inkjet printing as a deposition and patterning tool for
polymers and inorganic particles. Soft Matter 4:703
49. Tolstoi VP (2009) New routes for the synthesis of nanocomposite layers of inorganic com-
pounds by the Layer-by-Layer scheme. Russ J Gen Chem 79:2578–2583
50. Wang X, Yuan M, Qin M (2016) Surface energy-modulated inkjet printing of semiconductors.
In: Yun I (ed) Printed electronics – current trends and applications. INTECH, pp 5–24
51. Wu Y, Tamaki T, Volotinen T, Belova L, Rao KV (2010) Enhanced photoresponse of inkjet-­
printed ZnO thin films capped with CdS nanoparticles. J Phys Chem Lett 1:89–92
52. Wu Y, Li Y, Ong BS, Liu P, Gardner S, Chiang B (2005) High-performance organic thin-film
transistors with solution-printed gold contacts. Adv Mater 17:184–187
53. Xia Y, Zhai G, Zheng Z, Lian L, Liu H, Zhang D et al (2018) Solution-processed solar-blind
deep ultraviolet photodetectors based on strongly quantum confined ZnS quantum dots. J
Mater Chem C 6:11266–11271
54. Younus IA, Ezzar AM, Uonis MM (2020) Preparation of ZnTe thin films using chemical bath
deposition technique. Nanocomposites, 6:4:165–172
452 S. K. Mostaque et al.

55. Zhang Y, Hellebusch DJ, Bronstein ND, Ko C, Ogletree DF, Salmeron M, Alivisatos AP
(2016) Ultrasensitive photodetectors exploiting electrostatic trapping and percolation trans-
port. Nat Commun 7:11924
56. Zhou W, Shang Y, de Arquer PG, Xu K, Wang R, Luo S et al (2020) Solution-processed upcon-
version photodetectors based on quantum dots. Nat Electron 3:251–258
57. Zhouping Y, Yongan H, Ningbin B, Xiaomei W, Youlun X (2010) Inkjet printing for flexible
electronics: materials, processes, and equipment. Chin Sci Bull 55(30):3383
Chapter 19
Multicolor Photodetectors

Paweł Madejczyk

19.1 General

The requirement for improved target recognition and temperature estimation has led
to an increasing interest in infrared (IR) detection in more than one band. Multicolor
detectors have the ability to detect a number of different infrared bands or a number
of different wavelengths in the same band separately and independently [40].
The goal of multiband IR visualization is to overcome the contrast limitations
that exist with single band imagery [37]. For typical display of single band infrared
imagery, a monochrome display is used and the limitation is that a human viewer
has an instantaneous dynamic range equivalent to only 100 shades-of-gray, a 1%
contrast difference. By using two of more infrared bands to create a color space, an
associated composite color image can be created using visible color bands for dis-
play to human viewers. In a color image people can discriminate millions of colors
defined by varying hue, saturation, and brightness. Multicolor detectors can deter-
mine the absolute temperature of objects in the scene. They play many important
roles in Earth and planetary remote sensing, medicine, astronomy, and many others.
Multicolor detection is the basic requirement for third generation IR devices [35].
One of the earlier concepts for infrared dual band (DB) vision was a system con-
sisted of two single-color focal plane arrays and a beam splitter [27]. This worked,
but there was considerable difficulty in optical alignment to a precision such that the
exact same image feature could be accurately compared on the two focal planes at
the pixel level. It also had the drawbacks of dual vacuum enclosures and cooling
systems. Recent advances in material, electronic, and optical technologies have led
to the development of novel types of electronically tunable filters, including so-­
called adaptive or tunable Focal Plane Arrays (FPAs) [10].

P. Madejczyk (*)
Military University of Technology, Warsaw, Poland
e-mail: pawel.madejczyk@wat.edu.pl

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 453
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_19
454 P. Madejczyk

There are currently several types of detector technologies which offer multicolor
capability over a wide IR spectral range. In the wavelength regions of interest, such as
short-wavelength IR (SWIR: 1–3 μm), medium-wavelength IR (MWIR: 3–5 μm), and
long-wavelength (LWIR: 5–14 μm), there are four technologies that are developing mul-
tispectral detection: HgCdTe [34], quantum well IR photodetectors(QWIP) [13], anti-
monide-based type-II superlattices [31], and quantum dot IR photodetectors (QDIPs) [21].
In this chapter we focus on multicolor detectors from II-VI materials. Among
them only Mercury Cadmium Telluride (MCT, Hg1-xCdxTe, HgCdTe) offers multi-
color capability.

19.2 HgCdTe Multicolor Detectors

HgCdTe is a pseudo-binary alloy semiconductor that crystallizes in the zinc blende


structure. It has a large optical coefficient that enable a high quantum efficiency.
Since firstly synthesized over 60 years ago by British scientists [22] HgCdTe pro-
vides an unprecedented degree of freedom in IR detector design. Because of its band-
gap tunability with composition x, Hg1-xCdxTe has become the most versatile material
for detector applications over the entire IR range [19, 28]. For these reasons the
HgCdTe alloy is also an unique material for designing multicolor IR structures [36].
Historically, multicolor detectors using HgCdTe photoconductors have been
demonstrated in the early 1970s [14]. Worth to mention is a concept of Italian sci-
entists who designed very first photovoltaic multicolor HgCdTe detector integrating
the physically separated single-color samples of different wavelengths into a three-­
band multispectral detector [11]. However, this concept did not contribute to the
development of multicolor focal plane arrays and was only limited to the multicolor
detector as a single element.
New integrated multicolor (multiple cutoffs) detectors have been fashioned from
multiple layer structures, where two or more color detectors are integrated into a
single pixel. Figure 19.1 presents the general idea of a three-color detector pixel.
Shorter wavelengths of infrared flux are absorbed in the upper layer (Absorber 3),
while longer wavelengths are transmitted to lower placed layers and are absorbed
there, successively. Therefore, the energy gap Eg of individual absorbers should be
selected in such a way that the top layer (through which the stream incidences
firstly) has the widest energy gap, that is: Eg1 < Eg2 < Eg3. The barrier layers separate
absorbers preventing or minimizing the photocurrent leaking into the adjacent band.

19.2.1 Dual-Band HgCdTe Detectors

Considering the operation mode, there are two basic kinds of multicolor detectors:
(1) Sequential, called also a bias – selectable detector where the detector bias is
changed to activate one of the colors at a time and (2) Simultaneous, where all col-
ors are detected simultaneously [6].
19 Multicolor Photodetectors 455

Fig. 19.1 Structure of a three-color detector pixel. Shorter wavelengths of infrared flux are
absorbed in the upper layer (Absorber 3), while longer wavelengths are transmitted successively to
lower placed layers

Fig. 19.2 The operating modes of two-color detectors: (a) sequential and (b) simultaneous.
(Pictures present both simplified electrical diagrams and exemplary structures)

Figure 19.2 presents simplified electrical diagrams and exemplary structures of


different operating modes of two-color detectors. Sign plus “+” in layers description
denotes high doping and the capital letter denotes wider bandgap.
The sequential-mode detector has a single indium bump per unit cell that permits
sequential bias selectivity of the spectral bands associated with operating back-to-
back photodiodes. Alternatively, simultaneous detector typically contains multiple
electrical contact per unit cell and are grounded through the substrate on which
detector is grown. Photosignal is extracted directly from one of contacts, while the
second contact, common to both spectral bands, provides the sum of the two photo-
currents. While this architecture provides the benefit of real-time imaging, the
geometry of the pixel reduces the optical fill factor and reduces the maximum
456 P. Madejczyk

achievable quantum efficiency. The challenges involved with fabricating arrays hav-
ing multiple contacts per unit cell further limit the development of simultaneous
detectors. For the above reason sequential detectors have emerged as the favored
two-color technology.
Back-to-back photodiode two-color detector as an integrated pixel was first dem-
onstrated using quaternary III-V alloy (GaxIn1-xAsyP1-y) absorbing layers in a lattice-­
matched InP structure sensitive to two different SWIR bands [4]. A decade later the
original back-to-back concept was designed using HgCdTe at Santa Barbara
Research Center [5] and Rockwell [3]. First successful demonstration of two-color
128 × 128 elements FPA was implemented in liquid phase epitaxial (LPE)-grown
HgCdTe devices [42]. The real progress in the dynamic development of two-color
detectors has taken place since the mid-90s, when advanced technologies were used
to obtain complex HgCdTe heterostructures: molecular beam epitaxy (MBE) and
metal-organic vapor phase epitaxy (MOVPE). There are several scientific centers
that reported dual-band (DB) HgCdTe detectors using epitaxial techniques. MBE
was used at: Rockwell (today Teledyne) [2, 39], Chinese Academy of Sciences
(ChAs) [16], Sofradir-CEA-Leti [7, 32], Raytheon [38, 41] and AIM [9, 43]. In turn,
DB HgCdTe detectors using MOVPE were reported by: Lockheed Martin (today
BAE Systems) [26, 33] and British workers [1, 12, 25, 29].
Both technologies (MBE and MOVPE) reveal their specific capabilities and limi-
tations in the fabrication of sophisticated HgCdTe heterostructures for multi-color
detectors. The main disadvantage of MBE is the difficulty in activating acceptor
impurities in situ, while abrupt interfaces are hardly achievable in MOVPE depos-
ited heterostructures due to a high temperature growth (over 300 °C).
One example of earlier promising implementations of DB MW/LW HgCdTe
detectors was MOVPE grown P-n-N-P structure shown in Fig. 19.3. The cross sec-
tion and the energy band profiles present the dual-band (DB) detector composed

Fig. 19.3 Cross section and energy band profiles for the simultaneous P-n-N-P dual-band (DB)
HgCdTe detector. (Adapted with permission from Ref. [33]. Copyright 1998: SPIE)
19 Multicolor Photodetectors 457

electrically of two back-to-back HgCdTe photodiodes. It was designed for simulta-


neous operation mode. The LW photodiode is a P-on-n heterojunction, grown
directly on top of the MW photodiode, which is an n-on-P heterojunction. A thin
n-type compositional barrier layer is placed between the MW and LW absorber lay-
ers. This barrier layer forms isotype n-N heterojunctions at the interfaces, which
prevent MW photocarriers from diffusing into the LW absorber layer and prevents
LW photocarriers from diffusing into the MW absorber layer. Two indium bump
interconnects in each unit cell provide independent electrical access to the back-to-­
back MW and LW photodiodes, and allow the MW and LW photocurrents to be
separate and independent. Lockheed Martin demonstrated MOVPE grown 64 × 64
FPA with 75 μm pixel pitch reported by [26]. Simultaneous MW/LW DB HgCdTe
detectors were fabricated from a P-n-N-P films grown in situ by interdiffused mul-
tilayer process (IMP) MOVPE onto lattice-matched CdZnTe substrates. Detectivity
at λPEAK (f/2.9) of 4.8 × 1011 cmHz1/2/W for MWIR (4.3 μm cutoff wavelength) and
of 7.1 × 1010 cmHz1/2/W for LWIR (10.1 μm cutoff wavelength) were obtained at
78 K. Simultaneity, separate and independent MW and LW integration within each
unit cell, and full stare efficiency were maintained as key features. Improved R0A
for the HgCdTe detectors and lower operating temperatures allowed near-BLIP per-
formance at low backgrounds. There were promising plans to advance simultaneous
DB HgCdTe FPA technology to larger array sizes, smaller unit cells, and higher
performance but unfortunately this technology was abruptly suspended from
unknown reasons at the end of last century.
In turn Raytheon reported MBE-grown HgCdTe M/LWIR DB 649 × 480 FPA
with 20 μm pixel pitch operating in the sequential mode [41]. Triple-layer n-P-n
heterojunction (TLHJ) device structures were deposited on 100-mm (211)Si sub-
strates. The wafers showed low macrodefect densities (<300 cm2). Typical 81 K
cutoff wavelengths of 5.1 μm for MWIR and 9.6 μm for LWIR were obtained. The
FPAs exhibited high pixel operabilities in each band with NETD operabilities up to
99.98% for the MWIR band and 98.7% for the LWIR band at 81 K, at f/3 background.
Figure 19.4 presents scanning electron microscopy (SEM) images of chosen DB
HgCdTe detectors designed at different technologies and operating in different
modes. The achievements of German technology in the field of two-color detectors
with HgCdTe are also worth emphasizing. Figure 19.4a presents a fragment of
MBE-grown 640 × 512 FPA with 20 μm pixel pitch reported by AIM [43]. Sequential
MW/LW DB HgCdTe detectors were fabricated from a n-p-P-P-N films deposited
on CdZnTe substrates. NEDT was lower than 18 mK and 25 mK for MWIR and
LWIR ranges, respectively. ROIC technology based on analog CMOS let to operate
with frame rate equal 100 Hz at eight outputs. Authors emphasize, that one key
technology for processing narrow and deep mesa trenches an etching technique,
which satisfies, among others, several key requirements: anisotropy, high selectiv-
ity, stoichiometricity, and freedom from damage. Anisotropy and high selectivity
are preconditions to achieve a high aspect ratio and by this a high fill factor. A
requirement just as important for the processing of DB arrays is a technique provid-
ing damage-free etching to avoid layer damage and/or Hg-diffusion. A technique
which meets all these demands is dry etching by inductively coupled plasma (ICP).
458 P. Madejczyk

Fig. 19.4 Scanning electron microscopy (SEM) images of different dual-band (DB) HgCdTe
detector designs: (a) MBE-grown with 20 μm pixel pitch. (Reprinted with permission from Ref.
[43]. Copyright 2011: SPIE), (b) MBE grown with 20 μm pixel pitch. (Reprinted with permission
from Ref. [32]. Copyright 2011: SPIE), (c) MOVPE grown with 24 μm pixel pitch. (Reprinted with
permission from Ref. [29]. Copyright 2008: SPIE), (d) MOVPE grown with 12 μm pixel pitch.
(Reprinted with permission from Ref. [25] Copyright 2019: SPIE)

Another example of MBE possibilities for DB detectors are French scientists’


achievements. Figure 19.4b shows a fragment of MBE-grown 640 × 512 FPA with
24 μm pixel pitch reported by Sofradir-CEA-Leti [32]. Simultaneous or sequential
MW/LW DB HgCdTe detectors were fabricated from a n-p-P-P-N or n-p-P-N films
deposited on substrates issued from homemade 90-mm diameter Cadmium Telluride
(CdTe) ingots sliced into crystal-oriented rectangular wafers. The typical perfor-
mances in each band of the DB FPA are close to what is routinely obtained in single
detectors of the same cut-off wavelength. NEDT was lower than 20 mK and 25 mK
for MWIR and LWIR ranges, respectively. ROIC technology based on CMOS let to
operate with frame rate equal 90 Hz at two analog outputs per band. Spectral cross-
talk was lower than 1% and the operability higher than 99.5% was achieved.
Many original solutions of British scientists made a significant contribution to
the development of DB detectors. Figure 19.4c, d present the micrographs of
MOVPE grown FPAs with 24 μm and 12 μm pixel pitches, respectively reported by
Leonardo (formerly Selex ES). Sequential MW/LW DB HgCdTe detectors were
19 Multicolor Photodetectors 459

fabricated from a n-P-N structures deposited on low cost gallium arsenide (GaAs)
substrates. 640 × 512 FPAs were demonstrated with NEDT parameter lower than
25 mK and the spectral crosstalk lower than 0.2% due to the selective filters applica-
tion. The pixel operability of >99% in each waveband has been achieved. Constant
progress in pixel sizes reduction was reported from 30 μm at initial research to
12 μm at current implementation.
MCT multicolor heterostructures are extremely challenging material, which
presents many different problems to each growth technique. Epitaxial techniques
not only remove many of the hazards associated with the high pressure processes
required for conventional bulk growth but also improve diode performance router
by offering increased area and improved crystalline quality, alloy composition, and
dopant and thickness control, and by enabling the use of semiconductor manufac-
turing processing methods. The ability to control composition, thickness, and dop-
ants within the growth facilitates device engineering opportunities such that
multicolor heterostructure designs can be realized [8, 24]. MBE and MOVPE tech-
nologies compete against each other providing a high quality HgCdTe material for
the production of multi-color detectors. Today, both technologies can offer HgCdTe
heterostructures for DB FPA with comparable final parameters: NEDT <20 mK,
pixel size <20 μm, array size 640 × 512, crosstalk <1% covering IR spectral ranges
required by the industry.
Figure 19.5 shows examples of current – voltage characteristics of DB MCT
detectors. The current-voltage characteristics shown in Fig. 19.5a were measured in
a cryogenic prober on back-to-back test diodes to sample the complete DB diode
stack by changing the bias polarity of the diode from positive to negative. They

Fig. 19.5 Examples of current – voltage characteristics of DB HgCdTe detectors: (a) (50 × 57)
μm2 structures measured at 77 K. (Adapted with permission from Ref. [43]. Copyright 2011:
SPIE), and (b) (400 × 400) μm2 measured at elevated temperatures. (Adapted from Ref. [20].
Published 2019 by Springer as open access)
460 P. Madejczyk

concern (50 × 57) μm2 MBE grown-n-p-P-P-N structures. In Fig. 19.5a left-hand
side, the reverse biased part of the LWIR curve is graphed and on the right-hand
side, the reverse biased part of the MWIR diode. These combined current-voltage
curve represents the well-known shape of back-to-back diodes. The LWIR diode
show an expected small spread in reverse bias, but nevertheless demonstrates rea-
sonable diode characteristics. The performance of the LWIR diodes will be further
improved by appropriate measures concerning layer growth and array processing,
mainly in terms of a further reduction in dislocation density. The I-V curve of the
MWIR diode exhibits a broad plateau in reverse bias demonstrating low leakage
currents and by this an excellent diode performance. Figure 19.5b shows current –
voltage characteristics of (400 × 400) μm2 mesa structures measured at elevated
temperatures [20]. In left-hand side, the reverse biased part of the MW1 curves is
graphed and on the right-hand side, the reverse biased part of the MW2 diode. The
advanced calculations found, that I-V characteristics were shaped with different
generation- recombination mechanisms occurring through trap states related to
metal (mercury) vacancies and the dislocations. The diffusion current strongly
influences I-V characteristics at elevated temperatures.
Figure 19.6 illustrates the examples of normalized spectral responses of DB
HgCdTe detectors from MBE grown structures. Figure 19.6a shows example of
spectral responses of MW/LW detector [43]. The cut-off wavelengths are 5.4 μm for
the MWIR and 9.1 μm for the LWIR band at 60 K. As can be seen, the cross-talk
between the two bands is very low. By exclusion of the wavelength range between
5 and 7 μm, which is of no interest for applications due to atmospheric absorption,
the cross-talk of the two spectral bands is less than 1%. Figure 19.6b presents the
spectral response for DB HgCdTe detector operating at the temperature 230 K. The

Fig. 19.6 Examples of normalized spectral responses of DB HgCdTe detectors: (a) MW/LW type.
(Adapted with permission from Ref. [43]. Copyright 2011: SPIE), and (b) MW1/MW2 type.
(Reprinted from Ref. [23]. Published 2020 by Springer as open access)
19 Multicolor Photodetectors 461

measurements were performed without the bias voltage applying (Ub = 0 V). The
MW1 pins were connected to the spectrophotometer, and the spectral response
characteristic was obtained with the λ50%CO = 5 μm and next the connections were
switched to the MW2 pins and the spectral response was taken with the
λ50%CO = 6.5 μm. The MW2 responsivity is of several percentages lower than that
MW1, because the MW2 absorber is too thin and failed to capture all photons.
Moreover, there is evident gradient of the composition x in the MW2, which disor-
ders the absorption process. There is also the substantial gradient of the arsenic
concentration within the whole area of the MW2 absorber what can contribute to
higher recombination of the minority carriers that mutes the signal [23].
Table 19.1 presents the comparison of DB HgCdTe FPA parameters published
over last 20 years. Results of major scientific centers from USA, UK, France,
Germany and China are given. Worth to note is FPA’s operating temperature oscil-
lating around 77 K.
Raytheon Vision Systems (RVS) developed a 640 × 480 two-color ROIC type
SB-275 based on time division multiplexed integration (TDMI). A photograph of an
SB-275 developed on the basis of MBE grown DB MCT 640 × 480 FPA is shown
in Fig. 19.7. In that solution, the bias polarity on the detector is switched many times
within a single frame period. This allows interlacing of the integration times of the
two spectral bands. Fast subframe switching of less than 1 msec is typically
employed. This achieves good simultaneity between the two spectral bands with
minimal spectral latency. The MWIR band is integrated by summing the charge col-
lected from the individual subframe integration periods. The LWIR band is inte-
grated by averaging the charge collected from the individual subframe integrations.
This approach optimizes the effective charge storage capacities of the two bands.
The SB-275 operates at a frame rate of 30 Hz.
Equally advanced DB IR system was elaborated by British scientists on the basis
of MOVPE grown HgCdTe 640 × 512 FPA [29]. A new silicon readout integrated
circuit has been designed specifically for CONDOR II arrays at Leonardo (Fig. 19.8).
Two capacitors are incorporated into each pixel to allow integration and readout of
both bands in a single frame offering quasi-simultaneous operation. A 4:1 capaci-
tance ratio between the LW and MW bands enables equivalent thermal sensitivity to
be achieved at the same well fill. Several different modes of operation can be used
which make the ROIC highly versatile and suitable for many different applications.
Figure 19.9 presents imaging from CONDOR bispectral detector (F2, 4 ms/0.4 ms
integration time).
These different modes come under four main categories: MW band only, single
stare per frame; LW band only, single stare per frame; alternate wavebands between
frame or frames, one waveband per frame; and interleaved mode where MW and
LW bands are imaged and output within the same detector frame, in any order.
Performance in the two single band modes can be further improved by a feature
which integrates the charge on both capacitors. This feature provides increased
charge storage, improved NETDs and allows some matching to different optical
systems. Interleaved mode offers the greatest flexibility with a choice of integrating
a combination of wavebands several times within a single frame. This capability
462

Table 19.1 Comparison of DB HgCdTe FPA parameters


Center
Parameter Lockheed Martin Leonardo Rockwell ChAS Sofradir/CEA-Leti Raytheon AIM
Growth MOVPE MBE
Structure p-n-N-N-P n-P-N p-n-N-P-N n-p-P-P-N n-p-P-P-N/n-p-P-N n-p-n n-p-P-P-N
Mode Simultaneous Sequential Simultaneous Simultaneous Simultaneous/sequential Sequential Sequential
Spectral range 3–4.3 3–6.5 3–3.9 2–4.8 3–5 3–5.1 3–5
(Band 1) (μm)
Spectral range 4.3–10 6–11 3.9–5 5–9.7 8–9.5 5.1–9.6 8–9.5
(Band 2) (μm)
Array 64 × 64 640 × 512 128 × 128 128 × 128 640 × 512 649 × 480 640 × 512
Pixel pitch (μm) 75 12 40 50 24 20 20
Operating temperature 78 K 80 K 78 K 78 K 77 K 78 K 77 K
NEDT (mK) 12–20 25 9.3 10a 15–20 39 25
Band 1
NEDT (mK) 6.2–7.5 25 13.3 10a 20–25 26 18
Band 2
Cross talk 10% 0.2% 3–6% 0.7–1.25 <1% 1–10% <1%
Reference [33] [25] [39] [16] [32] [41] [43]
a
Estimated values on the basis of the Detectivity
P. Madejczyk
19 Multicolor Photodetectors 463

Fig. 19.7 MBE grown DB HgCdTe 640 × 480 FPA type SB-275 in which the detector bias polar-
ity is alternated many times within a single frame period. (Reprinted with permission from Ref.
[30]. Copyright 2005: SPIE)

Fig. 19.8 CONDOR II


dual waveband MW/LW
integrated detector cooler
assembly. MOVPE grown
HgCdTe 640 × 512 array is
inside the assembly.
(Reprinted with permission
from Ref. [29]. Copyright
2008: SPIE)
464 P. Madejczyk

Fig. 19.9 Imaging from CONDOR bispectral detector (F2, 4 ms/0.4 ms integration time).
(Reprinted with permission from Ref. [24]. Copyright 2020: John Wiley & Sons Ltd)

provides a mechanism to improve temporal alignment between the integration times


as MW integration times are invariably longer than LW. The system configures and
controls the detector operation and function dynamically through a high speed digi-
tal serial interface. The increased amount of signal information from the array is
output from the detector using eight buffered outputs to enable higher frame rates of
up to 120 Hz in single waveband modes and 60 Hz in dual waveband modes to be
achieved. The ROIC also provides other read out modes, windowing and with con-
figurable frame size, scan direction and comprehensive windowing functions.

