0% found this document useful (0 votes)
79 views211 pages

FaiqPaperIF6 2

The review article discusses the emerging trends in Metal-Organic Frameworks (MOFs) as photocatalysts for hydrogen energy production through water splitting. It highlights the advancements in MOFs, their structural characteristics, and the challenges of improving light absorption and stability. The article also covers synthetic techniques, performance optimization strategies, and future directions for research in this field.

Uploaded by

Zohaib Ahmad
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)
79 views211 pages

FaiqPaperIF6 2

The review article discusses the emerging trends in Metal-Organic Frameworks (MOFs) as photocatalysts for hydrogen energy production through water splitting. It highlights the advancements in MOFs, their structural characteristics, and the challenges of improving light absorption and stability. The article also covers synthetic techniques, performance optimization strategies, and future directions for research in this field.

Uploaded by

Zohaib Ahmad
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/ 211

See discussions, stats, and author profiles for this publication at: https://www.researchgate.

net/publication/375290581

Emerging Trends in Metal-Organic Framework (MOFs) Photocatalysts for


Hydrogen Energy Using Water Splitting: A State-of-the-Art Review

Article · November 2023


DOI: 10.1016/j.jiec.2023.10.055

CITATIONS READS

15 142

11 authors, including:

Samia Arain Faiq Saeed


Tianjin University Tianjin University
18 PUBLICATIONS 71 CITATIONS 14 PUBLICATIONS 121 CITATIONS

SEE PROFILE SEE PROFILE

Aneela Ahmad Fu Yikai


Tianjin University Tianjin University
4 PUBLICATIONS 24 CITATIONS 17 PUBLICATIONS 28 CITATIONS

SEE PROFILE SEE PROFILE

All content following this page was uploaded by Faiq Saeed on 10 May 2024.

The user has requested enhancement of the downloaded file.


Journal Pre-proofs

Review Article

Emerging Trends in Metal-Organic Framework (MOFs) Photocatalysts for


Hydrogen Energy Using Water Splitting: A State-of-the-Art Review

Samia, Faiq saeed, Li Jia, Musfira Arain, Aneela Ahmed, Fu Yikai, Chen
Zhenda, Ijaz Hussain, Ghulam Abbas Ashraf, Samia Ben Ahmed, Haitao Dai

PII: S1226-086X(23)00699-8
DOI: https://doi.org/10.1016/j.jiec.2023.10.055
Reference: JIEC 7039

To appear in: Journal of Industrial and Engineering Chemistry

Received Date: 11 July 2023


Revised Date: 22 October 2023
Accepted Date: 30 October 2023

Please cite this article as: Samia, F. saeed, L. Jia, M. Arain, A. Ahmed, F. Yikai, C. Zhenda, I. Hussain, G. Abbas
Ashraf, S. Ben Ahmed, H. Dai, Emerging Trends in Metal-Organic Framework (MOFs) Photocatalysts for
Hydrogen Energy Using Water Splitting: A State-of-the-Art Review, Journal of Industrial and Engineering
Chemistry (2023), doi: https://doi.org/10.1016/j.jiec.2023.10.055

This is a PDF file of an article that has undergone enhancements after acceptance, such as the addition of a cover
page and metadata, and formatting for readability, but it is not yet the definitive version of record. This version
will undergo additional copyediting, typesetting and review before it is published in its final form, but we are
providing this version to give early visibility of the article. Please note that, during the production process, errors
may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

The Korean Society of Industrial and Engineering Chemistry


Emerging Trends in Metal-Organic Framework (MOFs) Photocatalysts for Hydrogen
Energy Using Water Splitting: A State-of-the-Art Review

Samia a, Faiq saeed b, Li Jia c, Musfira Arain d, Aneela Ahmed e, Fu Yikai f,Chen Zhenda g, Ijaz
Hussain h, Ghulam Abbas Ashraf i, Samia Ben Ahmed j, Haitao Dai a*

a Tianjin Key Laboratory of Low Dimensional Materials Physics and Preparing Technology, School of Science, Tianjin University, Tianjin
300072, China

b School of Precision Instrument and Optoelectronics Engineering, Tianjin University, Tianjin 300072,

c School of Precision Instrument and Optoelectronics Engineering, Tianjin University, Tianjin 300072, China

d School of Environmental Science and Engineering Tianjin University, Tianjin 300072, China

e Tianjin Key Laboratory of Low Dimensional Materials Physics and Preparing Technology, School of Science, Tianjin University, Tianjin
300072, China

f Tianjin Key Laboratory of Low Dimensional Materials Physics and Preparing Technology, School of Science, Tianjin University, Tianjin
300072, China

g Tianjin Key Laboratory of Low Dimensional Materials Physics and Preparing Technology, School of Science, Tianjin University, Tianjin
300072, China

h Interdisciplinary Research Center in Refining and Advanced Chemicals-Research Institute, King Fahd University of Petroleum and
Minerals, Saudi Arabia

i New Uzbekistan University, Mustaqillik Ave. 54, Tashkent, 100007, Uzbekistan

j Department of Chemistry, College of Sciences, King Khalid University, Abha, 9004, Saudi Arabia

a* Tianjin Key Laboratory of Low Dimensional Materials Physics and Preparing Technology, School of Science, Tianjin University, Tianjin
300072, China

Graphical abstract

1
H2
H2
H2 H2 Fuel cells

Vehicle gas
H2O
H2

Biomass

H2 clean
energy carrier
Power
generation

Electricity H2
production
Industrial
feedstock

* Corresponding author: Haitao Dai

Tel:+8615620398753; E-mail addresses: htdai@tju.edu.cn (H. Dai)

Abstract

Various photocatalysts have been developed for photocatalytic water splitting—one of the most
important processes that produces dihydrogen as clean energy for fuel cells. The successful
achievements for this application are based mainly on transition metal oxides and some metal
sulfides/nitrides. Recently, metal–organic frameworks (MOFs), a class of hybrid functional
materials comprising organic backbone tethered infinitively in limitless way by metal-oxide
clusters, both of which can be customized accurately at the molecular level for targeted

2
applications, have been able to photocatalytically degrade water. Apart from representing an
array of intrinsic structural and physicochemical characteristics, MOFs are well susceptible for
various post-synthetic modifications to address specific challenges. Despite years of research
in this field and a good number of seminal studies, further efforts should be geared toward the
improvement of light absorption and stability of MOFs, which are the principal challenges that
should be overcome. This review includes the most recent research that has been committed to
MOFs materials for photocatalytic water-splitting applications. It also encompasses a variety
of synthetic techniques and post-modifications that have been employed to increase the
performance of MOFs. In addition, a brief discussion of the techno-feasibility analysis of water
splitting has been offered, extending the conversation to include current challenges and future
direction recommendations. The recent advancements of using MOF photocatalysts for water
splitting are further described in a way that benchmark achievements and limitations are
considered so that the readers can imagine the big picture in this field and pay considerable
attention to future solutions.
Keywords: Water splitting, Photocatalyst, Metal-Organic Frameworks, MOFs; Green
Hydrogen Energy.

Table of contents

Abstract
1. Introduction
1.1 Renewable hydrogen energy
1.2 A brief background of photocatalytic water splitting
1.3 Photocatalytic overall water splitting
2. Development of a catalytic Metal organic framework (MOFs) for photocatalytic
water splitting

2.1 Transition metals-based MOFs as photocatalyst

2.2 Metal organic framework composites as photocatalyst


2.3 MOFs-derived as photocatalysts

3. Factors affecting the photocatalytic activity

3.1 Surface area


3.2 Co catalyst
3.3 Heterostructure
3.4 Organic linker/functional group/ ligands
3.5 Synthesis methods
3.6 Process conditions
4. Strategies to optimize the photocatalytic water splitting
4.1 Ligand functionalization
4.2 Metal doping
4.3 Defect regulation
4.4 Morphology regulation
4.5 Surface and pore decoration

5. Techno-socioeconomic and feasible analysis


6. Current challenges and future perspectives
7. Summary
8. Reference

3
Abbreviations
Abbreviations Full form
MOFs Metal organic frameworks
HOMO Highest occupied molecular orbital
LUMO Least unoccupied molecular orbital.
Ru2(p-BDC)2] n Ru based MOF where p-BDC is 1,4-benzene
dicarboxylate
Bpy 2,20-dipyridyl
Ru2(MTCPP)BF4 M = H2 (PCL-1), Zn (PCL-2);
TCPP = Tetrakis(4-carboxyphenyl)porphyrin
TBP Tributyl phosphate
TON turnover number
ICP-MS Inductively coupled plasma mass spectrometry
CuI24(µ3-Cl)8(µ4-Cl)6- based (ZZULI-1)
POM@- MOF
POM Polyoxometalates
BODIPY Fluorinated Boron-Dipyrromethene Dyes
(DmgH)2 Dimethylglyoxime (neutral form)
TEOA Triethanolamine
Dcbpy 2,2′-bipyridyl-5,5′-dicarboxylic acid
Bpdc 4,4′-biphenyldicarboxylic acid
BDC 1,4-benzenedicarboxylate
TBAPy tetracarboxylate ligand
RhB Rhodamine B
MAA mercaptoacetic acid
PBA Phenylboronic acid
CCM 1-Methyl-1-Carboxy-Cyclopentane
UNH Unnilhexium
COF Covalent organic framework
CCNU 1-(2-chloroethyl)-3-cyclohexyl-1-nitrosourea (i.e.
lomustine)
H2SDB 4,40-sulfonyldibenzoic acid
LMCT ligand-to-metal charge transfer
HNTM zirconium-porphyrinic hollow nanotubes MOFs
H3BTC benzene-1,3,5-tricarboxylic acid
4-bpdh 2,5-bis(4- pyridyl)-3,4-diaza-2,4-hexadiene

1. Introduction

4
Over the past ten years, the average yearly growth in primary energy consumption has
been 1.7%, reaching 17.9 TW (terawatts) in 2017 [1]. The tremendous increase in population,
economic advances, and new technology development are responsible for the continuously
increasing request for energy worldwide. The most common types of fossil fuels, which include
coal, petroleum oil, and natural gas, are the principal types of energy sources that are utilized
all over the world [2], [3], [4]. With the continuous use of these fossils, we are not only
eliminating them but also providing severe damage to our Earth planet. These fossil fuels emit
highly toxic and destructive gases (NOx, COx, SOx) into the environment, causing problems
such as greenhouse gas emissions, global warming, water pollution, and land degradation.
Global energy consumption rate has significantly outpaced the rate of energy storage; hence it
seems likely that fossil fuel reserves will run out shortly. The heavy reliance on fossil fuels has
created worries about modern energy generation's long-term viability and environmental
impacts. As a result, it is pretty vital to search for a sustainable energy source that can save
money and is kind to the Earth [5], [6], [7]. Renewable energies are considered sustainable
alternatives to fossil fuels because they have significant advantages over fossil fuels as clean
energy sources. Both hydrogen and electricity can be used as storage mediums for renewable
energies [8]. Near-infrared, visible, and ultraviolet sunlight radiates a lot of energy and
intensity to the Earth; therefore, capturing it would help with our electrical and chemical
demands. Since the 1970s energy crisis, much research has gone into developing efficient
devices to absorb and transform solar light into chemical energy. Under sunlight, the

photocatalytic splitting of water into H2 and O2 is one of the most promising of these "artificial
photosynthesis" reactions [9], [10], [11].

1
𝐻2 𝑂 → 𝐻2 + 2 𝑂2 ∆𝐺 = 237 𝑘𝐽/𝑚𝑜𝑙 (1)

Fujishima and Honda made the announcement almost exactly half a century ago, and it
was based on the reaction of semiconductor material in water splitting under the irradiation of
light. TiO2 nanoparticles were used as photocatalyst for energy conversion, storing, and
cleansing the environment [12]. Influenced by their work on precursors in 1972, explosive
growth in photocatalysts has been reported in recent years for the generation of highly efficient,
stable, and low-cost H2 production systems to meet the target of large-scale solar to-fuel
conversion [13], [14], [15], [16], [17], [18]. Hydrogen as a fuel produces enormous amount of
energy, Hydrogen promises to preserve the quality of environment with almost no pollution
when an engine burns pure hydrogen. Liquid hydrogen has its use since the 1970s when NASA
(National Aeronautics and Space Administration) used Hydrogen fuel for propelling the space
shuttles and maneuver probes into space. Hydrogen has been used to provide the power to the
electrical systems of the spacecrafts, producing H2O as a cleaner by-product after the reaction
of H2 with oxygen. The generated byproduct (water) was utilized by crew members in space
for drinking purposes. Presently, hydrogen as a fuel is in its infant stage in terms of scalability
but researchers are working toward clean, safe, economical and environment friendly hydrogen
production and on its widespread distribution. Hydrogen is quite abundant in the environment
but it exists in the form of heterogeneous compounds like Water, Methane and other Organic
matter [19],[1], [20], [21], [22]. Although plants use an integrated molecular system to power

5
photosynthesis, semiconductors are often used in artificial water photolysis systems instead of
dye molecules [23]. Due to narrow band gap and peculiar electronic structure, materials like
TiO2, Fe2O3, WO3, ZnO, CeO2, CdS, Fe2O3, ZnS, MoO3, ZrO2, and SnO2 are frequently used
as photocatalysts. In semiconductor photocatalysis, electrons from a semiconductor's valence
band are stimulated to the conduction band by higher-energy light that matches the band gap,
resulting in eCB / h+ VB pairs [5], [24], [25], [26], [27], [28]. At the beginning of the 21st
century, researchers devoted more efforts to visible-light-driven hydrogen generation.
Perovskites and its derivatives encompass the plethora of knowledge and potential to get
channelized in sustainable hydrogen energy production. Perovskite photocatalysts can replace
the conventional materials such as TiO2 or platinum that possesses intrinsic drawbacks of
limited scope of solar absorption and whopping cost respectfully. During photocatalytic water
splitting, Perovskite oxides exhibit favorable states of conduction band possessing high
potential in comparison to the reduction potential of H+ ions, therefore, delivering adequate
over-potential to evolve H2. [29]. Following that, photocatalytic materials, which include metal
oxides, metal sulfides, metal (oxy)nitrides, and metal-free polymeric polymers, are new
effective techniques that were continually appearing as a result of the researchers' sustained
efforts [30], [31], [32]. Hydrogen can be produced as a renewable energy source using
nanomaterials, which are developing as a new form of photocatalyst due to their larger surface
areas and more varied morphologies than bulk materials [33]. Because of this, a significant
focus has recently been directed to researching and developing various methods for creating
hydrogen, considered the most appropriate renewable energy resource [34]. Several
investigations on using MOFs as photocatalysts in hydrogen generation via water splitting have
been undertaken during the last few decades. The number of articles on photocatalytic MOFs
for water splitting has greatly increased recently, as shown by the Web of Science in Clarivate
Analytics (Figure 1), demonstrating the importance and significance of this topic.

2015 2016
1% 3% 2017
3%
2023 2018
9% 7%

2019
2022 13%
24%

2020
17%

2021
23%

Figure 1. The number of photocatalytic studies involving MOFs that were published on the
Web of Science on October 17, 2022, with the keyword "photocatalytic water splitting with
MOFs."

6
In this scenario, several review studies have been executed to cover different aspects of
photocatalytic water splitting to produce hydrogen [35], [36], [37], [38], [39], [40], [41], [42],
[43], [50], [55], [213]. Even though there have been numerous studies on photocatalytic water
splitting to produce hydrogen, there is still a need for a state-of-the-art review, and more
research is anticipated in this area. In this article, the dominant theme is the production of
hydrogen via photocatalytic splitting of water using MOFs and their reaction processes. First
we look at the brief history, chemical reactions and development of a catalytic Metal organic
framework (MOFs) for photocatalytic water splitting. Next recent advances in photocatalytic
water splitting using different transition metals, supporting materials, derivatives and
composites of MOFs are discussed in detail focusing on their optimal synergistic relationship
between surface area and porosity, co-catalyst, heterostructure, and organic linker/functional
group/ligands towards enhanced photocatalytic activity. We also provide synthesis methods of
MOFs and the process conditions to improve photocatalytic activity. We summarize the critical
steps needed to be improved and then provide a clear train of thought that will give some
guidance for constructing an efficient water-splitting system. On a more tangible level, the cost
and efficiency goals of solar-to-fuel conversion are assessed, together with a techno socio-
viable study and at the end conclusion are also provided.

1.1 Renewable Hydrogen Energy

On Earth, hydrogen does not occur naturally; however, it can be produced from the
different components (naturally and manmade); the universe's most prevalent element,
hydrogen, accounts for around 75% of all stuff [44]. The demand for hydrogen production has
increased due to its qualities, such as clean, safe, and highly combustible gas, and all of this is
being used as a fuel in various fields as a source of energy. Although there are several

7
Figure 2: Applications of hydrogen in different fields.

8
drawbacks to hydrogen, such as its low energy content per volume, high cost, the need for
expensive infrastructure, and high rates of greenhouse gas emissions during production, the
notion that hydrogen is the "fuel of the future" remains debatable [45], [46]. Solar hydrogen is
a clean alternative fuel capable of substituting fossil fuels and ceases the addiction to oil and
gaseous fuels, abating CO2 emissions to save the world from global warming. Due to the
following factors, hydrogen is a perfect energy carrier or storage medium [47], [48]. To begin
with, it is the most common element and is present in both water and biomass [49]. Second,
compared to other fuels like gasoline (40 kJ/g), it has a high energy yield (122 kJ/g); Third, it
is favorable for the environment because it won't produce any pollutants or greenhouse gases
while in use, and it will not have any other unfavorable effects on the environment. Eventually,
hydrogen can be kept as a gaseous, liquid, or metal hydride and transported over great distances
via pipelines or tankers [5], [50], [51], as depicted in Figure 2. Although various technology at
different scales has been developed for hydrogen as an energy carrier, it is widely used as a
chemical feedstock in the industry [44], [52]. Methods that are available for hydrogen
production are from fossil fuels, especially from methane, the main constituent of natural gas
and coal, from biomass, water, compounds like metal hydrides, H2S, and biological sources
[44], [53], [54], [55]. The steps involved in producing hydrogen are listed down in Figure 3.
This scheme recommends two main tracks for producing hydrogen, i.e., fossil fuels and
renewable sources. Fossil fuels are not sources that can be replenished, and their availability is
dwindling fast. So, the primary goal of this study is to produce hydrogen by conserving our
fossil fuels. In this regard hydrogen production from water splitting is the future and current
promising method to produce renewable energy [53], [56], [57]. This technology can
manufacture hydrogen without depleting fossil fuels, making it suitable for an eco-friendly
setting. Splitting water is a thermodynamically uphill reaction, as opposed to a natural process
that occurs on its own and requires no external energy to proceed. Figure 4 describes the paths.

9
Figure 3. Different hydrogen production routes Reprinted with permission from [53]. Copyright 2017 Elsevier.

10
Figure 4. Different water splitting mechanisms for producing hydrogen.

through which water can be split using thermal, electrical, photonic, and biochemical energy.
These technologies include thermolysis, electrolysis, photolysis (such as photocatalysis), and
biolysis (such as dark fermentation). Hydrogen can be generated easily by combining two or
more types of these hybrid energy methodologies [58]. In the process of thermochemical water
splitting, water is heated to a high temperature, which causes it to break up into hydrogen and
oxygen [59], [60], [61], [62], as illustrated in Eq. 1. The main advantage of using this
technology is that no greenhouse gases are discharged; nevertheless, the ingredients' toxicity
and the high-cost requirements are downsides [63], [64]. For this purpose, Farid Safari et al.
[60] recommended thermochemical cycles as long-term, large-scale, hygienic, and economic
hydrogen production systems. The thermochemical cycles are solely driven by thermal energy,
as illustrated in Figure 5. First, water decomposes into H2 and O2 due to poor thermodynamics
and the high temperature required for the single step. Thermochemical cycles are proposed as
a repeating sequence of reactions in which water is divided using thermal energy at
temperatures below 2000 °C and often in two or more stages.

11
Figure 5. Basic diagram of pure and combined thermochemical cycles for splitting water
Reprinted with permission from [60]. Copyright 2020 Elsevier.

Scientists from a wide variety of countries and institutions have found and researched a wide
variety of cycles. Only a few of them, however, have been practically integrated with a
renewable energy source for producing hydrogen. The other method is photo biological water
splitting [57], [58], [65]. In this method, enzymes are added to the water because they are
necessary for transforming water into hydrogen and are therefore used to purify the water. This
process has two subunits i.e., indirect and direct water splitting [56]. Hydrogenous enzymes
are utilized in the direct technique, and the advantage of this process is that it is sensitive to
oxygen, allowing the synthesis of oxygen to be controlled. Although it is costly, it can be
considered as a drawback. The other technique is indirect method which is cheap as it uses
cyanobacteria and green algae for water splitting. Hydrogen is created by hydrogenase and
nitrogenase enzymes [61]. Both indirect and direct photolysis has the same efficiency;
however, the indirect method is significantly less expensive than the direct approach [75].
Moreover, photo electrolysis is a method that involves exposing electrodes to external
radiations while using a heterogeneous photocatalyst on one or both electrodes [66], [67], [68].
This has been proposed as a viable solution to the renewable energy intermittency problem as
an electrical current is used to split water electrolytically, and the electrical energy is converted
to chemical energy at the electrode-solution junction via charge transfer processes in a device
known an electrolyze. At the anode, water interacts to create oxygen and protons, however at
the cathode, hydrogen is evolved [69], [70]. Besides, it is estimated that electrolysis can provide
3.9 % of the world's hydrogen consumption [71]. Using photochemical or photocatalytic water
splitting to produce hydrogen is a potential alternative for hydrogen generation that is aimed
toward reducing CO2-emissions and making use of sustainable resources like water and
sunshine [72]. The steps required to generate hydrogen while using this method are i) allowing
a photon with sufficient energy to generate an electron-hole pair, (ii) creating an electrical
current, (iii) Splitting water into hydrogen and oxygen ions, (iv) forming gaseous hydrogen at
the through proton reduction, and (v) separating, processing, and storing the resulting fumes.
The stability of the light electrode and inefficient hydrogen production are drawbacks of
employing the photo electrolysis technology [58]. Water splitting is generally the breakdown
of water in hydrogen and oxygen, as shown in Eq. 1. During this burning of water, a tremendous
amount of energy is released, enough to meet the demands of the world economy. The above
processes can be used for hydrogen production, but in this review, we have focused on

12
photocatalytic water splitting, its factors, and strategies for improving catalytic activity. All
these aspects will explain in subsequent sections.

1.2 A brief background of photocatalytic water splitting

Photocatalysis is a promising technique for producing and retaining solar energy [73],
[74], which will contribute to the achievement of the United Nations' Sustainable Development
Goals (SDGs). Photocatalytic water splitting is the most basic water splitting technique, with
the potential to enable low-cost, large-scale H2 production. The origins of photocatalytic water
splitting can be traced back to Boddy's work in 1968 [76], where several investigations on TiO2
under illumination tracking O2 evolution have been carried out. Following the pioneering
exploration of H2 development using photoelectrochemical water splitting on type TiO2
electrodes under ultraviolet (UV) irradiation in 1972, significant research resources were
focused on solar hydrogen synthesis [77], [78], [79], [80]. Figure 6 depicts the current
successful photocatalytic systems for overall water splitting. In this particular setup, the
photocatalyst needs to possess a thermodynamic potential that is adequate for the process of
water splitting, a band gap that is narrow enough to allow for the collection of visible photons,
and stability against photo corrosion [81], [82], [83], [84]. This also shows how photocatalytic
activity depends on the photocatalyst's physicochemical parameters, which are very important
to run the chemical reaction. A photocatalyst is a material that speeds up the reaction under the
influence of light (UV light). They function as materials that deteriorate harmful chemicals
when exposed to UV-rayed sunlight [85]. Mainly, TiO2 is used as a photocatalyst; at present,
photocatalysts have been widely applied in photocatalytic water splitting for hydrogen
production in the presence of UV light and visible light irradiation. Recently, plasmonic
semiconductor-based photocatalysts have attracted significant interest owing to their broad
light-harvesting capacities, which range from visible to near-infrared light. Furthermore, they
can harvest low-energy photons to generate extra high-energy “hot electrons,” allowing for
full-spectrum-driven water reduction reactions [86]. Among the reported plasmonic
semiconductors, nonstoichiometric molybdenum oxide with abundant oxygen vacancies
(MoO3−x) demonstrates a broad and intense localized surface plasmon resonance (LSPR)
absorption centered at 700 nm, with adsorption tailing to 2000 nm. Accordingly, MoO3−x has
become one of the most attractive candidates for harvesting visible to near-infrared light for
photocatalytic H2 generation. In this regard, an LSPR-enhanced 0D/2D
CdS/MoO3−x heterojunction composed of 0D CdS nanoparticles grown on 2D plasmonic
MoO3−x elliptical nanosheets has been successfully fabricated. The 0D/2D
CdS/MoO3−x heterojunction exhibits improved photocatalytic H2 generation activity under a
wide range of visible light irradiation, including using 420, 450, 550, and 650 nm monochromic
light, and the optimal CdS/MoO3−x demonstrates 10.3 times higher photocatalytic H2 evolution
(7.44 mmol·g−1·h−1) than that of pristine CdS under visible light irradiation (>420 nm) [87].
He et al [88] synthesized plasmonic TiN nanobelts, as a novel cocatalyst, were coupled with
CdS nanoparticles to construct a 0D/1D CdS/TiN heterojunction. Utilization of the localized
surface plasmon resonance (LSPR) effect generated from TiN nanobelts was effective in
promoting light absorption in the near-infrared region, accelerating charge separation, and
generating hot electrons, which can effectively improve the photocatalytic H2 generation
activity of the 0D/1D CdS/TiN heterojunction over a wide spectral range. Furthermore, owing
to the high metallicity and low work function, an ohmic-junction was formed between the CdS
and TiN, favoring the transfer of hot electrons generated from TiN nanobelts the CdS
nanoparticles, followed by the reaction with water to generate H2. Figure 7 proposed a list of
qualities that a promising photocatalyst should have, and it also depicts some known

13
photocatalysts that have been used in the production of hydrogen through the process of water
splitting so far.

Figure 6. Photocatalytic water splitting produces hydrogen by using different semiconductor


catalyst. The gear with the number indicates the order of the photocatalytic process to be
successful for overall water splitting. Reprinted with permission from [89]. Copyright 2010
Elsevier.

A study of Wang et al. [90] explained the potential technology that can be achieved in a one-
step excitation framework using a single photocatalyst or via a Z-scheme approach based on a
pair of photocatalysts. In an ideal world, such photocatalysis would proceed without
recombination or trapping, and surface catalytic activities would not include undesired
reactions. There are two mechanisms exist for photocatalytic water splitting [91]: One-step and
two-step excitation. Figure 8a demonstrates a one-step photo-excitation mechanism in which a
photocatalyst is activated by visible light and has the requisite potential to split water into O2
and H2. The photocatalyst should have attributes such as a deceased band gap for producing
visible photons and stability and thermodynamic potential for hydrogen

14
Figure 7. Properties of a promising photocatalyst and different types of catalyst using in
water spitting.

production. It's challenging to find a material that meets all of this mechanism's needs; Figure
8b shows a two-step photo-excitation system that meets these requirements [92], [93]. This
strategy, known as Z-scheme [94], [95], [96], is similar to the "natural photosynthesis process."
The possibility of splitting up evolved hydrogen and oxygen and the widespread availability of

15
Figure 8: a) One-step and b) two-step excitation techniques for water splitting Reprinted with permission from [90]. Copyright 2019 Elsevier.

16
visible light makes this strategy more practical and effective than a one-step photo-excitation
process; in this method, the first step is electron and hole generation. When light is equal to or
higher than the photocatalyst's band gap energy cascades, electrons are excited from the
valence band to the conduction band, leaving holes in the VB [97]; the reaction mechanism is
described in Eqs. 4 and 5:

H2O + 2H+ + reduction → 2H2 + oxidation (4)


H2O + oxidation → O2 + 2H + + reduction (5)

Because the VB must be more favorable than the oxidation potential needed to convert H2O to
O2, and the CB must be greater than the H+ to produce hydrogen at reduction potential, this
process requires around 1.23 e.v of energy. There is, however, an activation barrier for electron
transport across photocatalyst particles and water, which necessitates higher energy than the
band gap energy to enable photocatalytic water splitting [98]. The utilization of semiconductor
materials as photocatalysts is widespread since they are promising materials with the
appropriate band gap [99]. In semiconductor photocatalysis, photocatalyst electronic structure
is critical, consisting of VB and CB energy bands. The energy band gap is the space between
CB and VB, where electrons typically reside. Photocatalytic water splitting with
semiconductor-based photocatalysts involves several steps. Band structure plays an essential
role in water splitting [100]. The determined band gap for a semiconductor used in water
splitting is known to be 2eV. Semiconductors having band gaps less than 1.23 eV are not
suitable for photocatalytic water splitting because an interfacial charge transfer over potential
must also be evaluated [90]. A smaller band gap restricts photocatalyst selectivity in total water
splitting. As a result, photocatalytic water splitting mechanisms include photon absorption,
charge separation, charge transport and diffusion, catalytic reaction, and mass transfer [101].
Heterojunction semiconductors, which consist of two different semiconductors along with
thermal and lattice matching, are commonly employed in water splitting [40], [102]. Tuning
the parameters of photocatalysts can boost the efficiency of photocatalytic water splitting [103].
Carbonate salts, noble metal loading, metal ion doping, anion doping, dye sensitization,
composite semiconductors, metal ion implantation, and other approaches were examined, and
some of them were shown to be beneficial in enhancing hydrogen generation [98]. The amount
of H2 or O2 evolved from water in a certain period when exposed to light can be used to
calculate the efficiency of photocatalytic water splitting., with units mol h-1g-1 catalyst used to
make a measurable comparison (Eq. 6) [57].