19.2.2 Three-Color HgCdTe Detectors

A natural path in the evolution of multi-color detectors is the development of three


color detectors. Collection of signals from more than two infrared bands provides
enhanced target discrimination and identification. So far, there are only a few works
devoted to the practical implementation of triple band detectors with HgCdTe pre-
sented by British researchers [15, 17]. The bias dependent cut-off is achieved by
employing three absorbers in an n-p-n structure with low p-doped electronic barri-
ers at the junctions, see Fig. 19.10. The first n-region (in direction of radiation)
defines the shorter wave (SW) side; the p-region, the intermediate wave (IW)
response; and the top n-layer the longer wave (LW) response. At low biases either
the SW or LW response would dominate, depending on which junction is reverse
biased, in a similar way to a 2-color detector. In this bias range, the barriers prevent
electron flow from the IW region from both the photogenerated carriers from IW
absorption, and by direct injection from the forward biased junction. As the barrier
region is low doped, any applied bias will predominately fall on this side of the
junction. Increasing reverse bias will, therefore, reduce the barrier until eventually
energetic electrons generated by IW photons can cross the junction.
Figure 19.11 shows the obtained spectral response of a three-color MCT detector
at various biases. The cut-on wavelength at 2.33 μm is given by a coated Ge win-
dow. In positive bias the LW/IW junction is in reverse bias and a bias independent
19 Multicolor Photodetectors 465

Fig. 19.10 Three-color concept and associated zero-bias energy band diagram. (Adapted with
permission from Ref. [15]. Copyright 2006: SPIE)

Fig. 19.11 Spectral response of a three-color MCT detector at various biases. (Adapted with per-
mission from Ref. [15]. Copyright 2006: SPIE)

LW spectrum is obtained above 0.2 V. The doping levels chosen result in no barrier
lowering at the LW/IW junction at these applied biases. The response below 4.2 μm
is due to incomplete absorption in the IW absorber resulting in carrier generation in
the LW absorber at these wavelengths. Carriers generated in the IW absorber have
insufficient energy to surmount the LW barrier.
466 P. Madejczyk

As the positive bias is reduced to below 0.2 V, the LW signal collapses and a
signal from the SW side begins to appear with the current flowing in the opposite
direction. In this regime the built-in fields dominate the behavior, and the largest
field is at the SW/IW junction due to the larger band-gaps. Further reduction in the
bias to 0 V causes the SW response to grow. Changing the bias polarity to negative
puts the SW/IW junction into reverse bias, giving the SW response with cut-off
2.8 μm. Increasing the bias magnitude lowers the electron barrier at this junction
and allows a response from the IW absorber, moving the cut-off out to 4.2 μm as
shown at −0.6 V in Fig. 19.11. The increase in the SW signal with increasing nega-
tive bias is due to incomplete SW absorption in the SW absorber.
Because of the complicated and expensive fabrication process, numerical simu-
lation has become a critical tool for the development of HgCdTe bandgap-­engineered
devices. It was shown, that the performance of a three-color detector is critically
dependent on the barrier doping level and position in relation to the junction [18]. A
small shift of the barrier location and doping level causes significant changes in
spectral responsivity. This serious disadvantage of the considered three-color detec-
tor should be eliminated in further investigation. To ensure adequate barrier lower-
ing, the barrier must be low doped and of the same polarity as the intermediate
absorber, and be located adjacent to the junction. The MOVPE growth process is
shown to readily achieve such control.

19.3 Conclusions

The presented analysis of the literature on infrared detectors shows a huge progress
in improving the parameters of multi-color detectors with HgCdTe over the last
30 years. Significant progress has been observed in the crosstalk reducing, increas-
ing the size of arrays, reducing pixel sizes and increasing the signal-to-noise ratio as
defined by the NEDT parameter. The dynamic development of multi-color detectors
is related to the progress in advanced epitaxial techniques: MBE and MOVPE. The
improvements in processing techniques: the photolithography, dry etching and pas-
sivation, as well as the readout electronics (ROIC) and the compound image analy-
sis tools achievements are equally important. Multi-color detectors developed so far
operate basically at cryogenic temperatures what is their general disadvantage
related to the cost, weight, size and reliability. So, future studies should focus on the
operation temperature increasing. The ultimate performance potential of MCT will
ensure that it is the material of choice for all high performance infrared systems
including multispectral detectors.

References

1. Abbott P, Pillans L, Knowles P, McEwen RK (2010) Advances in dual-band IRFPAs made


from HgCdTe grown by MOVPE. Proc SPIE 7660:766035-1-11
19 Multicolor Photodetectors 467

2. Almeida LA, Thomas M, Larsen W, Spariosu K, Edwall DD, Benson JD, Mason W, Stoltz AJ,
Dian JH (2002) Development and fabrication of two-color mid- and short-wavelength infrared
simultaneous unipolar multispectral integrated technology focal-plane arrays. J Electron Mater
30(7):669–676
3. Blazejewski ER, Arias JM, Williams GM, McLevige W, Zandian M, Pasko J (1992) Bias-­
switchable dual-band HgCdTe infrared photodetector. J Vac Sci Technol B 10:1626
4. Campbell JC, Dentai AG, Lee TP, Burrus CA (1980) Improved two-wavelength demultiplex-
ing InGaAsP photodetector. IEEE J Quantum Electron 16:601
5. Casselman TN, Walsh DT, Myrosznyk JM, Kosai K, Radford WA, Schultz EF, Wu OK (1990)
An integrated multispectral IR detector structure. In: Extended abstracts of the U.S. workshop
on the physics and chemistry of Mercury Cadmium Telluride, San Francisco, 2–4 October 1990
6. Casselman TN (1997) State of infrared photodetectors and materials. Proc SPIE 2999:1–10
7. Destefanis G, Baylet J, Ballet P, Castelein P, Rothan F, Gravrand O, Rothman J, Chamonal
JP, Million A (2007) Status of HgCdTe bicolor and dual-band infrared focal plane arrays at
LETI. J Electron Mater 36(8):1031–1044
8. Dvoretsky SA, Mikhailova NN, Remesnik VG, Sidorov YG, Shvets VA, Ikusov DG et al
(2019) MBE-grown MCT hetero- and nanostructures for IR and THz detectors. Opto-Electron
Rev 27(3):282–290
9. Eich D, Ames C, Breiter R, Figgemeier H, Hanna S, Lutz H, Mahlein KM, Schallenberg
T, Sieck A, Wenisch J (2019) MCT-based high performance bispectral detectors by AIM. J
Electron Mater 48(10):931–936
10. Faraone L (2005) MEMS for tunable multi-spectral infrared sensor arrays. Proc SPIE
5957:59570F
11. Fiorito G, Gasparrini G, Svelto F (1976) Multispectral Hg1-xCdxTe photovoltaic detectors.
Infrared Phys 16:531–534
12. Gordon NT, Abbott P, Giess J, Graham A, Halis JE, Hall DJ, Hipwood L, Jones CL, Maxey
CD, Price J (2007) Design and assessment of metal-organic vapour phase epitaxy-grown dual
wavelength infrared detectors. J Electron Mater 36(8):931–936
13. Gunapala SD, Bandara SV, Liu JK, Mumolo JM, Ting DZ, Hill CJ, Nguyen J, Rafol SB (2010)
Demonstration of 1024x1024 pixel dual-band QWIP focal plane array. Proc SPIE 7660:76603L
14. Halpert H, Musicant BI (1972) N-color (Hg, Cd)Te photodetectors. Appl Opt 11:2157–2161
15. Hipwood LG, Jones CL, Maxey CD, Lau HW, Fitzmaurice J, Catchpole RA, Ordish M (2006)
Three-color MOVPE MCT diodes. Proc SPIE 6206:620612
16. Hu W, Ye Z, Liao L, Chen H, Chen L, Ding R, He L, Chen X, Lu W (2014) 128 × 128
long-wavelength/mid-wavelength two-color HgCdTe infrared focal plane array detector with
ultralow spectral cross talk. Opt Lett 39(17):5184–5187
17. Jones CL, Hipwood LG, Price J, Shaw CJ, Abbott P, Maxey CD, Lau HW, Catchpole RA, Ordish
M, Knowles P (2007) Multi-colour IRFPAs made from HgCdTe grown by MOVPE. Proc SPIE
6542:654210
18. Jóźwikowski K, Rogalski A (2007) Numerical analysis of three-colour HgCdTe detectors.
Opto-Electron Rev 15:215–222
19. Kinch MA (2014) State-of-the-art infrared detector technology. SPIE Press, Bellingham
20. Kopytko M, Gawron W, Kębłowski A, Stępień D, Martyniuk P, Jóźwikowski K (2019)
Numerical analysis of HgCdTe dual-band infrared detector. Opt Quant Electron 51:62
21. Krishna S, Forman D, Annamalai S, Dowd P, Varangis P, Tumolillo T, Gray A, Zilko J, Sun K,
Liu M, Campbell J, Carothers D (2006) Two-color focal plane arrays based on self-assembled
quantum dots in a well heterostructure. Phys Status Solidi (c) 3(3):439–443
22. Lawson WD, Nielson S, Putley EH, Young AS (1959) Preparation and properties of HgTe and
mixed crystals of HgTe-CdTe. J Phys Chem Solids 9(3–4):325–329
23. Madejczyk P, Gawron W, Kębłowski A, Mlynarczyk K, Stępień D, Martyniuk P, Rogalski A,
Rutkowski J, Piotrowski J (2020) Higher operating temperature IR detectors of the MOCVD
grown HgCdTe heterostructures. J Electron Mater 49:6908–6917
24. Maxey CD, Capper P, Baker IM (2020) MOVPE growth of cadmium mercury telluride and
applications. In: Irvine S, Capper P (eds) Metalorganic Vapor Phase Epitaxy (MOVPE):
growth, materials properties, and applications. Wiley, Hoboken, pp 293–324
468 P. Madejczyk

25. McEwen KR, Hipwood L, Bains S, Owton D, Maxey C (2019) Dual waveband infrared detec-
tors using MOVPE grown MCT. Proc SPIE 11002:1100218-1–1100218-6
26. Mitra P, Barnes SL, Case FC, Reine MB, O’Dette P, Starr R, Hairston A, Kuhler K, Weiler
MH, Musicant BL (1997) MOCVD of bandgap – engineered HgCdTe p-n-N-P dual – band
infrared detector arrays. J Electron Mater 26(6):482–487
27. Norton P (2002) HgCdTe infrared detectors. Opto-Electron Rev 10(3):159–174
28. Piotrowski J, Rogalski A (2007) High-operating temperature infrared photodetectors. SPIE
Press, Bellingham
29. Price JPG, Jones CL, Hipwood LG, Shaw CJ, Abbott P, Maxey CD et al (2008) Dual-band
MW/LW IRFPAs made from HgCdTe grown by MOVPE. Proc SPIE 6940:69402S
30. Radford WA, Patten EA, King DF, Pierce GK, Vodicka J, Goetz P et al (2005) Third generation
FPA development status at Raytheon vision systems. Proc SPIE 5783:331–339
31. Razeghi M, Dehzangi A, Li J (2021) Multi-band SWIR-MWIR-LWIR Type-II superlattice
based infrared photodetector. Results Opt 2:100054
32. Reibel Y, Chabuel F, Vaz C, Billon-Lanfrey D, Baylet J, Gravrand O, Ballet P, Destefanis G
(2011) Infrared dual band detectors for next generation. Proc SPIE 8012:801238-1-13
33. Reine MB, Hairston A, O’Dette P, Tobin SP, Smith FTJ, Musicant BL, Mitra P, Case FC (1998)
Simultaneous MW/LW dual-band MOCVD HgCdTe 64 × 64 FPAs. Proc SPIE 3379:200–212
34. Rogalski A (2000) Dual-band infrared detectors. Proc SPIE 3948:17–30
35. Rogalski A (2019) Infrared and terahertz detectors, 3rd edn. CRC Press, Boca Raton
36. Rutkowski J, Madejczyk P, Piotrowski A, Gawron W, Jóźwikowski K, Rogalski A (2008) Two-­
colour HgCdTe infrared detectors operating above 200K. Opto-Electron Rev 16:321–327
37. Scribner D, Schuler J, Warren P, Satyshur M, Kruer M (1998) Infrared color vision: separating
object from backgrounds. Proc SPIE 3379:2–13
38. Smith E, Venzor G, Gallagher A, Reddy M, Peterson J, Lofgreen D, Randolph J (2011) Large-­
format. J Electron Mater 40(8):1630–1636
39. Tennant WE, Thomas M, Kozlowski LJ, McLevige WV, Edwall DD, Zandian M et al (2001)
A novel simultaneous unipolar multispectral integrated technology approach for HgCdTe IR
detectors and focal plane arrays. J Electron Mater 30(6):590–594
40. Vallone M, Goano M, Tibaldi A, Hanna S, Eich D, Sieck A, Figgemeier H, Ghione G, Bertazzi
F (2020) Challenges in multiphysics modeling of dual-band HgCdTe infrared detectors. Appl
Opt 59(19):5656–5663
41. Vilela MF, Olsson KR, Norton EM, Peterson JM, Rybnicek K, Rhiger DR et al (2013)
High-performance M/LWIR dual-band HgCdTe/Si focal-plane arrays. J Electron Mater
42(11):3231–3238
42. Wilson JA, Patten EA, Chapman GR, Kosai K, Baumgratz B, Goetz P et al (1994) Integrated
two-color detection for advanced FPA applications. Proc SPIE 2274:117–125
43. Ziegler J, Eich D, Mahlein M, Schallenberg T, Scheibner R, Wendler J et al (2011) The devel-
opment of 3rd gen IR detectors at AIM. Proc SPIE 8012:801237-1-13
Chapter 20
Flexible Photodetectors Based on II-VI
Semiconductors

Mingfa Peng and Xuhui Sun

20.1 Introduction

Photodetector as an important component that can invert incident light into electric
signals have attracted intensive research interest and exhibited various promising
applications, including optical communication, imaging technique, environment
monitoring, remote sensing, etc. [1–4]. Conventional photodetectors are usually
fabricated on rigid substrates or wafers using inorganic semiconductor materials as
functional materials, such as three-dimensional (3D) bulk crystalline Si, InGaAs,
HgCdTe and related heterostructures in high-performance visible and infrared pho-
todetectors [5, 6]. However, to obtain high-performance photodetectors, these func-
tional materials should be fabricated with large thickness due to their low light
absorption coefficients, and thus resulting in complex and expensive fabrication
process and hindering their application in flexible electronics. Compared to photo-
detectors on rigid substrates, flexible photodetectors exhibit many advantages such
as a good bendability, foldability and even stretchability as well as weight light,
which have triggered a widely concerned in wearable electronics including wear-
able monitoring, wearable image sensing, self-powered integrated electronics, etc.
[7–10]. For flexible photodetectors, not only the “5S” key parameters including
photoresponse sensitivity, signal-to-noise ratio, speed, selectivity, and stability, but

M. Peng
School of Electronic and Information Engineering, Jiangsu Province Key Laboratory of
Advanced Functional Materials, Changshu Institute of Technology,
Changshu, Jiangsu, People’s Republic of China
X. Sun (*)
Institute of Functional Nano & Soft Materials (FUNSOM), and Jiangsu Key Laboratory for
Carbon-based Functional Materials and Devices, Soochow University,
Suzhou, Jiangsu, People’s Republic of China
e-mail: xhsun@suda.edu.cn

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 469
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_20
470 M. Peng and X. Sun

also the mechanical flexibility should be taken into account to evaluate the perfor-
mance of the devices [11, 12]. Therefore, the flexible substrates, device structure
and functional materials are three important factors for flexible photodetectors. Up
to now, several flexible substrates have been employed to construct flexible photo-
detectors, for instance, paper, polyethylene terephthalate (PET), polyethylene naph-
thalate (PEN) and polyimide (PI), etc. [13–16]. The stability and flexibility of these
substrates should be paid more attention. In addition, the flexible Si based photode-
tectors can also be fabricated by reducing the thickness of Si film to tens to hundreds
of nanometres, leading to the low light absorption [7]. Moreover, different device
structures including planar photoconductive and vertical photodiode types have
been demonstrated in high-performance flexible photodetectors [15, 17].
Apart from the flexible substrates and device structures, the functional materials
as a building blocks play a key role in determining the photoelectronic performance
of the flexible photodetectors. In recent years, various functional materials have
been exploited and demonstrated in flexible photodetectors, including organic semi-
conductors, metal halide perovskites, two-dimensional (2D) nanostructures, etc. [7,
9, 18]. As an alternative, low dimensional II-VI semiconductors which can be
divided into zinc chalcogens (e.g. ZnO and ZnS, etc.) and cadmium chalcogens (e.g.
CdS, CdSe, and CdSxSe1-x, etc.) exhibit promising applications in optoelectronic
devices due to their unique characteristics, such as direct bandgap, excellent optical
and electric properties, and high quantum efficiency [19, 20]. For example, the opti-
cal properties and bandgap structure of the 0D II-VI nanostructures could be easily
tuned by adjusting the size and shape of the nanostructures [21, 22]. 2D II-VI nano-
structures in the nature with covalent bonds in all three dimensions exhibit many
distinct properties different from 2D layered materials, such as abundant surface
dangling bands, high-activity and high-energy surface [23, 24]. So far, different
dimensional II-VI semiconductors, such as 0 D nanocrystals (NCs) and quantum
dots (QDs) [25, 26], 1D nanowires (NWs) and nanoribbons (NRs) [27, 28], 2D
nanosheets and related heterostructures [29], have been successfully exploited and
demonstrated in flexible photodetectors due to their unique properties. However, the
photoresponse of the single II-VI nanostructures based flexible photodetectors just
only focus on UV (e.g. ZnS) or visible (e.g. CdS and CdSe) light region due to their
intrinsic optical bandgap resulting in narrow wavelength range of light absorption,
which will limit their potential application in some specific field, such as wide spec-
trum light detection, NIR light imaging, and night vision technique, etc. Furthermore,
the II-VI semiconductors hybrid structure or heterostructure can be easily fabricated
on flexible substrate, which can be employed as functional materials for flexible
photodetectors [30, 31], thus improving the photoelectric performance and realizing
broadband or NIR photoresponse.
With the rapid development of the Internet of Things (IOTs), flexible photodetec-
tors exhibit a promising application in wearable electronics. However, the progress
review of flexible photodetectors based on II-VI semiconductors are very few. In
this chapter, we introduce the most recent progress on low dimensional II-VI semi-
conductors (0D, 1D, 2D and their related heterostructures) based flexible photode-
tectors and their application in wearable electronic. Firstly, the sensing mechanisms
20 Flexible Photodetectors Based on II-VI Semiconductors 471

and key figures of merits for photodetectors have been summarized and discussed,
which will help us to better understand and study various photodetectors. Then, the
typical flexible photodetectors based on different dimensional II-VI semiconductor
nanostructures have been introduced, in which the functional materials synthesis
method have also been discussed. We mainly focus on the device structure and fab-
rication process of the II-VI semiconductors based flexible photodetector, as well as
the mechanical flexibility and electric stability. In this review, we also discuss the
design of various hybrid nanostructure or heterostructure for the flexible photode-
tectors to improve the photoelectric performance and realize broadband or NIR pho-
toresponse. Furthermore, various applications of the II-VI semiconductors based
flexible photodetectors have been summarized, including wearable monitoring sen-
sors, image sensors, and self-powered integrated wearable electronics. At last, we
summarize the review and discuss the challenges and perspectives of the II-VI semi-
conductors based flexible photodetectors.

20.2 Device Structure and Substrate Materials


of Flexible Photodetectors

It has witnessed many important progresses in architecture design and fabrication of


flexible photodetectors in the past decades. Generally, photodetectors can be classi-
fied as photoconductors, photodiodes and phototransistors according to their differ-
ent working mechanisms. Through architectural design, various featured
photodetectors could be obtained. In addition, the mechanical flexibility is a signifi-
cant factor to evaluate the performance of the flexible photodetectors. Therefore, the
device fabrication and the selection of flexible substrate materials are also very
important for flexible photodetectors. In this section, we will mainly introduce the
device structure and flexible substrate materials.