𝑛(𝐻2 ,𝑂2 )
𝑅= (6)
𝑡𝑖𝑚𝑒 (𝑐𝑎𝑡𝑎𝑙𝑦𝑠𝑡)

Conversely, the catalyst-based metrics do not consider light intensity, making it impossible to
compare studies across research facilities, whereas light-based metrics are more universally
applicable. Therefore, another method is given to find the efficiency called quantum yield (QY)
(Eq. 7).

17
𝑁𝑜 𝑜𝑓 𝑟𝑒𝑎𝑐𝑡𝑒𝑑 𝑒𝑙𝑒𝑐𝑡𝑟𝑜𝑛𝑠
𝑄𝑢𝑎𝑛𝑡𝑢𝑚 𝑦𝑖𝑒𝑙𝑑 % = ∗ 100 (7)
𝑁𝑜.𝑜𝑓 𝑖𝑛𝑐𝑖𝑑𝑒𝑛𝑡 𝑝ℎ𝑜𝑡𝑜𝑛𝑠

In addition, Eq. 8 can be used to compute the efficiency (STH) under the illumination of air
mass 1.5 global (AM 1.5G).

Output energy of hydrogen evolved


𝑆𝑇𝐻% = ∗ 100 (8)
energy of incident solar light

Mainly, two theoretical thermodynamic parameters are used to evaluate the HER efficiency of
MOFs over other. Either or the ΔG*H or Gibbs free energy for water dissociation energy
(ΔG*H2O) are found to be an important parameter to identify and evaluate catalyst efficiency.
The ΔGH* is usually used to describe the catalytic efficiency towards HER of the catalyst,
where a lower the ΔGH* value indicates stronger adsorption. Typically, the ΔGH* value ∼ 0 eV
for the catalyst demonstrates its highest catalytic activity towards HER. For the hydrogen
evolution reaction (2H₂O → 2H₂ + O₂), a negative ΔG indicates that the reaction is
thermodynamically favorable, meaning it can proceed spontaneously. If ΔG is negative, it
implies that the reaction can occur without requiring an external energy source. In the context
of water splitting, a negative ΔG means that hydrogen gas (H₂) can be produced from water
(H₂O) without additional energy input. The ΔGH* is also used to evaluate or compared the
relative performance of the different HER electrocatalysts [104]. Fan et al [105] successfully
constructed ZnCdS quantum dots using the own metal centers of MOF and avoid its structural
collapse through photoetching self-sulfurization and ion exchange methods. By adjusting the
photoetching time, the ZnCdS−3/PZH exhibits the highest H2 generation rate of
11.32 mmol h−1, which is about 56 and 8 times higher than that of S-ZnCdS and S-ZnCdS/PZH
(Figure 9 a-d). DFT calculation results confirm the |ΔGH*| of ZnCdS−3/PZH is closer to zero,
which can improve the hydrogen evolution activity.

18
Fig. 9. The structural models for calculating Gibbs free energy of (a) PZH, (b) S-ZnCdS/PZH
at (0 0 2) facet, (c) ZnCdS−3/PZH (Yellow: S atoms, Purple: Cd atoms, Dark gray: Zn atoms,
red: O atoms, Brown: C atoms, French grey: N atoms and White: H atoms), (d) The calculated
free energy for hydrogen production reaction and (e) Schematic illustration of charge transfer
pathways for ZnCdS−3/PZH [105].

The main challenge is to generate H2 from renewable sources, preferably H2O. But, water
splitting is thermodynamically an uphill reaction that involves a large Gibbs free energy
change of 237 kJ mol−1[89] With the help of an appropriate catalyst one can decrease the energy

19
barrier for water splitting and increase the H2 evolution rate. To overcome this large energy
barrier, researchers are utilizing semiconductor photocatalysts for water splitting, which can
mimic the natural photosynthesis process involving the following two half-reactions [106].

Water reduction:

2H+ + 2e− → H2, ΔEo = −0.41 V

Water oxidation:

2H2O → O2 + 4H+ + 4e−, ΔEo = +0.82 V

Overall water splitting

2H2O→H2 + ½ O2, ∆E∘ = +1.23V, ∆G∘ = +237 kJ mol-1

1.3 Photocatalytic Overall Water Splitting

Photocatalytic overall water splitting (OWS) is the process of dissociation of water to H2 and
O2 without the addition of sacrificial agents, which is a grail reaction in photocatalysis. There
are very limited reports on photocatalytic OWS over MOFs, in which the activity remains
rather low. Huang and co-workers reported the immobilization of the active hydrogen
production site of Ni2+ into an Al-BDC-NH2 MOF by coordination with -NH2. In the
photocatalytic process, the electrons produced by BDC-NH2 were transferred to Ni2+ sites to
participate in water reduction, while the holes reserved on the linker drove water oxidation.
The dual-cocatalyst strategy was found to be effective for photocatalytic OWS over MOFs. Pt
and RuOx were introduced to MIL-125-NH2, which successfully drive water splitting into
H2 and O2 [107]. Furthermore, Zhang et al. achieved photocatalytic OWS by loading Pt and
MnOx inside and on the surface of UiO-66-NH2, respectively [108]. The spatially separated
cocatalysts greatly accelerated the separation of electrons and holes and prolonged their
lifetimes, promoting the OWS performance. Surprisingly, IEF-11 can drive photocatalytic
OWS without activity loss over 10 days.

Single-component catalysts directly driving photocatalytic OWS remain a grand challenge


because the bands must simultaneously satisfy the positions of H2 and O2 production. The Z-
scheme is a two-step water splitting approach consisting of two different semiconductors that
drive oxidation and reduction reactions, respectively, which can overcome the thermodynamic
limitation of single-component photocatalysts [109]. Recently, the Wang group assembled two
MOFs for H2 production and water oxidation reaction (WOR) into one liposome constructing
a Z-scheme for photocatalytic OWS [110]. The MOF for H2 production was a Hf-based MOF
nanosheet containing Zn-porphyrin as photosensitizers and Pt-porphyrin as water reduction
sites, which was functionalized with pentafluoropropionic acid to create a hydrophobic surface.
The WOR-MOF with hydrophilicity was constructed from Zr-oxo clusters and 2,2′-bipyridine-
5,5′-dicarboxylic acid (BPYDC) linkers, where the [Ru] and [Ir] molecular catalysts were
incorporated by the BPYDC acting as photosensitizers and catalytic sites, respectively (Figure
10a and 10b). Both MOFs were installed in the hydrophobic bilayers and the aqueous phase of
one liposome, respectively (Figure 10c). The carrier transport between the two MOFs was
connected by the Fe3+/Fe2+ and TCBQ/TCBQH redox relays to avoid charge recombination.
The optimal hybrid system achieved photocatalytic OWS with an apparent quantum yield of
(1.5±1) %.

20
Figure 10. The structures of (a) MOF for H2 production and (b) WOR-MOF. (c) Schematic
illustration of assembling the two MOFs into liposome to construct Z-scheme. Adapted with
permission [110]. Copyright 2021, Nature Publishing Group.

2. Development of a catalytic MOFs for photocatalytic water splitting

Numerous semiconductors, including metal oxides and sulfides, have been employed
to produce hydrogen. However, due to their remarkable physical and chemical properties, this
study considers MOFs and MOFs-based semiconductors as a photocatalyst in water splitting.
MOFs or porous coordination networks are a class of advanced materials designed by
employing various metal ions and organic linkers [111]. Recently MOFs have been gathering
attention as the best semiconductor for hydrogen production due to their structural and physical
properties [112], [113]. Furthermore, they will surely draw a lot of attention in the future for
catalytic applications like water splitting owing to their wide variety of inorganic and organic
constituents. Following are some of the most common benefits of MOFs and their composites
for using as catalyst: (i) All of the building units, such as metal nodes, organic ligands, and
channels, could be modified, resulting in a variety of frameworks, morphologies, sizes, and
functional areas; (ii) organic ligands and metal nodes, including both are employed as catalytic
centers and light-harvesting centers, to increase active surface area and shorten the photo-
generated charges transmission distance; (iii) The porosity of the structure
allows encapsulation of additional functional elements in order to create a novel composites
and achieve synergetic catalysis; (iv) nanopores may provide the substratum with a unique
chemical environment that is able, like the secondary structure of the active site, to interact
with the catalytic active site and further enhances its activities and selectivity; (v) The
simulation at molecular level of the natural photosynthesis system and single-site catalysts
loading can be carried out over certain control of the contact between metal and ligand; (vi)
MOFs to create particular functional metal complexes might be utilized as a precursor to get
new structures which cannot be achieved using conventional techniques [127]. Unlike other
porous materials like zeolites, activated carbon, and metal-complex hydrides, etc., MOFs are a
class of materials with several advantages, including thermal stability, discrete ordered
structure, ultra-low densities, an enormous interior surface area exceeding 6000 m2/g, ease of
synthesis, and a wide range of characteristics suitable for biological and chemical applications.
To elicit opinions on the precise definition of MOFs, Batten et al. [114], carried out a thorough
survey in 2010. According to their research, MOFs were broadly defined as porous, 3D
networks with carboxylate-containing materials, but there is disagreement on the precise

21
definition [115], [116], [117]. One of the early demonstrations of MOFs' potential was in
heterogeneous catalysis, which was originally investigated all the way back in 1994 [118]. This
information provides a continually expanding study interest due to the chemical mutability,
customized pore architectures, and extensive, conveniently accessible internal surface areas in
the potential applicants of MOFs [119], [120], [121]. A comprehensive look into MOFs,
including their history and the development of novel MOFs-based catalysts up to the current
decade, is presented in Figure 11. As potential materials for the splitting process, chemically
stable MOFs with characteristics similar to those of semiconductors were examined [102],
[122], [123]. Furthermore, the tunable porous structure obtained by choosing the suitable
linkers and metal nodes (Figure 12a), considerably contributes to an efficient hydrogen
evolution rate. When MOFs are used as a sacrificial precursor or support to build MOF-derived
nanostructures and MOF-based composites, it is possible to see a reduction in both the specific
surface area and the pore volume [124], [125], [126]. MOFs are well-known for their incredible
capability of adding new functions, which they accomplish with the assistance of various
organic linkers and active catalytic metal sites [128]. They exhibit exceptional optical,
electronic and catalytic properties because of their high porosity (90% free volume), dynamic
structure, design tunability, ultrahigh surface area, and crystalline nature [129], [123]. MOFs
possess unique advantages over conventional semiconductor catalysts and can

22
Polyoxometalate- Metal–
Encapsulating organic
Cationic Metal– frameworks metal–organic
Organic Framework for selective
Two-Dimensional frameworks
(POM@MOF)as a catalysis in
Metal–Organic are used for
Heterogeneous biofuel
Homochiral Surfaces for the derivation Mixed‐metal MOFs:
Efficient Hydrogen Catalyst for
Metal–Organic of different unique opportunities
Desulfurization
Frameworks for Evolution from nanomaterials in metal–organic
119 121
Asymmetric Water 113 framework (MOF)
126
Heterogeneous functionality and
Manufacturing of MOF Catalysis 110 design. Functionalized metal-
Thin Films on 136 organic frameworks for
Structured Supports photocatalytic
for Separation and degradation of organic
Catalysis pollutants in
109 147
environment
Metal–Organic Metal oxide integrated
Frameworks used for 2015 metal organic frameworks
Improved (MO@ MOF) was used in
Photocatalytic emerging photocatalytic
Hydrogen Evolution 106 applications.
222

MOF-5 for Metal-organic


degradation of framework-derived
phenol in water
under UV light for
the first time. 105
M E T A L O R G AN I C multifunctional
photocatalysts.
234

F R A M E W O RK S
T I MELINE

Figure 11. This timeline shows the historical background of the invention of MOFs, their derivatives, and their roles in different eras.

23
Figure12. (a) Components of a MOF Reprinted with permission from [132]. Copyright 2016
Elsevier b) 3D MOFs c) ZIF- MOF d) MOF199 Reprinted with permission from [133].
Copyright 2010 Elsevier. e) MOF-5 Reprinted with permission from [133]. Copyright 2017
Elsevier.

be used as templates for producing functionalized materials (e.g., porous carbon, metal sulfides,
metal oxides, and carbon metal/metal oxide hybrids) [130]. The availability of 0, 1, 2, or 3D
superstructures provides practical conformational control over MOFs, which may translate
well to applications requiring thin films, hollow spheres, or membranes, as shown in Figures
12 (b-e) [131]. They have several distinct advantages over other photocatalysts in photoinduced

24
reactions, which can increase the transfer and diffusion of substrate molecules [134], [135],
[136]. These advantages were inherited from porous materials such as zeolite [137], carbon
nanomaterials [138], covalent organic frameworks (COFs) [139], and DNA hydrogels.
Moreover, the growing release of antibiotics into the environment has motivated extensive
studies on the elimination of these harmful substances and the development of advanced
adsorbents. Metronidazole is a major example of antibiotic pollutants due to its extensive
usage, high solubility in water and considerable resistance to degradation The adsorption of
MNZ by these MOFs can be considered as environmentally benign and the results of this
research are expected to further motivate the application of metal–organic frameworks as
promising adsorbents for removal of antibiotics from wastewater [140]. Shi et al. [141] have
highlighted the photocatalytic application of MOF-based composites and MOF-derived
materials for hydrogen production from water. However, there are several bottlenecks in the
pathways of MOFs for producing hydrogen by photocatalytic water splitting [142], [143], but
in our report, we find effective strategies to improve the catalytic efficiency of MOFs.

Using light, a semiconductor generates electrons and holes. Then, the MOF undergoes a charge
shift in which electrons migrate to the CB and spots go to the VB [144], [118]. The reduction
reaction happens in the CB of MOF, whereas the oxidation reaction occurs in the VB with the
presence of an oxidizing agent. MOFs use a similar photocatalytic process, with the energy
levels HOMO and LUMO. The response is depicted in Eqs. (9, 10, and 11), which explain how
it takes place.

𝑀𝑂𝐹𝑠 → 𝑒 − + ℎ+ (9)
1
𝐻2 𝑂 + ℎ+ → 2 𝑂2 + 2𝐻 + (10)
2𝑒 − + 2ℎ+ → 𝐻2 (11)

Figure 13 depicts the mechanism of H2 production by photocatalysis upon a photoactive MOF


semiconductor. At first, light is typically absorbed by the organic functional groups in MOFs
to form electron-hole pairs. In step 2, the metal nodes use the cluster charge-transfer process to
capture the photogenerated electrons, and in step 3, they mix with protons to make H2. The
holes in the HOMO return to the ground state after obtaining an electron from the electron
donor (Step 4). In most cases, a MOF will go through the reaction mechanism described above.
Several types of research have been conducted to demonstrate the mechanism of MOFs as a
semiconductor for the splitting of water to produce hydrogen as a source of energy [144], [145],
[129], [238], [369]. The formation of hydrogen through photocatalytic water splitting was the
topic of the next section, which focused on materials based on MOFs.

25
Figure 13. Shows the mechanism of MOFs in photocatalytic water splitting Reprinted with
permission from [369]. Copyright (2019) American Chemical Society.
The development of photocatalysts is very important to enable water splitting with visible light.
The main steps in the photocatalytic water splitting reactions should be tailored to meet the
requirements for photocatalysts capable of water splitting. To date, the application of MOFs
and their derivatives as photocatalysts for water splitting has become a burgeoning field.
Herein, we showcase several representative MOF-based photocatalytic materials for H2
evolution reactions (HER) which are being discussed in later sections.

2.1 Transition metals-based MOFs as photocatalysts


Transition-metal (Fe, Co, Ni) based metal-organic framework materials with
controllable structures, large surface areas and adjustable pore sizes have attracted wide
research interest for use in next-generation in different field of environment [146], [147], [148],
[149], [150]. As a class of multipurpose material, MOFs materials are obtained by utilizing
cheap starting materials, such as metal-salt of Fe, Co and Ni. In addition, the valence electron
structures of Fe, Co and Ni are all 3d6–84s2, so they share similar electrochemical properties.
Also, the transition-metal (Fe, Co, Ni) ions have many coordination numbers when they are
coordinated with organic ligands, so they can be used to design and synthesize various novel
MOFs. Mori et al. [151] first depict that a Ru-MOF made with paddlewheel diruthenium cores
([Ru2(p-BDC)2] n could perform as active sites for water reduction to H2 in the availability of
[Ru(bpy)3]2+. They found a significant amount of activity, which equated to an apparent
quantum yield of 8.16 %. Furthermore, this material enables the adsorption of various gases
into its pores and its hydrogen uptake capacity is 1.2 wt% at 77.4 K. Figure 14e shows [Ru2(p-
BDC)2]n R1(catalyst) as a 2-D square grid sheet configuration with 1-D channel pores which
is unusually stable in air, and has a strong degree of hydroscopicity. Zhang et al. [146] then
identified two isostructural Zr-based MOFs, UiO-66 and NH2-UiO66; when exposed to visible
light, they enable photocatalytic hydrogen generation in methanol or a water/methanol mixture.
MOFs in covalently coupled dimers, coordination polymers, and supramolecular assemblies
can also function as effective photocatalysts when combined with a photosensitizer [152]. The
26
production of photogenerated charge carriers began when the light was trapped by a
photosensitizer. The electrons are subsequently received by the MOF, which serves as a
photocatalytic site for H2, culminating in the reduction reaction and photosensitizer recovery
with the help of a sacrificial agent. Photosystem efficiency is ensured by good light, reactant,
and MOF photocatalyst interaction [145].
Ordinarily, a noble metal is thought of as a metallic chemical element that is typically
resistant to corrosion and can be found in nature in its unprocessed state [153], [154], [155].
Noble-metal-based catalysts are widely regarded as state-of-the-art catalysts due to their unique
intrinsic properties and irreplaceable catalytic activities [156]. Noble metals such as rhenium,
ruthenium, rhodium, palladium, silver, osmium, iridium, platinum, and gold can be utilized as
co-catalyst; these metals are categorized as belonging to group VIIb [157], [158], [159].
Recently several Ru noble metal-based MOFs with high hydrogen evolution rates have been
reported, including Rh2(p-BDC)2 and Ru2(MTCPP)BF4 [160]. Similarly, Lan et al. [161]
developed two advanced and efficient MOFs, Ru-TBP and Ru-TBP-Zn, constituting Ru2
secondary building blocks. The coordinating environment of Ru2 paddlewheel secondary
building units, as well as the prospective perspective and crystalline structure of the generated
MOFs, are depicted in Figure 14a. Upon visible-light (λ> 400 nm) irradiation, the excited
porphyrin ligands inject electrons into adjacent Ru2 SBUs to produce hydrogen from water
(Figure 14c). With an average size of 500 nm, the produced MOFs clearly feature a nanoplate-
like architecture. The higher HER catalytic activity of Ru-TBP-Zn than that of Ru-TBP is likely
due to the better photosensitizing ability of the TBP-Zn ligand than the TBP-Ru or TBP ligand.
The TON [defined as n(1/2H2)] reached 21.2 for Ru-TBP and 39.4 for Ru-TBP-Zn after 72 h
irradiation as illustrated in Figure 14b. In comparison, the Ru2-PD + H4TBP-Zn homogeneous
control exhibited a modest TON of 1.4. The significantly enhanced catalytic activity of Ru-
TBP-Zn over the homogeneous control confirmed the important role played by hierarchical
organization of photosensitizing ligands and catalytic SBUs in facilitating multielectron
transfer to drive photocatalytic HER. After photocatalysis, both Ru-TBP and Ru-TBP-Zn
showed the same XRD patterns as freshly prepared MOFs (Figure 14d) and showed only 0.34%
and 0.30% leaching of Ru (determined by ICP-MS), respectively, indicating their structural
stability during HER. This work provides a blueprint for designing multifunctional MOFs with
catalytic SBUs and photosensitizing ligands for photocatalytic and other applications.

27
b

Figure 14. (a) A cross-section of the Ru-TBP crystal structure along the (100) axis. (b) MOF
TEM, HRTEM, and SAED patterns for Ru-TBP and Ru-TBP-Zn. (c) Time-dependent
comparison (d) The visible-light-driven HER reaction process with Ru-TBP or Ru-TBP-Zn
Reprinted with permission from [161]. Copyright 2021 Elsevier. (e) Ru2 paddlewheel motif
structure and self-assembled molecular structure of [Ru2(p-BDC)2]n Reprinted with permission
from [151]. Copyright 2009 Elsevier.

Moreover, Maeda et al. [162] discovered Noble-Metal/Cr2O3 Core/Shell Nanoparticles as a


Co-catalyst for Photocatalytic water splitting. Further research by Liu and colleagues [163]
investigated the efficiency of carbon-supported noble metal nanoparticles as catalysts for the
process of water splitting. They demonstrate a flexible synthetic method for creating noble

28
metal nanoparticles/nanocarbon composites, which are indicated as Rh(nP)/nC, RhPt(nP)/nC,
and Pt(nP)/nC, respectively, by combining different noble metals, such as rhodium (Rh), RhPt,
and Pt nanoparticles, with carbon. In a related study, Sakamoto et al. [164] developed highly
distributed noble-metal/chromium (core/shell) nanoparticles as effective hydrogen evolution
boosters for overall photocatalytic water splitting under visible light. Although noble metals
evolve a large amount of hydrogen, they have a drawback affecting their ability; furthermore,
noble metals are not commissionable. Hydrogen can be efficiently produced by water
electrolysis, which is essential for the growth of renewable energy; currently, scale-up uses for
noble-metal catalysts are constrained by their high price and unavailability [165].
For economical and long-term hydrogen production, noble metal-free electrocatalysts
for water splitting are essential [166]. After the defect of noble metals, researchers developed
some noble-free metals which can be used as co-catalysts, i.e., MOFs composed of titanium,
zirconium, cadmium, copper, nickel, indium, and lanthanum. Shi et al. [167] demonstrated one
of the first examples of noble-metal-free photocatalysts for hydrogen and oxygen production.
As a result, they developed a high-nuclear CuI24(µ3-Cl)8(µ4-Cl)6-based POM@- MOF, which
shows a powerful dual-functionalized photocatalytic ability. According to this, using mixed
ZZULI-1 PMOs instead of oxygen generation is anticipated to provide a more efficient charge
transport mechanism toward oxidized [Ru(bpy)3] via the dual-functionalized ZZULI-1, as
shown in Figure 15. The oxidized [Ru(bpy)3] is regenerated by 3+ to the starting position. By
incorporating the oxidative [W12O40]8− and [W6O19]2− polyoxoanions within the pores of Cu I
mXn-based MOFs, they develop a new method to merge the photocatalytic water oxidation and
water reduction within a novel dual-functionalized POM@MOF. They envision that the mixed
Keggin- and Lindqvist-type POMs may work as oxidative photocatalysts to generate O2,
whereas the {CuI24(µ3-Cl)8(µ4-Cl)6} clusters of the cationic framework can act as photoelectron
generators to improve the activity of H2 generation.

29
Figure 15. POM@MOFs with dual-functionalized mixed Keggin- and Lindqvist-type CuI24(µ3-
Cl)8(µ4-Cl)6 as hydrogen and oxygen production photocatalysts Reprinted with permission
from [167]. Copyright (2019) American Chemical Society.

In addition, Yang et al. [168] suggested a noble-metal-free photocatalyst system developed


utilizing BODIPY-based MOFs for extremely effective visible-light-driven H2 evolution. It is
expected that this research will guide continuing efforts to use noble-metal-free photocatalytic
techniques based on BODIPY-based MOFs for highly efficient and long-term solar fuel
production, providing essential insights into the function of porosity in photocatalytic
performance. Interestingly, Luo et al. [169] established novel bifunctional NH2-MIL-
125/Co(dmgH)2 composite catalysts with varying Co(dmgH)2 concentrations that can achieve
photocatalytic NO elimination and hydrogen production simultaneously. Similar studies have
shown that Nobel-free metals are used widely as a co catalyst in photocatalytic water splitting
[170], [171], [172], [173]. Several noble free metals, such as Cu, Ti, and Zn, are covered in the
following sections because they play an essential role in visible-light-driven water splitting and
are also depicted in Table 1. Copper, the first metal discovered and used by humans, is abundant
in nature and available in reasonable price. Dong et al. [174] produced a Cu-MOF for the first
time, with redox-active copper centers with photoactive rhodamine-derived ligands (Cu-RSH).
Figure. 16(a) shows the perspective view of the coordination environment of the Cu II center
in Cu-RSH. The resultant Cu-RSH was verified for photocatalytic hydrogen evolution rate
under visible light irradiation with a wavelength greater than 420 nm, using TEOA as the
sacrificial agent. The evolved hydrogen amount was found to be 7.88 mmol g-1 h-1, which is
much higher in comparison to AlRSH, H2RSH, and Co-RSH (Fig. 16(b)). The remarkable
hydrogen evolution rate in the case of Cu-RSH was experimentally analyzed by measuring the
charge transportation rate using cyclic voltammetry, time-resolved fluorescence decay, and
transient photocurrent experiments. The working mechanism of photocatalysis over
rhodamine-based Cu-RSH is displayed in Fig. 16(c). The emission spectra showed that the
introduction of Cu-RSH resulted in a significant decrease in the band intensity of the Eosin Y
(EY) emission indicating electron transfer from EY* to Cu-RSH. Furthermore, from the
photocurrent measurement it was clearly noticed that greatly enhanced photocurrent was
observed on the electrode modified by EY + Cu-RSH (3.3 mA cm-2). This result indicated an
improvement in the charge transport from the excited state of EY to Cu-RSH and then to the
surface of the working electrode Fig. 16(d and e).

30
a b d
c b

c
c e

Figure 16. a) Perspective view of the Cu-RSH crystal structure. (b) Comparison of Cu-RSH
hydrogen evolution rate with Al-RSH, Co-RSH, and H2RSH as a function of time. (c)
Schematic illustration of the visible-light-driven HER reaction process using Cu-RSH as a
photocatalyst and TEOA as the sacrificial agent. (d) Emission spectra of EY as a function of
Cu-RSH. (e) Transient photocurrent/time profiles of the bare glassy carbon electrode (black),
pristine Cu-RSH (red), EY + Cu-RSH (blue), which were coated on the glassy carbon electrode.
Reprinted with permission from [174]. Copyright 2016 Elsevier.

Similarly, Hu et al. [175] synthesized a copper(II)-based MOF film (MOF-199/Ni) prepared


by electrodeposition shows exceptional high photocatalytic hydrogen production rates of 8000
μmol h-1 g-1 (based on the mass of MOF-199) and 24400 μmol h-1 g-1 with Pt as the cocatalyst
and eosin Y as photosensitizer. The activity is one of the highest among all the reported MOF-
based hydrogen production systems. With nickel foam as a substrate, the deactivation of MOF-
199 is largely alleviated, and the MOF199/Ni film can be easily recycled and reused after
photocatalytic hydrogen production. The obtained MOF-199/Ni film (Figure 17a) was
characterized by scanning electron microscope (SEM), powder X-ray diffraction (PXRD) and
Fourier transform infrared spectroscopy (FT-IR). SEM observation shows the uniform coating
of MOF-199 on the entire surface of nickel foam; the intergroup MOF-199 grains display a
narrow size distribution centered at ca. 50 nm (Figure 17b). The phase purity of the formed
MOF-199 on the nickel foam was evidenced by PXRD (Figure 17c). FT-IR measurement of
the film also displayed identical characteristic signals with pure MOF-199 (Figure 17d), further
confirming the formation of pure MOF-199 on the nickel foam substance. Figure 17e depicts
the eosin Y-sensitized photocatalytic process on the MOF-199/Ni film. Eosin Y first absorbs
visible light, triggering photoelectrons from the highest occupied molecular orbital (HOMO)
to the lowest unoccupied molecular orbital (LUMO) (-0.64V vs NHE). The electrons then
transfer to the MOF-199 conduction band (CB) due to its lower energy level (-0.03 V vs NHE);
using triethanolamine (TEOA) for the electron donor and Pt for a co-catalyst, hydrogen is
reduced from water by accepting photoelectrons from the MOF-199 CB.