20.2.1 Device Structure

Photodetectors can be divided into lateral architecture and vertical architecture in


terms of their different spatial layouts between electrodes and photoactive materi-
als. Compared to the lateral structure photodetectors, the vertical structure devices
exhibit shorter carrier transport lengths due to the smaller electrodes spacing.
Photodiode is a typical vertical configuration structure device (Fig. 20.1), in which
the photoactive materials were designed and fabricated between the bottom and top
electrodes. Therefore, the photodiodes usually exhibit a faster response speed, an
improved detectivity and high sensitivity [32]. However, the flexible transparent
conductive materials should be employed as electrodes for this vertical configura-
tion device and thus bring about complex processes for the flexible photodiode
fabrication.
472 M. Peng and X. Sun

Fig. 20.1 Architectures of the typical flexible photodetectors

Compared with photodiode, the device structure of the photoconductor with lateral
architecture is simple and easy to fabricate. The device architecture of the photocon-
ductor is composed with photoactive materials as channel and two symmetrical elec-
trodes as source and drain. Therefore, the flexible photodetector could be easily
fabricated by using various metal electrodes rather than only transparent electrodes. In
addition, photoconductors could obtain high photocurrent and gain but relatively low
detectivity due to the high dark current resulting from the formation of ohmic contact
and their unbalance transport process of the photogenerated charge carriers [18].
Phototransistor is similar with the configuration of the field-effect-transistors
(FETs) and can be regarded as a special case of the photoconductor [33], which is
also a lateral architecture device with a dielectric layer and gate electrode as floating
gate to modulate the conductivity of the semiconductors effectively. Therefore, the
lifetime of the photo-carriers of phototransistors has been prolonged and thus lead
to a higher gain but slower response speed than photoconductors.

20.2.2 Substrate Materials

In order to obtain high performance flexible photodetectors, substrate materials,


electrodes and photoactive materials should be taken into account and designed.
Recently, various materials including polymer, paper and fiber have been employed
as substrates for flexible photodetectors. In addition, silicon can be exploited as
flexible substrates by reducing the thickness to tens to hundreds of nanometres.
Furthermore, mica can also be used as substrate materials to construct flexible pho-
todetector. In this part, we will mainly introduce the substrate materials in flexible
photodetectors.
Polymer materials including polyethylene terephthalate (PET), polyimide (PI),
polyethylene naphthalate (PEN), and polydimethylsiloxane (PDMS) have been
20 Flexible Photodetectors Based on II-VI Semiconductors 473

Fig. 20.2 (a, b) CdS nanowires based flexible photodetector on PI substrate. ((a, b) Reprinted
with permission from Ref. [34]. Copyright 2015: American Chemical Society). (c, d) ZnS-MoS2
heterostructure based flexible photodetector on paper substrate. ((c, d) Reprinted with permission
from Ref. [36]. Copyright 2017: Wiley-VCH). (e, f) n-TiO2/p-CuZnS heterostructure based flexi-
ble photodetector on Ti fiber substrate. ((e, f) Reprinted with permission from Ref. [38]. Copyright
2018: Wiley-VCH)

widely utilized as substrates for flexible photodetectors due to their intrinsic flexi-
bility and mechanical stability. The photoactive materials can be directly placed on
the polymer substrates by various strategies including spin-coating, transferring,
magnetron sputtering, and printing, etc. For instance, Li et al. [34] directly trans-
ferred the synthesized 1D CdS NWs on PI substrates by the contact printing method
to fabricate flexible photodetector (Fig. 20.2a). After bending 1500 cycles, the per-
formance of the PI substrate based flexible photodetectors remained almost
474 M. Peng and X. Sun

unchanged compared with that of the device before bending (Fig. 20.2b). In addi-
tion, Lou et al. [35] proposed a flexible photodetector by transferring 1D ZnS/CdS
heterostructure on PET substrate. After bending at different conditions, the photo-
current of the flexible photodetectors exhibit no clear change indicating the good
mechanical and electrical stability of the PET substrate based flexible photodetec-
tors. The advantages of the polymer materials usually as the substrate of flexible
photodetectors are as following: (1) The intrinsic flexibility and easy availability,
(2) The potential large-scale fabrication, (3) The precisely controlled of the shape
and thickness.
Paper materials as another alternative have also been employed as substrates for
flexible photodetectors because of intrinsic properties, such as mechanical flexibil-
ity, low cost and easy availability, and environmental-friendly. The paper substrate
exhibits a good adhesion with the photoactive materials because of the inherent
porous structure and hygroscopicity of the paper, therefore the design and fabrica-
tion of the paper substrate based flexible photodetectors are similar to that of the
polymer substrates based devices. For instance, Gomathi et al. [36] demonstrated a
paper substrate based flexible broadband photodetector with the ZnS/MoS2 hetero-
structure as photoactive material by using a simple two-step hydrothermal method,
in which the layered MoS2 was firstly grown on cellulose paper followed by synthe-
sis of ZnS on MoS2 (Fig. 20.2c, d). Gou et al. [37] proposed a high-performance
flexible photodetector by depositing P3HT/CdS/CdS:SnS2 nanowires hybrid films
on the paper substrate. After bending 200 cycles, the photocurrent and responsivity
of the paper substrate based flexible photodetector only decline 9%, revealing good
mechanical flexibility and electrical stability of the paper substrate based flexible
photodetector.
Apart from the aforementioned substrate materials, fiber has also been exploited
as a promising candidate for constructing 1D flexible photodetector because of its
inherent flexibility and wearability. A typical fiber substrate based flexible n-TiO2/p-­-
CuZnS heterostructure photodetectors have been demonstrated by Xu et al. [38], in
which Ti microfiber can be employed as flexible substrate as well as one electrode
due to its flexibility and high electrical conductivity, n-TiO2/p-CuZnS/ heterostruc-
ture as photoactive materials and CNT fiber as another electrode (Fig. 20.2e, f).
After bending 200 cycles, optoelectronic properties of the flexible device remained
stable owing to the bendable properties of the fiber substrate based flexible photo-
detector. Furthermore, a double-twisted broadband perovskite flexible photodetec-
tor based on fiber substrate was also proposed by Sun et al. [39]. The double-twisted
flexible photodetector exhibited excellent flexibility and bending stability even after
bending 60 cycles at the bending angle of 90.

20.3 Flexible Photodetectors Based on II-VI Semiconductors

Various II-VI nanostructures have been widely employed in photodetectors as pho-


toactive materials due to their intrinsic direct band-gap characteristics and excellent
optical and electrical properties. In the following section, we will discuss the
20 Flexible Photodetectors Based on II-VI Semiconductors 475

flexible photodetectors based on low-dimensional (0D, 1D and 2D) II-VI semicon-


ductors and address their photoelectric performances.

20.3.1 0D II-VI Nanostructures Based Flexible Photodetectors

0D nanostructures including nanocrystals (NCs) and quantum dots (QDs) were usu-
ally synthesized by solution routing. The optical properties and bandgap structure of
the 0D nanostructures could be easily tuned by adjusting the size and shape of the
nanostructures during the synthesis process. For example, the PL lifetimes and their
absorption and emission spectra of the core-shell CdSe/CdS QDs could be success-
fully adjusted by altering the thickness of the CdS shells [40]. Therefore, the solu-
tion processed 0D II-VI nanostructures exhibits many advantages in the fabrication
of flexible photodetectors, including low-temperature synthesis process, large-scale
production, and compatibility with flexible substrates, etc. Recently, various 0D
II-VI nanostructures have been exploited as sensing materials for the flexible pho-
todetectors, such as ZnS NPs, CdS QDs, CdSe QDs, etc. For instance, Yang et al.
[41] proposed a flexible photodetector based on a hybrid film of graphene/CdS
quantum dots by using chemical bath deposition method. The fabricated visible
photodetectors exhibit high performance due to excellent interface between the gra-
phene and the in suit growth of CdS QDs. The photodetectors exhibit good photo-
electric properties with the responsivity up to 40 A/W at the wavelength of 450 nm.
In addition, some universal method such as spin-coating and dip-coating have been
explored to fabricate the flexible photodetector [42, 43]. For example, Hsiao et al.
[44] fabricated a flexible photodetector on PEN substrate with s with CdSe/ZnS
core-shell QDs coated on ZnO nanorod arrays as photoactive materials. The respon-
sivity of the heterostructure photodetectors increased from 3.2 to 73.2 A/W with
increasing the cover of the core-shell QDs from 0 to 1.00% at the wavelength of
380 nm (Fig. 20.3c). The significant enhancement of the optoelectronic perfor-
mance can be attributed to the high surface-to-volume ratios of the ZnO/QDs nano-
structure easily providing a directed path for electronic transport. Xia et al. [45]
proposed solar-blind deep ultraviolet photodetectors by spin-coating colloidal ZnS
quantum dots (QDs) solution on the electrode device. The photodetectors exhibit
fast response (tr = 0.35 s, td = 0.07 s) and good responsivity of ~0.1 mA/W at 265 nm
light illumination. The further study indicated that the response of the ZnS QD pho-
todetectors can be attributed to the drift of the photogenerated carriers due to the
sensitivity of the ZnS QDs thin film to their surface states. However, all these pro-
cess based flexible photodetectors will need expensive and time-consuming pro-
cess, such as photolithography and lift-off technique.
Motivated by these problem, various novel techniques including all inject print-
ing [44], screen printing [46], roll-to-roll printing [47], three-dimensional (3D)
printing [48], and laser direct writing method [49] as alternative routes have been
employed in flexible electronics and exhibit many advantages, such as rapid proto-
typing, high precision, and energy efficiency, etc. For instance, Lin et al. [49]
476 M. Peng and X. Sun

Fig. 20.3 (a) Schematic illustration of the device with double hole transport layer (HTL). (b)
Energy band diagram of the photodetector with double HTL (c) Responsivity and (d) detectivity
of the single HTL and double HTL devices as a function of wavelength. (Reprinted with permis-
sion from Ref. [50]. Copyright 2019: Royal Society of Chemistry)

demonstrated flexible ZnS/SnO2 ultraviolet photodetectors by using a mask-free


laser direct writing method with ZnS/SnO2 heterostructure as photoactive materials,
PI as flexible substrate, and highly conductive graphene obtained from the under-
neath PI by CO2 laser irradiation as the lateral electrodes. The devices exhibited
high optoelectronic performance resulted from the good interfaces between the gra-
phene electrodes and the photoactive materials of ZnS/SnO2 by this in suit growth
method. The photocurrent of the devices has no obvious degradation after 500 bend-
ing cycles, presenting the excellent flexibility due to the reduced thickness of the
device by lateral electrode structure. Subsequently, Shen et al. [50] proposed a verti-
cal architecture flexible self-powered CdSexTe1-x QD photodetector with a double
hole transport layer (HTL) of PEDOT:PSS/P-TPD (Fig. 20.3a, b). The photodetec-
tion capacity was extended significantly from UV to NIR region due to the well
matched energy lever between P-PPD layer and the PEDOT:PSS/QD layers. The
flexible QDs based photodetectors exhibit a low dark current density of
1.03 × 10−6 mA/cm2, and a large specific detectivity of ~2.6 × 1012 Jones at 450 nm
light illumination (Fig. 20.3c, d). More importantly, the flexible photodetectors
exhibited excellent mechanical and electrical stability even after 150 cycles bending
at different bending angles.
20 Flexible Photodetectors Based on II-VI Semiconductors 477

20.3.2 1D II-VI Semiconductors Based


Flexible Photodetectors

1D semiconductor nanostructures, including nanowires (NWs) [51], nanoribbons


(NRs) [52] and nanotubes (NTs) [53], have drawn much attention as functional
building blocks in nanoscale electronic and optoelectronic devices due to their
unique optical and electrical properties. Compared with other nanostructure, 1D
nanostructures have a simple dimensional characteristic, which make it easy to real-
ize quantitative analysis of some important parameters for optoelectronics. In addi-
tion, 1D nanostructures have been beneficial to the charge carrier transportation and
thus exhibit excellent photoelectronic performance because of their large surface to
volume ratio and high quality crystal structure as well. Furthermore, 1D nanostruc-
tures have an intrinsic mechanical flexibility and easy to fabricate flexible photode-
tectors due to their above characteristics. Up to now, various 1D II-VI semiconductors
have been employed as functional materials in flexible photodetectors, such as ZnS
NWs [36, 49, 54], CdS NWs [55, 56], CdSe NRs [57, 58], CdSxSe1-x NWs or NRs
[27, 59], and their hybrid nanostructures [60–62]. For instance, Zhang et al. [63]
proposed core-shell ZnS/InP nanowires based ultraviolet flexible photodetectors
with a high photoconductive gain of 4.36 × 102, a high detectivity of 5.10 × 1012
Jones, and fast response speed of 0.74 s/0.38 s, respectively. The excellent optoelec-
tronic performance can be attributed to the formation of core-shell heterostructures.
Furthermore, the results also indicated that the flexible photodetectors have excel-
lent mechanical flexibility and electrical stability by bending the device even for
1200 times at different bending angles.
Recently, various 1D visible light absorption sensor materials of CdS, CdSe or
ternary CdSxSe1-x have also been widely investigated in photodetectors. For instance,
Shen et al. employed hierarchical CdS nanowires as building block to fabricate flex-
ible photodetector with ultrahigh sensitivity [34]. The results indicate that the CdS
NWs based flexible photodetectors exhibit a high current on/off ratio larger than
2500, the response time of 0.2 s at 407 nm, as well as a super mechanical flexibility
and electric stability even after 1500 cycles of bending. Sun et al. synthesized 1D
ternary CdSxSe1-x NRs and demonstrated them in flexible visible photodetector with
an ultrahigh current on/off ratio of 4 × 106, a high responsivity of 1.24 × 103 A/W,
and excellent mechanical stability as well [27]. Moreover, organic-inorganic hybrid
nanostructure has also been demonstrated in flexible photodetectors by Shen’s
group. They proposed a hybrid flexible photodetectors by introducing high hole-­
transport rate and strong absorption materials of P3HT into high electrical conduc-
tivity materials of CdSe nanowires. The hybrid device exhibits an extremely high
on/off switching ratio larger than 500, fast response time of 10 ms, and excellent
mechanical flexibility and stability as well. Recently, Lin et al. [64] demonstrated
1D CdS NRs/2D WSe2 heterostructure with piezo-phototronic effect for a perfor-
mance enhanced flexible photodetector. Figure 20.4a shows the optical image of the
flexible 1D CdS/2D WSe2 heterostructure photodetector. The heterostructure device
fabrication processes are shown in Fig. 20.4b, c. Briefly, 2D WSe2 was precisely
478 M. Peng and X. Sun

transferred on one-terminal of the pre-dispersed CdS NW by using an accurate


transfer platform equipped with micromanipulators, and then the metal electrodes
were patterned on the surface of the NW and nanosheet with electron beam lithog-
raphy and lift-off process. Figure 20.4d shows the schematic illustration of experi-
mental setup for the applied strain and the piezo- phototronic process, in which the
strain is applied through a two-point bending apparatus. The results exhibit that the
photocurrent increase with increasing the compressive strain, while the photocur-
rent increased from 0.32 to 0.65 nA with the light intensity of 16.9 μW/cm2 at the
compressive strain of −0.73%. The optimized photoresponsivity of the device is of
33.4 A/W (Fig. 20.4e).
In addition, many other efforts have been devoted to realize the broadband pho-
todetector based on 1D II-VI semiconductors by integrating semiconductors of 1D
ZnTe, ZnS, or CdS semiconductors with strong light-absorbing materials such as
colloidal PbX (X = S, Se) quantum dots, perovskite or even other 2D materials [30,
35, 65–67]. For instance, Liu et al. [68] demonstrated flexible visible-light photode-
tector based on high-quality ZnTe nanowires on polymer substrate. The photocur-
rent of the flexible photodetectors maintained at 30 nA + 1.2 nA under different
bending radius. More importantly, the photocurrent of the flexible device remained

Fig. 20.4 (a) Optical image flexible 2D WSe2 nanosheet/1D CdS nanowire heterojunction photo-
detector. SEM image of (b) 2D WSe2/1D CdS heterojunction and (c) heterojunction based device.
(d) Schematic illustration of experimental setup for the applied strain and the piezo-­photoelectronic
process. (e) Strain dependence of the photoresponsivity of the flexible photodetector at a bias of
2 V. (Reprinted with permission from Ref. [64]. Copyright 2018: Royal Society of Chemistry)
20 Flexible Photodetectors Based on II-VI Semiconductors 479

unchanged even after 250 cycles bending. Subsequently, Wang et al. [60, 61] dem-
onstrated a piezo-phototronic effect enhanced UV and visible photodetector by
using ZnO/CdS core/shell nanowire or carbon-fiber/ZnO-CdS double shell nanow-
ire as sensor materials. For the flexible photodetector, CdS NWs were firstly synthe-
sized on the surface of ZnO NWs to form core/shell structure, and then the obtained
single CdS/ZnO NW was bonded on a polymer substrate to fabricate flexible pho-
todetector, which shows excellent photoresponse from UV to visible (372–574 nm)
with the responsivity of 11 A/W at 548 nm [61]. The photocurrent and sensitivity of
the core/shell structure photodetector is 1000 times higher than that of 1D CdS NRs.
Liang et al. [69] have also proposed flexible resistivity-type UV-visible photodetec-
tors based on CdS NWs with Ag NWs as electrode by using a non-transfer process.
The flexible photodetector exhibits a UV-visible light sensitivity with the maximum
sensitivity of 120 and the response time of 6 ms.

20.3.3 2D II-VI Semiconductors Based


Flexible Photodetectors

Since the discovery of graphene [70], various layered two-dimensional (2D) nano-
structures, such as transition metal dichalcogenides (TMDs) (e.g. MoS2 and WS2)
[71, 72], Indium selenide (InSe) [71, 73], black phosphorus (BP) have been explored
as building blocks in electronics and optoelectronics due to their unique dimen-
sional dependent properties. For these layered structures, the atoms are usually
bonded with strong covalent in-plane but with weak van der Waals interactions out-­
of-­plane direction [74]. This characteristic allows that high-quality multilayer or
even single layer 2D semiconductors can be easily obtained by simple mechanical
exfoliation from bulk crystals. In contrast, there are larger number of non-layered
semiconductors in the nature with covalent bonds in all three dimensions, and thus
make it difficult to obtain 2D non-layered semiconductors. However, these non-­
layered materials exhibit many their distinct properties different from 2D layered
materials, such as abundant surface dangling bands, high-activity and high-energy
surface [23, 75, 76]. Inspired by these unique features, various 2D non-layered
semiconductors including 2D II-VI groups semiconductors have been attracted con-
siderable interest and have been synthesized by wet-chemical synthesis, such as
CdS, CdSe, CdSxSe1-x, and ZnO etc [77–83]. However, the lateral sizes of the semi-
conductors are small than 300 nm, which limit their practical application in opto-
electronics. Thus, many efforts have been devoted to explore larger lateral size 2D
II-VI semiconductors by chemical vapor deposition (CVD) method and demon-
strated them in photodetectors.
For instance, Hu et al. [84] fabricated 2D non-layered CdS on MoS2 substrate by
epitaxial growth technique with the uniform thickness of 50 nm and the lateral size
larger than 1 μm. The formed vertical heterostructure of 2D CdS/MoS2 have been
demonstrated in photodetectors and the results exhibited a broad wavelength
480 M. Peng and X. Sun

photoresponse (larger than 680 nm) and the responsivity enhance over 50 time
(70.8 mA/W at 610 nm light illumination), which can be attributed to the epitaxial
growth of 2D CdS and photoinduced electrons transfer between the energy levels of
CdS and MoS2. Subsequently, Meng et al. [29, 85] synthesized 2D CdSe and
CdSxSe1-x on mica substrate via van der Waals epitaxy technique. Briefly, CdCl2
located at the centre of tube and mixture S and Se powder located at the upstream of
the CdCl2 were employed as powder source to synthesize 2D CdSxSe1-x by using a
vapor deposition method. SEM image shows that the shape of the as-synthesized 2D
CdSxSe1-x flakes is hexagon with isolated domains and the thickness is around 76 nm
(Fig. 20.5a). Figure 20.5b represents the schematic illustration of the flexible pho-
todetectors, and the linear I-V curves of the photodetector indicate that a good ohmic
contact has been formed between 2D CdSxSe1-x flakes and metal electrodes. The
photocurrent of the 2D CdSxSe1-x flakes based photodetector increase with

Fig. 20.5 (a) SEM image of the synthesized 2D CdSxSe1-x flakes. Inset is AFM image and height
profile of a representative 2D CdSxSe1-x. (b) Schematic illustration of 2D CdSxSe1-x based flexible
photodetector. Inset is the optical image of the flexible photodetector. I-Time characteristics of the
flexible photodetector (c) before and (d) after 50 times bending cycles. (Reprinted with permission
from Ref. [29]. Copyright 2017: Royal Society of Chemistry)
20 Flexible Photodetectors Based on II-VI Semiconductors 481

increasing the incident light intensities from 0.56 to 5.73 mW/cm2 under 450 nm
light illumination. The demonstrated 2D CdSxSe1-x flexible photodetectors exhibit a
high responsivity of 703 A/W, specific detectivity of 3.41 × 1010 Jones, external
quantum efficiency (EQE) of 1.94 × 103, and a good mechanical stability after
repeated bending, respectively. Although the photocurrent shows a slightly decrease
after bending, the flexible photodetector still exhibits a good mechanical stability
with clear on/off photoresponse (Fig. 20.5c, d).
Recently, Sun et al. presented 2D non-layered CdSxSe1-x nanosheets with lateral
size of 10 μm by one-step physical evaporation method. By introducing the strong
light absorption material of PbS QDs or perovskite [20, 86], the photoelectric per-
formance of the 2D CdSxSe1-x based heterostructure photodetectors was improved
significantly compared with that of pristine 2D CdSxSe1-x nanosheet. In addition,
Zhai et al. [24] proposed self-limited epitaxial growth of 2D non-layered CdS flakes
on mica substrate with ultrathin thickness of 6 nm and large lateral size of 40 μm by
using In2S3 as surface passivation agent. The growth mechanism of ultrathin 2D
CdS was also studied by the experiments and theoretical calculation. The 2D CdS
flakes based flexible photodetector was fabricated by a photolithography and lift-off
process and then was directly attached to the flexible PET film, exhibiting the
responsivity and detectivity of 0.05 A/W and 7.81 × 108 Jones, respectively, at the
light intensity of 26.6 mW/cm2 under 400 nm light illumination. The flexible photo-
detector also showed excellent mechanical stability with the current on/off ratio of
~104 even after bending 200 cycles, which can be attributed to the ultrathin 2D CdS
flakes with outstanding deformation tolerance.

20.4 Applications of Flexible Photodetector

The above mentioned II-VI semiconductors based flexible photodetectors are indi-
vidual device structure, which are not suitable for the practical application in moni-
toring [87], imaging [88], and optical communication [89] because their operation
are based on the integrated high-density device arrays. In this section, we will
mainly introduce the recent applications of II-VI semiconductors flexible photode-
tectors, including wearable monitoring sensors, image sensors, self-powered inte-
grated electronics.

20.4.1 Wearable Monitoring Sensors

The wearable monitoring sensors usually demand a precise detection of output sig-
nals due to the weak incident light intensity or low light intensity variation in many
practical applications [11]. Thus, the detectivity and responsivity of the photodetec-
tor should be larger enough to make sure that the monitoring systems work nor-
mally. Furthermore, the integrated photodetectors should be flexible, stable and skin
482 M. Peng and X. Sun

feeling for the practical application of wearable monitoring sensors. In order to


continuous monitor human blood waves, Kim et al. [90] successfully presented
stretchable optoelectronic sensors by integrated QDs based LEDs and QDs based
photodetectors with elastomeric substrates (Fig. 20.6). The optoelectronic proper-
ties of the flexible sensors were stable under different deformations due to the using
of graphene electrodes ensuring the excellent bendability of the devices. The

Fig. 20.6 (a) Schematic diagram of QDs based flexible photodetector. (b) The external quantum
efficiency (EQE) spectra of the QD photodetector. (c) Optical image of skin-mounted optoelec-
tronic sensor composed of QD-LEDs and QD photodetectors wrapped around the finger. (d) Real-­
time optoelectronic signal pulse from a stretchable QD photodetector. (e) Real-time record over
several pulse periods under illumination from the red QD-LED. (Reprinted with permission from
Ref. [90]. Copyright 2017: American Chemical Society)
20 Flexible Photodetectors Based on II-VI Semiconductors 483

emission of the LEDs could be efficiently controlled by adjusting the composition


of the QD, such as CdSe/CdS/ZnS (core/shell/shell) QDs for the red-emissive while
CdSe/ZnS QDs for the green and blue-emissive. The schematic diagram of QDs
based flexible photodetector is shown in Fig. 20.6a, in which the QDs and graphene
were used as photoactive materials and electrode by assembling them on elasto-
meric PEN substrates via transfer printing method. MoO3 and TiO2 layers were used
as hole and electron transport layers, respectively. The external quantum efficiency
(EQE) of the photodetectors can reach 12% at a wavelength of 618 nm (Fig. 20.6b).
To demonstrate the application of the flexible devices, the fabricated QDs based
LEDs and photodetectors had been stretched around the tip of a forefinger
(Fig. 20.6c). The pressure pulse signals of the sensors could be detected by monitor-
ing the variation of light absorption resulted from illumination in the skin. One
pulse wave from a stretchable QD photodetector could be obtained by using a
stretchable QDs based LED as light source under a luminous intensity of 520 cd m−2
(Fig. 20.6d), revealing that the flexible sensor showed two clearly distinguishable
peaks. Figure 20.6e showed the real-time record over several pulse periods under
illumination from the red QD-LED. The results indicated that the proposed stretch-
able optoelectronic sensors exhibited excellent optoelectronic properties as well as
health-monitoring capability.
Similarly, Yan et al. [91] proposed novel flexible photodetector arrays based on
all-solid-state organic electrochemical transistors (OECT) for photosensing and
imaging. The OECT based phototransistor with CdSe/ZnS quantum dots as photo-
active materials exhibit excellent responsivity and detectivity of 6.9 × 105 A/W and
5.8 × 1012 Jones under 355 nm UV light illumination, respectively. Furthermore, the
flexible image sensors have been fabricated by integrating the flexible OECTs based
phototransistor into 10 × 10 pixel on PET substrate. The demonstrated flexible
image sensors have demonstrated excellent detection ability for broadband and thus
can clearly distinguish the target images composed of red and white light even
under bending states. Most recently, Lee et al. [92] reported a stretchable wearable
electronics based on CdSe/ZnS quantum-dots for visual display of body movement
and skin temperature signals from skin-attached sensors. Firstly, an array of
CdSe/ZnS QD-LEDs was fabricated on top of the pre-formed NOA63 islands array
and then liquid metal Galinstan was spray-coated on the top of the QDs-LED array
as conductive electrode. Finally, the array of QD-LEDs on the islands was trans-
ferred onto a stretchable Eco-PDMS elastomer substrate. The fabricated QDs based
wearable electronics exhibited stable operation under 50% uniaxial and 30% biaxial
strains due to the architectural design minimizing the strain applied to the QD-LEDs
by transforming the strain onto the stretchable elastomer substrate and the liquid
metal electrode. More importantly, the stretchable QD-LED array were demon-
strated in the visual pattern display by monitoring the extent of knee bending and
the changes of human skin temperature.
484 M. Peng and X. Sun

20.4.2 Image Sensors

The other important application of the flexible photodetector is image sensing,


which can detect environments and objects and have a specific application accord-
ing to their different sensing wavelength range, such as UV imaging for skin health
monitoring and remote sensing, visible light imaging for face recognition system,
and infrared imaging for medical diagnosis and night vision, etc. [11, 93, 94] For
instance, Li et al. [95] presented ZnO quantum dot decorated Zn2SnO4 nanowire
heterojunction photodetectors for flexible ultraviolet image sensors. To investigate
the image sensing, 10 × 10 pixel flexible heterojunction photodetectors were inte-
grated to recognize the letters under the bending states (Fig. 20.7). Figure 20.7a
shows the optical image of the flexible photodetector arrays on a PET substrate. The
flexible photodetector shows excellent stability with almost identical I-V character-
istics even after bending 2000 cycles. According to the energy band diagram
(Fig. 20.7b), the photogenerated electron-hole pairs will be efficiently separated and
the holes migrate from NWs to QDs to neutralize the absorbed O2− and then increase
the electron concentration, thus enhancing the photoresponse performance of the
heterostructure photodetectors. Figure 20.7c, d presents the schematic illustrations
of using a 10 × 10 flexible photodetector array to sense the letters “F” and “E” under
different bending directions, in which the letter patterns were placed on the back
side of the devices and the incident light was also illuminated from the back side. It
is obviously that the corresponding output images of letters “F” and “E” can be
clearly observed under bending conditions (Fig. 20.7e, f). The results indicate that
the QDs based heterostructure has a promising application in flexible UV image
sensing.
In addition, a rectifying diode-type flexible photodetectors based on II-VI semi-
conductors have also been demonstrated in image sensing. This type device is usually
fabricated with a vertical structure, the top and bottom electrodes were used to
extracted electrons and holes, while the photoactive materials located in middle layer
and closely contact with electron and hole transporting layers, respectively. Brabec
et al. [96] proposed the solution-processed flexible photodetector in visible and near-
infrared imaging with P3HT/O-IDTBR as the photoactive and with ZnO as electron
transporting layer. The photodiode shows a record responsivity of 0.42 A/W and
external quantum efficiency (EQE) of 69% at 755 nm NIR light, respectively. More
importantly, the photodiodes based image sensor can realize objects detection under
visible and NIR light conditions with excellent imaging performance. Similarly, Kim
et al. [97] reported a matrix-type multiplexed photodiode array with perovskite film as
photoactive materials and with spin-coated ZnO NPs as electron transporting layer by
using a novel spin-on-patterning (SoP) process. By controlling the spin-coating rate
and the temperature of the precursor solution, the optimized pattern yield can be
obtained in SoP process. Subsequently, a 10 × 10 multiplexed ultrathin and flexible
image sensor array was fabricated to realize a high-resolution imaging of dot array
patterns. The device exhibited the responsivity and detectivity of 0.109 A/W and
1.35 × 1012 Jones at 530 nm, respectively, which can be compared with the commer-
cial Si photodetector of 0.01 A/W and 4 × 1012 Jones.
20 Flexible Photodetectors Based on II-VI Semiconductors 485

Fig. 20.7 Application of a flexible photodetector array as UV image sensor. (a) Optical images of
as-fabricated flexible photodetector arrays on a PET substrate. (b) Energy band diagrams of carrier
separation mechanism. (c, d) Schematic illustrations of using a 10 × 10 flexible photodetector
array to sense the letters “F” and “E” under different bending directions. (e, f) Corresponding
output images of letters “F” and “E” by the flexible photodetector array. (Reprinted with permis-
sion from Ref. [95]. Copyright 2017: American Chemical Society)
486 M. Peng and X. Sun

20.4.3 Self-Powered Integrated Wearable Electronics

With the rapid development of the Internet of Things (IOTs) technology [98], it is
more and more urgent for wearable electronic systems towards minimization, high
integration and multifunctional. However, most of the current wearable electronics
usually works by the external power source supply including battery and solar cell,
which will inevitably increase the system size and thus limit their practical applica-
tion. Consequently, various self-powered photodetectors have been developed to
integrate into wearable electronics, which can drive the device work directly by the
self-powered function without the demands of external power supply or series con-
necting wires. For instance, Dai et al. [31] proposed a self-powered flexible photo-
detector based on CdS/Si heterostructure by pyro-phototronic effect. The
self-powered photodetector exhibited a broadband response from 325 to 1550 nm at
a bias of 0 V and fast response time of 245 μs /277 μs at 325 nm light illumination,
which is over the optical bandgap lamination of the CdS and Si because of the pyro-­
phototronic effect in CdS and thus photoinduced heating or cooling. More impor-
tantly, the responsivity and specific detectivity were also improved 67.8 times.
Recently, Park et al. [99] reported self-powered ultra-flexible electronics via nano-­
grating-­patterned organic photovoltaics which can realize the detection of biometric
signals with high SNR when applied to skin tissue. The flexible biomedical devices
were integrated into a 1 μm thick flexible substrate in which the organic electro-
chemical transistors and organic photovoltaic (Fig. 20.8a) were used as sensors and
power sources supply, respectively. In order to increase the efficiency of the organic
photovoltaics, the charge transporting layers of ZnO were patterned to form a
760 nm periodic nano-grating morphologies by using a room-temperature moulding
process. When the device separated from the supporting substrate, 3 μm thick of
flexible OPVs were obtained (Fig. 20.8b). Figure 20.8c shows the circuit diagram of
cardiac signal recording and photograph of the self-powered integrated electronic
device attached to a finger, in which the potential difference between Gel electrode
and OECT was employed as gate bias. The biological signal can be monitored by
LED light illumination. The optimized organic photovoltaics exhibited a high
power-conversion efficiency of 10.5% and a high power-per-weight ratio of
11.46 W/g.
As an alternative, triboelectric nanogenerators (TENGs) based on the coupling
effect of triboelectrification and the electrostatic induction have been demonstrated
in self-powered flexible electronics by harvesting external mechanical energy and
converting it into electric power to driver the devices [100]. Recently, TENGs as a
power supply have attracted intensive interest in self-powered gas sensing [98],
flexible UV light detection [101], self-powered weighting [102], etc. For instance,
Zhang et al. [101] reported a self-powered wearable real-time UV Photodetector by
coupling effects of triboelectric and photoelectric. The schematic illustration of the
flexible real-time UV light detection system is show in Fig. 20.8d, and the inset is
the optical image of flexible self-powered photodetector on human body. The flex-
ible photodetection system is consist of flexible-film-based TENG working in
20 Flexible Photodetectors Based on II-VI Semiconductors 487

Fig. 20.8 (a) Structure diagram of the organic photovoltaic (OPV) device. (b) Photography of the
OVP device was wrapped around an object. (c) The circuit diagram of cardiac signal recording and
photograph of the self-powered integrated electronic device attached to a finger. ((a–c) Reprinted
with permission from Ref. [99]. Copyright 2018: Springer Nature). (d) Schematic diagram of the
flexible photodetector in tapping-mode. ((d) Reprinted with permission from Ref. [101]. Copyright
2020: American Chemical Society)

tapping mode as energy harvester, UV photodetector as sensor, and a series of com-


mercial LEDs as alarm. The output voltages of the constant resistance changes with
increasing the UV light intensities, which can be reflected directly through series
LEDs on/off display in circuit. The working mechanism can be attributed to imped-
ance matching effect between NPs-based UV photodetector and TENG. The pro-
posed self-powered flexible UV photodetector may pave a new way in self-­powered
wearable electronics.

20.5 Conclusion and Outlook

In this chapter, the most recent progress on different dimensional II-VI semiconduc-
tors based flexible photodetectors and their application in wearable electronics have
been introduced, including 0D nanostructures (QDs and NCs), 1D nanostructures
488 M. Peng and X. Sun

(NWs, NBs and NRs) and 2D non-layered nanostructures. The main sensing mech-
anisms and key figures of merits for photodetectors have been summarized and
discussed, which will help us to better understand and study various photodetectors.
The typical flexible photodetectors based on different dimensional II-VI semicon-
ductor nanostructures have been reviewed, in which the functional II-VI materials
synthesis method have also been discussed. We mainly focus on the device structure
and fabrication process of the II-VI semiconductors flexible photodetector, as well
as the mechanical flexibility and electric stability. Among these functional materi-
als, 0D nanostructures have been widely exploited in flexible photodetectors and
wearable electronics for their large scale manufacturing, low temperature solution
process, and easy compatible with flexible substrates. In addition, 2D II-VI nano-
structures (e.g. 2D CdS, CdSe, ZnO, CdSxSe1-x) belong to 2D non-layered semicon-
ductors in the nature with covalent bonds in all three dimensions and thus exhibit
many distinct properties different from 2D layered van der Waals materials, such as
abundant surface dangling bands, high-activity and high-energy surface, which
have also been developed and demonstrated most recently in flexible photodetec-
tors. However, the photoresponse of the single II-VI nanostructures based flexible
photodetectors just only focus on UV (e.g. ZnS) or visible (e.g. CdS, CdSe, and
CdSxSe1-x) light region due to their intrinsic optical bandgap resulting in narrow
wavelength range of light absorption, which will limit their potential application in
some specific filed, such as wide spectrum light detection, NIR light imaging, and
night vision technique, etc. In this chapter, we also discuss the designed various
II-VI hybrid nanostructure or heterostructure based flexible photodetectors to
improve the photoelectric performance and realize broadband or NIR photores-
ponse. Finally, various application of the II-VI nanostructures based flexible photo-
detectors have been summarized, including wearable monitoring sensors, image
sensors, and self-powered integrated wearable electronics.
Although significant research progresses on II-VI semiconductors based flexible
photodetectors have been achieved, there are still many challenges and opportuni-
ties in future studies, such as optimizing synthesis process to obtain more uniformly
and high quality crystalline nanostructure, further improving the adhesion proper-
ties between flexible substrate, functional materials and metal electrodes, and thus
improving the mechanical flexibility and electric stability of the flexible photode-
tectors. (1) For flexible photodetectors, various II-VI nanostructures have been
mainly synthesized by using solution process, chemical vapor deposition, or various
printing techniques, while different synthesis process have their unique characteris-
tics but also have their disadvantages. For instance, large lateral size (>1 μm) and
high-quality crystalline 2D non-layered II-VI nanostructures have been mainly syn-
thesize by CVD routing, but it will limit their application in large scale integrated
flexible photodetector arrays due to the difficult large scale synthesis by this rout-
ing. The solution process synthesis of 0D II-VI nanostructure exhibits potential
large scale manufacturing characteristics and can be easily spin-coated on flexible
substrate, but the post ligands-exchange step is further needed to improve the con-
ductivity of the NPs or QDs by organic solvent and thus resulting in the degradation
of the flexible substrate and damage the photoelectric performance of the flexible
photodetectors. Thus, the functional materials synthesis strategies should be
20 Flexible Photodetectors Based on II-VI Semiconductors 489

optimized to realize the penitential application for large scale and high-performance
flexible photodetectors. (2) The mechanical flexibility and electronic stability of the
flexible photodetector can be slightly affected after several hundred times bending
or twisting and resulting in the deterioration of contact property between the func-
tional materials and metal electrodes and the flexible substrate. Thus, it is urgently
to need further study and more endeavours to improve the adhesion property
between the functional materials and electrodes and substrate, thus ensuring the
flexibility and stability of the flexible photodetectors.
The II-VI nanostructures based flexible photodetectors exhibit promising and
inspiring application in wearable electronic including wearable monitoring sensors,
image sensors and self-powered integrated wearable electronic in Sect. 20.4. For
wearable monitoring sensors, flexible photodetector integrated into human body to
monitor the environment UV signals for skin disease prevention according to the
resistance variation caused by UV light illumination and thus lighting the LEDs.
Thus, the integrated flexible photodetectors with ultra-flexibility, high sensitivity
and skin like should be paid more attention for the practical application of wearable
monitoring sensors. For image sensors, few works including 10 × 10 pixel flexible
heterojunction photodetectors based on ZnO have been integrated and demonstrated
in UV image sensing system to sense the letters with bending states. However, the
visible and NIR image sensors are also very important for the practical application
in wearable electronics, such as visible light imaging for face recognition system,
and infrared imaging for medical diagnosis and night vision, etc. Therefore, we
should pay more attention not only to the fabrication of wearable image sensors in
different sensing wavelength range but also the precise and complicated imaging
recognition system for special application environments. For self-powered wearable
integrated electronics, various self-powered photodetectors have been developed to
integrate into wearable electronics to efficiently drive the photodetector by integrat-
ing p-n junction, solar cells or organic photovoltaics. These strategies have been
proved to be the effective ways while the complicated fabrication process will limit
their potential extensive application. Thanks to the development of TENG tech-
nique, TENG based flexible photodetector serve as another choice for the self-­
powered wearable integrated electronics. However, this strategy is still in its infancy,
there are many problems to solve in future research, such as the output electric sig-
nals stability in mechanical energy harvesting process, the integrated stability of the
whole systems. All in all, several significant progresses about the flexible photode-
tector integrated wearable electronic have been exploited in most recently, while
there are still tremendous challenges for the integration of wearable electronic in
real-time data monitoring, processing, and outputting.

References

1. Luo P, Zhuge F, Wang F, Lian L, Liu K, Zhang J et al (2019) PbSe quantum dots sensi-
tized high-mobility Bi2O2Se nanosheets for high-performance and broadband photodetection
beyond 2 μm. ACS Nano 13(8):9028–9037
490 M. Peng and X. Sun

2. Zhu T, Yang Y, Zheng L, Liu L, Becker ML, Gong X (2020) Solution-processed flexible
broadband photodetectors with solution-processed transparent polymeric electrode. Adv
Funct Mater 30(15):1909487
3. Xu Y, Lin Q (2020) Photodetectors based on solution-processable semiconductors: recent
advances and perspectives. Appl Phys Rev 7(1):011315
4. Konstantatos G (2018) Current status and technological prospect of photodetectors based on
two-dimensional materials. Nat Commun 9(1):1–3
5. Xie C, Yan F (2017) Flexible photodetectors based on novel functional materials. Small
13(43):1701822
6. Buscema M, Island JO, Groenendijk DJ, Blanter SI, Steele GA, van der Zant HS et al (2015)
Photocurrent generation with two-dimensional van der Waals semiconductors. Chem Soc
Rev 44(11):3691–3718
7. Chow PCY, Someya T (2019) Organic photodetectors for next-generation wearable electron-
ics. Adv Mater 32(15):1902045
8. Long M, Wang P, Fang H, Hu W (2018) Progress, challenges, and opportunities for 2D mate-
rial based photodetectors. Adv Funct Mater 29(19):1803807
9. Dong T, Simões J, Yang Z (2020) Flexible photodetector based on 2D materials: processing,
architectures, and applications. Adv Mater Interfaces 7(4):1901657
10. Yokota T, Fukuda K, Someya T (2021) Recent progress of flexible image sensors for bio-
medical applications. Adv Mater 33(19):2004416
11. Cai S, Xu X, Yang W, Chen J, Fang X (2019) Materials and designs for wearable photodetec-
tors. Adv Mater 31(18):1808138
12. Zheng L, Deng X, Wang Y, Chen J, Fang X, Wang L et al (2020) Self-powered flexible TiO2
fibrous photodetectors: heterojunction with P3HT and boosted responsivity and selectivity by
Au nanoparticles. Adv Funct Mater 30(24):2001604
13. Selamneni V, Raghavan H, Hazra A, Sahatiya P (2021) MoS2/paper decorated with metal
nanoparticles (Au, Pt, and Pd) based plasmonic-enhanced broadband (visible-NIR) flexible
photodetectors. Adv Mater Interfaces 8(6):2001988
14. Zhang Y, Huang P, Guo J, Shi R, Huang W, Shi Z et al (2020) Graphdiyne-based flexible
photodetectors with high responsivity and detectivity. Adv Mater 32(23):2001082
15. Wang M, Tian W, Cao F, Wang M, Li L (2020) Flexible and self-powered lateral photodetec-
tor based on inorganic perovskite CsPbI3–CsPbBr3 heterojunction nanowire array. Adv Funct
Mater 30(16):1909771
16. Schneider DS, Grundmann A, Bablich A, Passi V, Kataria S, Kalisch H et al (2020) Highly
responsive flexible photodetectors based on MOVPE grown uniform few-layer MoS2. ACS
Photonics 7(6):1388–1395
17. Fuentes-Hernandez C, Chou W-F, Khan TM, Diniz L, Lukens J, Larrain FA et al (2020)
Large-area low-noise flexible organic photodiodes for detecting faint visible light. Science
370:698–701
18. Hao D, Zou J, Huang J (2019) Recent developments in flexible photodetectors based on metal
halide perovskite. InfoMat 2(1):139–169
19. Majithia RY (2013) Microwave-assisted synthesis of II-VI semiconductor micro-and
nanoparticles towards sensor applications, vol 3572238. Texas A&M University. ProQuest
Dissertations Publishing
20. Peng M, Xie X, Zheng H, Wang Y, Zhuo QQ, Yuan G et al (2018) PbS quantum dots/2D
non-layered CdSxSe1-x nanosheets hybrid nanostructure for high-performance broadband
photodetectors. ACS Appl Mater Interfaces 10(50):43887–43895
21. Jang Y, Shapiro A, Isarov M, Rubin-Brusilovski A, Safran A, Budniak AK et al (2017)
Interface control of electronic and optical properties in IV–VI and II–VI core/shell colloidal
quantum dots: a review. Chem Commun 53(6):1002–1024
22. Kagan CR (2019) Flexible colloidal nanocrystal electronics. Chem Soc Rev 48(6):1626–1641
23. Zhou N, Yang R, Zhai T (2019) Two-dimensional non-layered materials. Mater Today Nano
8:100051
20 Flexible Photodetectors Based on II-VI Semiconductors 491