31
(e)

(c) (d)

Figure17. (a) Optical photograph, (b) SEM image and (c) PXRD pattern of the MOF-199/Ni
film prepared by electrodeposition, (d) FT-IR spectra of MOF-199 (black), MOF-199/Ni (red)
and MOF-199/Ti (blue), (e) Illustration of the eosin Y sensitized photocatalytic procedure on
the MOF-199/Ni film, Cu-based MOFs. Reprinted with permission from [175]. Copyright 2020
Elsevier.

Cu- MOFs produced roughly 4.6 times more H2 than Cd- MOFs under the same conditions,
according to a series of control studies. The primary cause of this is because, despite having
identical valence band (VB) energies, Cu-MOFs and Cd-MOFs have different conduction band
(CB) potentials. Cd- MOFs have a higher reduction ability [176], [177]. Zirconium (UiO-66,
UiO-67, UiO-68) and Ti-based MOFs have also gained popularity due to their high chemical
reaction activity and oxophilicity in the formation of solid crystal lattices [178], [144], [179],
[141]. Cavka and colleagues [180] present a novel zirconium-based inorganic building block
that synthesizes high surface area with unheard-of stability. Functional groups on the
backbones of MOFs allow the production of various materials with rapid diffusion rates inside
the MOF [181], [182]. Garcia and his colleagues [183] conducted one of the earliest
investigations into the photocatalytic capabilities of the UiO MOF family. They employed
UiO-66 and UiO-66 (NH2) as photocatalysts for H2 production. In addition, the inclusion of Pt
nanoparticles at the surface of the MOF (Pt@UiO-66) as co-catalysts resulted in a considerable
increase in H2 generation; this work marked the start of new research into these materials.
Additionally, Chao Hou et al. [184] proposed a molecular photosensitizer [Ru(dcbpy)(bpy)2]2+
and a proton reduction catalyst Pt(dcbpy)Cl2 which were successfully incorporated into a
highly robust metal–organic framework (MOF) of Zr(IV)6O4(OH)4(bpdc)6 by making use of a
mix-and-match approach. The molecular integrity of the Ru and Pt complexes within the MOFs
was demonstrated by a variety of techniques, including XRD, BET, TGA, SEM, TEM,
HAADF-STEM, EDX, DRUS and XPS. This di-component Ru–Pt@UIO-67 MOF assembly
allows a facile arrangement of the photosensitizer and the reduction catalyst with close spatial
proximity, promotes the electron transfer between them, and thus leads to a significantly
improved hydrogen evolution activity in aqueous solution at pH 5.0 upon visible light
irradiation. This Ru–Pt@UIO-67 system represents the first example of MOFs functionalized
with two different transition metal complexes, which can be used as a photosensitizer and
catalyst, respectively, for hydrogen generation from water. Direct reaction of ZrCl4, RuDCBPY
and PtDCBPY in DMF failed to produce highly crystalline Ru-Pt@UIO-67, probably due to
the steric demand of bulky Ru and Pt complexes. Then, they utilized the mix-and-match
approach and hypothesized that the Ru and Pt complexes could be doped into the framework

32
of UIO-67 by taking advantage of their matching ligand lengths (Figure 18). Indeed, they
experienced no problems in preparing highly crystalline Ru-Pt@UIO-67 MOFs from ZrCl4,
bpdc, RuDCBPY, PtDCBPY and acetic acid (as a crystal growth modulator) in DMF under
solvothermal reaction conditions for 24~76 h. Pt@UIO-67 with high quality was also obtained
by adopting the same synthetic strategy. The UiO-66(NH2) has also been modified with CdS
for the selective oxidation of certain alcohols [185], for the photocatalytic reduction of CO2
[186], and for H2 generation [187]. Inspired by this result, Yan and colleagues., [188]
introduced Rhodamine B into the MOF with Pt as a co-catalyzer and reported an increase in
activity up to 30 times that of the Pt@UiO-66 action, but the MOF was not responsible for the
photocatalytic activity. Ti is an appealing metal for constructing MOFs due to its low toxicity,
redox activity, and photocatalytic characteristics. However, there have been few reports of Ti-
MOFs since, under solvothermal conditions and in the presence of various ligands, the resultant
product is usually titanium dioxide or a weakly crystalline substance [189].

Figure 18. The hybrid material Ru-synthetic Pt@UiO-67's schematic diagram Reprinted with
permission from [184]. Copyright 2019 Elsevier.

In addition to this, four titanium-based MOFs were developed by Devic and his colleagues
(MIL-167, MIL-168, MIL-169, and "NTU-9-like"), and their photocatalytic activity was
considered accountable. The reactivity of 2,5-dihydroxyterephthalic acid (H4DOBDC) with
titanium(IV) precursors was thoroughly investigated for the synthesis of metal–organic
frameworks under solvothermal conditions. Four crystalline phases were isolated whose
structures were studied by a combination of single-crystal or powder X-ray diffraction and
solid-state NMR. The strong coordination ability of the phenolate moieties was found to favor
the formation of isolated TiO6 octahedra bearing solely organic ligands in the resulting
structures, unless hydrothermal conditions and precondensed inorganic precursors are used. It
is worth noting that these solids strongly absorb visible light, as a consequence of the ligand-
to-metal charge transfer (LMCT) arising from Ti–phenolate bonds. Preliminary photocatalytic
tests suggest that one compound, namely, MIL-167, presents a higher activity for hydrogen
evolution than the titanium carboxylate MIL-125-NH2 but that such an effect cannot be directly

33
correlated with its improved light absorption feature [190]. Due to the similarities between
zirconium (Zr) and titanium (Ti), several researchers have incorporated titanium into Zr-based
MOFs to increase photocatalytic activity. Using a post-synthetic exchange approach, Li's
group., [191] created the hybrid NH2-UiO-66(Zr/Ti) by substituting Ti for Zr ions in NH2-UiO-
66(Zr). Calculations based on the density functional theory (DFT) and an analysis of electron
spin resonance (ESR) revealed that the inclusion of titanium improved electron transport,
which was advantageous for increasing activity. In addition, Yuan and his coworkers. [192]
were able to produce MOFs based on Ti-oxo clusters (Ti-MOFs) which represent a naturally
self-assembled super lattice of TiO2 nanoparticles separated by designable organic linkers as
antenna chromophores, epitomizing a promising platform for solar energy conversion.
However, despite the vast, diverse, and well-developed Ti-cluster chemistry, only a scarce
number of Ti-MOFs have been documented. The synthetic conditions of most Ti-based clusters
are incompatible with those required for MOF crystallization, which has severely limited the
development of Ti-MOFs. This challenge has been met herein by the discovery of the
[Ti8Zr2O12(COO)16] cluster as a nearly ideal building unit for photoactive MOFs. These MOFs
demonstrate high porosity, excellent chemical stability, tunable photo response, and good
activity toward photocatalytic hydrogen evolution reactions. The discovery of the
[Ti8Zr2O12(COO)16] cluster and the facile construction of photoactive MOFs from this cluster
shall pave the way for the development of future Ti-MOF-based photocatalysts.
In contrast, the [Zr6O4(OH)4(COO)12] cluster (Figure 19a) has been shown to be a nearly ideal
inorganic building unit for the construction of MOFs. Various MOFs based on the
[Zr6O4(OH)4(COO)12] cluster have been synthesized under similar synthetic conditions. This
encourages them to find a Ti-containing cluster with behavior similar to [Zr6O4(OH)4(COO)12]
for MOF synthesis. Efforts were made in our group to build Ti-MOFs using existing Ti-oxo
clusters in the literature but to no avail. These clusters tend to dissociate or transform during
the MOF synthesis. While existing monometallic clusters do not meet the requirement for MOF
synthesis, bimetallic clusters have brought new opportunities. During the continued interest in
cluster chemistry, a [Ti8Zr2O12(COO)16] cluster (Figure 19c) was discovered as a promising
inorganic building unit for MOF synthesis. The [Ti8Zr2O12(COO)16] (R = Me, Et, Ph, p-Tol
and PTBB) clusters (Figure 19d−h) were synthesized by the solvothermal reactions of Zr4+,
Ti4+, and an excess amount of carboxylic acids, similar to the synthetic conditions for
[Zr6O4(OH)4(COO)12] clusters and most Zr-MOFs. Structure analysis reveals that the
bimetallic decanuclear cluster is composed of a Ti8-cube capped by two Zr4+ centers on the top
and bottom. Four μ2-O rest on the equatorial plane, each bridging a pair of Ti4+ to form a
[Ti8O4]24+ center. Two Zr4+ on the top and bottom were further connected to the [Ti8O4]24+
center by eight μ3-O, generating the [Ti8Zr2O12]16+ core. The [Ti8Zr2O12]16+ core was terminated
by 16 carboxylates fulfilling a neutral [Ti8Zr2O12(COO)16] cluster (Figure 19c). In fact, the
[Ti8Zr2O12(COO)16] cluster is structurally related to the [Zr6O4(OH)4(COO)12] cluster (Figure
19b). As a result, four carboxylate ligands within the equatorial plane that bridges the Zr4-
square are replaced by eight carboxylate ligands perpendicular to the equatorial plane, which
bridges the Ti8-cube
The solvothermal reaction between the [Ti8Zr2O12(MeCOO)16] cluster and BDC (1,4-
benzenedicarboxylate) gives rise to the desired MOF, PCN-415 (Figures 19i), under similar
synthetic conditions used for the syntheses of Zr-MOFs. The acetate terminated
[Ti8Zr2O12(MeCOO)16] cluster was adopted as the starting material because of its good
solubility in DMF. The preformed cluster as starting material is the key for the formation of
pure PCN-415 because it eliminates numerous competing side reactions. For comparison, the
synthesis of PCN-415 was attempted using a mixture of stoichiometric Zr4+ and Ti4+ metal salts
which leads to low crystallinity PCN-415 with UiO-66 impurity. In addition, an excess amount
of carboxylic acid as modulating reagent is also necessary for the formation of a highly

34
crystalline product. The carboxylic acid acts as a modulator that competitively coordinates with
the metals and slows down crystal growth to help produce highly crystalline products.
Furthermore, isoreticular expansion of PCN-415 can be realized by elongation of BDC into
2,6- naphthalenedicarboxylate (NDC), which gives rise to an isoreticular MOF, PCN-416
(Figures 19j).

Figure 19. Coordination compounds based on [Ti8Zr2O12(RCOO)16] cluster. (a)


[Zr6O4(OH)4(COO)12] cluster; (b) the relationship between [Zr6O4(OH)4(COO)12] and
[Ti8Zr2O12(RCOO)16] clusters; (c) [Ti8Zr2O12(RCOO)16] cluster; (d−h) discrete
[Ti8Zr2O12(RCOO)16] clusters formed with different carboxylate ligands; (i) and (j) MOFs
based on [Ti8Zr2O12(RCOO)16] clusters and different carboxylate linkers. Reprinted with
permission from [192]. Copyright (2018) American Chemical Society.

The [Ti8Zr2O12(COO)16] cluster is invented and the convenience with which photoactive MOFs
can be synthesized from this cluster will clear the path for developing future Ti-MOF-based
photocatalysts. Several efforts have been undertaken to create Ti crystalline materials for use

35
in photocatalysis: It has also been treated with nanoparticles of Pd, Pt, Cr, and Ag to boost the
MOF's catalytic activity or with organic dyes to improve its light absorption in the visible light
area [193]. Moreover, Assi et al. [190] created a variety of Ti-related MOFs, i.e., MIL-167,
MIL-168, MIL-169, and "NTU-9- like" for photocatalysts for great hydrogen generation. The
conclusion reveals that MIL-167 has a tremendous photocatalytic ability Figure 20b in contrast
to other MOFs. Figure 20a displays the crystal structures of NTU-9 and MIL-167, and MIL-
167 and NTU-9 have their own coordinating architecture and topological network.
a) b)

Figure 20 a) Crystalline structures. Coordination sphere around the titanium (IV) ions in NTU-
9 and MIL-167 and Coordination framework and the corresponding topological network for
MIL-167 and NTU-9, respectively. b) Photocatalytic productivity of MIL-167, MIL-169,
“NTU-9-like”, and MIL-125-NH2 (30 mg) for hydrogen production. Reprinted with permission
from [190]. Copyright (2019) American Chemical Society.

Guo et al. [194] investigated acet-dependent photocatalytic hydrogen production using NH2-
MIL-125(Ti) MOFs and discovered that the 110 facet has the highest evolution rate, roughly
three times that of the 111 aspects. Recently, Han et al. [195] developed x%-MIL-125- (SCH3)2
using MIL-125 as the host MOF and terephthalic acid linker (H2BDC-(SCH3)2), which shows
the cavity in the MOF structure and given a very high amount of hydrogen (3814 µmol g-1 h-1)
with a quantum yield of 8.9%. In addition, a porphyrin titanium MOF has been created by
employing Ti-oxo clusters as photoactive sites and the porphyrin ligands as photosensitizers to
increase the optical response of the MOF. This has allowed the MOF to display a greater degree
of photocatalytic activity [196]. Next to this, the ability of Cd and In-based MOFs to capture
visible light makes them attractive Wang et al. [197] adopted a simple and effective strategy
which was developed to immobilize ecofriendly dye inside a porous metal−organic framework
(MOF) built from Eosin Y with [Cd2(COO)2(μ2-H2O)] secondary building units for the first
time. The MOF exhibited efficient photocatalytic activity for H2 evolution under visible-light
irradiation with a turnover number of 13920 and an initial turnover frequency of 7433 h−1,
which was approximately 31 times the photocatalytic efficiency of Eosin Y. Xiao et al. [198],
for the first time, synthesize a Cd-based MOF (Cd-TBAPy) with a 2D layered framework using
1,3,6,8-tetrakis (p-benzoic acid) pyrene (H4TBAPy) as an organic linker. It was noted that the
hydrogen and oxygen evolution rates reached 4.3 and 81.7 µmol g-1 h-1, respectively, when
cobalt phosphate (CoPi) was used as a co-catalyst. As shown in Figure 21a, the Cd atom is

36
coordinated with eight oxygen atoms from four different linkers, and the bond length of Cd-O
is ≈2.348 Å. The coordination between Cd atoms and H4TBAPy linkers shapes the backbone
of Cd-TBAPy into 2D sheets (Figure 21b), and all single layers are orderly stacked to form a
2D eclipsed structure with an interlayer spacing of ≈4.28 Å (Figure 21c). Meanwhile, the BET
surface area and the Langmuir surface area of Cd-TBAPy is measured to be 27 and 38 m2 g−1
respectively. The experimental XRD pattern (Figure 21d) of Cd-TBAPy matches very well
with the simulated one which is generated on the basis of single crystal structure analysis,
showing a good phase purity of the bulk material. Moreover, the possible mechanism of Cd-
TBAPy for water reduction and oxidation is illustrated and shown in Figure 21e. First, the Cd-
MOFs absorb visible light to generate electrons and holes, which will be then transferred into
the cocatalysts Pt or CoPi due to their different energy alignment. Afterward the deposited Pt
and CoPi will act as active sites to carry out the catalytic reduction and oxidation of water,
respectively.
e)

Figure 21. The structure and analysis of Cd-TBAPy: a) the coordination mode of the Cd atom
in Cd-TBAPy, b) the single layer structure of Cd-TBAPy viewed along the b-axis, c) 2D
layered structure of Cd-TBAPy, and d) XRD patterns of Cd-TBAPy and Cd-TBAPy-simulated.
Color representation: red, O; gray, C; green, Cd. H atoms are removed for clarity, e) The
proposed mechanism for visible-light-driven photocatalytic H2 and O2 evolution over Cd-
TBAPy [198].

Moreover, Isostructural Cu and Cd MOFs [199] were studied to better understand


factors influencing photocatalyst behavior. The Cu MOF performance for hydrogen generation
was many orders of magnitude higher than the Cd counterpart under the same conditions. In
addition, Leng et al. [200] manufactured an unusual OOP (out-of-plane) porphyrin-based
MOF. This MOF was synthesized by controlling the metal ion release with an unprecedented
In(OH)3 precursor. Moreover, this possesses high stability and exhibits unexpectedly high
photocatalytic hydrogen production activity, which significantly outperforms the isostructural
in-plane porphyrin-based MOF counterparts. Under light activation, the OOP In(III) ions
quickly detach from the porphyrin rings, enhancing electron-hole separation efficiency and
photocatalytic performance. Although sharing the same LUMO potential, USTC-8 with
different metal ions at porphyrin centers exhibit distinctly different activities toward
photocatalytic H2 production (Figure 22a). Strikingly, USTC-8(In) displays the best
photocatalytic activity (341.3 µmol g−1 h−1) by introducing H2PtCl6 to give Pt nanoparticles as
co-catalyst under visible-light irradiation, which is around 18~37 times higher than the
activities of other USTC-8(M) (M = Co, Cu, Ni) as well as ~3.4 times of USTC-8(0.2In)
activity. Thanks to the high stability, all these MOFs present well retained framework integrity

37
and crystallinity during photocatalytic process. To elucidate the mechanism behind the
exceptional activity of USTC-8(In), the photocurrent and electrochemical impedance
spectroscopy (EIS) measurements have been carried out to unveil their e-h separation
efficiency and electronic conductivity, respectively. As expected, USTC-8(In) shows the
highest photocurrent response among all investigated MOFs (Figure 22b), revealing that OOP
porphyrin endows MOF with superior charge separation efficiency. The EIS curves are found
to be almost identical in the absence or presence of light irradiation for USTC-8(M) (M = Co,
Cu, Ni) (Figure 22c), which indicates that their high charge-transfer resistances are not affected
by light. Upon exposure to light for a while, the UV-Vis absorption spectrum of USTC-8(In)
suspension solution displays significant changes, e.g. the B-band redshift and the peaks from
three to four in Q-band, indicating the In3+ detachment (Figure 22d).

Figure 22. (a) Photocatalytic H2 generation by water splitting across several MOFs. (b) USTC-
8 series photocurrent responses. (c) US-8(M) (M = Cu, In) EIS Nyquist plots in the absence
and presence of light irradiation. (d) UV-vis absorption spectra of USTC-8(In) suspension
solution with and without illumination. Reprinted with permission from [200]. Copyright
(2018) American Chemical Society.

Lu et al. [201] proposed indium-based metal−organic framework which has been successfully
constructed and realizes efficient UV-light-driven H2 evolution in water as a photocatalyst.
They design a novel indium (III)-based MOF (In-MOF). Figure 23. shows that In-MOF may
be a potential photocatalyst for H2 evolution and can catalyze the decomposition of water into
H2 in the existence of Pt as a cocatalyst with a very trace amount. As shown in Figure 23a, the
absorption bands of In-MOF and the free H4L organic ligand are basically coincident, except
that In-MOF possesses a stronger absorption in the visible-light range of 420–570 nm because
of the low splitting energy of the d10 metal center. The CB potential of In-MOF is estimated to

38
be −0.20 eV. Considering that the band gap of 2.95 eV has been estimated by UV–vis DRS, its
valence band (VB) is calculated as 2.75 eV (Figure 23b). Its H2 evolution efficiency is up to
38.88 μmol h–1 (777.65 μmol g–1 h–1) for 4 h Figure 23c. The corresponding apparent quantum
yield is 1.63% (λex = 365 nm) shown in Figure 23d.

Figure 23. (a) Solid-state UV–vis absorption spectra of In-MOF and ligand H4L. (b) Mott–
Schottky plots of In-MOF. Inset: Energy diagram of the CB and VB levels. (c) Photocatalytic
H2 production of five samples. (d) PL emission spectra of In-MOF and the H4L ligand (λex =
320 nm) [201]

In particular, the excellent stability of In-MOF provides more opportunities for practical
application. Furthermore, Xu et al. [297] reported a key process of fabricating CdS-decorated
MOF composites, namely, CdS/UiO-66, which exhibit high H2 production activity from
photocatalytic water splitting, far surpassing the MOF and CdS counterparts, under visible light
irradiation. The close contact between CdS and UiO-66 would be favorable to the formation of
heterojunctions between them and facilitate their charge transfer and separation. The SEM and
TEM observations of CdS/UiO-66(10) indicate that the MOF is of great importance in the
formation of small CdS NPs, which present a good dispersion on the external surface of UiO-
66 octahedral crystals as shown in Figure 24. Interestingly, the photocatalytic activity of
CdS/UiO-66 composites decreases along with the increase of CdS content, where the CdS/UiO-
66(10) shows the best H2 production activity (1725 μmol g–1 h–1). This H2 production rate is
∼8.4 times higher than that of CdS. Moreover, the photocatalytic activity of CdS/UiO-66(10)
is higher than most of the reported CdS-based composites with other porous materials. This
report on charge-separation dynamics for CdS–MOF composites affords significant insights
the for future fabrication of advanced composite photocatalysts.

39
Figure24. SEM images of (a) CdS and (b) UiO-66. (c) SEM and (d) TEM images of CdS/UiO-
66(10 nm) [297].
Except than transition based MOFs, many other materials are using as a supporting material
incorporated with MOFs such as derived and composite substances which can increase the
catalytic activity drastically which are explained in later sections.

2.2 Metal organic framework composites as photocatalyst

Metal oxides are crystalline solids containing both a metal cation and an oxide anion.
In most cases, they react with water to generate bases or with acids to form salts. In addition to
it, the impregnation of MOs with MOFs is of great interest because of their enhanced features
such as thermal, optical and tensile strength stability. A vast range of MO based MOFs) (MO@
MOFs) such as Bi2O3@HKUST-1[202], CrMIL-101/TiO2 [203], Fe2O3/MIL-53(Fe) [204],
ZIF-8/Zn2GeO4[205], UiO-66/ HPW [206] etc.) has been reported to date having a wide range
of applications; however due to a huge number of publications on the synthesis and application
of (MO@ MOFs) it is quite difficult to cover the overall topic completely. In-situ preparation
of MO@MOF composites from pre-synthesized MOs has been explored by various synthesis
techniques such as hydrothermal, Sonochemical, electrochemical and mechano-chemical
methods [207], [208]. Furthermore, ZIFs have gained the most attention among the several
MOF materials because of their exceptional thermal stability and water tolerance. For example,
Wang et al. [209] developed yellow fluorescent Glutathione-Au nanoclusters/zeolitic
imidazolate framework-8 nanoparticles (GSH-Au NCs/ZIF-8 NPS) and their Au/ZnO NPs,
which demonstrated visible light photocatalytic hydrogen evolution as well as Rhodamine B
degradation (RhB). This is illustrated in Figure 25a. which briefly involves (1) the attachment

40
of the as-prepared NCs incorporating luminous GSH-Au on ZIF-8 to form GSH-Au NCs/ZIF-
8 NPs; (2) the fabrication of Au/ZnO NPs by the calcination of the GSH-Au NCs/ZIF-8 NPs;
(3) the photocatalytic hydrogen production as well as degradation of RhB using the resulting
Au/ZnO NPs. Other MOF-derived metal oxides have also been investigated for photocatalytic
applications. Moreover, Zhang et al. [210] offered NH2-MIL-125(Ti) as a precursor for the
preparation of hierarchical TiO2 with exposed active sites and a substantial particular surface
area, as shown in Figure 25b

a)

b)

Figure 25: a) The synthetic approach for the controllable construction of rhombic dodecahedral
GSH-Au NCs/ZIF-8 NPs and their derived Au/ZnO NPs for visible light photocatalytic
hydrogen generation Reprinted with permission from [209]. Copyright 2014 Elsevier. b)
Direct pyrolysis of Ti-containing MOFs using NH2-MIL-125(Ti) precursors and photo-
reduction resulted in a succession of Pd-adorned hierarchical TiO2. Reprinted with permission
from [210]. Copyright 2017 Elsevier.

41
Similarly, Dong's group., [211] used the new gel-to-crystal approach in 2015 to fabricate a
ternary photocatalytic material, CdS@Cd(II)-MOF@TiO2, that exhibited excellent
photocatalytic ability for H2 production from water. Also, Jin et al. [212] published a unique
hybrid material in 2017 for accelerating H2 synthesis from water splitting, named MoS2
quantum dots (MoS2 QDs) and graphene, which were placed onto NH2-UiO-66 to generate
MoS2 QDs/NH2-UiO-66/G. The photocatalytic activity of H2 evolution reach 186.37 μmol over
the EY-sensitized 5 wt% MoS2 QDs/UiO-66-NH2/G irradiated under visible light irradiation
(λ ≥420 nm) in the first 3 hours, and the apparent quantum efficiency (AQE) is 40.5% at 430
nm. The synergistic effect between MoS2 QDs and graphene together with UiO-66-NH2 is
corroborated by photo-luminescence spectra, electro-chemical and photo-electro-chemical
experiments. It demonstrates that the charge separation and the electrons transfer are more
efficient with the aid of the MoS2 QDs and graphene. MoS2 QDs might be a potential
photocatalyst for design new type of catalysts in photocatalysis proton reduction. The TEM
images of the UiO-66-NH2/G nanocomposite revealed that a large amount of rhombus-like
nanoparticles of UiO-66-NH2 grown on the surface of graphene sheets in Fig. 26A.
Fig.26 B and C clearly indicate that the increasing introduction of MoS2 QDs led to the
agglomeration of MoS2 QDs on the surface of UiO-66-NH2/G in the composite of MoS2
QDs/UiO-66-NH2/G. The energy dispersive X-ray (EDX) results further confirm the co-
existence of C, N, Zr, Mo and S elements in the area of red square of the MoS2 QDs/UiO-66-
NH2/G nanohybrid (Fig. 26D), which was supported by the XPS results of MoS2 QDs/UiO-66-
NH2/G. Moreover, the stability test of 5wt% MoS2 QDs/UiO-66-NH2/G was investigated over
four runs of totally 720 min, and the results were shown in Fig. 26F. It was observed that in the
first run, the maximum amounts of photocatalytic H2 evolution was 186.37 μmol. But in the
second run, the amounts of photocatalytic H2 evolution sharply declined. The H2 evolution
activity of MoS2 QDs/UiO-66-NH2/G could be revived by the concurrent addition of EY and
TEOA in the third run. The reason is that the stability of the dye sensitizer is not good because
of the adsorbed EY dye is easy to be desorbed and diffuse into the solution, and it could be
decomposed under light irradiation. In 2013, Zhang and co-coworkers., [213] fabricated a
novel multifunctional Fe3O4@MIL100(Fe) core-shell nanospheres via a versatile layer-by-
layer strategy. In a typical procedure, the mercaptoacetic acid (MAA)-functionalized Fe3O4
NPs were firstly synthesized by a simple stirring process. The product was fully dispersed in
FeCl3 ethanol solution and subsequently in benzene tricarboxylic acid (H3BTC) ethanol
solution under ultrasonic conditions, and Fe3O4@MIL-100(Fe) can be finally obtained after a
given number of cycles of the above two procedures. The as-fabricated Fe3O4@MIL-100(Fe)
exhibited remarkable photocatalytic activity for MB degradation, which is higher than some
typical photocatalysts, such as TiO2 and C3N4.

42
(E)

(F)

Figure 26. TEM images of (A) UiO-66-NH2/G and (B-C) 5wt% MoS2 QDs/UiO-66-NH2/G
composites; (D) EDX spectrum of 5wt% MoS2 QDs/UiO-66-NH2/G in which the Cu signals
originated from the Cu grid support for the TEM observation. (E) Photocatalytic mechanism
for H2 production over EY sensitized MoS2 QDs/UiO-66-NH2/G under visible light irradiation.
(F) Stability test of EY sensitized 5wt% MoS2 QDs/UiO-66-NH2/G for H2 generation under
visible light irradiation in TEOA-H2O solution (pH 7). The reaction was continued for 720 min,
with evacuation every 120 min. (I) First run; (II) evacuation; (III) add EY and TEOA, repeat
II; (IV) repeat III Reprinted with permission from [212]. Copyright 2017 Elsevier.

Besides several titanates-based metal oxides such as sodium titanates (Na2Ti3O7, Na2Ti6O13),
potassium titanates (K2Ti2O5, K2Ti4O9, K2Ti6O13), rubidium titanates (Rb2Ti6O13), caesium
titanates (Cs2Ti2O5, Cs2Ti5O11, Cs2Ti6O13), barium titanates (BaTi4O9), calcium titanates
(CaTiO3), lanthanum titanates and Ti-based MOFs such as MIL-125(Ti) with a chemical
formula of Ti8O8(OH)4(O2C-C6H4-CO2)6 have also been investigated for photocatalytic
reactions [214], [215], [216], [217], [218]. In addition to this Maggard's group., [219] has
reported several Cu1-based semiconductor materials for use in heterogeneous catalysis solar
water splitting processes. These materials include CuNbO3, CuNb3O8, CuNb13O33, Cu5Ta11O30,
and Cu3Ta7O19. In addition to it, Pan and co-workers [220] designed a heterostructure by in
situ growth of ZIF-9 on resynthesized amine functionalized ZnO nanorods. The MOF
encapsulation of ZnO was confirmed morphologically from SEM images of ZnO and the
composite. It is quite evident from UV-vis DRS and band gap analysis that the introduction of
ZIF-9 significantly improved the band structure of ZnO, hence enhancing its catalytic output.
In addition to these, Jin et al. [221] prepared CuO/HKUST-1 via a one-pot solvent-free route
from nano-CuO precursors in which CuO and H3BTC linkers were ground together followed
by heat treatment at 180°C leading to a hybrid where CuO NPs were trapped in the MOF
matrix. Metal oxide MOF (MO@MOF) based photocatalysts are a revolutionary field of
research towards photocatalysis. The morphological and flexible advantages of MOF structures
prompt the photocatalytic characters of MOFs. In addition to it, metal oxides also have a
superior capability towards their desired applications; however, the only shortcomings
associated with them are their low surface area and fast charge recombination. So, MO@ MOFs
are very effective at avoiding each other’s limitations [222].

43
Table 1: Different MOF based photocatalyst using in different fields.

Sr Photocatalyst Co catalyst Synthesis Light Sacrifici Recyclin Activit Key findings Ref
no /template method source al agent g time y µmol
. g-1 h-1

1 NH2-MIL- NiPd Solvotherma Xe lamp TEOA 3 times 8700 Depositing NiPd [558]
125(Ti)@graphitic carbon l method. (300 W) cocatalysts onto NH2-
nitride MIL-125(Ti)/CN
surface forms a
heterostructure
composite that’s why
NH2-MIL-
125(Ti)/CN/Ni15.8Pd2.1
heterostructure
exhibited the highest
photocatalytic activity
after sensitizing with
EY
2 UNiMOF(2D)@g-C3N4 - Solvo Xe lamp TEOA 3times 20.03 Constructing 2D-2D [559]
nanosheets thermal (300 W, heterojunctions
method 420 nm) provide a feasible
approach to obtaining
highly capable
catalysts for the
photocatalytic
decomposition of
water.
3 MIL-101(Cr) CdS Pt and PdS Photo Xe lamp Lactic 4 times ≈7500 Chalcogenide [224]
deposited (300 W) acid nanomaterials are
method and appealing because of
solvothermal suitable band edge
method positions for reducing
H2O while utilizing

44
visible the region of
sunlight such as CdS
which can be excited
by absorbing photons
of light having energy
exceeding its band gap
and then transferring
electrons directly to the
Pt to produce H2.
4 MIL-101(Cr) Au, CdS MoS2 or Xe arc Na2S and 4 times 25000 The Au@CdS/MIL- [560]
WS2 Hydrotherma lamp Na2SO3 101 heterostructure
l method (300W) implements an
excellent H2 production
and double rate higher than pure
solvents Cd45s due to strong
method surface Plasmon
resonance absorption
of Au, which lead to an
accelerated electron
separation and transfer
5 MIL-101(Cr) No One-step Xe lamp Lactic 4 times 488 This work provides a [333]
CDs, CdS cocatalyst double (300W) acid new strategy of a one-
solvents step double solvents
method method to co-
incorporate two
functional species to
the pores of MOFs to
improve the
photocatalytic H2
evolution activity of
host photocatalysts.
6 UiO-66 CdS, MOS2 Noble- Solvotherma Xe lamp Lactic 3 times 32500 The MoS2/U6–CdS [371]
metal-free l method (300 W) acid composites formed
MoS2 with tight spatial
integration have
remarkably hindered

45
the aggregation of CdS
nanoparticles because
of increased surface
area, reduced photo-
excited charge carriers
and abundant reactive
sites for H2 evolution.

7 NH2-MIL-125(Ti) Cu2O Cu2O Solvotherma 1.5 AM TEOA 5 times 11055. Cu2O cocatalyst [542]
l method. solar 5 encapsulation on NH2-
simulator MIL-125(Ti) MOF
(300W promoted Ti3+ ion
SS-EM, production and
Photo considerably boosted
emission electron-hole pair
tech, separation at the
3+
USA) junction of Ti sites on
the NH2-MIL-125(Ti)
MOF which increase
the activity.
8 TiO2@ZIF-8 - Sono Xenon Methanol 5 times 254.2 The morphology of [543]
Crystallizati lamp TiO2@ZIF-8 was a
on (300 W) double shell with high
specific surface area
and provide a good
stability and
recyclability.
9 UiO-66 CdS NPs Pt(0.5%) Solvotherma Xenon Ascorbic 235 The enhanced [561]
l method lamp acid photocatalytic H2
(300 W, evolution activity could
> 420 be caused by the
nm) effective interfacial
charge transfer
between CdS and UiO-

46
66 MOFs across the
heterojunction
10 MIL-101(Cr) CdS NPs No noble Three-step Xe lamp Na2S and 1900 It has been established [645]
cocatalyst solvothermal (300 W) NaSO3 that the combination of
is applied method TiO2 and CdS will help
surmount their weak
points by increasing
visible-light response
and facilitating
electron-hole
separation.
11 UiO-66-NH2 @ Cd0.2 - Solvotherma Xenon Na2S and 4 times 5846.5 MOF-based composite [564]
Zn0.8S NPs l method lamp NaSO3 materials hold great
(300 W) promise in the field of
energy conversion and
environmental
purification due to
Cd0.2Zn0.8S@UiO-66-
NH2 photocatalysts that
show excellent stability
during photocatalytic
hydrogen evolution and
CO2 reduction
12 UiO-66-NH2@Graphene Pt Visible TEOA 4 times 41400 The prepared graphene [646]
(50 wt%) Hydrotherma light was well-wrapped
l method UiO-66-
NH2 octahedrons, in
which every face was
in contact with
graphene, allowing
them to have the
highest H2 productivity
13 UiO-66@ graphene, Ni4S3 Ni4S3 Hydrotherma Perfect TEOA ≈2800 The hydrogen [647]
20wt% l method light production amount
over EY-sensitized

47
(PCX50 rGO/MOF/Ni4S3
A) photocatalyst is high
due to the rapid
interfacial charge
transfer and interfacial
photocatalytic reaction
of rGO nanosheets
while Ni4S3 provided
more effective active
sites
14 Cu-BTC ZnO Hydrotherma Mercury Methanol 191 [648]
Electrostatic interaction l method lamp The morphology of the
assembly of ZnO/GO and (500W) as-prepared ZnO/GO
Cu-BTC and (Cu-BTC)/GO, and
the structure of the
electrostatic interaction
assembly of ZnO/GO
and Cu-BTC
demonstrates that it
could stabilize and
extend the lifetime of
free radicals, thus
improving the
photocatalytic H2
evolution efficiency

15 UiO-66 CdS, RGO Pt Photo Visible Na2S and 138 CdS's H2 production [187]
deposition light NaSO3 performance was
method enhanced with the
antioxidant activities of
UiO-66 and RGO due
to increased catalytic
sites and reaction
centres and reduced

48
charge carrier
recombination.
16 UiO-66-NH2 MoS2 Solvotherma Xe lamp TEOA 4 186.37 The UiO-66-NH2/G [212]
quantum l method (300 W) possessed advanced
dots specific surface area,
(MoS2QDs/ which supported MoS2
graphene QDs formation
synergistic catalytic
effect and depicts the
excellent
photocatalytic activity
and stability for
hydrogen production.

17 Ni4P2@MOF-1 Ni Solvotherma Visible MeOH 4400 The loadings [656]


l method of Ni4P2 in Ni4P2@M
OF-
1 and Ni4P2@MOF-2
shows high catalytic
activity due to stability,
facile electron transfer
property and porosity.
18 MIL-125(Ti) (Ti3+ assisted Au, Pd, and Solvotherma Sunlight TEOA 154.72 MIL-125(Ti) has high [649]
method) Pt l method stability because it
comprises cyclic
octamers of TiO2
octahedral. This
synthetic method can
be used for various
porous and nonporous
materials.
19 UiO-66 Ni2P Hydrotherma LED TEOA 4 2000 The Ni2P nanoparticles [568]
l method white modified on the g-
light (5 C3N4 /UiO-66 display
W) the more excellent

49
active sites and
improve the
productivity of photo-
generated charge
separation, and the H2
evolution by providing
active sites.
20 NH2-MIL-125(Ti) Ni2P NPs Hydrotherma Xe lamp TEA 7 894 The H2 evolution rate [650]
l method (300W) increased because the
cocatalyst efficiently
drew electrons,
preventing unwanted
charge carrier
recombination.
Further, it is a low-cost,
stable and highly
effective photocatalytic
system
21 UiO-66-NH2 NiO Single-step Xe 3 2561.3 The elevated electrons [651]
nanoparticl sol-gel lamp(300 TEOA 2 of eosin Y and UiO-66-
es process W) NH2, the rapid
separation by the
feedforward band
structure, the more
negative CB of NiO,
and the excellent
dispersion on the
octahedral skeleton
depicts that they are
used as a cocatalyst for
UiO-66-NH2 -based
visible-light
photocatalytic H2
evolution.

50
22 UCNPs-Pt@MOF/ Au Pt NPs, Au Hydrotherma Laser TEOA 4 280 Exposure to UV light, [652]
NPs l method light (980 the MOF is excited and
nm) the electrons would be
transferred to Pt, which
behave as electron
reservoirs and active
sites for subsequent
proton reduction.
23 Cd-TBAPy Pt NPs, Hydrotherma Xenon TEOA 86 A cadmium-based [198]
CoPi l method lamp MOF single crystal
(300W) with a 2D layered
framework have
inherent band edge
positions, the loading
of suitable cocatalysts
and the π-conjugated
2D layered structure of
Cd-TBAPy integrally
contribute to the
extraordinary
photocatalytic
performance of Cd-
TBAPy.
24 MIL-101(Cr) Erythrosine Hydrotherma Xe lamp TEOA 6250 The Ni/NiOx particles [653]
B l method (300W) were photo deposited
in situ on MIL-101
MOFs as a catalyst
with cheap cost, ease of
operation, and good
stability for water
splitting.
25 UiO-66 Pt Hydrotherma Xenon L- 3 460 Owing to the [654]
l method lamp ascorbic advantages of low cost,
(300W) acid easy operation, and
abundance in nature,
the dye-sensitized

51
MOFs show excellent
potential for
photocatalytic H2
production.
26 UiO-66-N H2 Pt Hydrotherma Xe lamp Methanol 3 1528 Calix-3 was selected [655]
l method (300W) for durable
photocatalytic H2
production because
cone conformation is
beneficial to impede
dye aggregation and
has four −COOH
groups in one
molecule, and thus it is
expected that a strong
hydrogen bonding.

27 SMOF-1 Wells- Slow Solid Methanol 4 times 3353 When SMOF-1 is [547
Dawson- evaporation state light exposed to visible light ]
type POM (300 W, (500 nm), it allow for
500 nm) rapid multi-electron
injection from
photosensitive
[Ru(bpy)3]2+ units to
redox-active WD-POM
units, resulting in
efficient hydrogen
generation in aqueous
and organic mediums.
28 Pt One-pot Xe lamp TEOA 3 25578 Immobilizing
MIL-101(Cr) reaction (300 W) polyoxoanions into [663]
porous framework
materials improves
their ability to adsorb
cationic PSs.

52
29 Pt/UiO-66/CdS Pt Photo Xe lamp Na2S and 1700 Porous materials with [187]
deposition (300 W) Na2SO3 high specific surface
method areas have been
included to boost CdS
catalytic sites and
reaction centers Due to
enhanced catalytic sites
and reaction centers,
UiO-66 and RGO
improved CdS' H2
generation.
30 Pt/UiO-66/CdS/1%RGO Pt Photo Xe lamp Na2S and 2100 This ternary composite [187]
deposition (300 W) Na2SO3 showed advantage over
and binary hybrid UiO-
solvothermal 66/CdS and the
method hydrogen production
rate of UiO-
66/CdS/1%RGO was
higher than that of
UiO-66/CdS due to
interface charge
separation efficiency,
smaller radius of
curvature and more
efficient charge
separation.
31 Pt/UiO-66/g-C3N4 Pt Solvotherma Xe lamp Ascorbic 1141 The excellent stability [541]
l. method (300 W) acid in aqueous solution
enables UiO-66 MOF
to be a good candidate
for photocatalytic
hydrogen production.

32 NH2-MIL-101(Cr) Pt Hydrotherma Visible TEOA 5 times 600 Visible light-induced [359]


Cro cluster nanoparticl l method light photocatalytic H2
es production resulted

53
from the combined
promoting effects of
the Pt deeply
embedded MOF
catalysts and the dye
solution.

33 HNTM-Ir/Pt Porphyrin Solvotherma Visible TEOA 200 HNTM-Ir/Pt has [362]


Zro cluster l method light remarkable
photocatalytic
efficiency for H2
evolution due to its
hollow structure and
organic porphyrin
linkers, resulting in
generating electron-
hole pairs.

34 PMOF Pt Xe lamp Ascorbic 3 times 85200 The long-term course [360]


Tio cluster (300 W) acid reaction confirms the
Solvo excellent water
thermal resistance and photo-
stability of PMOF in
method photocatalytic
hydrogen evolution. In
LMCT, donor-to-
acceptor facilitated
electron transport from
porphyrin-based
ligands to Ti-oxo
clusters.

35 Cd-Metal organic pt Xe lamp TEOA 5 times 1200 Both Cd-MOFs and [199]
framework Cd-O cluster (300 W) Cu- MOFs had the
same VB energy;
however, the CB level

54
Hydrotherma of Cu-Metal organic
l framework (MOFs)
was more negative than
method that of Cd-(MOFs.
Metal ions were
incorporated into these
MOF materials to
change their CB energy
and MLCT character.
36 Cu-MOFs pt Xe lamp TEOA 8 times 7300 Cd-MOFs and Cu- [199]
Cu-O cluster Solvo (300 W) MOFs had the same
thermal valence band (VB)
energy, the conduction
method band (CB) level of Cu-
MOFs was more
negative than that of
Cd-MOFs. Therefore,
the reduction ability of
Cu-MOFs was greater
than that of Cd-MOFs.
37 Ti-MOF-NH2 pt Photo Visible- TEOA 3 times ≈400 [556]
deposition light The results obtained
method from several control
reactions revealed that
the light irradiation, as
well as the presence of
Pt(1.5)/Ti-MOF-NH2
and TEOA as a
sacrificial electron
donor, are essential for
the progression of the
reaction

38 {[CuICuII2 H2PtCl6 (1 Hydrotherma Methanol 320 Cu I/II serves as the [660]


(DCTP)2]NO3.1.5DMF}n mL, 0.1 l Xe lamp active point to provide
wt%) (200 W , holes and have superior

55
320 nm - band alignment. The
780 nm) results revealed that 1
is an efficient catalyst
in the photocatalytic
hydrogen evolution
reaction without adding
photosensitizers.
39 Hydrotherma MeOH 3 The as-prepared MOF
[Cu2I2(BPEA)](DMF)4 l Xe lamp 2 108 demonstrates high [661]
(500 W) 055 chemical stability,
which could be stable
in pH 2−13. This Cu(I)-
based MOF exhibits
high reactivity for the
reduction of Cr(VI) in
water, with 95% Cr(VI)
converting to Cr(III) in
10 min by using MeOH
as a scavenger under
visible light irradiation.
40 ZIF-67 Solution- TEOA 5 ZIF-67 is the key
based LEDs 40500 parameter that controls [662]
method. (450 nm) HER efficiency
because it as important
catalytic material
having guide light
harvesting, charge
separation, and HER
kinetics during the
photoinduced catalytic
reaction.
41 Pt@UiO-66(Zr) Pt Solvotherma Xe lamp TEOA UiO-66(Zr) was [188]
l method (300 W) 3900 considered as a
photocatalyst for
hydrogen production
due to its intrinsic

56
photoactivity5 and
good thermal stability.
42 MIL-125/Au Pt Solvo Xe lamp TEOA 3 1743 Separation of Pt and Au
thermal (300 W) in MIL-125, where Pt [288]
method/pot NPs are incorporated
synthesis uniformly into the
method MOF and Au NRs are
assembled on the MOF
external surface –
avoiding the light
shielding effect,
effectively guiding the
electron migration
direction and
promoting the rapid
charge transfer.
43 EY/Cu-RSH Porphyrin Solvo Xe lamp TEOA 7880 The remarkable[][160]
thermal (300 W , l stability of Cu-RSH is
reaction > 420 very rare for CPs based
nm) on divalent metal
cations, enabling them
to serve as an acid- and
base-tolerant catalyst,
such as in water-
splitting reactions.
44 [Al3(OH)3 (HTCS)2] Hydrotherma Xe lamp TEOA 50 Al-MOFs show [671]
l (300 W) exceptional water and
thermal stability due to
the low atomic number
of aluminum and
morphology.
45 MIL-125-NH2 -57 Mo3S132- Hydrotherma Xe lamp TEA Can be 2094 The high performance [664]
l (300W) recycled of Mo3S132- can be
attributed to the good
contact between these

57
clusters and the MOF
and the large number of
catalytically active
sites, whilst the high
activity of 1T-MoS2
NPs is due to their high
electrical conductivity
leading to fast electron
transfer processes.
46 EY/MoS2-MOF/ Hydrotherma Xe lamp TEOA 9002.6 To increase H2 [665]
Co3O4 l synthesis. (300 W) production, dye-
sensitizing materials
such as MoS2 with
good electrical
characteristics, MOF
with a large specific
surface, and Co3O4
with an appropriate
band gap for visible
light sensitivity are
utilized. The MoS2-
MOF/ Co3O4 catalyst
has a high fluorescence
emission rate and an
increased electron
lifetime.
47 NH2-UiO-66/g-C3N4 Pt Solvotherma Xe lamp Sodium 4 2930 Carbon nanodots (CDs) [642]
CDs l method (300 W) ascorbate have attracted
enormous attention in
photocatalytic area for
their high light
harvesting and
outstanding electron
transfer ability.

58
48 Pt/ZIF-8/g-C3N4 Pt Solvo Xe lamp TEOA 309.5 The structural merits of [643]
deposition (300 W) the ZIF-8/g- C3N4
composite lead to its
optimized physical and
chemical properties,
visible light absorption
and utilization
becoming more
effective by integrating
g- C3N4 with ZIF-8.
49 ZIF-67 Hydrotherma LED TEOA 4000 The MoS2 acted as a [644]
l method white charge transmission on
light (5 the surface of g- C3N4
W) /MOF, providing more
active sites and altering
the charge transmission
route. ZIF-67 provided
a large number of
active sites and
surfaces, which gave
places for the
adsorption of EY dye.
50 PNP MOF Pt Hydrotherma Xe lamp Ascorbic 2 TON The POM has four [670]
l (150 W , acid =66 targeted functions in
> 400 this hybrid material: it
nm) reduces H2PtCl6 to Pt
NPs, stabilizes the Pt
NPs, induces a strong
electrostatic
association of the
negatively charged Pt
NPs with the
protonated NH2-MIL-
53 sites on the particle
surfaces, and facilitates

59
the catalytic reduction
reaction itself.
51 MIL-125-NH2@ Post- Xe lamp TEOA 3 496 The MIL-125- [562]
TiO2 solvothermal (300 W) NH2@TiO2 coreshell
method particles combine the
advantages of highly
active TiO2 nanosheets,
MIL-125-NH2
photosensitizer, plenty
of linker defects and
oxygen vacancies and
mesoporous structure,
which have been
utilized as
photocatalysts for the
visible-light-driven
hydrogen production
reaction.
52 MIL-101(Cr)–Au/TiO2 TiO2 Hydrotherma Halogen Methanol 3 903 The crystalline anatase [563]
l lamp shell is created utilizing
(400 W) titanium(IV)
isopropoxide as a
precursor. MIL-101
(Cr) is used as the core
material, which
controls the size of the
core-shell complex and
ensures overall
stability. Furthermore,
its (good) reversible
big molecule sorption
behaviour enables
material production.
53 Pt/CPM-2-NH2 Pt 300 W Methanol 2000 The introduction of [669]
Xe lamp three-rings could allow

60
access to new zeolite-
like MOFs and MOPs.
These new materials
possess structural and
topological features
(such as three-rings,
large 12-ring channels,
and expanded MOP
cages) that are the
direct results of the
synthetic strategy.
54 ZnIn2S4@NH2-MIL-125(Ti) Pt One-step Xe lamp Na2S and 5 2204.2 The enhanced [565]
solvothermal (300 W) Na2SO3 photocatalytic activity
method. of is attributed to the
well-matched band
structure and intimate
contact interfaces
between ZnIn2S4 and
NH2-MIL-125(Ti),
which led to the
effective transfer and
separation of the
photogenerated charge
carriers.
55 Pt/CdS@NU-1000/RGO Pt Hydrotherma Xe lamp Na2S and 5 240 The hydrogen [566]
l method (300 W) Na2SO3 generation
performance of CdS
enhanced with the
combined action of
NU-1000 and RGO,
i.e., ternary composite,
due to increased
catalytic sites and
reaction centres by
avoiding
agglomeration of CdS

61
nanoparticles due to
NU-1000's large pore
size and minimal
charge carrier
recombination.
56 g-C3N4/MIL-53(Fe) Simple Xe lamp TEOA 905.4 Fabrication of g‐ C3N4 [567]
grinding (300 W) /MOF heterojunction
method might be a good
strategy for gaining the
desired photocatalyst
with superior
photocatalytic
performance due to an
intimate interfacial
contact, more active
sites and well‐ matched
energy level.
57 Al-TCPP-Pt Pt Hydrotherma Xe lamp TEOA 4 181 A highly stable [668]
nanoparticl l method (300 W) aluminum-based
e porphyrinic MOF was
employed as single Pt
atoms exhibit a superb
activity, giving a
turnover
frequency during
catalytic reaction.
58 Cd0.5Zn0.5S@UiO-66@g- Pt In situ Xe lamp Na2S and 3 1281.1 Cd0.5Zn0.5S@UCN [567]
C3N4 hydrotherma (300 W) Na2SO3 shows significantly
l method effective photocatalytic
activity for hydrogen
production and MO
degradation due to its
large surface area, the
enhanced visible-light
absorption ability and
high-efficiency

62
separation of
photogenerated
electron-hole pairs.
59 FI/[Ni2(PymS)4]n Pt Solvotherma White TEA 4 TOF [667]
l LED or =10.6 Microscale crystals of
direct h-1 a 2D layered MOF
sunlight shows high TOF due
to the heterogeneous
nature and
recyclability of the
catalyst which has
been also proposed,
based on fluorescence
spectra and cyclic
voltammetry.

60 PCN-9-Pt Pt Solvotherma 300 W TEOA 33.5 Platinum complex [666]


l method Xe lamp plays an important role
in improving the
separation efficiency of
photogenerated charge
carriers.
61 Al-ATA-Ni Ni Solvotherma Visible Methanol 1200 This MOF can be [582]
l method. modified by
incorporating Ni2+
cations into the pores
through coordination
with the amino groups,
and the resulting MOF
is an efficient
photocatalyst for
overall water splitting.
62 Pt/[Zn2(bmimb)(HPO4)2].2 Pt Solvo Xe lamp Na2S and 753 The compound exhibits [583]
H2O thermal (300W) Na2SO3 an unprecedented 3D
method 5-nodal 3,3,3,4,4-

63
connected
supramolecular net.
The topological
structure and
luminescent property
of Zn based MOF
shows good
performance.

64
In the realm of solar energy conversion, metal oxide-based semiconductors have garnered
much attention since it was reported that TiO2 may be used for photoelectrochemical water
splitting. Using metal oxide semiconductors has led to significant breakthroughs in
photocatalytic water-splitting reactions. So far, the best STH efficiency achieved on metal
oxide photocatalysts is only around 1%. This is a deficient number compared to what is needed
in real life. Thus, much work must be done before photocatalytic water splitting for solar energy
conversion may be used on a big scale [223].

2.3 MOFs-derived photocatalysts

Materials that are generated from MOFs have advantages such as hereditary shape,
controlled dimensions, and high crystallinity; their energy gap position could be controlled by
using external doping, and their programmable surface functionality has made them an
attractive choice for applications involving photocatalysts [225]. Due to their complex
structure, huge pores, and remarkable stability, MOFs are widely used in catalytic reactions;
MOF-calcined composites are more stable and can manufacture hydrogen in water [141]. A
wide variety of derivate MOFs can be utilized successfully in photocatalytic water splitting.
Jiang et al [226] has chosen thermally stable MOFs as hard porous templates to synthesize
hierarchically porous metal sulfides; porous nanostructured materials are demonstrated to be
very promising in catalysis due to their well accessible active sites. Significantly, as a
representative, the hierarchically porous CdS obtained by using MIL-53(Al) template exhibits
excellent photocatalytic performance that is superior to the bulk and nanosized CdS
counterparts toward hydrogen production by water splitting, manifesting the advantages of the
porous nanostructure in photocatalysis. Upon MOF template removal, the replica of porous
metal oxides/sulfides can be successfully obtained. The photocatalytic hydrogen production
was conducted in Na2S/Na2SO3 aqueous solution under visible light irradiation. Compared to
the poor activity (115.4 µmol g−1 h−1) of bulk CdS, nano CdS exhibits a relatively higher
photocatalytic efficiency (350.1 µmol g−1 h−1). Remarkably, the HP-CdS exhibits a
significantly enhanced activity (634.0 µmol g−1 h−1), which is ≈2 and 5 times higher than that
of nano CdS and bulk CdS, respectively. The superior photocatalytic efficiency of HP-CdS
should be ascribed to the nanosized effect and porous structure. Following that, Yong and their
companions [227] synthesized MOF-derived yolk-shell CdS microcubes (Cd-Fe PBA micro
cubes) with enhanced visible-light photocatalytic activity and stability for hydrogen evolution
through a facile microwave-assisted hydrothermal process. Benefitting from structural merits
including 3D open structure, the small size of primary nanoparticles, high specific surface area,
and good structural robustness, the obtained yolk-shell CdS microcubes manifest excellent
performances for photocatalytic hydrogen evolution from H2O under visible-light irradiation.
The photocatalytic H2 evolution rate is 3051.4 µmol h−1 g−1 (with an apparent quantum
efficiency of 4.9 % at 420 nm), which is ~2.43 times higher than that of conventional CdS
nanoparticles. Furthermore, the yolk-shell CdS microcubes exhibit remarkable photocatalytic
stability. Zhang's group also published a paper on CdS/ZCO composites for photolytic water
to generate hydrogen [228]. The ZnCo-ZIF was synthesized using the solvothermal process
suggested in Figure 27e and then calcined at various temperatures to create ZnxCo3-xO4 (ZCO).
Figure 27a shows that there is no hydrogen generation with bare ZCO and Co3O4, indicating
the zeolites materials of ZCO or Co3O4 are not directly excited by the visible light. Although
the CdS with lactic acid as sacrifice agent can generate the hydrogen, the evolution rate of H2
is still very low (about 1000 µmol/g h), possibly due to lack of active sites and the fast carrier
recombination. It can be noted that there is no hydrogen generation during the first hour only
CdS/Co3O4 without Zn atoms embedded with Co3O4 (Figure. 27b), and the phenomenon
indicates that the Co3O4 plays a role of the template agent and co-catalyst, which restrains the

65
hydrogen evolution performance of CdS on account of the altered band gap of CdS. When the
CDS content in CdS/ZCO composites reached 30% by weight, the H2 generation rate increased
to 3978.6 µmol g-1 h-1 under visible light irradiation as shown in Figure 27c. Figure. 27d
exhibits the stability of 30wt% CdS/ZCO for H2 evolution. During a long time, irradiation
under the visible light, it can be observed that there is no significant decreasing for H2 evolution
after four runs, revealing the good stability and durability. The results predicted that
incorporating Zn atoms into ZCO boosted photocatalytic activity and the porous structure and
larger surface area of CdS/ZCO composites are favorable for a wide range of absorption and
the enhanced photocatalytic activity.

66
(e)

67
Figure27. (a) The H2 evolution rates of ZCO, Co3O4 and CdS, (b) The H2 evolution rates 6 of
Co3O4 loaded with different CdS, (c) The percentages effect of CdS on the H2 generation of
ZCO, (d) The cycling runs of 30 wt% CdS/ZCO for photocatalytic H2 8 evolution at room
temperature under visible light (λ > 420 nm). (e) Schematic diagram of the synthesis of
CdS/ZCO nanocomposite. Reprinted with permission from [228]. Copyright 2018 Elsevier.