24. Jin B, Huang P, Zhang Q, Zhou X, Zhang X, Li L et al (2018) Self-limited epitaxial growth
of ultrathin nonlayered CdS flakes for high-performance photodetectors. Adv Funct Mater
28(20):1800181
25. Mitra S, Aravindh A, Das G, Pak Y, Ajia I, Loganathan K et al (2018) High-performance solar-
blind flexible deep-UV photodetectors based on quantum dots synthesized by femtosecond-­
laser ablation. Nano Energy 48:551–559
26. Peng M, Wang Y, Shen Q, Xie X, Zheng H, Ma W et al (2019) High-performance flexible and
broadband photodetectors based on PbS quantum dots/ZnO nanoparticles heterostructure.
Sci China Mater 62(2):225–235
27. Peng M, Wen Z, Shao M, Sun X (2017) One-dimensional CdSxSe1−x nanoribbons for high-­
performance rigid and flexible photodetectors. J Mater Chem C 5(30):7521–7526
28. Park J, Lee J, Noh Y, Shin K-H, Lee D (2016) Flexible ultraviolet photodetectors with ZnO
nanowire networks fabricated by large area controlled roll-to-roll processing. J Mater Chem
C 4(34):7948–7958
29. Xia J, Zhao YX, Wang L, Li XZ, Gu YY, Cheng HQ et al (2017) van der Waals epitaxial two-­
dimensional CdSxSe(1-x) semiconductor alloys with tunable-composition and application to
flexible optoelectronics. Nanoscale 9(36):13786–13793
30. Wang S, Zhu Z, Zou Y, Dong Y, Liu S, Xue J et al (2019) A low-dimension structure strategy
for flexible photodetectors based on perovskite nanosheets/ZnO nanowires with broadband
photoresponse. Sci China Mater 63(1):100–109
31. Dai Y, Wang X, Peng W, Xu C, Wu C, Dong K et al (2018) Self-powered Si/CdS flexible pho-
todetector with broadband response from 325 to 1550 nm based on pyro-phototronic effect:
an approach for photosensing below bandgap energy. Adv Mater 30(9):1705893
32. Saran R, Curry RJ (2016) Lead sulphide nanocrystal photodetector technologies. Nat
Photonics 10(2):81–92
33. Liu CH, Chang YC, Norris TB, Zhong Z (2014) Graphene photodetectors with ultra-­
broadband and high responsivity at room temperature. Nat Nanotechnol 9(4):273–278
34. Li L, Lou Z, Shen G (2015) Hierarchical CdS nanowires based rigid and flexible photodetec-
tors with ultrahigh sensitivity. ACS Appl Mater Interfaces 7(42):23507–23514
35. Lou Z, Li L, Shen G (2016) Ultraviolet/visible photodetectors with ultrafast, high photosen-
sitivity based on 1D ZnS/CdS heterostructures. Nanoscale 8(9):5219–5225
36. Gomathi PT, Sahatiya P, Badhulika S (2017) Large-area, flexible broadband photodetector
based on ZnS-MoS2 hybrid on paper substrate. Adv Funct Mater 27(31):1701611
37. Gou G, Dai G, Wang X, Chen Y, Qian C, Kong L et al (2017) High-performance and flex-
ible photodetectors based on P3HT/CdS/CdS: SnS2 superlattice nanowires hybrid films. Appl
Phys A Mater Sci Process 123(12):1–8
38. Xu X, Chen J, Cai S, Long Z, Zhang Y, Su L et al (2018) A real-time wearable UV-radiation
monitor based on a high-performance p-CuZnS/n-TiO2 photodetector. Adv Mater
30(43):1803165
39. Sun H, Tian W, Cao F, Xiong J, Li L (2018) Ultrahigh-performance self-powered flexible
double-twisted fibrous broadband perovskite photodetector. Adv Mater 30(21):1706986
40. Bae WK, Padilha LA, Park YS, McDaniel H, Robel I, Pietryga JM et al (2013) Controlled
alloying of the core-shell interface in CdSe/CdS quantum dots for suppression of auger
recombination. ACS Nano 7:3411–3419
41. Chan Y, Dahua Z, Jun Y, Linlong T, Chongqian L, Jun S (2020) Fabrication of hybrid gra-
phene/CdS quantum dots film with the flexible photo-detecting performance. Physica E Low
Dimens Syst Nanostruct 124:114216
42. Zheng Z, Zhuge F, Wang Y, Zhang J, Gan L, Zhou X et al (2017) Decorating perovskite quan-
tum dots in TiO2 nanotubes array for broadband response photodetector. Adv Funct Mater
27(43):1703115
43. Shen T, Yuan J, Zhong X, Tian J (2019) Dip-coated colloidal quantum-dot films for high-­
performance broadband photodetectors. J Mater Chem C 7(21):6266–6272
492 M. Peng and X. Sun

44. Hsiao Y-J, Ji L-W, Lu H-Y, Fang T-H, Hsiao K-H (2016) High sensitivity ZnO nanorod-­
based flexible photodetectors enhanced by CdSe/ZnS core-shell quantum. IEEE Sensors J
17(12):3710–3713
45. Xia YX, Zhai G, Zheng Z, Lian L, Liu H, Zhang D et al (2018) Solution-processed solar-­
blind deep ultraviolet photodetectors based on strongly quantum confined ZnS quantum dots.
J Mater Chem C 6(42):11266–11271
46. Chu L, Hu R, Liu W, Ma Y, Zhang R, Yang J et al (2018) Screen printing large-area organo-
metal halide perovskite thin films for efficient photodetectors. Mater Res Bull 98:322–327
47. Tong S, Yuan J, Zhang C, Wang C, Liu B, Shen J et al (2018) Large-scale roll-to-roll printed,
flexible and stable organic bulk heterojunction photodetector. Npj Flex Electron 2(1):1–8
48. Park SH, Su R, Jeong J, Guo SZ, Qiu K, Joung D et al (2018) 3D printed polymer photodetec-
tors. Adv Mater 30(40):1803980
49. Zhang C, Xie Y, Deng H, Tumlin T, Zhang C, Su JW et al (2017) Monolithic and flexible
ZnS/SnO2 ultraviolet photodetectors with lateral graphene electrodes. Small 13(18):1604197
50. Shen T, Binks D, Yuan J, Cao G, Tian J (2019) Enhanced-performance of self-powered
flexible quantum dot photodetectors by a double hole transport layer structure. Nanoscale
11(19):9626–9632
51. Cui Y, Zhong Z, Wang D, Wang W, Lieber CM (2003) High performance silicon nanowire
field effect transistors. Nano Lett 3:149–152
52. Duan X, Niu C, Zaanen J, Sahi V, Chen J, Parce JW et al (2003) High-performance thin-film
transistors using semiconductor nanowires and nanoribbons. Nature 425:274–278
53. Cho N, Roy Choudhury K, Thapa RB, Sahoo Y, Ohulchanskyy T, Cartwright AN et al (2007)
Efficient photodetection at IR wavelengths by incorporation of PbSe–carbon-nanotube con-
jugates in a polymeric nanocomposite. Adv Mater 19(2):232–236
54. Tian W, Zhang C, Zhai T, Li SL, Wang X, Liu J et al (2014) Flexible ultraviolet photodetec-
tors with broad photoresponse based on branched ZnS-ZnO heterostructure nanofilms. Adv
Mater 26(19):3088–3093
55. Jie JS, Zhang WJ, Jiang Y, Li YQ, Lee ST (2006) Photoconductive characteristics of single-­
crystal CdS nanoribbons. Nano Lett 6:1887–1891
56. Xing X, Zhang Q, Huang Z, Lu Z, Zhang J, Li H et al (2016) Strain driven spectral broaden-
ing of Pb ion exchanged CdS nanowires. Small 12(7):874–881
57. Jiang Y, Zhang WJ, Jie JS, Meng XM, Fan X, Lee ST (2007) Photoresponse properties of
CdSe single-nanoribbon photodetectors. Adv Funct Mater 17(11):1795–1800
58. Wu P, Dai Y, Sun T, Ye Y, Meng H, Fang X et al (2011) Impurity-dependent photores-
ponse properties in single CdSe nanobelt photodetectors. ACS Appl Mater Interfaces
3(6):1859–1864
59. Takahashi T, Nichols P, Takei K, Ford AC, Jamshidi A, Wu MC et al (2012) Contact printing
of compositionally graded CdS(x)Se(1-x) nanowire parallel arrays for tunable photodetectors.
Nanotechnology 23(4):045201
60. Zhang F, Niu S, Guo W, Zhu G, Liu Y, Zhang X et al (2013) Piezo-phototronic effect enhanced
visible/UV photodetector of a carbon-Fiber/ZnO-CdS double-shell microwire. ACS Nano
7:4537–4544
61. Zhang F, Ding Y, Zhang Y, Zhang X, Wang ZL (2012) Piezo-phototronic effect enhanced vis-
ible and ultraviolet photodetection using a ZnO-CdS core-shell micro/nanowire. ACS Nano
6:9229–9236
62. Rai SC, Wang K, Ding Y, Marmon JK, Bhatt M, Zhang Y et al (2015) Piezo-phototronic
effect enhanced UV/visible photodetector based on fully wide band gap type-II ZnO/ZnS
core/shell nanowire array. ACS Nano 9(6):6419–6427
63. Zhang K, Ding J, Lou Z, Chai R, Zhong M, Shen G (2017) Heterostructured ZnS/InP nanow-
ires for rigid/flexible ultraviolet photodetectors with enhanced performance. Nanoscale
9(40):15416–15422
64. Lin P, Zhu L, Li D, Xu L, Wang ZL (2018) Tunable WSe2–CdS mixed-dimensional van der
Waals heterojunction with a piezo-phototronic effect for an enhanced flexible photodetector.
Nanoscale 10(30):14472–14479
20 Flexible Photodetectors Based on II-VI Semiconductors 493

65. Hu L, Yang J, Liao M, Xiang H, Gong X, Zhang L et al (2012) An optimized ultraviolet-a


light photodetector with wide-range photoresponse based on ZnS/ZnO biaxial nanobelt. Adv
Mater 24(17):2305–2309
66. Zheng Z, Gan L, Zhang J, Zhuge F, Zhai T (2017) An enhanced UV-vis-NIR and flexible pho-
todetector based on electrospun ZnO nanowire array/PbS quantum dots film heterostructure.
Adv Sci 4(3):1600316
67. Gao T, Zhang Q, Chen J, Xiong X, Zhai T (2017) Performance-enhancing broadband and
flexible photodetectors based on perovskite/ZnO-nanowire hybrid structures. Adv Opt Mater
5(12):1700206
68. Liu Z, Chen G, Liang B, Yu G, Huang H, Chen D et al (2013) Fabrication of high-quality
ZnTe nanowires toward high-performance rigid/flexible visible-light photodetectors. Opt
Express 21(6):7799–7810
69. Pei Y, Pei R, Liang X, Wang Y, Liu L, Chen H et al (2016) CdS-nanowires flexible photo-­
detector with ag-nanowires electrode based on non-transfer process. Sci Rep 6(1):21551
70. Xia F, Mueller T, Lin YM, Valdes-Garcia A, Avouris P (2009) Ultrafast graphene photodetec-
tor. Nat Nanotechnol 4(12):839–843
71. Tamalampudi SR, Lu YY, Kumar UR, Sankar R, Liao CD, Moorthy BK et al (2014) High
performance and bendable few-layered InSe photodetectors with broad spectral response.
Nano Lett 14(5):2800–2806
72. Chhowalla M, Shin HS, Eda G, Li LJ, Loh KP, Zhang H (2013) The chemistry of two-­
dimensional layered transition metal dichalcogenide nanosheets. Nat Chem 5(4):263–275
73. Feng W, Zheng W, Cao W, Hu P (2014) Back gated multilayer InSe transistors with enhanced
carrier mobilities via the suppression of carrier scattering from a dielectric interface. Adv
Mater 26(38):6587–6593
74. Koppens FH, Mueller T, Avouris P, Ferrari AC, Vitiello MS, Polini M (2014) Photodetectors
based on graphene, other two-dimensional materials and hybrid systems. Nat Nanotechnol
9(10):780–793
75. Xie Z, Xing C, Huang W, Fan T, Li Z, Zhao J et al (2018) Ultrathin 2D nonlayered tellurium
nanosheets: facile liquid-phase exfoliation, characterization, and photoresponse with high
performance and enhanced stability. Adv Funct Mater 28(16):1705833
76. Wang F, Wang Z, Yin L, Cheng R, Wang J, Wen Y et al (2018) 2D library beyond gra-
phene and transition metal dichalcogenides: a focus on photodetection. Chem Soc Rev
47(16):6296–6341
77. Ithurria S, Dubertret B (2008) Quasi 2D colloidal CdSe platelets with thicknesses controlled
at the atomic level. J Am Chem Soc 130(49):16504–16505
78. Ithurria S, Bousquet G, Dubertret B (2011) Continuous transition from 3D to 1D confinement
observed during the formation of CdSe nanoplatelets. J Am Chem Soc 133(9):3070–3077
79. Schlenskaya NN, Yao Y, Mano T, Kuroda T, Garshev AV, Kozlovskii VF et al (2017) Scroll-­
like alloyed CdSxSe1–x nanoplatelets: facile synthesis and detailed analysis of tunable optical
properties. Chem Mater 29(2):579–586
80. Kelestemur Y, Dede D, Gungor K, Usanmaz CF, Erdem O, Demir HV (2017) Alloyed het-
erostructures of CdSexS1–x nanoplatelets with highly tunable optical gain performance. Chem
Mater 29(11):4857–4865
81. Sun Z, Liao T, Dou Y, Hwang SM, Park M-S, Jiang L et al (2014) Generalized self-assembly
of scalable two-dimensional transition metal oxide nanosheets. Nat Commun 5(1):1–9
82. Xu Y, Zhao W, Xu R, Shi Y, Zhang B (2013) Synthesis of ultrathin CdS nanosheets as
efficient visible-light-driven water splitting photocatalysts for hydrogen evolution. Chem
Commun 49(84):9803–9805
83. Maiti PS, Meir N, Houben L, Bar-Sadan M (2014) Solution phase synthesis of homoge-
neously alloyed ultrathin CdSxSe1−x nanosheets. RSC Adv 4(91):49842–49845
84. Zheng W, Feng W, Zhang X, Chen X, Liu G, Qiu Y et al (2016) Anisotropic growth of non-
layered CdS on MoS2 Monolayer for functional vertical heterostructures. Adv Funct Mater
26(16):2648–2654
494 M. Peng and X. Sun

85. Zhu DD, Xia J, Wang L, Li XZ, Tian LF, Meng XM (2016) van der Waals epitaxy and photo-
response of two-dimensional CdSe plates. Nanoscale 8(22):11375–11379
86. Peng M, Ma Y, Zhang L, Cong S, Hong X, Gu Y et al (2021) All-inorganic CsPbBr3 perovskite
nanocrystals/2D non-layered cadmium sulfde selenide for high-performance photodetectors
by energy band alignment engineering. Adv Funct Mater 31:2105051
87. An J, Le T-SD, Lim CHJ, Tran VT, Zhan Z, Gao Y et al (2018) Single-step selective laser
writing of flexible photodetectors for wearable optoelectronics. Adv Sci 5(8):1800496
88. Lee W, Liu Y, Lee Y, Sharma BK, Shinde SM, Kim SD et al (2018) Two-dimensional mate-
rials in functional three-dimensional architectures with applications in photodetection and
imaging. Nat Commun 9(1):1–9
89. Rein M, Favrod VD, Hou C, Khudiyev T, Stolyarov A, Cox J et al (2018) Diode fibres for
fabric-based optical communications. Nature 560:214–218
90. Kim TH, Lee CS, Kim S, Hur J, Lee S, Shin KW et al (2017) Fully stretchable optoelectronic
sensors based on colloidal quantum dots for sensing photoplethysmographic signals. ACS
Nano 11(6):5992–6003
91. Yan Y, Wu X, Chen Q, Liu Y, Chen H, Guo T (2019) High-performance low-voltage flexible
photodetector arrays based on all-solid-state organic electrochemical transistors for photo-
sensing and imaging. ACS Appl Mater Interfaces 11(22):20214–20224
92. Lee Y, Kim DS, Jin SW, Lee H, Jeong YR, You I et al (2022) Stretchable array of CdSe/ZnS
quantum-dot light emitting diodes for visual display of bio-signals. Chem Eng J 427:130858
93. Chen X, Shehzad K, Gao L, Long M, Guo H, Qin S et al (2019) Graphene hybrid structures
for integrated and flexible optoelectronics. Adv Mater 32(27):1902039
94. Yip S, Shen L, Ho JC (2019) Recent advances in flexible photodetectors based on 1D nano-
structures. J Semicond 40(11):111602
95. Li L, Gu L, Lou Z, Fan Z, Shen G (2017) ZnO quantum dot decorated Zn2SnO4 nanowire
heterojunction photodetectors with drastic performance enhancement and flexible ultraviolet
image sensors. ACS Nano 11(4):4067–4076
96. Gasparini N, Gregori A, Salvador M, Biele M, Wadsworth A, Tedde S et al (2018) Visible
and near-infrared imaging with nonfullerene-based photodetectors. Adv Mater Technol
3(7):1800104
97. Lee W, Lee J, Yun H, Kim J, Park J, Choi C et al (2017) High-resolution spin-on-patterning
of perovskite thin films for a multiplexed image sensor array. Adv Mater 29(40):1702902
98. Shen Q, Xie X, Peng M, Sun N, Shao H, Zheng H et al (2018) Self-powered vehicle emis-
sion testing system based on coupling of triboelectric and chemoresistive effects. Adv Funct
Mater 28(10):1703420
99. Park S, Heo SW, Lee W, Inoue D, Jiang Z, Yu K et al (2018) Self-powered ultra-flexible elec-
tronics via nano-grating-patterned organic photovoltaics. Nature 561:516–521
100. Fan F-R, Tian Z-Q, Lin Wang Z (2012) Flexible triboelectric generator. Nano Energy
1(2):328–334
101. Zhang Y, Peng M, Liu Y, Zhang T, Zhu Q, Lei H et al (2020) Flexible self-powered real-time
ultraviolet photodetector by coupling triboelectric and photoelectric effects. ACS Appl Mater
Interfaces 12(17):19384–19392
102. Xie X, Wen Z, Shen Q, Chen C, Peng M, Yang Y et al (2018) Impedance matching effect
between a triboelectric nanogenerator and a piezoresistive pressure sensor induced self-­
powered weighing. Adv Mater Technol 3(6):1800054
Chapter 21
Self-Powered Photodetector

Hemant Kumar and Satyabrata Jit

21.1 Introduction

The development in the very-large-scale-integration (VLSI) technology has enabled


the Internet of Things (IoT) networks to find applications in remotely controlled
monitoring of the forest, civil works, agricultural lands, wireless capsule endoscopy,
telepresence robots etc. [1]. These monitoring devices can capture and stream data
containing high-resolution images. It consists of processing and transmitting a large
volume of data wirelessly through IoT devices. A large number of various sensors
are required for converting various physical phenomena into electrical signals in the
IoT devices and systems. Photodetector may be viewed as an optical sensor which
is an important component of the IoT based optical image processing networks [2].
In general, photodetector is a device used to convert optical signals of some desired
wavelengths into electrical signals in the form of a current or a voltage. In addition
to IoT applications, photodetectors are used in the field of cameras, non-destructive
testing, mimicking artificial eye, optical computing systems, biological and envi-
ronmental applications [3].
The very basic form of the photodetectors is simply a pn junction operated under
reverse bias condition by using an externally connected power supply. When optical
signal is incident on the device, excess electron-hole pairs are generated in the
device due to the photoelectric effect. The applied reverse bias creates a large

H. Kumar
Department of Electronics and Communication Engineering, Jaypee Institute of Information
Technology, Noida, India
e-mail: hemant.kumar@jiit.ac.in
S. Jit (*)
Department of Electronics Engineering, Indian Institute of Technology (Banaras Hindu
University), Varanasi, India
e-mail: sjit.ece@iitbhu.ac.in

© The Author(s), under exclusive license to Springer Nature Switzerland AG 2023 495
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1_21
496 H. Kumar and S. Jit

electric field in the depletion region of the pn junction which helps in drifting out
these excess photogenerated electrons and holes in opposite directions at a very
high speed. The collection of the carriers results in an illumination-dependent cur-
rent component, called the photocurrent, in the photodetectors. The external power
supply in the form of a battery is thus an integral part of the conventional photode-
tectors for their operation [4]. However, self-powered photodetectors under discus-
sion do not require any external power supply in the form of a battery for their
operation. This class of photodetectors senses the incident photons and generates
sufficient power internally by photovoltaic effect for their operation. The self-­
powered photodetectors are very much suitable for future generation IoT devices to
be operated with optimal or minimal power requirements.
The working of self-powered photodetectors is closely related to that of the pho-
tovoltaic devices (i.e. solar cells) operated under short-circuited mode or open-­
circuited mode [5]. The self-powered photodetectors can be classified into two
types. The first type of self-powered photodetectors independently generates suffi-
cient power on their own and they do not require any external power supply or a
separate energy harvesting unit for their operation [6]. The second type of self-­
powered photodetectors is the integration of the conventional photodetectors with a
separate energy harvesting unit which supplies energy required for the operation of
the photodetectors connected to it [7]. The comparative illustration between self-­
powered photodetector and traditional photodetector is shown in Fig. 21.1. The
minimal circuit requirements with ideal conditions have been used to demonstrate
the difference between a traditional photodiode (which requires an external power

Fig. 21.1 Differentiation between the working of traditional and self-powered photodetector
under minimal conditions
21 Self-Powered Photodetector 497

supply for its operation) and a self-powered detector (which requires no power sup-
ply for its operation). Both the devices under dark conditions behave as an open
circuited element due to zero reverse bias current/dark current. Under illumination,
both the conventional and self-powered photodetectors act as an illumination-­
dependent current source as shown in the Fig. 21.1.
The tremendous development in the synthesis and fabrication of low-­dimensional
materials with a variety of new electrical and optical properties have enabled the
researchers to explore the nanostructured materials for the fabrication of self-­
powered photodetectors. Different nanostructures such as nanobelts (NBs), nanorib-
bons (NRs), and other one-dimensional nanostructures of high carrier mobility and
optical absorption coefficient have been used for the fabrication of self-powered
photodetectors [8, 9]. However, the requirement of sophisticated and expensive
equipment for the fabrication of the inorganic nanostructures puts major restriction
in achieving low-cost inorganic nanomaterials based self-powered photodetectors.
To fabricate the low-cost photodetectors, researchers have explored simple and
cost-effective solution-processed quantum dots (QDs) for photodetectors [6, 7], and
solar cells [10]. Various organic semiconducting materials have also been used for
the fabrication of low-cost self-powered photodetectors [5]. The major limitations
of the organic materials are their low carrier mobility and poor stability. That is why,
the organic photodetectors suffer from poor performance over inorganic photode-
tectors. Therefore, some researchers have investigated hybrid type self-powered
photodetectors using both the organic and inorganic semiconductors in same the
device [5, 11, 12].
This chapter presents various types of self-powered photodetectors of different
kind of materials and device structures. The discussion is restricted to the Schottky-­
based self-powered photodetectors and heterojunction-based self-powered photode-
tectors. Finally, spectrum selective self-powered photodetectors have also been
introduced.