Similarly, Lan et al. [229] designed a Co-doped Zn1-xCdxS photocatalyst for the photocatalysis
of water; ZIF-8 and ZnCo ZIF serve as templates to allow for the interaction with Cd(NO3)2
and thiourea, which ultimately results in the production of ZnCdS and Co-ZnCdS, respectively.
Additionally, Kim and co-coworkers [230] reported Co4S3/CdS nanoparticles for
photocatalytic hydrogen production. Owing to its low cost and high efficiency, this
photocatalytic system should hold great potential for the development of highly efficient
photocatalytic materials for use in various fields. Further, Cheng and his mates [231] applied
ZIF-8 as a template to harvest N-doped graphene analogues (ZNG) for hydrogen generation.
The temperatures at which the ZNGs were calcined were varied, and consequently, three
uniform polyhedrons of ZNGs were produced. These NGOs have been designated as ZNG-
800, ZNG-900, and ZNG-1000. Among them, ZNG-1000 had the more extraordinary
photocatalytic performance because it had the most graphite nitrogen, which allowed it to have
a different charge transfer rate and speed up the hydrogen evolution reaction. Subsequently,
another nanoscale heterojunction, ZnO/ZnS, for photocatalytic H2 generation was described in
a report by Cheng et al. [232]. It is shown that by controlling the calcination temperature and
time with MOF-5 as a host and guest thioacetamide as a sulfur source, the chemical
compositions of the formed heterojunctions of ZnO/ZnS can be tuned to further enhance the
visible-light absorption and photocatalytic activity. The nanoscale heterojunction ZnO/ZnS
structural feature serves to reduce the average free path of charge carriers and improve the
charge separation efficiency, thus leading to significantly enhanced HER activity under visible-
light irradiation (λ > 420 nm) with high stability and recyclability without any cocatalyst. Due
to the unstable characteristics of MOF, which are sometimes referred to as MOF derivatives,
MOF can be utilized effectively as a precursor or template in synthesizing nanomaterials. The
created MOF derivative retains the original morphology of the parent MOF materials and
inherits many of the parent MOF's capabilities, making it a good candidate for a photocatalyst.
Because of this, over the past several years, there has been a substantial increase in the number
of studies conducted on the application of MOF derivatives in a variety of photocatalytic
applications [173]. Homometallic metal organic frameworks (MOFs), an emerging class of
porous crystalline materials with tunable porosity and structural controllability, chemical and
thermal stabilities, and recyclability, are proclaimed as potent semiconducting photocatalysts
for water splitting reactions. Despite several advantages, majority of the homometallic MOFs
reported for hydrogen production suffer from the limitations of their activity in UV light and
low catalytic efficiency. However, heterogeneity can be devised into these frameworks by
introducing either the different metallic building units or the organic linkers in which the
functionalities can be tuned according to the specific application. Mixed metal-metal organic
frameworks (MM-MOFs) are the class of MOFs that contain two or more metal ions anywhere
in the framework, and endow these with an added degree of structural diversity and a variety
of innovative applications [233, 234] . MM-MOFs have attracted a great consideration in recent
years for water splitting reactions because, (1) the introduction of two metal ions causes
synergistic interaction between the metal ions which helps in reducing the optical band gap for
the desired reaction, (2) introduction of photoactive or metal-containing organic linker makes
the frameworks visible light active and improves the properties compared to the homometallic
MOFs, (3) presence of two metal ions greatly helps in reducing the recombination of photo-
generated electrons and holes, as one of the metal ions behaves as an electron sink and helps

68
in transferring electrons to the protons for the generation of molecular hydrogen [235]. Lin et
al. [236]. reported a simple, inexpensive, tunable, and scalable MOFs-templated strategy for
the synthesis of a mixed metal oxide nanocomposite with useful photo physical properties. This
facile method produces crystalline octahedral nanoshells composed of hematite Fe2O3
nanoparticles embedded in anatase TiO2 with some Fe doping. This material has interesting
photophysical properties as it enables photocatalytic hydrogen production from water using
visible light, while neither component alone is able to do so. The versatile MOF-templated
nanocomposite synthesis procedure can be readily modified, by varying the type of MOF (both
the metal ions and coordinating moieties) and the coating material, to prepare new
nanocomposite materials with desirable synergistic properties [236]. Moreover, Zhang et al.
[237]; examine an in-situ calcination method to construct porous CoOx-carbon hybrids this
method has created various products using various thermal treatments of ZIF-67. In another
work, the MOF-derived method was employed to synthesize bimetallic Fe-Ni-P nanotubes
from MOF Fe-Ni-MIL-88 nanorods acting as precursor. The synthesized Fe-Ni-P nanotubes in
combination with various dyes are able to act as photocatalyst for efficient hydrogen
generation. The parent MOF material was rod like however, the phosphating led to the
preservation of the rod like structure in MOF Fe-Ni-MIL-88 with hollow structure. Upon
phosphating treatment, a shallow hallow morphology having a rough shell was formed and the
shallow hollow became completely hollow with the passage of time. This morphological
change can be attributed to the Kirkendall effect. In Kirkendall effect, the surface of MOF was
dominant by the phosphating reaction for the formation of thinned shell of Fe-Ni-P which was
later acted as a precursor for the deposition of substrate for the Fe-Ni-P nuclei. When a MOF
particle is completely changed into a single phosphide particle, a hollow interior is generated
because the Fe-Ni-P crystal has a larger density than its MOF template [238]. Development of
water-stable metal–organic frameworks (MOFs) for promising visible-light-driven
photocatalytic water splitting is highly desirable but still challenging. Here Liu et al [257]
report a novel p-type nickel-based MOF single crystal (Ni-TBAPy-SC) and its exfoliated
nanobelts (Ni-TBAPy-NB) that can bear a wide range of pH environment in aqueous solution.
Both experimental and theoretical results indicate a feasible electron transfer from the
H4TBAPy ligand (light-harvesting center) to the Ni–O cluster node (catalytic center), on which
water splitting to produce hydrogen can be efficiently driven free of cocatalyst. Compared to
the single crystal, the exfoliated two-dimensional (2D) nanobelts show more efficient charge
separation due to its shortened charge transfer distance and remarkably enhanced active surface
areas, resulting in 164 times of promoted water reduction activity. The optimal H2 evolution
rate on the nanobelt reaches 98 μmol h–1 (ca. 5 mmol h–1 g–1) showing benchmarked apparent
quantum efficiency (AQE) of 8.0% at 420 nm among water-stable MOFs photocatalysts.
These works reflected the rapid development of MOFs in photocatalytic water splitting.
However, the bottlenecks of overall water splitting and water-splitting half reaction for
practical application are still unrevealed. Additionally, the functionalization and decoration
strategies of MOFs for further improvement in photocatalytic water splitting application are
urgently required to be clarified. In the next sections the strategies for the improvement are
explained. Although there are many challenges in the production of hydrogen from MOF-based
photocatalysts, the rapid development of new MOF-based photocatalysts in recent years has
ensured that in future, MOF-based photocatalysts can replace the conventional inorganic
semiconductor photocatalysts for efficient conversion of solar energy into hydrogen energy,
and replace the harmful fossil fuels. Apart from hydrogen production, designing MOFs capable
of both hydrogen productions. Table 2 shows different others MOF derived photocatalyst.

69
Table 2: Different kinds of MOF derived photocatalyst.

Sr Photocatalyst Co catalyst Synthesis Light Sacrificial Recycling Activity Key findings Ref
no. /template method source agent time µmol
g-1 h-1

1 Co3O4@C/ Co3O4@C Carbonization UV-LED Methanol 4 times The


TiO2 and one-pot lamp (25 (20%) 11400 photocatalytic H2
ZIF-67 method W, 365 evolution was [551]
nm) significantly
increased by
adding a small
amount of
carbonized ZIF-
67 (Co3O4@C) to
TiO2, due to the
high separation
efficiency of the
photo-generated
electrons/holes.
Xe lamp No obvious loss
2 ZIF-8 derived Simple (300 W, Methanol 3 times of photocatalytic [552]
ZnO/rGO/carbon dipping- PLS- 14.6 performance was
sponge pyrolysis SXE300, observed after
method Beijing being used
Perfectlight repetitively for
Co. Ltd.) three times,
indicating that
the ZRCs has
good stability
and recyclability,
thus making such
materials
promising for

70
industrial
application
NH2-MIL-101(Fe)/
3 Ni(OH)2 CN/FeNiP Annealing Xe lamp TEOA 5 times 13810 The enhanced [553]
method (300W) photocatalytic
activity can be
attributed to the
efficient spatial
separation of
photo-induced
electrons from
excited EY and
g-C3N4 to
CN/Fe2P and
Ni2P owing to
their good
interface,
staggered band
energy, more
surface-active
sites presented on
CN/Fe2P and
promoted
H+ reduction
reactions on Ni-
Fe phosphides
4 N-doped graphene Pt Visible TEOA 4 times 370 Graphene, a 2D [231]
analogues, Hydrothermal light sp2 hybridized
ZIF-8 method carbon network
sonication possessing
unique
. monolayer
structure and
high electron
mobility which is

71
benefited in
energy
conversion and
storage and
ZNGs exhibit
significant
hydrogen
evolution rates
due to the highly
nitrogen-doped
graphene
structure on the
surface.
5 ZnO/Co3O4 Pt Xe lamp Methanol 5 times 7800 This work [267]
Template of MOF = Hydrothermal (300W ) provides a new
ZnCo-ZIF method approach that
uses bimetallic
MOF as a
template which
exhibit excellent
photocatalytic
activity due to
exposed catalytic
active sites,
suitable band
matching and
strong electron
coupling.
6 Au/ZnO Glutathione-Au Ultraviolet Na2S and 4 times 29.8 Au nanoclusters [209]
nanoclusters/ZIF- Two-step (UV) lamp NaSO3 (Au NCs) have
8 hydrothermal (365 nm) gained lot of
method. attention because
of their
distinctive
fluorescence

72
properties, small
size and high
catalytic activity.
In addition, the
resulting Au/ZnO
NPs can be used
for visible light
photocatalytic
hydrogen
production and
degradation of
RhB.
7 C-ZIF/g-C3N4 Pt Facile thermal Xe lamp TEOA 4 times 32.6 The carbon [657]
condensation (300W) times nanostructure
method higher derived from
than pure ZIF-8 formed an
graphene. intimate contact
with g-C3N4 and
had large surface
area, which could
function not only
as an electron
acceptor for the
separation of
photogenerated
charge carriers,
but also as an
effective
cocatalyst to
promote the
hydrogen
evolution
reaction.

73
8 HP-CdS, - Xe lamp Na2S, 4times 630 Porous [226]
Template of Nano casting (300 W) NaSO3 photocatalysts
MIL-53(Al)/CEO process are very
attractive as the
porous structures
facilitate the
exposure of more
active sites and
fast charge
separation, as
well as mass
transfer. The
porous structure
of CdS is
expected to
greatly improve
the charge
separation and
photocatalytic
performance.
9 Ni-BTC MOF Ni2P Visible Lactic 5 times 33480 Ni2P's capacity to [658]
Wet chemical light acid remove
method photoexcited
charge carriers
from CdS and
provide strong
electrical
conductivity
contributes to the
composite's
enhanced
activity.
Moreover, the
optimized
composite

74
material also
exhibited better
photocatalytic
activity than Pt
cocatalyzed
CdS.
10 NH2-MIL-125(Ti), Pd Pyrolysis Xe lamp Methanol 48985 The hierarchical [577]
Hierarchical TiO2 (300 W) TiO2 was
composed of
small TiO2
nanoparticles
porous structure.
The enhanced
photocatalytic
performance was
resulted from
efficient optical
absorption,
promoted
separation and
migration of
photo-induced
electron-hole
pairs.
11 NiS/Zn0.5Cd0.5S NiS Solvothermal Xe lamp Na2S and 5 16 780 NiS/Zn0.5Cd0.5S [571]
sulfidation and (300 W) Na2SO3 possesses an
thermal extremely high
annealing stability and
recyclability due
to its high light
absorption
capacity and
optimized band
gap and

75
conduction band
edge
12 [Ir(ppy)2(bpy)]+- [Ni4(H2O)2 Hydrothermal Visible Methanol 3 4400 Transition-metal- [656]
derivied UiO-MOF (PW9O34)2] 10- method light substituted
POM POMs have
recently been
shown to possess
higher visible-
light
photocatalytic
activity because
of their extensive
tunability, rich
redox chemistry,
and synergism
between the
heterometallic
ions.

13 NH2-MIL-125(Ti) - Hg/Xe TEA 3 times 22500 Co@MOF is an [555]


lamp (500 efficient and
Solvo thermal W) fully recyclable
noble metal-free
method catalyst system
for light-driven
hydrogen
evolution from
water under
visible light
illumination.
14 Ti-MOF-Ru(tpy)2 Pt nanoparticles Solvothermal Visible TEOA 3 times ≈ 400 The production [557]
Ti-O cluster method. light of hybrid
materials
expands
chemical and

76
structural
alternatives due
to porous
coordination
polymers (PCPs),
which have high
specific surface
areas, well-
ordered porous
architectures, and
structural design
ability.

15 Ru-TBP-Zn Pt/NPs/ Solvo thermal Solid state TEOA 240 MOFs built from [659]
POMs reaction light (230 Ru2 SBUs and
W, 400 porphyrin-
nm) derived ligands
shows 28 times
higher HER
activity due to
photosensitizing
and multi-
electron transfer.
16 ErB/GOWPt@U6N- Pt NPs Xe lamp TEOA The designed [673]
8 (300 W) 18 150 structure was
found to possess
superior H2-
generation ability
compared to only
inside or outside
decorated
samples,
highlighting that
the enhanced
property strongly

77
correlates with
the inhibited
recombination of
electrons and
holes by the
inside/outside
modification
strategy.
17 Pt/PCN-777 Pt Microwave- Xe lamp Benzyl 586 PCN-777 has a [672]
assisted (300 W) amine much higher
activity than
MOF-808 due to
stable band gap.
18 Pt/NH2-UiO- Pt Hydrothermal 300 W Xe Sodium 23 410 The resulting [278]
66/TpPa-1-COF method lamp ascorbate hierarchical
porous hybrid
materials show
efficient
photocatalytic
H2 evolution
under visible
light irradiation
due to large
surface areas,
outstanding
visible light
absorbance, and
tunable band
gaps.
19 CdS/UiO-66 UiO-66 Solvo thermal Xe lamp Lactic 211.83 CdS/UiO-66(10) [318]
method (300 W) acid demonstrates an
initially high H2
generation rate
despite a low
percentage of

78
CdS due to the
preferred
creation of fine
CdS
nanoparticles and
heterojunction
between CdS and
UiO-66.
20 Nax-C3N4/Pt@UiO- Pt Thermal Xe lamp TEA 471.4 When we [570]
66 polymerization (300 W) increased Na
and thermal contents, the
oxidation band gap of Nax-
etching C3N4 becomes
method narrow, and
visible-light
harvesting is
enhanced, while
electron transfer
efficiency is
more complex,
which is affected
by electron trap
sites and the
driving force of
electron transfer
having higher
hydrogen
production.
21 CdS/YS Pt Facile Xe lamp Na2S and 4 3051.4 CdS has [227]
microwave- (300 W) Na2SO3 distinctive three-
assisted dimensional (3D)
hydrothermal open design,
process. Yolk-shell
structure that
provide

79
significant
benefits for
developing
improved
photocatalysts.
22 CdS/C@CoS2 C@CoS2 and Hydrothermal Xe lamp Lactic 4 87 730 Co-loading of [572]
TFA method (300 W) acid both photo-
generated
electrons and
hole-transporting
cocatalysts
(C@CoS2 and
TFA) on light-
harvesting
semiconductor
heterophase
homojunction
CdS (OD-2D
CdS) is
established as a
productive way
to ameliorate the
photocatalytic
water splitting
efficiency.
23 CdS/Co–C@Co9S8 Co–C@Co9S8 Hydrothermal Xe lamp Lactic 4 26 690 Formation of [573]
(150 W) acid open-pore 3D
CdS mesoporous
networks at the
nanoscale is an
ideal strategy to
overcome this
difficulty
because it will
provide a large

80
number of
catalytically
active reactive
sites by
combining the
high reactivity of
mesoporous
semiconductor
nanocrystals into
the same
material.
24 CdS/MCTMPs Ni co Hydrothermal Xe lamp Lactic 5 136 770 MCTMPs [574]
method (150 W) acid nanostructures
benefiting from
the rich structural
features, high
specific surface
area, and active
multicomponents
in the
composition, the
MCTMPs
manifest greatly
enhanced
photocatalytic
hydrogen
evolution
properties when
integrated with
CdS
semiconductor
nanorods.
25 CdS/ Co4S3 CdS Wet chemical Xe lamp Lactic 5 12 360 The [230]
nanoparticles method (150 W) acid incorporation of
Co4S3 with the

81
CdS particles
effectively
accelerated
charge separation
and transfer in
photocatalytic
reactions due to
the low density,
hollow interior,
and shell
permeability of
the onion-type
composite.
26 MOF-199 Cu/ TiO2- Cu–Cu2O Solvothermal F-lamp Glycerol 4 151300 The [575]
400 method (100 W and photocatalyst Cu-
365 nm) Cu2O/ TiO2-400
produced by
calcining the
TiO2-MOF
composite at 400
°C had the
highest activity
for hydrogen
generation from a
5% glycerol-
water mixture
because of low
cost, large
surface area, and
high pore size.
27 MOF template Cu Hydrothermal Newport TEOA 5 10 046 The high [576]
Cu/CuO@ TiO2 method xenon lamp photocatalytic H2
(450 W) production
activity is
attributed

82
predominantly to
the presence of
surface deposited
Cu0 species and
the small size
heterojunction
(1-2 nm) between
CuO and TiO2,
which synergizes
the interfacial
charge carrier
transfers from
TiO2
nanoparticle.
28 Pt/Cu–TiO2/C Pt LPE layer-by- Xe lamp Methanol 3 14 049 Hollow Cu- [578]
layer method. (300 W) TiO2/C
nanospheres
were thin,
uniform, and
homogenous,
allowing for
appropriate use
of the light
source and
efficient
transport of
reactant
molecules,
guaranteeing the
highest catalytic
activity.
29 ZnOS-30 Solvothermal Xe lamp Na2S and 8 435 The nanoscale [232]
method (300 W, λ > Na2SO3 heterojunction
420 nm) ZnO/ZnS
structural feature

83
reduces the
charge carriers'
average free path
and improves the
charge separation
efficiency, thus
leading to
significantly
enhanced HER
activity under
visible-light
irradiation with
high stability and
recyclability
without any
cocatalyst.
30 EY/C–ZrO2/g- C3N4 Ni2P Solvothermal Xe lamp TEOA 10 040 The presence of [579]
/Ni2P method (300 W) the Ni2P co-
. catalysts
accelerates the
surface reaction.
ZrO2 has been
used in the H2
generation and
organic
pollutants
degradation,
because its
surface can be
oxidized and
reduced and
contains both
acid and base
sites

84
31 EY/FeO3.3C0.2H1.0 Pt Hydrothermal White-light TEA 4167 Despite the
method LED absence of a [580]
noble metal
cocatalyst, the
MOF-derived
catalyst generates
H2. Using earth-
abundant and
cost-effective
substitute for Pt,
Pd, Ru, Ir, and Rh
in photocatalytic
water reduction
enables a more
practical
hydrogen energy
production
Solvothermal UV–vis TEOA 10.6 Dye-like-based [581]
32 Gd-(MOFs Ag (1.0−2.0%) method. Gd-MOFs are
more stable at
varying PH
levels. Moreover,
Gd-MOF does
not demand a
cocatalyst which
demonstrates
substantial
photocatalytic
activity for
hydrogen
generation under
UV-vis
irradiation.
33 NiMo@MIL-101 Ni Double solvent Visible TEOA 4 740.2 NiMo@MIL- [584]
method 101, as a

85
cocatalyst,
exhibited the
highest
photocatalytic
activity after
being sensitized
by EY due to
stability and high
apparent
quantum
efficiency. It also
shows a higher
transient
photocurrent,
lower over-
potential (-
0.51V) and
longer
fluorescence
lifetime.
34 Cu/TiO2-AA Hydro thermal Xe lamp Methanol 62.16 As compared
(300 W, with the widely [554]
>420 nm) used calcination
approach, the
simultaneous
etching and
reduction can
better preserve
the octahedral-
shaped shells and
crystal phase as
well as prevent
the formation of
carbon residues
and cocatalyst

86
aggregation,
which all help
improve the
efficiency of
photocatalysis.

87
High photocatalytic performance is hampered by the inherent poor light harvesting capacity of
many MOF materials, which is a major bottleneck. The soft structure of MOFs is susceptible
to breaking during the solar water splitting reaction, highlighting the issue of MOF materials'
stability. Additionally, charge carrier separation and transfer are restricted by the intrinsically
weak conductivity of MOF materials. As photocatalysts for water splitting, a variety of
materials can be used, but MOF-based materials have some benefits over traditional
photocatalysts. For instance, organic linkers and metal ions or metal cluster nodes in the
structures of MOFs can absorb visible or UV light, causing an electron shift from HOMO to
LUMO and exhibiting semiconductor behaviour. In some circumstances, MOFs can even
capture NIR light, and the charge transfer from metal to ligand can take place in their structures
at low energies. MOFs can also have their architectures altered by changing their linkers,
doping them with metal ions or nanoparticles, or both. We can be confident that MOFs are the
best for photocatalytic water splitting to create clean and environmentally friendly hydrogen
thanks to the strategies that are being presented in the leading sections.

3. Factor affecting the photocatalytic activity of MOFs

To initiate the photocatalytic water-splitting reaction, a photocatalyst must have two


properties: catalytic property and photosensitivity [239]. Regarding the ability of MOFs to
catalyze reactions, some of them have been classified as microporous semiconductors because,
when exposing them to light, they split electrons and holes into two distinct populations [113].
Furthermore, they have the disadvantage of poor electrical conductors, but this disadvantage
can be overlooked because of their porosity. These holes enable effective reactant diffusion
and close interaction among active sites. Some MOFs, such as MIL-101, MIL-100, and MIL-
53, have unusual photo activity and are even able to react to light in the visible spectrum [240].
Furthermore, because of its excellent design potential, component functionalization may be
used as an efficient approach to augment MOF's visible light and even NIR light reaction,
indicating MOFs' superiority over traditional semiconductors such as ZnO, CdS, and TiO2
[173]. Many factors affect the photocatalytic behavior of MOFs; some of them are discussed
in Figure 28.

88
Figure 28. Factor affecting the catalytic properties of MOFs for photocatalytic water spitting.

3.1 Surface area

The total surface area is commonly known for significant impact on the reaction rate
during the chemical processes. Several studies reported that the catalytic activity was highly
dependent on the analytical models for size-dependent surface energy [241], [242], [243],
[244], [245], [246], [247]. The surface area has a major effect on the ability of photocatalytic
water splitting; increasing the surface defect sites or specific surface area by simply regulating
the morphology is, to some extent, an effective method to improve photo activity [248], [249].
[250]. According to Gong et al. [251] semiconductors having narrow band gap will show
greater catalytic activity rather than those which have larger band gap. Moreover, they also
suggest some surface modifications to increase the performance such as surface protection,
cocatalyst loading, surface energetics tuning, and surface texturization. Thus, tailoring the
surface properties and energetics of semiconductors stands out as the particularly important
and effective approach to improve overall efficiency. Because of the increased surface area,
there will be more reactive sites, resulting in higher photocatalytic activity [252], [253], [254].

89
Increasing the number of surface defect sites or the specific surface area of the material by
merely altering its shape is, to some extent, an efficient strategy for enhancing photo activity
[250], [255]. Typically, 2D MOFs are utilised as a photocatalyst in catalytic water splitting
[142], [256], [257], [258]. Furthermore, MOF-derived 2D Nano sheets or 3D Nano rod arrays
are used to increase surface area, expose more active sites, and accelerate charge and mass
transfer for water splitting [259]. Constructing Nano 2D/3D MOFs to expand their surface
areas is one effective technique for HER to expose additional catalytic performance [239]. The
fabrication of 2D MOF Nanosheets can accelerate electron transfer and thus enhance
conductivity. Due to the large surface area, MOFs can attach more substrate, increasing the
catalytic reaction rate. Generally, they possess a significant area of 1,000 m2/g up to 7,000 m2/g
because of their uniquely porous nature, which is critical for enhancing water-splitting activity
[260]. A porous structure means a surface having small holes [261], [262], [263], [264], [265],
and therefore, MOFs are widely used in photocatalytic water splitting because of its porosity
[266], [111], [267], [268]. For the idea of a photocatalyst where catalytic reactions frequently
occur, hollow structures with large exposed surface areas are recommended [269], [270]. If the
porosity and homogeneity of the active catalytic sites can be preserved by following pyrolysis,
then high-performance heteroatoms-doped catalysts with unique surface atomic structures and
bonding arrangements can be manufactured [271]. MOFs have highly tunable porous
structures, which play a role in size-selective catalysis; the pore space provides special catalytic
microenvironments, such as a chiral environment, appropriate hydrophilic and hydrophobic
pore nature, and electron-deficient or electron-rich circumstances, which greatly affect the
interaction and reactivity of reactant molecules [118]. MOFs have high surface areas and
porous structures enable the introduction of photosensitizers or co-catalysts, such as
polyoxometalates (POMs), metal nanoparticles (NPs), metal complexes, semiconductors, etc.
They are leading to the spatial separation of charge carriers and, thus, achieving an enhanced
photocatalytic efficiency [260], [272], [121]. Jin et al. [273] proposed MOF into flaky CdS
photocatalyst forming special structure and active sites for efficient hydrogen production. The
obtained composite photocatalyst can not only make use of the specific surface area of ZIF-67
to provide a larger reaction place, but also can positively promote the reaction by binding sites
between CdS and ZIF-67. Figure 29 showed the hydrogen production activity test curves of
the respective photocatalysts under visible light irradiation (l:420 nm). It can be seen that the
hydrogen production activity of ZIF-67 was very weak, which was only 6.4 mmol H2 after 5
h. Under the same conditions, the activity of the composite photocatalyst encapsulated by CdS
was significantly improved, and can reach 308 mmol after 5 h. For comparison, pure CdS
activity was tested. Under the same conditions, only 42 mmol H2 was produced. It can infer
that the interface between CdS and ZIF-67 was the key factor in improving HER (hydrogen
evolution reaction) activity. According to the existing literature, the increase in photocatalytic
activity demonstrated that highly efficient charge path was formed in the composite
photocatalyst by coordination of CdS boundary with the active site of the unsaturated metal,
which can accelerate the mutual transfer of photogenerated electrons. Then life of the
photocatalyst was tested. It can be seen that under the condition that the activity of the
photocatalyst can keep stable within 20 h with lactic acid working as sacrificial reagent. The
amount of hydrogen produced per cycle was maintained at about 300 µmol. On the other hand,
the activity of catalyst depended on the active site of the catalyst, which was generally
considered to be uniformly distributed on the surface area of the catalyst. So the larger the
surface area can provide the more active sites and the greater the catalytic activity.

90
Figure 29. (A) The hydrogen production activity curves of the photocatalysts and (B) stability
test of CdS/ZIF-67.Reprinted with permission from [273]. Copyright 2019 Elsevier.

Similarly, Li et al [275] synthesize a Unique Ternary Ni-MOF-74/Ni2P/MoSx composite for


efficient photocatalytic hydrogen production. The BET specific surface area of the catalyst was
characterized by the N2 adsorption-desorption isothermal at 77 K. The surface structure and
microstructure of Ni-MOF-74, Ni-MOF-74/P, MoSx and 10%-NPMS were characterized SEM
and high-resolution transmission electron microscopy (HRTEM), respectively. Similarly,
Composite catalyst WO3/CoP was synthesized by a simple high temperature in situ method by
Jin et al [276]. When MOFs are employed as a sacrificial precursor or support to fabricate MOF
derived nanostructures and MOF-based composites, reduced specific surface area and pore
volumes may be noticed. In the case of MOF derived nanostructures, the reduced porosity may
be due to heating at high temperatures. During the heat treatment the crystalline porous
structure collapses by the breaking of coordination bonds and the removal of nanocarbon
elements, which greatly suppress the porosity [277], [124], [261], [141], [35], [262], [263]. In
the case of MOF supports porosity may be acquired from the bare MOFs but reduced specific
surface area and pore volume depend on pore occupation or blocking by the guest particles.
Overall, the reduced porosities are not necessarily detrimental to the photocatalytic hydrogen
evolution rate [361], [274]. In a report Lan et al. [278] successfully integrated two dimensional
(2D) COF with stable MOF. By covalently anchoring NH2 -UiO-66 onto the surface of TpPa-
1-COF, a new type of MOF/COF hybrid materials with high surface area, porous framework,
and high crystallinity was synthesized as shown in Figure 30. The resulting hierarchical porous
hybrid materials show efficient photocatalytic H2 evolution under visible light irradiation.
Especially, NH2-UiO-66/TpPa-1-COF (4:6) exhibits the maximum photocatalytic H2 evolution
rate of 23.41 µmol g -1 h-1 (with the TOF of 402.36 h-1), which is approximately 20 times higher
than that of the parent TpPa-1- COF and the best performance photocatalyst for H2 evolution
among various MOF- and COF-based photocatalysts.

91
Figure 30 Schematic illustration of the synthesis of NH2 -UiO-66/TpPa-1-COF hybrid material
[278].

Some of the MOFs that have been created up to the present have astonishingly large surface
surfaces. The efforts that were made to achieve even higher surface areas appeared to come to
a standstill; this was not primarily due to the difficulty in creating novel candidate materials;
rather, it was due to the progressively greater tendency of these materials to collapse upon the
removal of the solvent. Benefiting from the large surface area, adjustable chemical
components, tunable pore structure, controllable topology, and well-defined surface
functionality, various MOF-based/derived materials with excellent water splitting performance
have been developed. In conclusion, MOFs have the potential to be used as photocatalysts in
theory. This would allow one to maximise the photocatalytic surface area and precisely adjust
the active sites in a straightforward manner. This would be accomplished through the rational
selection of organic ligands and the modification of metal nodes.

3.2 Co-catalysts

A substance or agent that brings about catalysis in conjunction with one or more others
is known as a co-catalyst. Most semiconductors cannot give high H2 evolution activities
without a cocatalyst even in the presence of sacrificial electron donor. A reason might be due
to the facile recombination of electron-hole pairs before migrating to the surface for reactions.
Another major reason is that the surface reaction is too slow to efficiently consume the charges.
Traditionally, in order to “take” the electrons out to the surface, metals, especially noble metals
such as Pt, are used as the cocatalysts. Noble metals not only serve as electron sinks, but also
provides effective proton reductions sites, hence dramatically facilitates proton reduction
reaction [279]. Trasatti et al [280] found a volcano relationship between the exchange current
for H2 evolution and the MH bond strength (M is a transitional metal). Pt was on the peak of
the volcano, showing the lowest activation energy for H2 evolution. Therefore, from the
viewpoint of both electronic and catalytic properties, Pt is usually considered as the most
suitable H2 evolution cocatalyst. In most cases, cocatalysts are required to achieve higher
photocatalytic activity [281], [282], [283]. Moreover, Feng et al [284], reported a novel dual

92
co-catalysts of black phosphorene (BP) and single Pt atoms on CdS nanospheres. BP/CdS
heterostructures are prepared by grinding and sonication and then single Pt atoms are deposited
onto BP/CdS through a photo-reduction procedure. In addition to being anchored on the surface
step edges of CdS nanospheres, Pt single-atoms with positive charge are embedded on Cd
vacancies and stabilized by Pt-S covalent bonds. Single Pt atoms are immobilized on the
surface of BP as well. The as-prepared Pt-BP/CdS composites are evaluated toward visible-
light-driven hydrogen generation. In a range of Pt and BP loading contents, 0.5 wt% Pt-5 wt%
BP/CdS composites display the greatest photo activity and outperform pristine CdS
nanospheres by a factor of 96 in terms of hydrogen evolution rate, alongside a remarkable
apparent quantum efficiency of 46% at 420 nm. In-depth analyses on interfacial electronic
interactions and photoexcited charge-carrier dynamics demonstrate that both single Pt atoms
and BP strongly interact with CdS and synergistically steer spatial charge separation, thereby
boosting photocatalytic performance. Co-catalysts in a photocatalytic hydrogen evolution
system carry out four essential roles, each of which contributes to the overall enhancement of
the activity of the semiconductor photocatalyst [284], [285], [286], [287]. Co-catalysts play
four critical roles in photocatalytic water splitting, enhancing the efficiency and stability of
semiconductor photocatalysts, as illustrated in Figure 31. (i) Co-catalysts can reduce the
activation energy for H2 or O2 production on semiconductor surfaces. (ii) They can facilitate
electron-hole isolation at the interface between the co-catalyst and the semiconductor.
Although photogenerated electrons on a photocatalyst's conduction band (CB) convert protons
to H2 molecules, photogenerated holes on the valence band (VB) oxidize H2O to O2. iii) Co-
catalysts can significantly decrease photo-corrosion and enhance photocatalyst stability. iv) As
an active site, co-catalysts provide spaces for other substances to attach.

93
Figure 31. Illustrate some functions of the co-catalyst in the enhancement of catalytic activity.

Some plasmon metals, such as gold and copper, can better absorb visible light because of a
phenomenon known as localized surface plasmon resonance, or LSPR [288], [289], [290]. Co-
catalyst can affect the separation ability of electron-hole pair very rapidly because when a co-
catalyst is induced over a photocatalyst, it enhances the catalytic activity and makes possible
an early way to get charges separated for the reaction. Co-catalysts can act as active sites in
photocatalytic reactions [291], [292], [293], and [294]. Most co-catalysts can act as electron
sinks and proton reduction sites [295]. One of the most effective methods for improving the
performance of photocatalysts is the deposition of metal on the surface of semiconductors
[296], [297]. Many noble-metals such as Pt [298], [299], Pd [300], [301], Ag [302] etc have
recently been demonstrated which act as an efficient cocatalyst. Despite their high
photocatalytic activity, noble metal co-catalysts practical applications are limited by the
paucity of resources and expensive pricing [303], [304]. Non-noble metals are prevalent in the
Earth's crust, and some of their qualities, in terms of boosting photocatalytic efficiency, are
similar to those of noble metals [305], [306]. Furthermore, non-noble metals have large work
functions; they can co-catalyze photocatalytic processes to capture and transport electrons [58],
[307]. Zn, Ni, and Co have recently been applied to drive photocatalytic water splitting [173],
94
[308]. Ni is an excellent choice for use as a co-catalyst in photocatalytic H2 evolution because
its work function is comparable to platinum's and is even greater than gold [285], [309]. In
another article Cheng et al [310] studied the role of NiS on ZnO/ZnS Nanorod Heterostructures
for Enhanced Photocatalytic Hydrogen Production. NiS is regarded as a promising cocatalyst
to replace the noble metals due to its low cost and equivalent or even better performance.
However, there is a huge controversy over whether the NiS cocatalyst is used to trap electrons
or holes in the photocatalytic process. The Zn species in the bimetallic–MOFs can spatially
separate the Ni species to restrain their aggregation, which is beneficial for the formation of
NiS with a small enough size. The optimal heterostructure photocatalysts exhibit an excellent
hydrogen production rate of 27 mmol g−1 h−1, which is about seven times higher than that of
the ZnOS heterostructure. Moreover, NiS-loaded photocatalysts always display the equivalent
or even better H2 production performance than their Pt-loaded counterparts [311], [312]. The
photocatalytic hydrogen evolution activity over Cd0.5Zn0.5S is significantly increased by in situ-
loaded Ni2P due to the intimate interface contact and efficient charge transfer from Cd0.5Zn0.5S
to Ni2P. The 4 wt%Ni2P-loaded Cd0.5Zn0.5S photocatalyst shows five times higher H2 evolution
rate than that of the pristine Cd0.5Zn0.5S as shown in Figure 32. The highest H2 evolution rate
of 65.6 μmol/h and the apparent quantum yield (AQY) of 29% at 420 nm is achieved under the
optimized conditions; the noble metal-free Ni2P is stable during 20 h photocatalytic reaction.

Figure 32: Photocatalytic activity of H2 evolution on CZS and mNP/CZS photocatalysts of NP


Reprinted with permission from [312]. Copyright 2018 Elsevier.

Other than this, some metal oxides and hydroxides like TiO2, CuO, Cu2O, and NiO co-catalysts
for photocatalytic H2 production are well known [313], [314]. Figure 33a. shows a simplified
explanation of the photocatalytic production of hydrogen under visible light using TiO2.
Additionally, MOFs are extensively employed as electrocatalysts and photocatalysts due to
their substantial specific surface areas, tunable porosity channels, and high degrees of design

95
freedom [315], [316]. These properties expose many active sites to drive redox reactions and
reduce the distance that photo-excited carriers must travel to reach the pore surface [113]
through heat processing, phosphorylation, and sulfidation, materials formed from MOFs act as
precursors for porous semiconductors such as metal phosphide, metal oxides, carbon materials,
and metal sulfides [317]. Investigate the utilization of materials produced from MOF in
photocatalytic water splitting has received a lot of attention. The doping of C and N into MOF-
derived semiconductors can be aided by MOF-based materials acting as co-catalysts in the
hydrogen evolution half-reaction system, but more crucially, they can give concrete surfaces
to show a vast number of active sites [285], [102], [317]. UiO-66 as a representative stable
MOF owns intrinsic photocatalytic H2 evolution activity. Jiang's group., [318] synthesized the
composite of CdS-decorated UiO-66 that not only has taken the advantage of both CdS and
MOF, but also avoid the drawbacks of each component. CdS/UiO-66 (Figure 33 e) exhibits
high H2 production activity for water splitting of H2 evolution under visible light irradiation,
the key contribution of the composite is that CdS has the capability to adsorb visible light and
UiO-66 provides a large surface area to disperse the CdS on the MOF surface, so that providing
more adsorption sites and photocatalytic reaction centers. Figure 33e. Shows that UiO-66
crystals were generated utilizing ZrCl4 and terephthalic acid (H2BDC) as reactants. At 10, 20,
and 40% wt.% CdS loadings, CdS NPs, were in situ deposited onto the surface of UiO66
crystals with Cd(CH3COO)2H2O and thioacetamide (TAA) as cadmium and sulfur sources,
respectively, to yield CdS/UiO-66(X) composites, where X denotes the CdS theoretical
content. In addition, the g-C3N4/MOF heterojunctions that were synthesized by combined
MOFs (UiO-66) also displayed outstanding photocatalytic performance toward H2 production,
as well as CO2 reduction and organic pollutants degradation [319]. This design will provide a
new path to construct efficient photocatalysts for hydrogen evolution. Semiconductor co-
catalysts can improve photocatalytic water splitting by promoting charge collection and gas
development. The approach is most active when a co-catalyst is effectively loaded onto a
semiconductor. Some relatively plentiful and affordable transitional metals, such as Co, Ni,
and Cu, have been utilized as co-catalysts in photocatalytic H2 generation in addition to
expensive and rare noble metals. When these metals are loaded onto MOFs, they form a
Schottky barrier at the metal/semiconductor interface [303].

96
(e)

Figure 33. (a) Diagram displaying movement of the electrons and dissociation in CuO-TiO2 (b)
CB and VB edge energy levels vs normal hydrogen electrode (c) Charge transfer and separation
in a plate-like Co(OH)2/TiO2 nanosheets system; (d) Photocatalytic H2 production mechanism
Reprinted with permission from [313]. Copyright 2017 Elsevier. (e) Method for fabricating
CdS/UiO-66 Composites Artificially. Reprinted with permission from [318]. Copyright (2018)
American Chemical Society.

In 1980, Domen et al. [320] successfully embedded NiOx as a co-catalyst onto the surface of
SrTiO3. They found that the NiOx/SrTiO3 photocatalytic activity for water vapour degradation
was strongly influenced by the state of the loaded co-catalyst. To increase the amount of H2
that can be produced, the co-catalysts utilized in photocatalytic hydrogen evolution systems
have progressed from using noble metals to using transition metals and alloys. In the
aforementioned transitions, researchers have achieved numerous studies in photocatalytic
hydrogen production [285]. Apart from the abovementioned works, Table 3 summarizes other
recently published MOF based and derived photocatalysts for photocatalytic HER.

Cocatalyst loading onto the surface of semiconductor photocatalysts is frequently


used as a strategy to increase the efficiency of photocatalytic water splitting, particularly to
enhance the hydrogen evolution reaction (HER). The cocatalyst improves the separation of

97
photoexcited carriers and reduces the overpotential of the surface reaction, thereby enhancing
the performance of water splitting. In stark contrast, the unique properties of molecular metal
complexes, such as redox potentials, catalytic activity, and selectivity, can be altered by varying
the central metal and/or ligand molecule, allowing them to be utilised as cocatalysts. However,
molecular cocatalysts can be deactivated by redox degradation, ligand exchange, and
photodegradation. As a consequence, molecular cocatalysts are rarely used in water splitting
as a whole.

98
Table 3: Summary of MOF-based cocatalysts for photocatalytic HER in recent publications

Sr no Cocatalyst Light source Solvent H2 evolution rate (mmolh–1·g–1) Ref.

1 CdS/MIL-101(Cr) Visible-light Lactic acid 1.1 224

2 1TMoS2/MIL-125-NH2 UV-Vis light N/A 2.094 646

3 g-C3N4/UiO-66/Ni2P 5 W LED white light TEOA 0.2 mmol 5 h–1 g–1 568

4 Ni2P/CdS Simulated solar Lactic acid 33.48 640

5 MOF-Derived MoS2@TiO2 UV-Vis light TEOA 10. 046 321

6 NixMo1-xS2/MOF-5@g-C3N4 UV-Vis light TEOA 0.319 322

99
3.3 Heterostructured MOFs photocatalysts

Recently, MOFs have been doing great in photocatalytic water splitting due to their
several properties discussed above, but like other semiconductors, they also face some
bottlenecks in catalytic activity. The separation of photoelectrons from vacancies in
semiconductors has been supported by several techniques, including doping, loading co-
catalysts, changing interfaces, facet engineering, manipulating size and morphology, and
introducing heterostructure [142]. Making a heterostructure by fusing two semiconductors with
compatible band gap locations is one of the most promising techniques for encouraging charge
separation [323], [324]. This is because the interface potential difference creates an electric
field that can effectively and efficiently separate photoinduced electrons and holes from one
another and prevent their recombination. Two crucial factors must be considered while
constructing a highly-efficient heterostructure catalyst based on MOFs for the photolysis of
water [143]. First, an ideal photocatalyst should have a valence band potential higher than the
oxidation potential of O2/H2O (1.23 V vs NHE) conduction band potential that is lower than
the reduction potential of H+/H2 (0.0 V vs NHE). This difference must be more significant than
1.23 eV to provide more driving force to split water [84]. Secondly, the interface's nature
significantly impacts charge separation efficiency [325], [326]. There are three types of
heterostructure photocatalysts based on the semiconductor band gap as discussed below.

3.3.1 Type-I heterostructured MOFs as photocatalysts

In a type I heterostructure, semiconductor A's valence, and conduction bands cross


over semiconductor B's. In this scenario, the surface of semiconductor B will accumulate
with both photoelectrons and holes [142]. The schematic diagram is presented below
Figure 34a.

3.3.2 Type-II heterostructured MOFs as photocatalysts

The type II heterostructure, in contrast to the type I heterostructure, typically consists


of two semiconductors containing valence and conduction bands, as illustrated in Figure 34b.
In this interface, a potential built-in field will be used to push the photogenerated electrons and
holes on them, allowing for highly efficient spatial separation [142]. Due to the internal electric
field (IEF) created at the interface between these two semiconductors, the photogenerated
electron-hole pairs can be successfully separated, increasing the photocatalytic activity. As a
result, type II heterostructures are far more successful than type I or III heterostructures at
improving the spatial charge separation of photogenerated carriers [327], [297]. Three main
techniques can be used to produce type II heterostructures between MOFs and other
semiconductors: the one-pot method, pre-synthesis of other semiconductors. Followed by
post- synthesis modification with MOFs, and pre-synthesis of them are followed by post-
synthesis modification with other semiconductors [328]. Many more heterojunction

100
Figure 34. Band structure of multiple heterojunction types (a) type I heterojunction, (b) type II heterojunction. Reprinted with permission from
[328]. Copyright 2021 Elsevier

101
photocatalysts with 1D core-shell architectures have been discovered, such as TiO2/C3N4 [329],
C3N4/CdS [330], Ni3S2/VO2, and ZnO/WS2 [331] improve photo-induced pair dissociation and
transference. Moreover, Parida et al [332] fabricate a composite of citrate capped Fe3O4 and
UiO-66-NH2 which has been designed to remediate Cr (VI) by adsorption and harvest photons
from visible light for clean energy (H2) conversion. The composite of CCM with MOF (MU-
2) was studied through sophisticated analysis techniques; PXRD, FT-IR, BET, UV–Visible
DRS, PL, TG, HRTEM and XPS etc. to reveal the inherent characteristics of the
material. Additionally, the synergistic effect of CCM and UNH in MU-2 resulted in high
photocatalytic H2 evolution rate that followed Z-scheme mediated charge transfer
pathway. Meng et al. [333] proposed co-immobilization of CdS quantum dots and carbon
nanorods (CDs) in MIL-101 cages by a one-step multiple solvents technique followed by
heating treatment. This method prevents nanoparticle agglomeration during production and
reaction. CDs nanoparticles can form heterostructures with MOFs to separate photogenerated
charges, improving the photocatalytic activity of the resultant material with Pt as a co-catalyst.
Similarly, Zhao et al. [334] investigated the combined effects of octahedron NH2-UiO-66 and
flower-like ZnIn2S4 microspheres for Photocatalytic dye degradation and hydrogen evolution
under visible light. It was discovered that the characterization of the one-of-a-kind ZIS/NU66
that is created by using a type-II heterostructure can effectively promote the separation and
transfer of light-induced charges, as well as the exposure of rich active sites that are used for
photo catalyst redox reaction. In addition, Chen et al. [335] synthesized an additional
heterostructure COF-based core-shell framework for improved photocatalytic hydrogen
evolution. As a result of this research, the idea of a hetero-framework structure encourages the
logical design of heterostructured photo catalysts and broadens the family of organic photo
catalysts that can be used for the conversion and utilization of solar energy. Type-II
heterojunction photocatalysts generally have strong electron-hole separation efficiency, a
broad light absorption range, and fast mass transfer rates [336].

3.3.3 Type-III heterostructured MOFs as photocatalysts

Furthermore, in a type III heterostructure, it is challenging to transfer electrons from


semiconductor A to semiconductor B because semiconductor B has a larger negatively charged
valence band than semiconductor A's positively charged conduction band [328]. For example,
the heterostructure of NH2-MIL-125(Ti) with ZnCr-LDH was established by in-situ
solvothermal cultivating NH2-MIL-125(Ti) onto exfoliated ZnCr-LDH nanosheets [337] as
shown in Figure 35. NH2-MIL-125(Ti) and ZnCr-LDH nanosheets (NSs) hybrid composites
were created by growing NH2-MIL-125(Ti) in the presence of exfoliated ZnCr-LDH NSs.
LDHs have been reported to be prepared as a colloidal suspension of individually exfoliated
NSs in a non-aqueous solvent (i.e., formamide). Dimethyl formamide or another comparable
solvent is typically needed to produce NH2-MIL-125(Ti). Despite having lower surface areas
than the pure NH2-MIL-125(Ti), which has a high surface area, the hybrid composites gave
improved photocatalytic activity and increased stability for hydrogen production. These
improved photocatalytic activities can be due to the solid electrical contact between the two
components, extended light absorption, and band alignment that prolongs the lives of photo-
induced electrons and holes. According to these findings, the current synthesis method can
produce a wide variety of novel photocatalysts dependent on MOFs [337], [152].

102
b)b) b)
a) a) a)
a) +
Exfoliation Ti and ligand
ZnCr-LDH DMF/MeOH NH2-MIL-125 (Ti) / ZnCr-
Solvothermal
LDH

c) c) c)
b) b)
b)
b)
DMF/MeOH
+ LDH nanosheets

Solvothermal
Ti and ligand NH2-MIL-125 (Ti)

Figure 35. Display of the process of (a) pristine NH2-MIL-125 and (b) hybrid composites of
NH2-MIL-125(Ti) and ZnCr-LDH NSs (Ti). Additionally, provided the relevant FE-SEM
images for each image. Reprinted with permission from [337]. Copyright 2019 Elsevier.

Coordinate connections between organic ligands in MOFs and inorganic semiconductor


nanoparticles provide a more efficient charge transport channel, allowing for better
heterostructured interface charge transfer and separation [152], [142]. Xu and colleagues [338]
developed a coordination-driven method for accurately epitaxially growing metal sulfides
(CdS, ZnS, CuS, and Ag2S) nanoparticles onto MIL-101 in mild conditions to address this
constraint. Most CdS nanoparticles can be successfully placed on MIL-101 octahedral surfaces
using a solvothermal process, as elaborated in Figure 36. MIL101 nanoparticles, which have
been prepared through a hydrothermal method, are dehydrated at 120 °C to expose their CUSs
under vacuum treatment. Subsequently, the exposed CUSs react with cysteamine in anhydrous
ethanol, an organic molecule that features strong coordination ability with different types of
metal ions to form cysteamine-grafted MIL-101 (denoted as MIL-101-S). The MIL-101-S then
experiences a S 2- anion exchange process to grow CdS NPs on MIL-101, thereby forming the
MIL-101@CdS composites. However, some of them inevitably aggregated and grew out of the
MIL-101's exterior surface, resulting in heterophase separation due to the low solubility
product constant (Ksp) of CdS during solvothermal treatment.

103
Figure 36. Epitaxial growing metal on MIL-101 MOFs by heterojunction under mild conditions
Reprinted with permission from [338]. Copyright 2013 Elsevier.

Another critical factor to examine was the matching of MOF and g- C3N4 dimensions.
Because g- C3N4 is a graphite-like layered polymeric semiconductor, ultrathin 2D g- C3N4
materials are produced with better photocatalytic efficiency than bulk g- C3N4 materials. g-
C3N4 is an ideal candidate for photocatalytic water splitting as a result of the excellent
alignment of its band edges with water redox potentials. Visible-light-driven photocatalytic
properties of CuFe2O4/g-C3N4 heterostructures with different CuFe2O4 loadings have been
examined with two sacrificial agents. An up to 2.5-fold enhancement in catalytic efficiency is
observed for CuFe2O4/g-C3N4 heterostructures over g-C3N4 nanosheets alone with the apparent
quantum yield of H2 production approaching 25%. The improved photocatalytic activity of the
heterostructures suggests that introducing CuFe2O4 NPs provides more active sites and reduces
electron–hole recombination [339]. Oxide based heterostructured photocatalysts have shown
enormous progress in the past few years for diverse photocatalytic applications ascribed to the
facilitated separation and transfer of photo-induced charge carriers which marvelously advance
the rate of reaction. Photocatalytic CO2 reduction and H2 evolution are two major sustainable
applications which are environmentally benign in nature and are at the cutting edge of
technology. Oxide based heterostructured catalytic systems have the potential to efficiently
catalyze the rate of hydrogen evolution reaction and the product formation from CO2 reduction

104
reaction [340]. Ultrathin 2D MOFs nanosheets have a low electron transfer distance, abundant
exposed catalytic active surfaces with unsaturated metal sites, and help improve photocatalytic
activity compared to three-dimensional (3D) MOFs bulk 2D MOF materials [341]. In contrast,
the electrostatic interaction between UNiMOF and g-C3N4 resulted in the formation of an
advanced heterostructure photocatalyst. Type III heterojunctions are nearly identical to type II
heterojunctions, except that type III heterojunctions have mismatched band gaps due to the
tremendously staggered opening [342].
The construction of MOFs-based heterostructures has been proved to be an effective
way to enhance the hydrogen evolution activities of MOFs materials due to the improved
separation efficiency of photogenerated charges in comparison with independent MOF. The
synergistic effects of MOFs and semiconductors can play important roles in separation and
transfer of photogenerated charges as well as the enhanced light absorption of photocatalysts.
The positive effect of constructing MOFs-based heterostructures is apparent in enhancing the
photocatalytic hydrogen evolution activity of MOFs materials, including the photocatalytic
systems of metal nanoparticle/MOF, inorganic semiconductors/MOF, MOF/g-C3N4 and
MOF/COF heterostructures

3.3.4 Z-scheme heterostructured MOFs as photocatalysts

The formation of Z-scheme heterostructured photocatalysts is a promising approach for


efficient and scalable H2 production from water splitting due to wide absorption range, high
charge-separation efficiency and strong redox ability of the Z-scheme heterostructured
photocatalysts. Bard et al. [343] are credited with being the pioneers of the Z-scheme, and ever
since then, the Z-scheme has been frequently used to circumvent problems associated with the
traditional photocatalytic system. Because of their unique energy structure, electrons and holes
in this class have no chance of recombination. Due to their shallow recombination rate, such
heterojunctions are highly desired for photocatalytic processes. Although the type II
heterostructure enhances charge separation efficiency, it reduces the overall system's redox
capabilities since redox reactions often occur on semiconductors with lower reduction or
oxidation potentials. This semiconductor uses a "Z" shape transmission line to maximize the
system's redox capacity [142]. There are three types of Z-schemes: classic or redox-mediated
Z-schemes, direct Z-schemes, and Z-schemes using solid-state electron linkers [344]. The
traditional Z-scheme photocatalytic system is depicted in Figure 37. This system consists of
two separate photocatalysts that are coupled together with the assistance of an appropriate
shuttle electron mediator [345]. Because of the limited range of solar energy that can be used
to power each photocatalyst, it is possible to capture photons more effectively, and PS I and
PS II can produce electrons and holes with enhanced reduction and oxidation capabilities [84].
Li et al [346] approved that Z-scheme photocatalytic systems that mimic natural photosynthesis
is a promising strategy to improve photocatalytic activity that is superior to single component
photocatalysts. The connection between photosystem I (PS I) and photosystem II (PS II) are
crucial for constructing efficient Z-scheme photocatalytic systems using two photocatalysts
(PS I and PS II). They concisely summarize and highlights recent state-of-the-art
accomplishments of Z-scheme systems, which was categorized through diverse modes,
including i) with shuttle redox mediators, ii) without electron mediators, and iii) with solid-
state electron mediators. As shown in Figure 37, this kind of Z-scheme photocatalytic system
consists of two different photocatalysts and an acceptor/ donor (A/D) pair (so-called shuttle
redox mediator). No physical contact exists between PS I and PS II. The electron acceptor (A)
and donor (D) react with the photogenerated electrons in the CB of PS I and holes in the VB
of PS II, respectively, resulting in the obviously decrease in the effective number of
photogenerated electrons and holes. Thus, it is critical to suppress the backward reactions

105
involving redox mediators that are thermodynamically more favorable than water splitting. In
Z-scheme water splitting, reversible redox mediators (e.g., Fe3+/ Fe2+, IO3 – /I–, NO3 – /NO2 –)
are typically utilized as electron transport chains. In contrast, reverse reactions often occur in
the reversible redox mediator Z-scheme photocatalytic system, which is thermodynamically
heading in the other direction in most instances. However, the Z-scheme with a redox mediator

Figure 37. Forward and backward reactions in a Z-scheme system with shuttle redox media
Reprinted with permission from [346]. Copyright 2016 Elsevier.

has a drawback when the redox mediator pairs cause a backward reaction; in addition to this,
the redox mediators exhibit deactivation as a result of their instability. Furthermore, they
generate a light-shielding effect in an aqueous reaction medium and offer reduced diffusion
[95], [347]. Consequently, the reaction efficiency of the reversible redox mediator Z-scheme
photocatalytic system may be negatively impacted. Domen et al. achieved water splitting in
2010 utilizing a Z-scheme system assembled of Pt-loaded ZrO2/TaON as the H2 and O2
production photocatalysts, respectively, in the presence of an IO3-/I- redox mediator [348]. Z-
scheme heterostructures can be indirect or direct depending on whether an electron transport
intermediary is involved in the electron transfer. Many inorganic semiconductor-based Z-
scheme heterostructures have been widely reported to improve photocatalytic efficiency [346].
Zhou et al. [349] analyzed the synthesis and characterizations of metal-free
semiconductor/MOFs with good stability and high photocatalytic activity for H2 evolution.
They combined the MOFs of NH2-MIL-125(Ti) with g-C3N4 functionalized by benzoic acid
(CFB) to synthesize a novel composite catalyst of CFB/NH2-MIL-125(Ti) (CFBM) by covalent
bonds for the first time. The benzoic acid in the CFBM acts as electron mediator to well
separate photogenerated electrons and holes, leading to excellent photocatalytic performance
of photocatalytic hydrogen generation from water splitting under visible light irradiation.
Experimental results show that the H2 production rate of the 10CFBM is 1.123 mmol·h−1·g−1,
which is about 6 times of the NH2-MIL-125(Ti). Meanwhile, the simple physical mixture of
NH2-MIL-125(Ti) with 10 wt% g-C3N4 and the 10wt%g-C3N4/MOFs heterostructured catalyst
all show much smaller H2 evolution rate and worse stability than that of the 10CFBM.
Moreover, Liu et.al., [350] conducted photocatalytic coproduction of H2 and industrial
chemical direct Z-scheme heterostructure (Co9S8/CdS) constructed by self-recognition of
uniformly distributed metal cations during the sulfidation of a multivariate metal-organic
framework. The intimate contact of in situ formed CdS and Co9S8 into direct Z-scheme
heterostructure can facilitate interfacial charge separation and extend the oxidation and
reduction potentials to 2.09 V and −0.74 V (vs. RHE), respectively. An excellent H2 evolution
activity of 61,924 μmol g−1 with high selectivity for benzyl-alcohol (BA) oxidation in pure

106
water is achieved, ca. 21 and 16 times higher than that of isolated CdS and physical mixture of
Co9S8/CdS. This work supplies a new strategy for constructing Z-scheme heterostructure from
a solid solution to dramatically boost coproduction of H2 fuel and valuable chemicals without
any sacrificial agents. However, the construction of heterostructures is mainly confined to
limited classical MOFs semiconductors, expanding the range to other novelty MOFs is still
challenging due to the requirement for small size, well conductivity, and appropriate band gap
to form efficient heterostructures between MOFs and other semiconductors [142]. Because
from the perspective of synthesis, the MOFs materials are mostly crystalline bulk materials
with larger sizes, which creates certain difficulties for the composite with other nano-
semiconductor materials. Therefore, nano- or micro-sized MOFs maybe more conducive to the
realization close combination with other semiconductor materials. In addition, from the
perspective of improving the efficiency of the catalytic reaction, the small size can increase the
dispersion of the catalyst [351] and the contact area between the catalyst and the reaction
substrate, and expose more catalytic sites. Besides, good conductivity can improve the
separation efficiency of charges in the heterostructure. These factors are all very important for
improving the catalytic efficiency. The recent booming of nano and conductive MOFs may
provide a new solution to these bottlenecks [352].