21.2 II–VI Materials Based Self-Powered


Schottky Photodetectors

The Schottky junction or rectifying junction is basically a junction between a metal


and a n-type (p-type) semiconductor with the work function of the metal is higher
(lower) than the n-type (p-type) semiconductor. Otherwise, the metal-­semiconductor
junction is known as the ohmic junction. The Schottky junction is considered to be
equivalent to that of a p+-n (n+-p) junction if the semiconductor on which the contact
is formed is a n-type (p-type) material. In other words, the Schottky junction behaves
similarly as a p-n junction diode which conducts current only in the forward biased
condition. On the other hand, a metal-semiconductor ohmic junction simply behave
as a resistor which transports current in both directions irrespective of its biasing
condition. Both types of junctions are used in both the conventional and
498 H. Kumar and S. Jit

self-­
powered photodetector devices. Two major photodetector structures are
described below:
1. Photoconductor Type Photodetectors: This type of photodetectors consists of
a semiconductor with two ohmic contacts on it [13]. This structure requires a
continuous high potential between the two ohmic contacts made on the
­semiconductor. If an optical illumination is applied on the device, excess elec-
tron-hole pairs are generated in the semiconductor which changes the conductiv-
ity of the material thereby increasing the current in the external circuit. Since
both contacts are of ohmic type, polarity of biasing does not affect the device
characteristics. The device’s response is expressed in terms of the variation of the
resistivity under illumination. These structures provide very high photoconduc-
tive gain and high responsivity.
2. Schottky Junction Photodiodes: This type of photodetectors use one Schottky
contact and one ohmic contact on the semiconductor working as the photoactive
material in the device [14]. This kind of structure will produce rectifying behav-
ior due to the Schottky junction. The Schottky photodiodes are operated under a
reverse bias condition. Thus, a very small current flows in the external circuit
under dark condition. When the device is exposed to illumination, excess
electron-­hole pairs are generated due to photoelectric effect in the device. These
photo-generated excess electrons and holes are drifted in the opposite directions
by the high electric field in the depletion region of the Schottky junction resulted
due to the applied reverse bias voltage. The extraction of photogenerated elec-
trons and holes results in a significant current in the circuit proportional to the
intensity of the incident illumination.
In general, the Schottky junction photodiodes are faster and more sensitive to inci-
dent photons than the photoconductor type photodetectors. However, the selection
of both the metals and photoactive materials plays crucial roles in determining the
performance of the Schottky photodetectors. Some important works on the Schottky
contact based self-powered photodetectors using II-VI group materials are dis-
cussed in the following sub-sections.

21.2.1 Graphene or Carbon Based Self-Powered


Schottky Photodetectors

Graphene is one of the most versatile 2D materials used for making the Schottky
contacts on CdSe and ZnO materials [8, 9, 15]. Jin et al. [8] fabricated a graphene-­
based Schottky junction on a CdSe nanobelt (NB) at 1000 ° C using the chemical
vapor deposition (CVD) method shown in Fig. 21.2a. They achieved a very high
photosensitivity of 3.5 × 105 and the transient behavior of 82 μs rise time (tr) and
179 μs of decay time (td) shown in Fig. 21.2c with the corresponding experimental
schematic shown in Fig. 21.2b. The CdSe NB/graphene-based Schottky device
21 Self-Powered Photodetector 499

Fig. 21.2 (a) FESEM image of the photodetector showing CdSe NB with Graphene; (b) Schematic
diagram for obtaining a transient response using chopper; (c) Photodetector’s transient response
under 633 nm laser at 1000 Hz light switching frequency. (Reprinted with permission from [8].
Copyright 2012: The Royal Society of Chemistry)

showed a very high photoconductive gain of 28 at 633 nm laser light. Gao et al. [15]
fabricated a similar structure on PET flexible substrates and reported an improved
transient response with 70 μs rise-time and 137 μs fall-time. Though they did not
achieve any significant improvement in the transient response, but the realization of
the CdSe based self-powered photodetectors on the PET flexible substrates could be
viewed as a great achievement.
Graphene is also used for Schottky contact formation on the ZnO for fabricating
the fast-responsive self-powered UV photodetectors. Boruah et al. [9] sandwiched a
layer of ZnO nanorods (NR) between two graphene layers to demonstrate a very
fast photodetector with the rise-time of tr = 3 s and fall-time of td = 0.47 s under UV
illumination. The significant improvement in the response speed encouraged other
researchers to work on ZnO/graphene based photodetectors. Chen et al. [16] fabri-
cated a ZnO/graphene Schottky contact based self-power photodetector by treating
the ZnO with H2O2 to improve the barrier height at the ZnO/Graphene interface.
They observed a significant improvement in the transient characteristics of the
device with a rise-time of only 32 ms. Duan et al. [17] used the graphene-based
Schottky contact on a thick film of Al-doped ZnO nanorods to achieve a significant
improvement in the UV-to-Visible rejection ratio of 1 × 102 with a responsivity of
0.039 A/W. A self-powered Schottky photodiode based on carbon dot-decorated
ZnO Nanorods on a graphite-coated paper substrate was fabricated by Sinha et al.
500 H. Kumar and S. Jit

[18] for UV Detection applications. The transient response of the device was very
poor with a rise-time tr = 2 s and a fall-time td = 3.2 s.

21.2.2 Other Self-Powered Schottky Photodetectors

Being a noble metal, Gold (Au)-based electrode is frequently used for the fabrica-
tion of Schottky junction photodetectors. Chen et al. [19] used an asymmetric
metal-semiconductor-metal (MSM) self-powered photodetector structure with two
Au Schottky contacts of different widths on the same semiconductive material
(MgZnO). The asymmetric self-powered photodetector showed an increased
responsivity upto 20 mA/W. They [19] also demonstrated the impact of asymmetric
Au contact structures on the performance of the device in the UV region. Their
proposed photodetector showed the self-powered characteristics when asymmetric
ratio was maintained above 20:1. Benyahia et al. [20] fabricated a ZnO-ZnS
microstructure-­based broadband self-powered photodetector using Au Schottky
contact. The device showed a wideband response ranging from 300 nm to 900 nm.
Au Schottky electrode is also used on composite materials for the fabrication of
self-powered photodetectors. Ebrahimi et al. [21] fabricated an Au Schottky contact
on SnxZn1 − xS composite material prepared by using the low-cost solvothermal pro-
cess. They observed a reduction in the bandgap of the materials with increased Sn
content in the composite material. Schematic illustrations of the device fabrication
and variation of bandgap with Urbach energy (for varying concentration) are shown
in Fig. 21.3a, and b, respectively. The device was shown to possess the self-powered
feature in the UV-B region. The I-V characteristics of pure ZnS based device and
band-alignment diagram are shown in Fig. 21.3c, and d, respectively. In another
work [22], the same group investigated the transient response characteristics with
rise-time of tr = 1.9 ms and fall-time of td = 2.6 ms of the same device structure for
varying bandgaps of SnxZn1 − xS obtained by varying the Sn concentration in the
active material.
Besides the gold, silver (Ag) with a work function of 4.2 eV [23] is also used for
the fabrication of the Schottky junctions with ZnO for self-powered photodetectors.
Purushotaman et al. [23] reported Ag/ZnO nanorods (NRs) Schottky junction pho-
todetector shown in Fig. 21.4a. They developed floral ZnO NRs over the flexible
PVDF substrates and explored the piezotronics property of the ZnO to obtain the
self-powered feature of the photodetectors. The SEM image in Fig. 21.4b shows the
surface morphology of the PVDF film while the SEM images of Fig. 21.4c–e show
the morphology of the film containing the floral ZnO NRs. The flexible nature of the
self-powered photodetector is illustrated in Fig. 21.4f.
ZnO-based self-powered photodetectors are primarily designed to operate in the
UV region. For the self-powered photodetectors operating in the visible or NIR
region, researchers either use doped ZnO or use different materials for the active
region of the device. Selenium-based microrods were used to achieve wideband
photo-response ranging from UV to the entire visible region [24] as shown in
21 Self-Powered Photodetector 501

Fig. 21.3 (a) Schematic indicating the fabrication process for obtaining SnxZn1 − xS thin film based
Self-Powered photodetector; (b) Variation in band-gap against Urbrach energy for varying concen-
tration of Sn and Zn; (c) I-V curves under the dark and UV illumination for the MSM UV detector
(as-grown and annealed) based on Au/pure ZnS thin films/Au planar structure; (d) Band-Alignment
of MSM device under dark and under UV illumination. (Reprinted with permission from [21].
Copyright 2020: Elsevier)

Fig. 21.4 (a) Schematic for the fabrication of flexible self-powered photodetector; (b) Indicates
the untreated PVDF film; (c) SEM image of floral ZnO NRs; (d, e) magnified results, and (f)
Fabricated flexible self-powered photodetector. (Reprinted with permission from [23]. Copyright
2018: American Chemical Society)

Fig. 21.5a, b. A Schottky junction between the Se and Ga-In alloy was created to
obtain a very low dark current of 200 fA. The fabricated device showed a very high
responsivity and detectivity of 408 mA/W and 1.30 × 1013 J, respectively, with
tr = 124 μs and td = 146 μs as shown in Fig. 21.5c.
502 H. Kumar and S. Jit

Fig. 21.5 (a) SEM image of single Se rod; (b) Obtained broadband responsivity and detectivity of
the self-powered photodetector, and (c) Transient response of the device at 532 nm wavelength
under different bias. (Reprinted with permission from [24]. Copyright 2019: American Chemical
Society)

To enhance the photo response of the self-powered Schottky photodetectors,


some researchers have used a dielectric layer between the metal and semiconductor
in the device [25]. Zhang et al. [25] introduced a dielectric layer of Al2O3 between
Pt and ZnO to increase the Schottky barrier height. The device’s overall efficiency
was improved by 2.77 times as compared to the device without a dielectric layer due
to the Piezotronic effect.

21.2.3 II–VI Quantum Dots-Based Self-Powered


Schottky Photodetectors

A very limited amount of research is reported on the quantum dots (QDs) based
self-powered photodetectors. Kumar et al. [6] fabricated a self-powered Schottky
junction photodiode using Au on CdSe QDs deposited on ZnO QDs based electron
transport layer as shown in Fig. 21.6a, b. The device showed an optical response
over 350–750 nm (i.e. covering the complete visible region) with an ultrafast
response of tr = 18 ms and td = 17.9 ms. It was shown that the inclusion of the ZnO
QDs layer improved the photoresponse of the fabricated device by nearly 17 times
as shown in Fig. 21.6c. The same group [26] also fabricated ITO/ZnO QDs/CdSe
QDs/(Pd, Au) based self-powered photodetectors and compared the photoresponse
21 Self-Powered Photodetector 503

Fig. 21.6 (a) Schematic of QD-based self-powered photodetector; (b) Optical image of the fabri-
cated device; (c) I-V characteristics under dark and illumination, and inset showing I-V character-
istics without charge transport layer. Reprinted with permission from [6]. Copyright 2017: IEEE;
(d) Schematic of the device, and (e, f) Transient behavior device with Au and Pd electrode respec-
tively. Reprinted with permission from [26]. Copyright 2019: IEEE

of the Au/CdSe QDs and Pd/CdSe QDs Schottky contact photodetectors as illus-
trated in Fig. 21.6d. They observed that the metal electrode greatly affected the
spectral coverage of the device. The Pd based electrode increased the quantum effi-
ciency by 2.1 times by reducing the FWHM (full width at the half maximum) from
190 nm to 61 nm. Further, the Pd based device showed better transient response of
tr = 18.5 ms and td = 15.8 ms over the Au based device. Comparative results are
shown in Fig. 21.6e, f. Moreover, the spectral coverage of the device reported in
[26] was observed to be reduced (380–600 nm) as compared to the device reported
in [6]. In case of [6], light was incident from the top side of the device whereas the
light was allowed to enter into the device from the backside of the device considered
in [26]. The improved performance in [26] was attributed to the filtering action of
the ZnO QDs in addition to its normal electron transportation from the active region
of photodetector.

21.3 Heterojunction Based Self-Powered Photodetectors

A heterojunction is formed between two different types of semiconducting materi-


als of different bandgap energies. In this section, we will discuss some state-of-the-­
art heterojunction based self-powered photodetectors. We will first consider some
conventional heterojunction photodetectors. Then some special self-powered photo-
detectors using the 2D transition metal dichalcogenide (TMD) heterostructure
504 H. Kumar and S. Jit

materials and organic/inorganic hybrid heterojunction based self-powered photode-


tectors will be discussed in the following.

21.3.1 Some Conventional Heterojunction Based


Self-Powered Photodetectors

Bie et al. [27] fabricated a self-powered visible-blind UV photodetector by using


heterojunctions between a single n-type ZnO nanowire and a p-type GaN film. They
observed an ultrafast response with a rise time of ~20 μs and a fall-time of ~219 μs
which was two orders of magnitude faster than the conventional ZnO
photoconductivity-­based photodetectors. They also integrated the ZnO/GaN hetero-
junction with a CdSe nanowire device to achieve a selective multiwavelength pho-
todetector as shown in Fig. 21.7a. The SEM images of the n-ZnO/p-GaN structure
and CdSe NW device are shown in Fig. 21.7b, c. The maximum output power of
1.1 μW was obtained from the self-powered photodetector device [27].
Yamada et al. [28] fabricated a transparent self-powered photodetector using
p-CuI and n-InGaZnO heterojunction. The device worked as a self-powered device

Fig. 21.7 (a) Schematic of n-ZnO/p-GaN with derived CdSe NW photodetector, (b) SEM image
of ZnO NW and GaN, and (c) SEM image of CdSe NW. (Reprinted with permission from [27].
Copyright 2010: John Wiley and Sons)
21 Self-Powered Photodetector 505

under photovoltaic mode and showed a spectral sensitivity under UV region only.
Rana et al. [29] fabricated a transparent self-powered photodetector using a p-NiO
and n-ZnO heterojunction. Pyro-phototronic effect and photovoltaic mode were
explored to improve the responsivity and detectivity by 5460% and 6063% in the
UV region, respectively. The proposed device showed a tr = 3.92 μs and td = 8.90 μs.
Authors also tried to increase the spectral coverage of the detector by integrating
narrowband semiconductors.
Huang et al. [30] reported Si NW/InGaZnO self-powered photodetector to
achieve the spectral coverage in UV-NIR region by integrating the wide bandgap
material InGaZnO with the lower bandgap material Si. The device showed a peak at
~400 nm with a response time of 0.2 ms. Similarly, a self-powered photodetector
was fabricated by Hasan et al. [31] by integrating n-ZnO NRs (wide bandgap) with
p-Si (low bandgap). They obtained a rise-time tr = 25 ms and fall-time td = 22 ms.
Heterojunctions between the carbon and II-VI group semiconductors are also
reported for self-powered photodetectors. Huag et al. [32] reported a self-powered
UV–visible photodetector based on ZnO/graphene/CdS/electrolyte heterojunctions.
The device showed major response in the UV region with some tails extended to the
blue part of the visible spectrum. Though the authors considered the detector under
self-powered category, the results were not encouraging enough for such a claim.
Hatch et al. [33] reported a n-ZnO/p-Copper Thiocynate (CuSCN) heterojunction
based self-powered photodetector for operating in the UV region. The device
showed an ultrafast response of tr = 500 ns and td = 6.7 μs. In another work, the same
group attempted to improve the response speed by reducing tr from 500 ns to 25 ns
of the device in the same device structure by just modifying the ZnO annealing
temperature. Ghamgosar et al. [34] fabricated a self-powered photodetector using
n-ZnO and p − Co3O4 in core-shell structure with an Al2O3 buffer layer. The buffer
layer was shown to improve the photoresponsivity by six times of the device.
However, the results were not encouraging enough to consider the device under the
self-bias category. Guo et al. [35] fabricated a ZIF-8@H:ZnO core-shell NRs array/
Si based self-powered broadband photodetector. This device also showed a very
poor self-biased nature. Researchers have also tried to deplete the active layer by
introducing various heterostructures to improve the photoresponse characteristics of
the photodetectors. Mishra et al. [36] used ZnO/GaN heterostructure to improve the
dark to photocurrent ratio. They analyzed a variety of structures, though the devices
did not offer any promising results under self-biased conditions. Effect of doping on
the performance characteristics of the self-powered photodetectors have also been
investigated by some researchers. Ga doping in ZnO is one of the most common
techniques employed for expanding the photoresponse of the ZnO based photode-
tectors. In this direction, Saha et al. [37] prepared Ga doped ZnO NW coated with
Ag and deposited the same over the p-Si substrate as shown in Fig. 21.8a, b. The
fabricated Ga-ZnO/p-Si heterojunction offered a spectral coverage from 320 nm to
400 nm. A varying photo-response with red-shifted characteristics was observed
with increased Ga doping as shown in Fig. 21.8c.
Zhou et al. [38] fabricated a Perovskite/Ga-doped ZnO (GZO) NRs heterojunction-­
based self-powered photodetector with a broad spectrum covering the entire visible
506 H. Kumar and S. Jit

Fig. 21.8 (a, b) Schematic of the fabricated device using Ga-doped ZnO NW with Ag coating also
indicating with SEM image, and (c) EQE of different Ga-doped ZnO devices. (Reprinted with
permission from [37]. Copyright 2020: Elsevier)

Fig. 21.9 (a) I–V characteristics under dark and illumination, and vertical SEM image indicating
device structure is shown in inset and (b) Transient response under self-bias for the different illu-
mination power densities. (Reprinted with permission from [40]. Copyright 2020: John Wiley
and Sons)

band (400–800 nm). The device did not exhibit any variation in the transient
response for varying Ga doping concentrations. Chen et al. [39] fabricated a self-­
powered photodetector using n − MgxZn1 − xO alloy/p-Si heterojunctions. The photo-­
response characteristics of the device were optimized by modifying the pizeopotential
MgxZn1 − xO by controlling its Mg content. Jiang et al. [40] fabricated a SnS2/ZnO1 − xSx
heterojunction based self-powered photodetector shown in the inset of Fig. 21.9a
with its current-voltage characteristics under dark and under illumination (365 nm)
shown in the Fig. 21.9a. The fabricated self-powered photodetector exhibited a
broadband detection capability over 365 nm to 850 nm. The transient response of
the detector is shown in Fig. 21.9b with varying power intensity. The device
showed tr = 49.51 ms and td = 25.93 ms.
21 Self-Powered Photodetector 507

21.3.2 2D Semiconducting Transition Metal Dichalcogenide


(TMD) Heterostructure Materials Based
Self-Powered Photodetectors

Researchers have also explored 2D semiconducting transition metal dichalcogenide


(TMD)-based van der Waals (vdW) heterostructures for self-powered operations
[41, 42]. Zou et al. reported a vertical GaSe/MoS2 pn heterojunction self-powered
photodetector grown by liquid gallium (Ga)-assisted chemical vapor deposition
method [41]. The device exhibited a sizeable open-circuit voltage (0.61 V) with a
broadband detection capability over 375–633 nm. Under self-powered operation, the
device showed a high responsivity of 900 mA/W and a fast response speed of 5 ms.
A self-powered broadband photodetector based on vertically stacked WSe2/Bi2Te3
p–n heterojunctions was reported by Liu et al. [42]. The device exhibited a maximum
short-circuited current of 18 nA and an open circuit voltage of 0.25 V at 633 nm
(with an incident power density of 26.4 mW/cm2). The device’s detection range was
extended from the visible to near-infrared (375–1550 nm). The photodetector showed
a fast response time (∼210 μs) and high responsivity (20.5 A/W at 633 nm and
27 mA/W at 1550 nm) under zero bias voltage [42]. A flexible self-­powered photo-
diode using MoS2/WSe2 vertically stacked vdW heterostructures was reported by
Lin et al. [43]. Though they failed to measure the device’s response time accurately,
the detector was expected to provide a fast response speed. Ahn et al. [44] fabricated
a MoTe2/MoS2 semi-vertically stacked vdW heterojunction based self-powered pho-
todetector with a photo detection capability ranging from the violet (405 nm) to
near-infrared (1310 nm) [44]. A WSe2/Bi2O2Se 2D/2D vdW heterostructure-­based
self-powered photodetector with a broadband spectrum over 365 to 2000 was
reported by Luo et al. [45]. The schematic of the fabricated self-­powered photodetec-
tor showing vdW based heterostructure between WSe2 and Bi2O2Se is shown in
Fig. 21.10a. The transient response of the detector under 532 nm wavelength
(20 kHz) at Vds = − 5V gave the rise time of 2.4 μs and fall time of 2.6 μs (Fig. 21.10b).
The electrical power generated by the WSe2/Bi2O2Se heterostructure under different
illumination intensities is shown in Fig. 21.10c as a function of bias voltage. Recently,
Zheng et al. [46] demonstrated the growth of quasi-van der Waals epitaxial (QvdWE)
based vertically aligned one-dimensional (1D) GaN nanorod arrays (NRAs) on
TMDs/Si substrates for high-performance self-powered photodetection applications.
A competitive photovoltaic photoresponsivity over 10 AW−1 under a weak detectable
light signal was demonstrated under self-bias conditions.