3.4 Organic linker/ functional group/ ligands

MOFs include both organic and inorganic building blocks in their construction;
carboxylates or anions, such as phosphonate, sulfonate, and heterocyclic chemicals, make up
the organic units (linkers/bridging ligands) [353], [354]. It is essential to choose the organic
linkers to prevent interpenetrating structures from forming because MOFs with wide gaps have
the potential for developing interpenetrating designs. The selective introduction of organic
functional groups into the framework of MOF materials can effectively reduce their optical
bandgap and broaden the visible light response range. Therefore, ligand functionalization
methods have been widely used to improve the photocatalytic activity of the MOF materials
for water reduction. Both Figures 38 and 39 provided an explanation of some organic linkers
and functional groups that are used in MOFs to achieve high levels of photocatalytic activity
[132]. SBUs are a type of inorganic unit composed of clusters of metal ions or metal ions
themselves [355]. The coordination number, coordination geometry, and kind of functional
groups of the metal ions all affect this structure's geometry [356].

107
Figure 38. Give some instances of organic ligands and include the carboxyl functional group
[132]. Copyright 2016 Elsevier.

108
Figure 39. Some examples of ligands containing nitrogen, sulfur, phosphorous, and
heterocycles are employed in preparing MOFs. Reprinted with permission from [132].
Copyright 2016 Elsevier.

It has been discovered that introducing organic functional groups with high polarity diminishes
the bandgap of MOFs [144]. Apart from the aforementioned examples, photosensitizer units
can also be applied to construct MOF-based photocatalysts with visible light absorption ability.
For example, pyridine-functionalized boron dipyrromethene (BODIPY) units possess large
electron delocalization systems and have excellent response for visible light. Inspired by this,
Wen et al. synthesized two 3D MOFs, named, CCNU-11 and CCNU-12, through a mix-ligand
method, in which 4,40-biphenyldicarboxylic acid (H2BPDC) or 4,40-sulfonyldibenzoic acid
(H2SDB) was used to form the layer structure, and the BODIPY-based ligand was applied as
the pillar to form the 3D structure [168]. In virtues of the excellent light-harvesting ability, both
two materials displayed high performance for photocatalytic H2 evolution. During the
photocatalytic reaction, the BODIPY photosensitizer first excited electrons under visible light
irradiation, then were injected into the [Cd2(COO)2] cluster, and finally transferred to
[Co(bpy)3]Cl2, which resulted in a slower photogenerated carrier current sub combination rate.
Finally, the water was efficiently reduced to hydrogen on the surface of the catalyst. Similar to
earlier work, they also used this BODIPY ligand to successfully construct novel Zn-based
MOFs, which also exhibited excellent activity for photocatalytic H2 production. This is capable
of efficiently absorbing visible light from 200 to 800 nm which is thought to be the first
BODIPY-based MOF decorated with Pt nanoparticles as the co-catalyst for efficient
photocatalytic H2 generation under visible-light illumination. Impressively, Pt/CCNU-1
exhibited a remarkable H2 production rate as high as 4680 μmol g−1 h−1 with the use of L-
ascorbic acid as a sacrificial reagent from an aqueous medium, and a high apparent quantum
efficiency of 9.06% at 420 nm. To our knowledge, Pt/CCNU-1 is the most visible-light
109
photoactive Pt/MOF composite for H2 generation from water hitherto, highlighting the
promising future of MOF materials in solar-to-chemical energy conversion. Apart from
BODIPY ligand, other photosensitizers, such as Eosin Y-based ligands, can also be introduced
into MOFs as organic linkers to improve the visible light absorption range of MOF-based
materials. For instance, through combining [Cd2(COO)2(μ 2-H2O)] secondary building units
and Eosin Y-based ligands, a porous Cd-based MOF was successfully constructed by Wu's
group [197]. Due to excellent photo activity of the Eosin Y-based ligands, such MOFs
displayed high photocatalytic activity for water reduction even under visible light irradiation.
The MOF exhibited efficient photocatalytic activity for H2 evolution under visible-light
irradiation with a turnover number of 13920 and an initial turnover frequency of 7433 h–1,
which was approximately 31 times the photocatalytic efficiency of Eosin Y. Based on earlier
works, we can recognize that ligand functionalization method is a versatile and reliable tool for
improving the photocatalytic activity for H2 evolution of the MOF-based photocatalysts.
Moreover, there is also essential views on the role of ligands in maintaining the stability of
MOFs in unfriendly environments. Seo et al [358] propsed Fe-MOFs, MIL-100(Fe) and MIL-
101(Fe) that possess 3D structures, where only MIL-100(Fe) have stability in thermal and
water where MIL-100(Fe) with a tricarboxylate linker was theoretically more stable than other
Fe-MOFs (MIL-53, MIL-88 and MIL-101) with a dicarboxylate linker. Similarly, Wen and
colleagues explored the photocatalytic hydrogen productivity of amino-functionalized MIL-
101(Cr) [359]. They found that adding rhodamine B (RhB) photosensitizer resulted in
hydrogen production from pure NH2-MIL-101as depicted in Figure 40A. After the absorption
of visible light by the RhB molecule, an excited photoelectron is formed from the LUMO state
of the RhB molecule. The generated photo-excited electron transfer occurs not only to the Pt
NPs (path 1) but also to the unsaturated Cr(III) sites (path 2). The injected electrons to the
unsaturated Cr(III) sites are occasionally transferred to the Pt NP sites (path 3). Such an electron
transfer pathway can be evidenced by the significant decrease of emission intensity of RhB in
the presence of NH2-MIL-101(Cr) with and without Pt. Moreover, the incident visible light
absorption by the organic linker of NH2-MIL-101(Cr) produces the photo-excited electrons
which firstly transfer to chromium-oxo through a linker-to-cluster charge-transfer mechanism
(LCCT) and then to the Pt NPs (path 4). Finally, the proton which gathered on the surface of
NH2-MIL-101(Cr) by hydrogen bonding are transferred to the Pt NPs or unsaturated Cr(III)
sites in NH2-MIL-101(Cr) by water and react with photoelectrons to produce H2 (paths 5 and
6). In this case, the porous MOFs not only served as electric conductors to promote the electron
transfer in facilitating the separation of photoelectrons from RhB, but also act as photo-electron
generators to enhance the activity of H2 production.
Porphyrins are used in conjunction with amino groups to increase the photosensitivity
of MOFs. Wang et al [360] synthesised ultrathin porphyrin-based metal-organic framework
(PMOF) with Ti-oxo clusters. In the photocatalytic process, the photo-induced electrons
transfer from the porphyrin-based ligands to the Ti-oxo clusters through π bond significantly
enhancing the charge carriers’ separation efficiency and thereby, exhibiting a photocatalytic
hydrogen evolution rate of as high as 8.52 mmol g−1 h−1 in the presence of Pt as cocatalyst
under a wide range of visible light irradiation up to 700 nm. The photocatalytic reaction over
PMOF is a typical LMCT process, i.e., the photoinduced electrons transfer from porphyrin-
based ligands (light harvester) to Ti-oxo clusters (catalytic center), and thereby the charge
carriers’ separation is significantly enhanced. Moreover, the long-term course reaction
confirms the excellent water-resistance and photo-stability of PMOF in photocatalytic
hydrogen evolution [360]. In addition to serving as an organic ligand in MOFs, porphyrin
molecules can be metalized to boost the efficiency of photocatalysts [324], [173], [361]. For
example, He et al. [362] used a solvothermal method to synthesize zirconium-porphyrinic
hollow nanotubes MOFs and then they mounted a single noble-metal atom (Pd, Ru, Pt, Au, Ir)

110
on a porphyrinic linker. It has been discovered that HNTM-Ir/Pt has a superior performance in
photocatalytic hydrogen production (201.9 mol h-1g-1), where the porphyrinic Pt units acted as
catalysts, and the porphyrinic Ir units acted as photosensitizers. Additionally, Ti-MOF-NH2
was developed by Matsuoka and coworkers [218] that describes the hydrogen production from
an aqueous medium over amino functionalized Ti(IV) metal–organic framework (Ti-MOF-
NH2) under visible-light irradiation. Ti-MOF-NH2, which employs 2-amino-
benzenedicarboxylic acid as an organic linker, has been synthesized by a facile solvothermal
method. The action spectrum of TiMOF-NH2 along with the results of in-situ ESR experiments
demonstrate that hydrogen production takes place efficiently through the LCCT mechanism
illustrated in Figure 40B. Pt nanoparticles as co-catalysts are then deposited onto Ti-MOF-NH2
via a photo deposition process (Pt/Ti-MOF-NH2). The XRD (Figure 40C) and N2 adsorption
measurements reveal the successful formation of a MOF framework structure and remaining
its structure after deposition of Pt nanoparticles. The observable visible-light absorption up to
around 500 nm can be seen in the DRUV−vis spectrum of Ti-MOF-NH2, which is associated
with the chromophore in the organic linker. Ti-MOF-NH2 and Pt/Ti-MOF-NH2 exhibit
efficient photocatalytic activities for hydrogen production from an aqueous solution containing
triethanolamine as a sacrificial electron donor under visible-light irradiation. The longest
wavelength available for the reaction is 500 nm. The results obtained from wavelength
dependent photocatalytic tests and photocurrent measurements as well as in-situ ESR
measurements demonstrate that the reaction proceeds through the light absorption by their
organic linker and the following electron transfer to the catalytically active titanium-oxo
cluster.
In addition, carbonyl-containing functions include urea, amide, oxalamide, squaramide,
ketone, imide, carboxyl, and some different groups like hydrazine serve the purpose of organic
linkers in MOFs [363], [364]. Utilizing the concepts of deploying organ catalyst hydrogen bond
donors and limiting homogeneous urea catalyst self-association, Hupp and colleagues., [365]
published the first urea-containing MOF in their study. In addition, carboxylate-based MOFs
have a large surface area, homogeneous pore size distribution, highly ordered crystallinity, and
clearly defined reticular chemistry.

111
A

C
B

Figure 40. a) Possible process for visible light-induced photocatalytic H2 from RhB solution
over Pt/NH2-MIL-101. Reprinted with permission from [359]. Copyright 2014 Elsevier. b) A
schematic representation of a photocatalytic hydrogen generation reaction using the LCCT
mechanism over Pt-supported Ti-MOF-NH2. c) XRD patterns of (a,b) Ti-MOF and (c,d) Ti-
MOF-NH2 before (a,c) and after (b,d) photo deposition of Pt nanoparticles. Reprinted with
permission from [218]. Copyright 2012 Elsevier.

Margaritis et al. [366] introduced the initial combination of pyridine-2,6-dimethanol with


benzene-1,4-dicarboxylic acid has provided access to three new mixed-ligand MOFs based on
trinuclear SBUs. They are the first MOFs containing H2pdm in its neutral or anionic form.
Moreover, carbon-based materials also serve as an excellent organic ligand in the progression
of catalytic hydrogen production using MOFs [144]. Furthermore, sulfur-based functional
groups such as thiol, sulfides, sulfonate, sulfonic acid, and some additional functionalities such
as thiourea, thiadiazole, and thiocathecole are employed as organic linkers inside MOFs [367],
[316]. A thiol-functionalized MOF with the same isostructural profile as UiO-66, Zr-dmbd,
was designed and synthesized by reacting ZrCl4 and mercapto-1,4-benzene dicarboxylate
(dmbd) [368]. Similarly, as a family of significant functional materials, quantum dots (QDs)

112
offer unique physicochemical properties such as a limited luminescence spectrum, long carrier
lifetimes, and high light absorption quantum yield because QDs are tiny in size, the porous
structure of MOFs can serve as an excellent host for functional QDs [369], [370]. Ling Wu et
al. [371] develop the highly competitive photocatalyst MoS2/UiO-66/CdS by transforming CdS
twice with UiO-66 and MoS2. This technique has two distinct features: the synthetic strategy
is simple and straightforward, and the MoS2 is directly deposited upon the surface of UiO-66
without the assistance of surfactants, establishing an intimate interfacial interaction between
MoS2 and U6-CDs, which would be beneficial for improving interfacial electron transfer
efficiency and increasing photocatalytic H2 evolution activity. Increasing the charge of the
metal typically results in increased strength for a given linker which, in turn, contributes to an
increase in the hydrothermal stability of the MOF [372]. Shen et al. [373] embarked a detailed
study of the structure–photo activity relation with an isoreticular series of UiO-66-X (X = H,
NH2, NO2, Br) and firstly showed that the electronic effect of the ligand substituents has a
systematic and important impact on the photocatalytic activity of the MOFs. UiO-66 and its
functionalized compounds are zirconium-based microporous MOFs, which are built from
Zr6O4(OH)4 oxo-clusters related together by 1,4-benzene-dicarboxylate linkers. These
materials stand out among the numerous MOFs due to their high stability [180], [374]. XRD
has been used to evaluate the crystallinity and structural topology of UiO-66-X (X = H, NH2,
NO2 and Br) and optical absorption of the UiO-66-X samples has been investigated using
UV/vis spectroscopy as shown in Figure 41. It is worth noting that in many cases, the
introduction of functional groups into the organic scaffold would alter the topology of the
desired MOFs. Fortunately, comparison between the diffraction patterns of the parent UiO-66
and those of the functionalized samples demonstrates the formation of pure, topologically
equivalent phases for UiO-66-X (X = NH2, NO2 and Br). The BET surface areas of UiO-66-X
(X = H, NH2, NO2 and Br) are determined to be 1141.1, 732.2, 464.8 and 455.9 m2 g−1,
respectively, suggesting the large surface areas of UiO-66-X. In conclusion, the photocatalytic
activity of MOF catalysts can be strongly affected by using functionalized linkers.

a) b)

UiO-66
UiO-66-NH2
UiO-66-NO2
Absorbance

UiO-66-Br
Intensity

200 300 400 500 600 700 800


Wavelength 2 thetha/degree

Figure 41. a) UV-vis absorption spectra of UiO-66-X (X = H, NH2, NO2 and Br). b) XRD
patterns of the UiO-66-X (X = H, NH2, NO2 and Br) [374].

The functionalization of ligands and encapsulation of co-catalysts within MOF frameworks


are commonly suggested as the best PSM options to improve the visible-driven photocatalytic
113
activities of MOF photocatalysts. From the perspective of the photocatalytic capability of
MOFs, a technical breakthrough in the application is highly dependent on improving redox-
activities and visible-light responsiveness towards various applications (e.g., N2/CO2 reduction,
H2 production, organic photo redox reactions, and pollutants (aqueous and gaseous)
degradation) as discussed in the review [150]. Such limitations can be overcome by a number
of technical options such as the substitution of visible-light responsive moieties (such as ligand
functionalization and photosensitization). In another study, Wang et al. [375] unraveled the
impact of ligand composition on photocatalytic hydrogen evolution over Ti-based MOFs. Ti-
based MOFs have drawn attention because of the following merits: a) the small ionic radius of
Ti4+ yields a strong Ti-O bonding; b) the high coordination number of Ti4+ increases the
structural diversity of the metal-oxo clusters; c) the reversible redox transition between
Ti4+/Ti3+ under photo-excitation enables the Ti-based MOFs to store charge effectively. The
MIL-125-Ti compound has been reported in 2009 as the first crystalline porous carboxylate-
based Ti-MOF [217]. They compared triethanolamine with methanol as hole scavengers in the
photocatalytic hydrogen evolution reaction. The XRD results of the as-prepared MOF powders
(Figure 42a) show that both single-ligand MOFs exhibit high crystallinity with sharp
distinguishable peaks that can be attributed to the reported crystal structures of MIL-125-Ti
and NH2-MIL-125-Ti [376], [217].

Figure 42 (a) XRD patterns and (b) Kubelka-Munk plots of the as-prepared single-ligand
MOFs: MIL-125-Ti in grey, NH2-MIL-125-Ti in blue. The inset in (b) corresponds to the
Kubelka-Munk function of the ligands, H2BDC (grey) and NH2BDC (blue). The white and
yellow powders correspond to MIL-125-Ti and NH2-MIL-125-Ti, respectively [376], [217],
[377].

Figure 42b summarizes the light absorption properties (measured via DRS) of the two MOFs
as well as their corresponding organic ligands. As can be seen in the inset, molecular H2BDC
absorbs only in the UV region with an absorption edge at 341 nm, while the absorption of
NH2BDC exhibits an absorption peak centered around 400 nm, which can be explained by
excitation of lone-pair electrons from N via n-π* transition [377]. Further This work provides
a comprehensive comparison of photoactive MIL-125-Ti, NH2-MIL-125-Ti and their mixed-
ligand variations with respect to their photo stability, charge separation characteristics and
photocatalytic performance toward hydrogen evolution in TEOA and MeOH solutions.
The MOF sector has been a hot research spot and has been actively investigated during
the past twenty years. Introducing an organic linker made it easy to use MOFs for
photocatalytic production. Although MOF synthesis through linker design is sometimes tricky

114
from an experimental point of view, many MOFs with predesigned channels or cavities have
been accomplished [354]. It can be concluded that ligands are quite important to preserve the
porous structure of the MOFs during the catalytic process. Besides, the ligand engineering of
MOFs has a great influence on their water splitting performance, especially for the
photocatalytic water splitting. Furthermore, the synthesis methods are explained in detail in the
next sections, which greatly affect how MOFs function.

3.5 Synthesis methods of MOFs

The geometrically robust crystalline MOF structures, with conventional porosity in up


to 93% of the MOF crystal volume, are generated when the aromatic organic acids are
connected to metal-containing components [378], [379]. The hydrogen production experiments
followed the reported procedures by El Naggar et al. [380], which included semiconductors,
aqueous solution suspension, a gas-collecting vessel, and a visible-light lamp as the radiation
source. The MOFs were successfully synthesized using different metals, organic linkers, and
solvents. They have other textural characteristics and different solvothermal reliabilities [180].
MOFs are synthesized under two main branches 1. Liquid phase synthesis 2. Solid phase
synthesis [381]. The liquid phase is further divided into different types, as shown in Figure 43.
In general, the conditions for MOF synthesis should be chosen so that it is possible to establish
metal-ligand connections, but they should also be able to be broken and reformed to allow for
structure propagation [382].

115
Figure 43. Different synthesis methods of metal-organic frameworks

3.5.1 Liquids phase synthesis

The wet chemistry process is increasingly being deployed to manufacture porous


materials such as metal oxides [383] and zeolites [384]. This method has recently been adopted
to prepare MOFs, where three chemical constituents (selected solvent, metal salt, and
appropriate organic linker) of high purity are required [385]. Typically, the metal salt and
organic ligand are combined in a suitable container with the selective solvent, or the metal salt
and organic ligand are absorbed individually in the solvent and then blended in a single pot
[386]. There are many dependable methods for liquid-phase synthesis, which are described as
follows; the most basic and extensively used approach is solvothermal synthesis. The term
"Hydrothermal" corresponds to any homogeneous or heterogeneous chemical reaction that
occurs in the presence of water and certain physical circumstances, whereas "solvothermal"
refers to a chemical reaction that occurs in the presence of an organic solvent [387], [381].
Usually, a high boiling point solvent (such as DMF, DEF, or DMSO) is used to combine a
metal salt and a multitopic organic linker. This synthesis needs to be carried out in closed
containers, such as a glass vial or an acid digestion vessel, so that the appropriate conditions
may be created for the reaction to be effective [388], [389]. In this classic synthesis method
metal precursor and organic linker should be slightly soluble when the mixture reaches the

116
desired temperature. In the solvothermal process, the precursor is combined with the
appropriate solvent before being sealed in an autoclave and heated above the boiling point of
the selected solvent [390]. In contrast to this, Zhang et al [454] synthesized Co-MOF-74 in
low-temperature NH3-SCR process. The most optimal hydrothermal temperature for synthesis
Co-MOF-74 is 100 °C. MOFs are made up of metal-based nodes (e.g., Zn2+, Zn4O6+,
Zr6O4(OH)412+, Cr3(H2O)2OF6+) that are linked by organic linkers, are also synthesized by
hydrothermal process [391]. In most cases, solvothermal interactions between metal precursors
(inorganic salts) and organic linkers occur in a single pot during de novo MOF synthesis. Many
typical MOFs with exciting properties, including MIL-53, MIL-100, MIL-101, MOF-74, UiO-
66, and PCN series, have been synthesized using a similar synthetic technique [392]. The
hydrothermal synthesis of MOFs is viable for two reasons. First of all, the issue of solubility
can be minimized, Second, using the same experimental settings, the nucleation process can be
started quickly to create uncommon complexes [393]. The main disadvantages of this
technology include waste generation, the dangers and expenses connected with anions such as
nitrates and chlorides, and the risks of oxidation and corrosion [381]. In their study, Schlesinger
et al. [394] analyze various synthesis methods, one of which is the solvothermal approach to
synthesize Cu-HKUST-1 using Cu (NO3)2 3H2O, H3BTC as a precursor. Moreover, Hausdorf
et al. [395] studied zinc oxide/zinc hydroxide 1,4-benzene dicarboxylate MOFs synthesis
(MOF-5, MOF-69c), which has been used in various fields. The most significant negatives of
this technique include excessive volumes of trash, the risks, and expenses connected with
anions like nitrates and chlorides, as well as the potential for oxidation and corrosion. Other
researchers are working hard to improve the photocatalytic potential of MOFs by either Solvo
or hydrothermal synthesis [396], [217].
Moreover, the prefabrication of inorganic building blocks is another method for
producing MOFs. These prefabricated polynuclear coordination complexes have structures and
functions similar to those of the inorganic MOF bricks. By way of illustration, the trinuclear
oxo-bridged iron (III) acetates MIL-88 and MIL-89 were manufactured by exchanging the
monocarboxylate (acetate) ligand for dicarboxylate molecules [397], [398]. Guillermet al.,
[399] synthesized the precursor to manufacture porous zirconium dicarboxylate with the UiO-
66 architecture, as shown in Figure 44. This cluster is a member of the family of organically
functionalized tetravalent transition metal oxo clusters. These clusters are surrounded by an
inorganic metal oxide M-O-M core (M = Ti, Zr, Hf...) by a carboxylate ligand shell. These
clusters have widely applied as molecularly and structurally well-defined building blocks in
synthesizing inorganic-organic hybrid materials. This technology not only makes it possible to
synthesize new porous Zr dicarboxylate and put them to use in applications such as catalysis,
gas storage, thin film production, and drug delivery, but it also makes it possible to find new
structures in MOFs.

117
methacrylates

Dicarboxylate

T= 25-100°C, DMF

Figure 44. A schematic depiction of the synthesis of zirconium dicarboxylate MOF Reprinted
with permission from [399]. Copyright 2010 Elsevier.

The next method is Sonochemistry which is the study of molecules that undergo
reactions as a result of the use of solid ultrasonic radiation [400]. Sonochemical methods can
lead to homogeneous nucleation and a substantial reduction in crystallization time compared
with conventional oven heating when nanomaterials are prepared [403], [401]. The
Sonochemical approach typically takes advantage of the energy provided by ultrasound
(frequency range: 20–1,000 kHz) to speed up the reaction in the solvent between metal centers
and organic ligands [402]. When preparing nanomaterials, Sonochemical approaches can result
in homogeneous nucleation and a significant reduction in crystallization time compared to
standard oven heating [403]. For the first time, Seung et al. [404] manufactured a high-grade
MOF-5 crystals with diameters ranging from 5 to 25 nm which were synthesized using a
Sonochemical technique that required much less time (approximately 30 minutes) than
conventional solvothermal synthesis. Here, the effects of various Sonochemical synthesis
parameters on the MOF-5 crystal properties were examined, and the properties of the sample
were compared with those of a MOF-5 sample prepared using a conventional solvothermal
method. To do this, MOF-5 is used, which has a straightforward cubic architecture composed
of zinc oxo clusters and terephthalate linkers and serves as a precursor material for several
reticular MOFs. Figure 45 (a1) and (a2) show the SEM images of the S-MOF-5 and C-MOF-5
samples. Both samples were single-phase cubic crystals but S-MOF-5 was 5–25 mm in size,
which is approximately 60 times smaller than C-MOF-5 (ca. 900 mm) on average. Such a size
difference is commonly observed in Sonochemistry [405]. As shown in Figure 45(b), both S-
MOF-5 and C-MOF-5 samples exhibited type I isotherms with no hysteresis. S-MOF-5 had
virtually the same textural properties to those of C-MOF-5. The Langmuir surface area of S-
MOF-5 was 3208 m2 g-1 with a total pore volume of 1.26 cm3 g-1, while the corresponding
values for C-MOF-5 were 3200 m2 g-1 and 1.21 cm3 g-1, respectively. Figure 45(c) shows the
thermogravimetric analysis results of S-MOF-5 and C-MOF-5.

118
a)

b) c)
1000
Absorbed N2 (cm3/g)

100

800
80

Weight loss (%)


600
60

400
40

200
20

0
0
0.0 0.2 0.4 0.6 0.8 1.0 300 400 500 600 700 800 900

Relative pressure (P/P0) Temperature (K)


Figure 45. Comparison of the physicochemical properties of S-MOF-5(1) and C-MOF-5(2):
(a) SEM images, (b) N2 adsorption isotherms, and (c) thermogravimetric analyses. Reprinted
with permission from [404]. Copyright 2008 Elsevier.

Furthermore, three-dimensional coordination polymer nanostructures, [Pb(4-


bpdh)(NO3)(H2O)]n (HMTI-1), and [Pb(4-bpdh)(NO3)]n (HMTI-2), {4-bpdh = 2,5-bis(4-
pyridyl)-3,4-diaza-2,4-hexadiene}, were produced at various dosages using a Sonochemical
technique [400]. The process is considered a quick, cost-effective, and ecologically benign
approach for generating MOFs. MOF-5, Fe-MIL-53[406], and MOF-177 [407] are a few of the
MOF materials have been synthesized using this ultrasonic method. This approach has unique
advantages over other synthetic processes, such as it is believed to be a speedy, cost-effective,
and environmentally friendly way to synthesize MOFs. In addition to the short time needed for
the reaction, uncommon reaction conditions, such as high pressure and temperature, are not
required. Compared to the standard hydrothermal procedures, the Sonochemical processes can
produce homogeneous nucleation centers with a significant reduction in the time required for
crystallization [412], [409]. Ultra-sonication has the potential to be utilized in some
circumstances in the solvothermal/hydrothermal synthesis technique in the role of an assisted
pre-primary mixing tool [410]. Further different studies demonstrate the use of this method in
MOFs synthesis [408], [411], [406], [412].
Other than this, microwave method is proposed. This approach was first employed to
produce nanosized metal oxides because it is a quick, clean, and cost-effective method [413].
Microwave (MW) irradiation for synthesis has been used because it reduces crystallization
periods, controls phase and shape, and distributes particle size [414]. Irradiation with
microwaves provides the energy necessary for the production of MOFs. The interaction

119
between electromagnetic waves and mobile electric charges in the solution, such as polar
solvent molecules or ions, determines how this phenomenon manifests itself. This method
offers many advantages, including its high efficiency, phase selectivity, ability to reduce
particle size, and control over morphological characteristics [415]. The basic idea behind the
microwave approach is to create an electromagnetic field that stimulates the interaction of
moving electric charges in the solution with the area (polar solvent, metal ions, deprotonated
organic ligand) [416]. On the other hand, concentrated reagent solutions are not recommended
since they can produce metal oxides [417]. This standard approach was used by Zhao et al.
in 2011 to manufacture Cr-MIL-101, which had a size of around 100 nm, absorbed benzene at
a rate of up to 16.5 mmol g-1 at 288 K, and had a pressure of 56.0 mbar. This study demonstrates
that MOF materials can be synthesized in less time using microwave irradiation [418]. In 2013,
Liang et al. [419] described the synthesis of MIL-140 by this method. In this investigation, they
could generate a purer phase and higher quality goods in a significantly shorter time (>95%)
than when they used the conventional electrical heating method. Figure 46 displays SEM
images of the frameworks of CE heating at 220°C. Compared to the considerably more
disordered aspect of the materials produced by CE heating, the materials formed by MW
heating showed plate-like crystallites, as shown in (Figures 46 a-e). A comparison of SEM
images reveals that the materials generated with MW irradiation (Figure 46 a and c) had a
higher degree of crystallinity than those formed with CE heating. In a related manner, Sabouni
et al. [420] announced 2012 their success with a microwave-assisted synthesis of CPM-5
Figure 46 (f, g), which illustrates the SEM micrographs of the CPM-5M (1) and CPM-5(OV)
samples. Comparing these photos, it is evident that CPM-5M, i.e., microwave synthesis, has a
significantly more uniform morphology and particle sizes that are 10 times smaller than those
of CPM-5(OV). Rapid and consistent heating of the reactants accelerates the reaction rate and
produces highly pure products with a narrower particle size dispersion and smaller particle size.
Microwave
heating
Conventional
heating

Figure 46. SEM views of (a) MIL-140A-MW, (b) MIL-140A-CE, (c) MIL-140B-MW, (d)
MIL-140B-CE and (e) MIL-140A-NH2-MW (scale bar = 3 mm). Reprinted with permission
from [419]. Copyright 2013 Elsevier. Scanning electron microscopy (SEM) views at two

120
various magnifications of (f) CPM-5M and (g) CPM-5(OV) Reprinted with permission from
[420]. Copyright 2012 Elsevier.