21.3.3 Inorganic-Organic Hybrid Heterostructures Based


Self-Powered Photodetectors

Organic semiconductors are suitable for the flexible electronics devices. But the
major drawbacks of the organic semiconductors are their poor carrier mobility, low
optical absorption coefficient and poor stability over longer period. That is why,
508 H. Kumar and S. Jit

Fig. 21.10 (a) Schematic of the WSe2/Bi2O2Se 2D/2D vdW heterostructure-based self-powered
photodetector, (b) Transient response of the device showing rise time of 2.4 μs and fall time of
2.6 μs under 532 nm (20 KHz) at Vds = − 5 V, and (c) Generated electrical power by the hetero-
structure as a function of Vds. (Reprinted with permission from [45]. Copyright 2020: John Wiley
and Sons)

many researchers have explored the hybrid heterojunctions formed between an


organic semiconductor and an inorganic semiconductor for the self-powered photo-
detectors. Sarkar et al. [47] fabricated the PEDOT:PSS/ZnO@CdS NR core-shell
heterojunction based self-powered photodetectors. The photo-response spectrum
covered the UV and visible region with a transient response of 20 ms. Zhan et al.
[48] fabricated a rGO/ZnO hybrid heterojunction based self-powered photodetector.
Game et al. [12] explored the hybrid heterojunction between n-type Nitrogen-doped
ZnO and p-type Spiro-MeOTAD for fabricating self-powered visible photodetectors
with rise time tr = 200 μs and fall-time or decay time td = 950 μs. The self-biasing
characteristic of the detector is shown in Fig. 21.11a, b. The experimental setup and
transient characteristics of the device are shown in Fig. 21.11c, d, respectively.
Kumar et al. [49] fabricated CdSe QDs/PQT-12 based hybrid self-powered photo-
detector shown in Fig. 21.12a. The photovoltaic effect in the fabricated detector is
shown in Fig. 21.12b. The device exhibited a rise-time tr = 15.32 ms and a fall-time
td = 12.01 ms as shown in Fig. 21.12c.
Fig. 21.11 Current N: ZnO-SPD self-powered photodetector (a) under dark, (b) under illumina-
tion, (c) Schematic of the experimental setup for obtaining a transient response, and (d) transient
behavior under self-bias condition. (Reprinted with permission from [12]. Copyright 2014: The
Royal Society of Chemistry)

Fig. 21.12 (a) Schematic of fabricated CdSe QD/PQT-12 self-powered photodetector; (b) I-V
characteristics of hybrid photodetector, and (c) Transient response of the self-powered hybrid pho-
todetector. (Reprinted with permission from [49]. Copyright 2017: IEEE)
510 H. Kumar and S. Jit

21.4 Spectrum Selective Self-Powered Photodetectors

Spectrum selective photodetectors with high detectivity, high signal-to-noise ratio,


and fast response speed are important for various applications such as fluorescent-­
based biomedical imaging [50] and mimicking the human eye [51]. Spectrum selec-
tivity of the photodetectors is defined as its ability to detect the incident photons of
a single desired wavelength (or photon in a very narrowband of wavelengths cen-
tered about the desired wavelength) from the incident photons of different wave-
lengths on the device. In this section, we will discuss some self-powered
photodetectors with a very narrow full-width at half maximum (FWHM). Chen
et al. [52] reported a self-powered UV photodetector using Schottky junction
between MgZnO and asymmetric Au electrode. The fabricated device showed a
very narrowband spectral response in the UV region with a peak at ~320 nm. The
responsivity of the device is shown in Fig. 21.13a. The asymmetric Au contact-
based device structure and its SEM image are shown Fig. 21.13b, and c, respectively.
Ni et al. [53] fabricated an Au/i-ZnO/n-ZnO/Al2O3 structure based self-powered
spectrum selective photodetector shown in Fig. 21.14a. The n-ZnO layer was used
to act as a filter layer, whereas the i-ZnO layer was used to act as a multiplier via
impact ionization. The fabricated photodetector showed a responsivity with FWHM
of 9 nm under the self-biased condition. Fig. 21.14b shows the responsivity at −5 V
reverse bias voltage, while the inset of Fig. 21.14b shows the increasing trend of the
responsivity with increased reverse bias voltage. High responsivity was achieved at

Fig. 21.13 (a) Responsivity of the Au/MgZnO device at self-bias, (b) schematic of device struc-
ture with asymmetric Au electrode, and (c) SEM image confirming the asymmetric Au electrodes.
(Reprinted with permission from [52]. Copyright 2019: IOP Publishing)
21 Self-Powered Photodetector 511

Fig. 21.14 (a) Responsivity of the i-ZnO/n-ZnO device at self-bias with inset showing device
structure, and (b) Improved responsivity of the fabricated device under −5 V indicating reduced
FWHM with a variation of responsivity against applied reverse bias voltage shown in inset.
(Reprinted with permission from [53]. Copyright 2012: American Chemical Society)

high reverse bias voltage at the cost of reduced spectrum selectivity, as shown in
Fig. 21.14b. Shen et al. [54] fabricated a homojunction between n-ZnO and p-ZnO
for UV detection applications. They deposited both n-ZnO and p-ZnO on a sapphire
substrate by using the molecular beam epitaxy method. Li, and N dopants were used
to achieve the p-ZnO to act a filter layer, whereas the n-ZnO acted as the active
layer. The combined effect of the filter layer and the active layer provided a very
narrow spectral width of 9 nm. The detection capability of the detector was extended
from the visible to the NIR region. Li et al. [55] fabricated CdSe/p − Sb2(S1 − xSex)3
heterojunction based self-powered narrowband photodetectors without using any
filter layer to achieve a FWHM of ~35 nm. The device was capable to detect light in
650–900 nm. It was shown that the FWHM of the device could be controlled by
varying the concentration of Se. Wei et al. [56] fabricated a p-NiO/Al-doped n-ZnO
heterojunction based self-powered photodetector using the low-cost chemical bath
deposition technique. The device was shown to possess a narrow FWHM at
~400 nm. Kumar et al. [57] fabricated a TiO2/Si-based self-powered photodetector
demonstrates its application in digital communication by sending optical pulses.

21.5 Conclusion

Advances in different types of self-powered photodetectors using Schottky junc-


tions, heterojunctions, 2D heterostructure materials, colloidal quantum dots, and
hybrid organic/inorganic heterojunctions have been discussed in this chapter.
Among various II-VI group materials, most of the researchers have mainly explored
ZnO and CdSe for the fabrication of self-powered photodetectors to primarily cover
the UV and visible regions. It is observed that only a limited amount of work on
512 H. Kumar and S. Jit

QD-based self-powered photodetectors has been reported. Particular emphasis must


be given to developing self-powered photodetectors with detection capability in the
NIR and IR regions.

References

1. Ayaz M, Ammad-Uddin M, Sharif Z, Mansour A, Aggoune E-HM (2019) Internet-of-Things


(IoT)-based smart agriculture: toward making the fields talk. IEEE Access 7:129551–129583.
https://doi.org/10.1109/ACCESS.2019.2932609
2. Masoud M, Jaradat Y, Manasrah A, Jannoud I (2019) Sensors of smart devices in the Internet
of Everything (IoE) era: big opportunities and massive doubts. J Sens 2019:1–26. https://doi.
org/10.1155/2019/6514520
3. Jansen-van Vuuren RD, Armin A, Pandey AK, Burn PL, Meredith P (2016) Organic photo-
diodes: the future of full color detection and image sensing. Adv Mater 28(24):4766–4802.
https://doi.org/10.1002/adma.201505405
4. Rawat G, Somvanshi D, Kumar Y, Kumar H, Kumar C, Jit S (2016) Electrical and ultraviolet-a
detection properties of E-beam evaporated n-TiO2 capped p-Si nanowires heterojunction pho-
todiodes. IEEE Trans Nanotechnol:1–1. https://doi.org/10.1109/TNANO.2016.2626795
5. Bera A, Das Mahapatra A, Mondal S, Basak D (2016) Sb2S3/Spiro-OMeTAD inorganic–
organic hybrid p–n junction diode for high performance self-powered photodetector. ACS
Appl Mater Interfaces 8(50):34506–34512. https://doi.org/10.1021/acsami.6b09943
6. Kumar H, Kumar Y, Mukherjee B, Rawat G, Kumar C, Pal BN et al (2017) Electrical and
optical characteristics of self-powered colloidal CdSe quantum dot-based photodiode. IEEE J
Quantum Electron 53(3):1–8. https://doi.org/10.1109/JQE.2017.2696487
7. Lu H, Tian W, Cao F, Ma Y, Gu B, Li L (2016) A self-powered and stable all-perovskite
photodetector-­ solar cell nanosystem. Adv Funct Mater 26(8):1296–1302. https://doi.
org/10.1002/adfm.201504477
8. Jin W, Ye Y, Gan L, Yu B, Wu P, Dai Y et al (2012) Self-powered high performance photodetec-
tors based on CdSe nanobelt/graphene Schottky junctions. J Mater Chem 22(7):2863. https://
doi.org/10.1039/c2jm15913a
9. Boruah BD, Mukherjee A, Misra A (2016) Sandwiched assembly of ZnO nanowires
between graphene layers for a self-powered and fast responsive ultraviolet photodetector.
Nanotechnology 27(9):095205. https://doi.org/10.1088/0957-­4484/27/9/095205
10. Sargent EH (2012) Colloidal quantum dot solar cells. Nat Photonics 6(3):133–135. https://doi.
org/10.1038/nphoton.2012.33
11. Xie Y, Wei L, Li Q, Wei G, Wang D, Chen Y et al (2013) Self-powered solid-state photodetec-
tor based on TiO2 nanorod/spiro-MeOTAD heterojunction. Appl Phys Lett 103(26):261109.
https://doi.org/10.1063/1.4858390
12. Game O, Singh U, Kumari T, Banpurkar A, Ogale S (2014) ZnO(N)–Spiro-MeOTAD hybrid
photodiode: an efficient self-powered fast-response UV (visible) photosensor. Nanoscale
6(1):503–513. https://doi.org/10.1039/C3NR04727J
13. Kumar Y, Kumar H, Rawat G, Kumar C, Pal BN, Jit S (2018) Spectrum selectivity and respon-
sivity of ZnO nanoparticles coated Ag/ZnO QDs/Ag UV photodetectors. IEEE Photon Technol
Lett 30(12):1147–1150. https://doi.org/10.1109/LPT.2018.2836978
14. Kumar Y, Kumar H, Mukherjee B, Rawat G, Kumar C, Pal BN et al (2017) Visible-blind Au/
ZnO quantum dots-based highly sensitive and spectrum selective Schottky photodiode. IEEE
Trans Electron Devices 64(7):2874–2880. https://doi.org/10.1109/TED.2017.2705067
15. Gao Z, Jin W, Zhou Y, Dai Y, Yu B, Liu C et al (2013) Self-powered flexible and transpar-
ent photovoltaic detectors based on CdSe nanobelt/graphene Schottky junctions. Nanoscale
5(12):5576. https://doi.org/10.1039/c3nr34335a
21 Self-Powered Photodetector 513

16. Chen D, Xin Y, Lu B, Pan X, Huang J, He H et al (2020) Self-powered ultraviolet photovoltaic


photodetector based on graphene/ZnO heterostructure. Appl Surf Sci 529:147087. https://doi.
org/10.1016/j.apsusc.2020.147087
17. Duan L, He F, Tian Y, Sun B, Fan J, Yu X et al (2017) Fabrication of self-powered fast-response
ultraviolet photodetectors based on graphene/ZnO:Al nanorod-array-film structure with sta-
ble Schottky barrier. ACS Appl Mater Interfaces 9(9):8161–8168. https://doi.org/10.1021/
acsami.6b14305
18. Sinha R, Roy N, Mandal TK (2020) Growth of carbon dot-decorated ZnO Nanorods on a
graphite-coated paper substrate to fabricate a flexible and self-powered Schottky diode for
UV detection. ACS Appl Mater Interfaces 12(29):33428–33438. https://doi.org/10.1021/
acsami.0c10484
19. Chen H-Y, Liu K-W, Chen X, Zhang Z-Z, Fan M-M, Jiang M-M et al (2014) Realization of a
self-powered ZnO MSM UV photodetector with high responsivity using an asymmetric pair
of Au electrodes. J Mater Chem C 2(45):9689–9694. https://doi.org/10.1039/C4TC01839G
20. Benyahia K, Djeffal F, Ferhati H, Bendjerad A, Benhaya A, Saidi A (2021) Self-powered pho-
todetector with improved and broadband multispectral photoresponsivity based on ZnO-ZnS
composite. J Alloys Compd 859:158242. https://doi.org/10.1016/j.jallcom.2020.158242
21. Ebrahimi S, Yarmand B (2020) Solvothermal growth of aligned SnxZn1-xS thin films for tun-
able and highly response self-powered UV detectors. J Alloys Compd 827:154246. https://doi.
org/10.1016/j.jallcom.2020.154246
22. Ebrahimi S, Yarmand B (2021) Tunable and high-performance self-powered ultraviolet detec-
tors using leaf-like nanostructural arrays in ternary tin zinc sulfide system. Microelectron J
116:105237. https://doi.org/10.1016/j.mejo.2021.105237
23. Purusothaman Y, Alluri NR, Chandrasekhar A, Vivekananthan V, Kim S-J (2018) Direct in situ
hybridized interfacial quantification to stimulate highly flexile self-powered photodetector. J
Phys Chem C 122(23):12177–12184. https://doi.org/10.1021/acs.jpcc.8b02604
24. Chang Y, Chen L, Wang J, Tian W, Zhai W, Wei B (2019) Self-powered broadband
Schottky junction photodetector based on a single selenium microrod. J Phys Chem C
123(34):21244–21251. https://doi.org/10.1021/acs.jpcc.9b04260
25. Zhang Z, Liao Q, Yu Y, Wang X, Zhang Y (2014) Enhanced photoresponse of ZnO nanorods-­
based self-powered photodetector by piezotronic interface engineering. Nano Energy
9:237–244. https://doi.org/10.1016/j.nanoen.2014.07.019
26. Kumar H, Kumar Y, Mukherjee B, Rawat G, Kumar C, Pal BN et al (2019) Effects of optical
resonance on the performance of metal (Pd, Au)/CdSe quantum dots (QDs)/ZnO QDs optical
cavity based Spectrum selective photodiodes. IEEE Trans Nanotechnol 18:365–373. https://
doi.org/10.1109/TNANO.2019.2907529
27. Bie Y-Q, Liao Z-M, Zhang H-Z, Li G-R, Ye Y, Zhou Y-B et al (2011) Self-powered, ultrafast,
visible-blind UV detection and optical logical operation based on ZnO/GaN nanoscale p-n
junctions. Adv Mater 23(5):649–653. https://doi.org/10.1002/adma.201003156
28. Yamada N, Kondo Y, Cao X, Nakano Y (2019) Visible-blind wide-dynamic-range fast-response
self-powered ultraviolet photodetector based on CuI/In-Ga-Zn-O heterojunction. Appl Mater
Today 15:153–162. https://doi.org/10.1016/j.apmt.2019.01.007
29. Rana AK, Kumar M, Ban D-K, Wong C-P, Yi J, Kim J (2019) Enhancement in performance
of transparent p-NiO/n-ZnO heterojunction ultrafast self-powered photodetector via Pyro-­
Phototronic effect. Adv Electron Mater 5(8):1900438. https://doi.org/10.1002/aelm.201900438
30. Huang C-Y, Huang C-P, Chen H, Pai S-W, Wang P-J, He X-R et al (2020) A self-powered
ultraviolet photodiode using an amorphous InGaZnO/p-silicon nanowire heterojunction.
Vacuum 180:109619. https://doi.org/10.1016/j.vacuum.2020.109619
31. Hassan JJ, Mahdi MA, Kasim SJ, Ahmed NM, Abu Hassan H, Hassan Z (2012) High sensitiv-
ity and fast response and recovery times in a ZnO nanorod array/p-Si self-powered ultraviolet
detector. Appl Phys Lett 101(26):261108. https://doi.org/10.1063/1.4773245
32. Huang G, Zhang P, Bai Z (2019) Self-powered UV–visible photodetectors based on ZnO/
graphene/CdS/electrolyte heterojunctions. J Alloys Compd 776:346–352. https://doi.
org/10.1016/j.jallcom.2018.10.225
514 H. Kumar and S. Jit

33. Hatch SM, Briscoe J, Dunn S (2013) A self-powered ZnO-Nanorod/CuSCN UV photodetector


exhibiting rapid response. Adv Mater 25(6):867–871. https://doi.org/10.1002/adma.201204488
34. Ghamgosar P, Rigoni F, Kohan MG, You S, Morales EA, Mazzaro R et al (2019) Self-powered
photodetectors based on Core–Shell ZnO–Co3O4 nanowire heterojunctions. ACS Appl Mater
Interfaces 11(26):23454–23462. https://doi.org/10.1021/acsami.9b04838
35. Guo T, Ling C, Li X, Qiao X, Li X, Yin Y et al (2019) A ZIF-8@H:ZnO core–shell nanorod
arrays/Si heterojunction self-powered photodetector with ultrahigh performance. J Mater
Chem C 7(17):5172–5183. https://doi.org/10.1039/C9TC00290A
36. Mishra M, Gundimeda A, Garg T, Dash A, Das S, Vandana et al (2019) ZnO/GaN heterojunc-
tion based self-powered photodetectors: influence of interfacial states on UV sensing. Appl
Surf Sci 478:1081–1089. https://doi.org/10.1016/j.apsusc.2019.01.192
37. Saha R, Karmakar A, Chattopadhyay S (2020) Enhanced self-powered ultraviolet photo-
response of ZnO nanowires/p-Si heterojunction by selective in-situ Ga doping. Opt Mater
105:109928. https://doi.org/10.1016/j.optmat.2020.109928
38. Zhou H, Yang L, Gui P, Grice CR, Song Z, Wang H et al (2019) Ga-doped ZnO nanorod
scaffold for high-performance, hole-transport-layer-free, self-powered CH3NH3PbI3
perovskite photodetectors. Sol Energy Mater Sol Cells 193:246–252. https://doi.org/10.1016/j.
solmat.2019.01.020
39. Chen Y-Y, Wang C-H, Chen G-S, Li Y-C, Liu C-P (2015) Self-powered n-MgZnO/p-Si pho-
todetector improved by alloying-enhanced piezopotential through piezo-phototronic effect.
Nano Energy 11:533–539. https://doi.org/10.1016/j.nanoen.2014.09.037
40. Jiang J, Huang J, Ye Z, Ruan S, Zeng Y (2020) Self-powered and broadband photodetector
based on SnS2/ZnO1−xSx heterojunction. Adv Mater Interfaces 7(20):2000882. https://doi.
org/10.1002/admi.202000882
41. Zou Z, Liang J, Zhang X, Ma C, Xu P, Yang X et al (2021) Liquid-metal-assisted growth of
vertical GaSe/MoS2 p–n heterojunctions for sensitive self-driven photodetectors. ACS Nano
15(6):10039–10047. https://doi.org/10.1021/acsnano.1c01643
42. Liu H, Zhu X, Sun X, Zhu C, Huang W, Zhang X et al (2019) Self-powered broad-band
photodetectors based on vertically stacked WSe2/Bi2Te3 p–n heterojunctions. ACS Nano
13(11):13573–13580. https://doi.org/10.1021/acsnano.9b07563
43. Lin P, Zhu L, Li D, Xu L, Pan C, Wang Z (2018) Piezo-Phototronic effect for enhanced flex-
ible MoS2/WSe2 van der Waals photodiodes. Adv Funct Mater 28(35):1802849. https://doi.
org/10.1002/adfm.201802849
44. Ahn J, Kang J-H, Kyhm J, Choi HT, Kim M, Ahn D-H et al (2020) Self-powered visible–
invisible multiband detection and imaging achieved using high-performance 2D MoTe2/MoS2
Semivertical heterojunction photodiodes. ACS Appl Mater Interfaces 12(9):10858–10866.
https://doi.org/10.1021/acsami.9b22288
45. Luo P, Wang F, Qu J, Liu K, Hu X, Liu K et al (2021) Self-driven WSe2/Bi2O2Se Van der Waals
Heterostructure photodetectors with high light on/off ratio and fast response. Adv Funct Mater
31(8):2008351. https://doi.org/10.1002/adfm.202008351
46. Zheng Y, Cao B, Tang X, Wu Q, Wang W, Li G (2022) Vertical 1D/2D heterojunction architec-
tures for self-powered Photodetection application: GaN Nanorods grown on transition metal
Dichalcogenides. ACS Nano 16(2):2798–2810. https://doi.org/10.1021/acsnano.1c09791
47. Sarkar S, Basak D (2015) Self powered highly enhanced dual wavelength ZnO@CdS
Core–Shell Nanorod arrays photodetector: an intelligent pair. ACS Appl Mater Interfaces
7(30):16322–16329. https://doi.org/10.1021/acsami.5b03184
48. Zhan Z, Zheng L, Pan Y, Sun G, Li L (2012) Self-powered, visible-light photodetector based
on thermally reduced graphene oxide–ZnO (rGO–ZnO) hybrid nanostructure. J Mater Chem
22(6):2589–2595. https://doi.org/10.1039/C1JM13920G
49. Kumar H, Kumar Y, Rawat G, Kumar C, Mukherjee B, Pal BN et al (2017) Colloidal CdSe
quantum dots and PQT-12-based low-temperature self-powered hybrid photodetector. IEEE
Photon Technol Lett 29(20):1715–1718. https://doi.org/10.1109/LPT.2017.2746664
50. Guo Z, Park S, Yoon J, Shin I (2014) Recent progress in the development of near-infrared
fluorescent probes for bioimaging applications. Chem Soc Rev 43(1):16–29. https://doi.
org/10.1039/C3CS60271K
21 Self-Powered Photodetector 515

51. Jansen-van Vuuren RD, Pivrikas A, Pandey AK, Burn PL (2013) Colour selective organic pho-
todetectors utilizing ketocyanine-cored dendrimers. J Mater Chem C 1(22):3532. https://doi.
org/10.1039/c3tc30472h
52. Chen H, Sun X, Yao D, Xie X, Ling FCC, Su S (2019) Back-to-back asymmetric Schottky-­
type self-powered UV photodetector based on ternary alloy MgZnO. J Phys Appl Phys
52(50):505112. https://doi.org/10.1088/1361-­6463/ab452e
53. Ni P-N, Shan C-X, Wang S-P, Li B-H, Zhang Z-Z, Zhao D-X et al (2012) Enhanced responsiv-
ity of highly Spectrum-selective ultraviolet photodetectors. J Phys Chem C 116(1):1350–1353.
https://doi.org/10.1021/jp210994t
54. Shen H, Shan CX, Li BH, Xuan B, Shen DZ (2013) Reliable self-powered highly spectrum-­
selective ZnO ultraviolet photodetectors. Appl Phys Lett 103(23):232112. https://doi.
org/10.1063/1.4839495
55. Li K, Lu Y, Yang X, Fu L, He J, Lin X et al (2021) Filter-free self-power CdSe/Sb2(S1−x,Sex)3
near infrared narrowband detection and imaging. InfoMat 3(10):1145–1153. https://doi.
org/10.1002/inf2.12237
56. Wei C, Xu J, Shi S, Bu Y, Cao R, Chen J et al (2020) The improved photoresponse properties of
self-powered NiO/ZnO heterojunction arrays UV photodetectors with designed tunable Fermi
level of ZnO. J Colloid Interface Sci 577:279–289. https://doi.org/10.1016/j.jcis.2020.05.077
57. Kumar M, Park J-Y, Seo H (2021) High-performance and self-powered alternating cur-
rent ultraviolet photodetector for digital communication. ACS Appl Mater Interfaces
13(10):12241–12249. https://doi.org/10.1021/acsami.1c00698
Index

A Broadband photodetectors, 243, 505, 507


Auger Buffer layers, 219, 225
cooling, 92, 93
mechanism, 37, 39
Peltier element, 93 C
process, 35, 37, 39 Cadmium Selenide (CdSe), 231–247
recombination, 37 Cadmium sulfide (CdS), 210, 212, 214, 216,
Split-Stirling cycle, 93 217, 219–224
suppressed structure, 38, 39 Cadmium Telluride, 205, 210
Auger generation-recombination, 147 Carrier dynamics, 234–237
Auger recombination, 185, 193 Carrier lifetimes, 234, 235
Autodoping, 63, 64 Carrier multiplication (CM), 186
CdS film deposition
atomic layer deposition, 259, 262, 263
B chemical vapour deposition, 258,
Back contacts, 214, 215, 217, 222, 224–225 259, 262–265
Background limited infrared photodetector pulsed laser deposition, 265
(BLIP) limit, 29, 45, 48, 49 rf-sputtering, 259, 266
Back-to-back photodiode, 456 spin coating, 259, 264, 265
Bandgap, 183–185, 187–191, 194, 281, 282, thermal evaporation, 256, 259, 262, 265
288–292, 295, 454, 455 CdS nanostructures
Bandgap engineering, 62, 67, 71, 341 hydrothermal synthesis, 264
Band structure modifications, 340 seed growth, 264
Bandwidth, 54, 58, 60, 70 CdS photodetector
Barrier IR detectors figures of merit, 266–274
advantages, 114 flexible, 259, 265
nBn structure, 113, 116 multispectral, 252
nBp structure, 115 self-powered, 264, 266
unipolar barrier, 113–116 Chalcogenide thin film, 337, 340
Barrier structures, 135–149 Charge carriers, 53, 54, 58, 61
Bilayer graphene (BLG), 186, 188–192, 194 Chemical vapor deposition (CVD), 186, 189
Bispectral, 461, 464 Chlorine treatment, 222, 223

© The Editor(s) (if applicable) and The Author(s), under exclusive license to 517
Springer Nature Switzerland AG 2023
G. Korotcenkov (ed.), Handbook of II-VI Semiconductor-Based Sensors
and Radiation Detectors, https://doi.org/10.1007/978-3-031-20510-1
518 Index

Close-spaced sublimation (CSS), 214, 215, Excess noise, 53–55, 58, 60, 67, 71
219, 221 External quantum efficiency (EQE), 236, 237,
Close-spaced vapor transport (CSVT), 215, 241–244, 506
219, 221
Colloidal particles
core-shell, 435, 436 F
quantum dots, 448, 449 Fabrication, 55, 57, 62, 63, 67, 71
synthesis, 429, 447 Field of view (FOV), 29, 49
wet chemical methods, 428, 437, 441 Flexible photodetector
Color vision, 453 image sensor, 471, 481, 483–485, 488, 489
Compound semiconductors, 53–71 monitoring sensor, 471, 481, 488, 489
Cooling self-powered, 469, 471, 476, 481, 486,
Peltier element, 93 487, 489
Split-Stirling cycle, 93 wearable electronics, 469, 470,
Core/shell, 282, 283, 286, 291–293 483, 486–489
Cross talk, 460, 462 Fluorescence Emission, 282, 283
Current-voltage characteristic (CVC), 142 Focal plane array (FPA), 55, 56, 58, 62, 64,
66, 67, 71, 168, 173–174, 176,
456–459, 461–463
D cooled, 85, 94, 95
Dark current, 53–55, 59, 61, 67–71, 135–138, HgCdTe, 82–96
142, 143, 147, 148, 185, 186, 191, HOT, 101, 102
194, 195 monolithic, 99, 100
Defects, 63, 66, 71 multi-spectral, 100
Detection, 156, 173, 176 photovoltaic, 86–96
Detectivity, 26, 28, 29, 37, 39, 41–50, 140, pixel, 80, 85, 90, 91, 94–96, 98, 99, 101, 102
141, 148, 149, 236, 241–244, 501, two-color, 100
502, 505, 510 FPA technology
Detectors, 183–195, 454–466 HDVIP architecture, 98
barriers, 32, 36, 38, 39 heterostructures, 82, 87
dember, 40, 43 mesa structure, 96
high operating temperature (HOT), 24, molecular beam epitaxy (MBE), 86
35, 37–39 Si-based substrate technology, 87, 88
infrared (IR), 23, 24, 29, 38, 43 Front contact, 217–219
magnetoconcentration, 42 Full-color detection, 246
photoconductive (PC), 23–49
photoelectromagnetic (PEM), 40–42
photon, 23–27, 37, 39, 43 G
photovoltaic (PV), 23–49 Gain-bandwidth product, 54, 58, 60, 70
Diffusion, 62, 64, 67, 69, 70 Grain boundaries passivation, 207, 221, 225
Doped-ZnS Nanostructures, 334 Graphene, 183–195
Doping, 235, 505, 506 Growth mechanism, 341
Drift, 53, 61
Dual-band detector, 454–464
Dual-junction, 189, 190 H
Dual waveband, 463, 464 Heterojunctions, 183–195, 239–244, 503–511
Heterostructures, 55, 62, 185, 186, 195,
456, 459
E HgCdTe, 55–58, 60–68, 70, 71, 136, 138–149,
Electric field, 186, 189–194 183–195, 454–466
Electron transport layer, 236, 502 High operating temperature (HOT), 136,
Electro-optical modulator, 208–211 138, 147–149
Energy band diagrams, 141, 146 Hole transport layer (HTL), 414, 415,
Epitaxial layers, 62 443, 476
Index 519

Homojunction, 189, 190 Low dimensional structures


Homostructures, 186 0D quantum dots, 470, 475, 488
HOT detectors 1D nanowires, 470
non-equilibrium mode of operation, 108 2D nanosheets, 470
HOT device, 136, 138, 148 2D non-layered nanostructures, 470, 488
Hybrid, 234, 242–244, 497, 504, 507–511
Hybrid PDs
0D-2D hybrid structures, 417–419 M
Graphene, 417, 418 Mechanism of photosensitivity
MXene, 403 band-bending, 372
Polymer, 413, 414 Debye length, 372
QDs-polymer composites, 414–417 electron-holes pairs, 390
Hybridization oxygen molecules, 372
flip-chip bonding, 90 phototransistive gain, 392
indium bump technique, 90, 99 surface trap states, 372
loophole technique, 84 Medium-wave infrared (MWIR), 454, 457,
458, 460, 461
Mercury cadmium telluride (MCT), 136, 138,
I 144, 149, 454, 459, 461, 464–466
Impact ionization, 53, 54, 58, 60, 62, 68, 71 Mercury chalcogenides
Impurities, 239, 284, 286, 290 HgS, 156, 174
Infrared (IR), 135–149, 155–176, 183–195, HgSe, 156, 174
453–455, 459, 461, 464, 466 HgTe, 155, 160, 174
Interband transitions, 174 Metal-organic chemical vapor deposition
Intraband transitions, 174, 176 (MOCVD), 69, 138, 144, 145
Ion beam irradiation, 258, 274, 275 Metal-organic vapor phase epitaxy (MOVPE),
IR detectors, 183–195 62–64, 456–459, 461–463, 466
achievement, 6 Metal-semiconductor-metal (MSM)
applications, 4, 8, 11, 15 detectors, 334
comparison, 12 Mid-wave IR (MWIR), 55, 56, 58, 66, 67,
limits to detector performance, 19 69–71, 135, 138–140, 142, 143,
parameters, 5, 10, 18, 19 148, 149
IR radiation Molecular beam epitaxy (MBE), 55, 56, 62,
FIR, 3 63, 66, 138, 185, 187, 456,
LWIR, 20 458–463, 466
MWIR, 20 Multiband IR, 453
NIR, 20 Multicolor detectors, 453–466
Plank’s law, 3 Multicolor IR detectors
SWIR, 20 bias-selectable detector, 121
transparency windows, 20 MEMS optical filters, 121, 122
VLWIR, 3, 14 two-color pixels, 120
William Herschel, 3 Multi-stage/cascade IR detectors
IR Windows, 207, 208 advantages, 120

L N
Lattice-matched growth, 62, 63 Nanobelts, 239, 242, 497, 498
Ligands, 156–159, 168, 171, 176 Nanocrystals, 155–176
Light matter coupling, 170, 176 Nanomaterials, 186
Lights, 281, 282, 284, 286–289, 291, 292, Nanoplatelets, 241, 244
294, 295 Nanoribbons (NRs), 232, 237, 497
Liquid phase epitaxy (LPE), 62, 63 Nanorods, 499, 500, 507
Long-wave infrared (LWIR), 55, 58, 63, 66, Nanosheets, 285, 287, 288
67, 69, 71, 135, 138, 142, 143, 148, Nanotechnology, 211, 212
149, 454, 457, 458, 460, 461 Nanotubes (NTs), 235, 236, 240
520 Index

Nanowire-based photodetectors P
2D materials, 125, 126 Passivation, 35–37
graphene, 125, 128 Patterning, 258, 266, 268
transition metal dichalcogenides, Photoconductor (PC), 162, 164, 234, 235,
125, 129 237–240, 245, 246, 334, 335, 498
Nanowires, 285, 286, 291, 292 Photocurrent, 183, 189–191, 195, 284,
Narrowband photodetectors, 231, 511 286–288, 291, 292
nBn, 136–149 Photodetection performance, 345
NBνN, 147–149 Photodetectors, 56, 57, 184, 187, 191,
NIR detector, 231–247, 511 231–247, 281–296,
Noise 333–345, 495–512
equivalent power (NEP), 28 band gap engineering, 352
1/f, 31, 32, 47–49 buffer layers, 361
generation–recombination, 31, 34 core-shell nanobelt-based
Johnson–Nyquist, 31, 32, 43 photodetectors, 355
Noise equivalent differential temperature detectors for visible range, 353–355
(NEDT), 457–459, 462, 466 injection photodetectors, 354
NW-based heterostructures lattice matching, 359–361, 364–365
core-shells, 382, 384 MSM photodetectors, 354, 357
crossed, 385 nanowire-based detector, 355
NW-substrate, 385–387 photonic
1D axial, 383, 385 avalanche photodiodes (APD), 88
NW-based photodetectors dual-band detector, 89, 90
contacts, 393 QDIP, 82, 83
direct integration, 375 QWIP, 82, 94, 95, 99
fabrication, 373, 374 photoresponse rate, 356
individual, 373 Schottky barrier detectors, 357, 358,
in-plane geometry, 373 362, 364
NW array, 372, 373, 379, 388, 392 selectivity, 354, 355
NW bridges, 374 semiconductor materials
performance, 373, 376, 388–390 CdZnTe (CZT), 86, 87, 91
single NW, 375–381 GaAs/AlGaAs, 82
tunable, 379 HgCdTe, 82–92, 94, 95, 98, 102
vertical architecture, 373 HgTe/CdT, 99
InAs/GaSb, 82
thermal
O microbolometric, 81
1D nanostructures pyroelectric, 81
doping, 373 UV detectors, 361
hierarchical, 371 Photodetectors parameters
nanobelts, 371 decay time, 380
nanoribbons, 371 detectivity, 380
nanotubes, 371 external quantum efficiency, 380, 390
nanowires, 371 gain, 380, 393
synthesis, 393 light intensity, 378
Optical absorption, 62 photoresponse, 380
Optical bandgap energy, 337 response time, 380, 390
Optical immersion technique, 48 responsivity, 380, 390
Optical properties, 282–283 sensitivity, 378
Optoelectrical properties, 334, 335, 337 spectral photoresponse, 377, 389
Optoelectronic Devices, 283, 284, 286, 290, Photodetectors types
291, 296 heterostructures, 378
Over-time stability, 215 photoconductive, 375–378, 380
Index 521

phototransistors, 380, 381 CdSe/ZnS/CdS, 406


Schottky barrier, 390, 391 growth, 404, 407
Photodiode (PD), 136–138, 143, 148, 149, type I, 401, 403–407
166–168, 171, 175, 176, 455, 457, type II, 401, 403–407
496, 498, 502, 507 types of core/shell QDs, 404, 405
Photodiode QD-based ZnO/ZnS, 407
CdS, 402, 409, 410, 412 ZnTe/ZnSe, 406
Fabrication, 401, 403, 410 Quantum confinement effect, 282, 286, 295
flexible PDs, 400, 410, 411 Quantum-dimension structures
heterostructure-based, 410–412 quantum dots (QDs), 82, 99, 100
photoresponsivity, 407 quantum wells, 81, 82
ZnS, 401–403 Quantum dot IR photodetectors
ZnSe, 402 advantages, 118
Photolithography, 260–262, 264 Quantum dots, 236, 243, 497, 502, 511
Photon counting, 55, 61, 71 applications, 400, 409, 419
Photon detectors doping, 409
advantages, 9 properties, 409
focal plane array (FPA), 12 synthesis, 401, 402, 407, 409, 414
limitations, 12 Quantum efficiency, 23, 28, 31, 36, 37, 39, 40,
photoconductors, 8, 9 58, 67, 185, 188, 191
photodiode, 9, 10 Quantum well IR photodetectors
photoresistor (PR), 8, 9 advantages, 110
phototransistor, 10 GaAs/AlGaAs structures, 110
Schottky barrier, 10 limitations, 110
Photon trapping detectors
immersion lenses, 123, 124
mesa-structures, 124 R
Photoresponse, 284, 288, 291 Readout integrated circuits (ROICs), 187,
Photosensitive device, 135, 146 188, 193
Photosensitive materials charge-coupled devices (CCD), 85
GaSb, 55 complementary metal-oxide-
Ge:Au, 9 semiconductor (CMOS), 85, 100
HgCdTe, 14 multiplexer, 83, 85, 90, 99
InAs, 14 Response time, 53, 55, 58, 60, 70, 71,
InGaAs, 14 186, 191
InSb, 14 Responsivity, 28, 31, 36, 40, 41, 43, 234–236,
PbS, 14 239, 241, 242, 244, 498–502, 505,
PbSnTe, 14 507, 510, 511
PtSi-Si, 10 RF sputtering, 219, 220
Si:As, 14, 16
Photosensitivity, 191
Phototransistor (PT), 162–166, 171 S
Photovoltaic efficiency, 206, 216 Schottky junctions, 342, 497, 498, 500–502,
Photovoltaic modules, 216 510, 511
Piezoelectric, 271, 275 Schottky photodiodes, 498, 499
Pixel pitches, 457, 458, 462 Self-powered photodetector, 495–511
Pyroelectric, 251, 266–268, 275 Self-powered UV detectors, 341, 342, 345
Semiconductors, 281–283, 286, 288–292, 295,
454, 459
Q CdS, 352, 355, 356, 358
QD core/shell structures CdTe, 350, 351, 353, 356, 363, 364
band gap alignment, 404 GaAs, 361
CdSe/CdS, 406 GaN, 359
CdSe/ZnS, 406 GaP, 357, 361, 362
522 Index

Semiconductors (cont.) successive ionic layer adsorption and


InSb, 361 reaction (SILAR), 435
MnTe, 351 successive ionic layer deposition
Si, 350, 360–363 (SILD), 429
ZnS, 349, 352, 356–358 tape casting, 431
ZnSe, 361, 364 Spatial resolution, 62
ZnTe, 351, 363 Spectral resolution, 62
ZnTe:Cu, 364 Substrates, 62–67, 71
Sensitivity, 55–58, 61, 62 Superlattices, 145–147, 149
Sensor, 428 nBn barrier structure, 102
Sequential-mode detector, 455 type-1 (T1SL), 82
Shockley–Read–Hall (SRH) process, 37 type-II (T2SL), 82
Short-wave infrared (SWIR), 55, 58, 71, type-III, 82, 99
454, 456 Surface leakage currents, 142, 143
Signal-to-noise ratio (SNR), 53, 57, 67 Surface plasmon, 232
Silvaco TCAD, 194
Simultaneous detection, 454, 455
Solar cells, 205–226, 292–295 T
bottom cell, 352 Terahertz, 281, 288–290, 295
optimization, 352 Thermal detector
solution processing, 352 advantages, 6
tandem solar sell, 351, 352 bolometers, 5, 7, 20
top cells, 351 ferroelectric, 6
wide bandgap window, 352 limitations, 7
Solution-processed technology MEMS technology, 6
casting, 430 microbolometers, 6, 7
chemical bath deposition (CBD), 429, 432, pyroelectrics, 5, 6, 20
433, 435, 440 thermistors, 5, 20
combined approach, 439–443 thermopiles, 5, 7, 20
dip coating, 429, 444, 446 Thermal imaging camera
inkjet printing optical-mechanical system, 80
application, 448 scanner, 80
conductive inks, 447 scanning, 79, 80
inkjet ink composition, 448 SWaP, 101
photodetectors thermograph, 79
detectivity, 437 Thin Films, 206–208, 211, 212, 214–219, 222,
flexible, 428 223, 225, 284–285, 294
IR, 442 Third generation IR devices, 453
multicolor, 446, 447 Three-color detector, 454, 455, 466
semiconductor-polymer Transparent and contacting oxides (TCOs),
composites, 439 218, 219
sensitivity, 436, 443 Transient response, 499, 500, 502,
solar-blind deep ultraviolet, 437 503, 506–509
SWIR, 442 Transport, 156–160, 166–168, 172, 176
upconversion, 442, 443 Tunable Schottky barrier, 344
UV detectors, 435 Two-color detector, 456
roll-to-roll technique, 429, 444 II-VI semiconductor optimization
screen printing, 431 diluted magnetic semiconductors
slip casting, 430 (DMS), 363
solar cells, 428–431, 437, 440, doping, 363, 364
441, 444 spintronics, 363
spin coating, 429, 430, 437–439, 441, II-VI semiconductor-based heterostructures
443, 444 CdS/CdSSe, 390
spray coating, 430 CdS/CdS:SnS2, 389, 390
Index 523

CdSe/ZnTe, 383 ZnSSe, 357


CdS/Si, 390 ZnSTe, 357, 358, 361
CdS/ZnTe, 383, 390 Type-II strained-layer superlattice
Ge/CdS, 383 advantages, 127
ZnO/ZnS, 388 HgCdTe, 111
ZnO/ZnSe, 383, 390 InAs/GaSb, 111, 112
ZnS/CdS, 389 limitations, 128
ZnSe:Sb/ZnO, 390
ZnSe/Si, 387
ZnSe/ZnO, 388 U
ZnS/InP, 383, 390 Ultraviolet, 281, 282, 292, 295
ZnS0.49Se0.51/ZnSe, 384, 390 Unipolar barrier, 135–149
ZnS/Si, 390 Unipolar structures, 145–149
II-VI semiconductors, 206, 207 UV detectors, 231, 232, 239, 241, 244,
CdS, 375–378, 380, 381, 383, 384, 245, 501
388–391, 432–434, 439, 442–445, UV-vis photodetector, 211–213
447, 470, 473–475, 477–481, 483,
486, 488
CdSe, 375–378, 380, 382, 383, 388, 389, V
391, 470, 475, 477, 479, 480, Vacuum Evaporation, 284, 289, 290
483, 488 Valence band offset, 140, 144
CdSxSe1-x, 470, 477, 479, 480, 488 Visible, 281, 282, 286, 288, 290, 291, 295
CdTe, 375, 376, 378, 383, 388, 390, 434, Visible detector, 231–247, 505, 508
437, 438, 440, 441, 449
HgTe, 446
wide band compounds, 429 W
ZnO, 435, 436, 443, 446, 447 Wide bandgap nanostructures, 334
ZnS, 375, 377, 380, 383, 388–391, 427, Width of Urbach Tail, 342
432, 435–437, 442–445, 470,
473–478, 483, 488
ZnSe, 375, 383–388, 390, 435, 440, 441 Z
ZnTe, 375, 382, 383, 388–390, 432 ZnS, 333–345
II-VI ternary alloys ZnSe photodetectors
CdSxSe1-x, 378 advantages, 301, 324
ZnSxSe1-x, 380 heterostructure-based photodetectors
ZnxCd1-xSe, 380 advantage, 320
II-VI-based multicomponent alloys superlattice, 310, 322
Cd1-xCrxTe, 364 UV-transparent conducting
Cd1-xMnxTe, 351, 352, 364 polymer, 328
Cd1-x-yHgxMnyTe, 364 wide-gap window, 320
Cd1-xZnxTe (CZT), 351, 352, 360, 363 ZnSTeSe photodiodes, 322
CdS-CdSxTe1-x-CdTe, 355, 356 hybrid structures
CdSTe, 354, 355 organic–ZnSe hybrid structures, 326
CdZnS, 352 PEDOT:PSS, 328
Cu2ZnSnS4, 352 ZnSe-graphene structure, 326
CZT:Cu, 351 metal–semiconductor–metal
Hg1-xCdxTe, 352 photodiodes, 316
Zn1-xCrxTe, 364 advantages, 316
Zn1-xMgxS, 357 interdigital electrode, 316
Zn1-xMnxS, 356 ITO, 317, 319
ZnBeSe, 361 metals, 317
ZnMgBeSe, 358, 359, 361, 362 quantum efficiency, 316, 319
ZnMnSe, 364 nanowire-based photodetectors
ZnSeTe, 357 1D nanostructures, 324, 325
524 Index

ZnSe photodetectors (cont.) ZnSSe APDs, 310, 311


doped NWs, 325 p–n junction photodiodes
limitations, 325 diffusion region, 309
nanowires, 324, 325 diffusion tail, 312
performances, 313, 317 disadvantages, 308
responsivity, 324, 325 GaAs substrates, 306–308
photoconductive detector MBE, 307, 308
decay of photocurrent, 306 responsivity, 306
persistent photoconductivity, 303 structure, 309–312
structure, 302 Schottky photodiodes
weakness, 303 advantages, 312
phototransistors hydrogen peroxide, 314
graphene, 322, 323 ideality factor, 314
monolithic, 322 performances, 313
p–i-n junction photodiodes responsivity, 312–314
advantages, 312 Schottky barrier, 313–314
implementation, 310 ZnMgBeSe PD, 313, 314
responsivity, 308, 309 ZnTe, 282–296
superlattice buffer layer, 310

You might also like

pFad - Phonifier reborn

Pfad - The Proxy pFad of © 2024 Garber Painting. All rights reserved.

Note: This service is not intended for secure transactions such as banking, social media, email, or purchasing. Use at your own risk. We assume no liability whatsoever for broken pages.


Alternative Proxies:

Alternative Proxy

pFad Proxy

pFad v3 Proxy

pFad v4 Proxy