CPM-5 generated by microwave-assisted synthesis had an extremely high surface area of 2187
m2 g-1 compared to conventionally synthesized samples, which had a surface area of 580
m2g-1. The microwave approach has several drawbacks: i) using a microwave reactor raises
particular concerns about health and safety [421]. ii) A high temperature of the reaction can be
obtained by fast microwave heating in a sealed vessel, iii) solvents with lower boiling points
should be employed under pressure (closed vessel conditions), and iv) Temperatures are
significantly higher than their boiling points [381]. The microwave approach produces MOFs
with textural characteristics, shape, and adsorption affinity comparable to those obtained from
solvothermal synthesis [422]. Despite mild fabrication conditions, quicker crystal growth, and
ease of scale-up, the electrochemical approach has emerged as a promising and
environmentally friendly method for MOF synthesis [423], [424]. Because of the ability to
control the reaction conditions during the synthesis process, electrochemical synthesis has a
significant advantage in industrial domains where metal salts can be created continuously
[425]. The general principle is dependent on three pre-steps: mixing an organic linker and an
electrolyte in a selective solvent, anodic dissolution of metal ions, and increasing the pH of the
solution by increasing the concentration of HO [426], [427], [428]. Anodic dissolution causes
MOF crystals to start forming at the electrode surface, which produces a thin MOF layer [429].
The first report on the synthesis of MIL-100(Fe) via electrochemical deposition at elevated
temperatures and pressures was published in 2013 by Campagnol et al. [430]. In addition,
Rosenthal et al. [431] constructed direct electrochemical syntheses of four iron-based MOFs
via potential controlled oxidation of dissolved metal cations, as shown schematically in
Figure 47a. Although electrochemical synthesis is an active route for the generation of MOF
layers, anodic electrodeposition, which consists of anodic generation of the metal ions required
for MOFs formation in a solution containing organic linkers, has the limitation of being able
to synthesize MOFs with the same metal as the substrate [430]. Stassen et al. [432] recently
reported anodic and catholic electrochemical membrane of lamination of UiO-66 utilizing
zirconium foil as the sole metal source. At first, the solution (BDC: HNO3: H2O: AA: DMF =
1:2:4:5/10/50:130) was produced and heated to 383 K, as shown in Figure 47b. An oxide bridge
layer was then developed that improved the adherence of the MOF layer to the zirconium
substrate during anodic deposition. Cathodic deposition, on the contrary side, had the
advantage of broad substrate versatility. This is useful for online analytical sampling and
diluting volatile organic chemical concentration applications.

121
a)

b)

Ligand pool NO 3-
ZrO
H2O

UIO-66

Favored at low AA DFM = 383K Favored at high AA

Figure 47. a) Fe-BTC MIL-100(Fe) layer created using solution A at 190 °C for the Fe
substrate. Reprinted with permission from [431]. Copyright (2021) American Chemical
Society. b) Illustration of the anodic and cathodic electrochemical deposition processes
Reprinted with permission from [432]. Copyright 2015 Elsevier.

More specifically, the electrochemical process involves submerging an electrode in a


linker solution containing an electrolyte material to produce metal ions. Numerous known
MOFs were produced by electrochemical synthesis via anodic dissolution at milder
temperatures and ambient pressure utilizing the same metals as electrodes [426]. Fe-MIL-100,
on the other hand, was not synthesized at average temperature by electrochemical synthesis,
but it was successfully created at high temperature and high pressure by anodic dissolution
[430]. Electrochemical processes are used to create numerous more MOFs [433], [434], [435];
slow evaporation is the oldest traditional method for producing crystals of various chemical
compounds, and it has recently been applied in a relatively simple procedure to synthesize
several MOFs. Because it is a low-temperature process, it is occasionally used, but more time
is required during MOF syntheses [381]. The products formed under these conditions may not
be identical to those achieved by other methods that produce highly crystalline materials [436].
The slow evaporation method concentrates an initial material solution by slow evaporation of

122
the solvent at a constant low temperature. Co-solvents are sometimes used to increase reagent
solubility and speed up the process by allowing low-boiling solvents to evaporate faster [381].
For instance, using a slow evaporation process, Ameloot et al. [437] analyzed direct patterning
of oriented metal–organic framework crystals (HKUST-1) via control over crystallization
kinetics in Clear Precursor Solutions. Moreover, David j. [438] synthesized four MOFs using
the same method: MOF-5, MOF-74, MOF-177, and MOF-199. In addition, a significant
number of publications make use of this technology to produce various types of MOFs [439],
[440], [387].
Furthermore, Spray-drying was formerly utilized to create dry suspensions and micronize and
crystallize active species for various applications in the food, medicine, and construction
industries [441]. The spray drying process begins with atomizing a precursor solution into a
spray of microdroplets using a two-fluid nozzle [442]. This is performed by infusing a solution
of precursors at a specific flow rate via the diffuser's central port while supplying compressed
air or nitrogen gas through the surrounding ports. A droplet of a precursor is suspended in a
gas stream and heated to a specific temperature, causing the solvent to evaporate and the
precursors to spread radially to the surface [381]. This technology has been developed to allow
the collection of dry MOFs in the form of micro-spherical concrete structures as a continuous
process [443]. Furthermore, the used solvent is easily recoverable; hence, this procedure can
be employed efficiently to reduce consumption costs and waste. Additionally, researchers have
conducted additional research on this technique for producing MOFs [444], [445], [446].

3.5.2 Solid phase synthesis

Grinding solid reagents in solvent-free or low-solvent settings has developed as a more


straightforward and efficient synthetic methodology that can replace traditional solvent-
intensive procedures [447]. It is challenging to acquire single crystals and thus determine
product structure in solid-state/phase synthesis; otherwise, solution-phase synthesis is a simple
approach [381]. The synthesis of porous MOFs has been dramatically enhanced by applying
mechano-chemical techniques [448]. The functional requirement for mechano-synthesis has
developed to create some MOFs, which cannot be manufactured using the usual wet synthesis
methods due to their low solubility [449]. Most reports of mechano-chemical synthesis use
mills with small balls (weight 1 g) to achieve interactions between metal salts and organic
linkers that yield enough MOF with high surface area [450]. Braga et al. [451] have reported
several one-dimensional coordination polymers through this method; they used powder X-ray
diffraction patterns to establish the crystalline nature and simulation studies to confirm the
structural identity. The technique of mechanochemical synthesis to create MOFs with high
surface areas is proven for two model systems [450]. Mechanochemistry is an approach for
synthesizing one, two, and three-dimensional metal-organic compounds currently employed
by several groups, mainly focusing on synthesizing new structures [453]. Beldon et al. [454]
synthesized Zeolite Imidazolate Frameworks using Mechanochemistry. They used ZnO as the
starting material and established a mechanochemical technique for room-temperature and
topologically selective ZIF production and discovery. Using these procedures, the rapid and
quantitative synthesis of practically all significant MOF families, including MOF-5, HKUST-
1, MOF-74, pillared MOFs, ZIFs, and UiO-66, has been proven. James et al. [452]
demonstrated the synthesis of MOFs by a mechanochemical reaction for the first time. The
synthesis of copper (II) isonicotinic MOF came from the neat grinding of a dry combination of
copper acetate and isonicotinic acid (Figure 48), Cu(INA)2 (INA ¼ isonicotinic), with acetic

123
acid and water byproducts partially occluded in the pores. This work demonstrates that
Mechanochemistry is effective for a fast, convenient, and less expensive synthesis of MOFs.

Neat grinding

10 min

copper(II) acetate
Isonicotinic acid
monohydrate
copper(II)
isonicotinate MOF,
Figure 48. First mechanochemical MOF synthesis reaction between copper acetate and
isonicotinic acid, resulting in the formation of copper (II) isonicotinic MOF, Cu(INA)2 MOF.
Reprinted with permission from [452]. Copyright 2006 Elsevier.

N O OH
a) ZnO + fum b)
+
LAG LAG LAG LAG N HO O
EtOH or MeOH 1:1EtOH : H2O 4 eq. H2O 3 eq. H2O

NO3- SO4-2
ILAG

1 1.2 H2O 1.5 H2O 1.4 H2O


3-D polymer 2-D polymer 1-D zig zag polymer 1-D linear polymer

Figure 49. a) The mechanochemical LAG arrangements of zinc oxide-based small-pore pillared
MOFs. Reprinted with permission from [455]. Copyright 2009 Elsevier. b) Zinc oxide
mechanochemical ILAG conversion into large-pore MOFs guided by residues of nitrate or
sulfide salts. Reprinted with permission from [456]. Copyright 2010 Elsevier.

Figure 49 a demonstrates that the presence of guests is significant for mechanochemical MOF
synthesis and in the synthesis of pillared MOFs by grinding of zinc oxide (ZnO), fumaric acid,
and symmetrical linker molecule 4,40 -dipyridyl or trans-1,2-bis(4-pyridyl) ethylene, even so,
if the reaction was carried out in the absence of a quantity of DMF sufficient to fill the pores
of the expected structure in 30 min [455]. Furthermore, tiny concentrations of alkaline metal
or ammonium sulfate (down to 300 ppm based on Zn) were particularly successful in activating
the reaction mixture, yielding a topological isomer of the same MOF with a hexagonal structure
based on Kagome-topology sheets of zinc terephthalate [456] as shown in Figure 49b.
Furthermore, Tanaka et al. demonstrated a simple grinding synthesis of zinc-based ZIF-8 [457].
Water was formed as a byproduct of the conversion of ZIF-8 from ZnO and an imidazole
ligand. This method involves mixing a powder mixture of ZnO and imidazole ligand at the
stoichiometric ratio at room temperature and pressure. There is no need for a solvent, and ZIF-
8 can be made simply by physically combining ZnO with the imidazole ligand at the
stoichiometric ratio shown in Figure 50. The kinetic energy supplied during milling can have

124
several effects on the reaction including reduction in particle size, formation of defects and
dislocations in crystal lattices, local heating and local melting. In addition, collisions between
particles can lead to agglomeration formation and local deformations. The polycrystallization
of ZIF and inclusion of ZnO are attributable to three competitive formation processes: (1)
breakage of nanoparticle agglomerates, (2) the complex forming reaction between ZnO and an
imidazole ligand, which may be facilitated by the autogenous water, on the surface of ZnO and
(3) the agglomeration of the nanoparticles, resulting in the formation of a ZIF coveroverlying
ZnO. The volume change during the conversion of ZnO to ZIF-8 may be accompanied by the
generation of particle distortion. When the deposition of ZIF-8 on the ZnO surface and
agglomeration of particles proceed to some extent, the mechanochemical dry conversion of
ZnO to ZIF-8 does not occur further.

Breakage of agglomerates

Conversion of ZnO to ZIF ZnO ZIF-8 Case i


ZIF-8 Un transformed ZnO

Case ii

ZIF-8 ZIF-8 polycrystalline

Agglomeration between
nanoparticles

Figure 50. Proposed mechanism for the mechanochemical dry conversion of ZnO to ZIF-8
Reprinted with permission from [457]. Copyright 2013 Elsevier.

This approach has also been used to generate Zn-MOF-74, Cd-MOF, Zr-UiO-66, Co-MOF,
Ni-MOF, Fe-MOF, and Y-MOF [458], [459], [460], [461].

We have outlined our view of the current best practices for the most common techniques and
methods used to synthesize, activate, and characterize MOFs. Different ways are used
according to the type of MOFs and the ease of reactions.

3.6 Process conditions

The significant conditions that should be maintained all through catalytic reactions are
reagent concentration, heating temperature and duration, pH of the solution, and the nature of
the precursors; which influence the morphological characteristics, crystal sizes, topologies, and
phase purity of the produced material, light source, and porosity as shown in Figure 51 [462],
[463].

125
Process
conditions

Figure 51. Process conditions that are necessary for the water splitting to produce safe and
clean hydrogen

3.6.1 Effect of pH

pH has significantly impacted heterogeneous photocatalytic efficiency due to particle


aggregation, which reduces accessible surface-active sites and blocks excitation events within
the catalyst [464], [465]. The pH also significantly impacted the ionization state of the catalyst
surface and the organic compounds, which has been demonstrated to influence the sorbate
surface interactions that affect performance due to OH radical scavenging [466], [467]. It is
regarded that the production of H2 from water splitting is dependent upon the proton
concentration, i.e., the pH of the solution [468]. Since proton reduction by the photogenerated
electron is generated throughout water splitting, the solution’s pH is an essential component
affecting system efficiency in photocatalytic water splitting [469]. As a result of the coupling
of an excited electron and an H+ ion adsorbed on the photocatalyst surface, a partial reaction
of hydrogen production occurs [470], [471]. In a report, Zhong et al. [472] proposed the
stability of MOFs in fluoride solutions and studied 11 water-stable MOFs (MIL-53(Fe, Cr, Al),
MIL-68(Al), CAU-1, CAU-6, UiO-66(Zr, Hf) and ZIFs-7, -8, -9). The initial pH of the fluoride
solution has no influence in the neutral range but has a negative effect when the pH is more
than 9.0. In the acidic solution of pH = 2.0, the XRD pattern of the material fits well with that
of H2-BDC. Based on the observation that UiO-66(Zr) is stable in a solution of pH = 2.0 without

126
F-, we suggest that the existence of fluorine leads to the collapse of the framework. It could be
due to solid hydrogen-bonding interactions between the acidic HF and the carboxylate groups
of the BDC linkers, causing the Zr-O-C bonds to dissolve. Bi et al. [212] presented MoS2
quantum dots (MoS2 QDs) and graphene on MOFs for photocatalytic hydrogen production.
Moreover, it was previously established that the pH of the solution considerably impacts
photocatalytic activity, as illustrated in Figure 52 a, b Acidic and highly alkaline environments
do not promote H2 generation. The rate of photocatalytic hydrogen generation was most
fantastic at pH 7 when the pH of the TEOA aqueous solution was increased from 5 to 9. The
status of TEOA, which was altered by the pH of the solution, was most likely the cause of the
decreased rate of hydrogen evolution. More H+ at acidic pH causes protonation of TEOA, a
less efficient electron donor, resulting in a shorter lifespan of the excited EY and reduced
efficiency of the excited dye species.

70 70

a) 62.12 b) 62.12
60 Rate of H 2 evolution 60
Rate of H 2 evolution

52.75
50 50
43.46
40 40
35.33

30 30 26,9
25.26

20 20 16.88

10.09 9.71
10 10
0.66
0
0
5 6 7 8 9 0 1 2 3 5 7
Weight ratio of MoS 2 QDs
pH
Figure 52. (a) Impact of the pH on the photocatalytic ability of MoS2 QDs/UiO-66-NH2/G (b)
performance of the different MoS2 QDs mass ratios from 1 to 7 on the photocatalytic activity
of MoS2QDs/UiO-66-NH2/G. Reprinted with permission from [212]. Copyright 2017 Elsevier.

The photocatalytic activity of MOF-253-Pt for H2 evolution was evaluated using MOF-253-Pt
at pH 8.5 under visible-light irradiation (l > 420 nm) [473]. Mariya and coworkers [474] have
studied the annulation reaction of phenol and isoprene using different types of MOFs, such as
MIL-100 (Fe), MIL100 (Cr), Cu3(BTC)2, and Fe(BTC) as catalysts. The study found that MIL-
100 (Fe) exhibited the highest activity with an intermediate Lewis acidic nature, and MIL-100
(Cr) showed meagre catalytic activity with high Lewis acidic centres. They claimed that MIL-
100 (Fe) was a potential candidate for annulation reactions and that the catalytic activity of
MOFs was in the order of MIL-100(Cr) < Fe(BTC) < Cu3(BTC)2 < MIL-100(Fe).
Moreover, Du's group designed a 2D Ni-MOF [Ni2(Pyms)4] photocatalyst for hydrogen
production [475]. In their work, the effects of particle size, original pH value, electron donor,
and dye on the photocatalytic hydrolysis efficiency were investigated. The pH values were
measured by a Sartorius PB-10 pH meter. As illustrated in Fig. 53a, compound [Ni2(PymS)4 ]n
does not decompose in the presence of either air or nitrogen until the temperature rises up to
300 °C and can be stored in air at ambient temperature for months. Moreover, after being
soaked in water with a wide pH range for five days, little change was observed in the XRD
spectra (Fig. 53b), which indicated the noticeable durability of [Ni2(PymS)4]n to both acid and
basic solutions.

127
Figure 53. (left) Thermal gravimetrical analysis of [Ni2(PymS)4]n in the presence of air (black
line) and nitrogen (red line); (right) powder XRD patterns of [Ni2(PymS)4]n after soaked in
water for five days with varied pH values adjusted by concentrated hydrochloric acid and TEA
[475].

Several photocatalytic experiments were conducted in water with the pH values varied from 7
to 10. When the pH value was equal or less than 8, no hydrogen was measured. This could be
because more TEA is protonated at low pH, which diminished its electron-donating ability in
the photocatalytic circulation and led to no hydrogen production [476]. This MOF shows a high
stability in water over a wide range of pH and efficient catalytic activity in aqueous solution
for the photocatalytic hydrogen production under illumination of LED or direct sunlight.
Electrostatic interactions between the reactants are determined by the pH value, however, each
catalyst may work more efficiently at a different pH [477]. Later on Jin et al [478], synthesized
Zr–Metal–Organic Framework as Efficient Photocatalyst for Hydrogen Production. The
solution pH had significant influence on photocatalytic activity [479]. Different pH values of
the solution were studied from 5 to 11. It can be distinctly seen that the rate of hydrogen
evolution maximized at pH 7 (18.10 mmol), while the rate of H2 evolution was decreased under
both more acidic and more alkaline TEOA aqueous solutions. The rate of hydrogen evolution
was only 0.78 mmol at pH 5 because of the protonation of TEOA at acidic pH, which resulted
in a shorter lifetime and efficiency of excited EY, and the rate of hydrogen evolution was
decreased. However, the activity of the photocatalyst exhibited a decrease following increase
of the basicity. This is because the strong alkaline conditions reduce the concentration of
H+ and lead to a decrease in the kinetics of hydrogen production [478].
Based on the findings of the studies described above, it is abundantly obvious that PH
plays a significant role in increasing the activity of catalysts. Each type of catalyst has a unique
PH range at which it operates at its most efficient and produces the highest output. It is
imperative that we keep the PH at the intended level so that we can improve performance.

3.6.2 Light source

128
The light source is directly related to the performance of the photocatalyst. Generally,
xenon (Xe) lamps, mercury (Hg) lamps, AM1.5 G, and light-emitting diode (LED) lights are
the most widely used light sources in the lab [480], [481]. However, there are big differences
between all of these artificial light sources and it is difficult to unify the standard between
different labs. For different light sources, both spectral range and light intensity vary greatly.
Different incident photon fluxes can lead to different photocatalytic performances even for the
same photocatalyst. However, most of the reports do not provide sufficient information on the
light source, which leads to difficulties in impartial comparison and evaluation. Three main
insufficiencies exist:

1. Ambiguous or incomplete information regarding light. Most reports only provide the
type, power, or wavelength band of the light source (such as 300 W Xe lamp, 500 W
Hg lamp, UV light or visible light) rather than the actual light intensity and the accurate
light spectrum or wavelength used to irradiate the photocatalyst [482].

2. Attenuation of incident photon flux after a period of operation. Each light source has
its own working life-span. A decrease in radiation should be noted, and it is best to
measure the light intensity at both the start and end of the test, making sure of the
stability of light source. If the light source is in a decay process, the photon flux may
decrease over the hours of operation.

3. Configuration of light source and reactor. The actual light intensity irradiated on the
reaction surface greatly depends on the material of the reactor, the distance between the
light and the reactor, and the irradiated area of the reactor. Even for the same weight of
photocatalyst, different irradiation areas with different suspension conditions can lead
to diverse photocatalytic performances.

Owing to the above-mentioned insufficiencies, accurate information regarding the light source
is often not available and inconsistent, unrepeatable experimental results are obtained. It has
been confirmed by that irradiation at different wavelengths, even within the same visible range,
can lead to various photocatalytic performances [483]. For example, a light source with a
wavelength of 420 nm and illumination intensity of 100 mW cm−2 was used. The wavelength
or spectrum and the light intensity can be measured by special spectrometers and the type of
intensity meter used should be provided. Besides, the correct use and calibration of the light
intensity meter for the spectrum is important. Considering that the different light intensity
meters purchased from different companies have different usage regulations, it is imperative to
follow the instructions carefully and faithfully to assure the accuracy of data. Calibration can
rely on the use of a spectroradiometer and radiometer (standard CEN/TS 16980-1-2016), or
professional institutions, such as national metrology institutes, or a third party can be engaged
to calibrate instruments [484].

Solar energy used as light is the most efficient solar path to hydrogen since it does not
have the inefficiencies associated with thermal transformation or with the conversion of solar
energy to electricity followed by electrolysis. Therefore, the most promising method of
hydrogen generation using a source of renewable energy is that based on water decomposition
by means of photoelectrochemical or photocatalytic technologies using solar energy (sunlight)
[485]. When supplying photocatalyst activity, the wavelength or spectrum of the light source
and the light intensity irradiated on the photocatalytic system are the critical parameters that
must be determined [486], [487]. There are two types of light irradiations i.e., UV-light which
have wavelength between 200 and 400 nm, while visible light has wavelength between 400
and 800 nm [488]. The effectiveness of photocatalytic water splitting can be enhanced by

129
increasing light intensity with energies more significant than the activation threshold [489].
The sun emits radiation from X-rays to radio waves, but the most intense solar radiation occurs
in the visible light range [490]. According to some reaction models, when light intensity
increases, the reaction order reduces from one to one-half [491], [492], [493], [494], [495].
Light intensity is significant because our whole setup depends upon it, as we need it as a starting
material to proceed with the reaction. When the light falls on the surface of the photocatalyst,
it breaks down, and holes and electrons are separated. The light source used in recent studies
is 300,500,250 Xe lamps, white light, bulbs, 300 W solid-state light, etc. [496].

3.6.3 Sacrificial agent

Sacrificial agents are introduced to promote photocatalytic performance [497], [484].


The performance can be increased by adding organic species such as methanol, ethanol, phenol,
glycerol, Na2S, and Na2SO3 [498], [499]. Moreover, ions, such as I-, CN-, and Fe3+, have been
used as sacrificial agents which act as hole scavengers compared to the water alone [500].
When alcohols are used to boost H2 generation, the process is known as photo-reforming,
which causes semiconductors to promote the oxidation of organic molecules and the reduction
of H+ to H2 [501]. Increased hydrogen production can be obtained by adding a sacrificial agent,
which functions as a reducing agent and enables water reduction by photoelectrons from the
conduction band [502]. Due to the low boiling point and high H/C ratio, methanol (CH3OH) is
a good source of high-purity hydrogen [503]. Methanol is a good hole-scavenger in a water-
alcohol mixture and usually has higher effectiveness than other chemicals [504]. According to
Melian et al., using methanol as a sacrificial agent in the hydrogen production reaction serves
a dual purpose: it increases H2 production while also decomposing/mineralizing into less toxic
substances due to the methanol degradation in the process. Different ethanol concentrations
were recorded in the literature, and the amount of sacrificial reagent ranged from as low as 2%
v% [505] and 5 v% [506] to higher values of 10 v% [507], 20 v% [508] and 50 v%. Strataki et
al. [509] and Cui et al. [510] argued that the mass transfer of methanol from the solution to the
catalyst surface reduces the photocatalytic reaction rate at low methanol concentrations.
However, with more significant attention, the reactions at the interface dominate the entire
process due to methanol adsorption saturation on the catalyst surface. Moreover, Roseller et al.
[511] also proposed that adding methanol has a more beneficial effect on hydrogen production.
Appropriate hole/electron sacrificial agents are required to maximize the activity of the Z-
scheme system. In addition, TEOA and TEA is also used as sacrificial agent in different
materials [111], [141], [512] during the catalytic reaction.

3.6.4 Catalyst amount

As mentioned above, the catalyst is an additional reagent to speed up the reaction. In


photocatalytic water splitting using MOFs, different types of triggers and cocatalysts are
introduced to enhance the catalytic performance of hydrogen production [513], [514]. This
process's improvement is primarily credited to the creation of catalysts with higher activity.
Several approaches, such as microporous supports, can boost the photocatalytic activity of
semiconductors in water splitting [515]. We should consider several factors while selecting a
catalyst, including its high availability, photochemistry stability, and non-toxicity [516]. The
catalyst quantity is one of the primary physical characteristics of a photocatalytic process's
reaction rate [517]. The reaction rate increases with increasing catalyst quantity up to a

130
maximum value corresponding to the system's ultimate photon absorption condition. After that,
the reaction rate decreases gradually due to light obstructing caused by an excess catalyst in
the reaction medium [518]. Although the reactor contains a primary motivation and some
cocatalysts, including pt, pd, NiPs, metals, MoS2, etc. For example, UiO-66-NH2 is a
photocatalyst MOF that is used along with pt nanoparticles as a cocatalyst to produce hydrogen
from water splitting [183]. Kong et al. [519] synthesized UiO-66(COOH)2/ ZnIn2S4 and state
that this can also have a large amount of hydrogen (18 794µmol g-1 h-1) with the loading of
MoS2. The process conditions listed above show just how important they are for the
photocatalytic reaction to proceed and produce hydrogen. They assist the catalytic activity in
raising the productivity rate in various ways. We summarize a few elements that significantly
affect this part's hydrogen production activity. We provide tactics and procedures in the next
piece to ensure these factors operate well.

4 Strategies to optimize photocatalytic water splitting

Photocatalytic water splitting is one of the significant daily concerns for producing safe
and eco-friendly hydrogen gas, an excellent energy source [520], [521]. Although MOFs have
been regarded as types of prospective candidates for use in photocatalytic water reduction [173]
several problems still need to be resolved. (1) The bulk of metal frameworks have a limited
light absorption region that can only be triggered by UV light. (2) Some MOF materials'
bandgap makes them unsuitable for photocatalytic water reduction activity. (3) There is still
room for development in the ability to separate electron-hole pairs [522]. Due to these factors,
we must develop specific techniques to increase their effectiveness [523]. Using the two
significant advantages of MOF materials as a foundation, approaches for improving MOF
photocatalytic hydrogen production performance, including configurable structure, are
provided (ligand functionalization, metal doping, defect engineering, and morphology
regulation) [522]. In addition, they differ from other materials in that they have a porosity that
is all their own and a high BET, which enables them to load guest molecules as cocatalyst
(metal or non-metal), photosensitizers, and templates for heterostructures that produce
conventional porous semiconductors [524], [525], [218].

4.1 Ligand functionalization

Organic functionalization has the potential to effectively lower the optical bandgap of
MOFs and increase the range of visible light to which they respond [526], [527]. Due to this
characteristic, ligand functionalization has been readily used in the recent era [522]. This has
already been mentioned in the preceding article. Figure 54A depicts a new Methylthio-
functionalized MOFs photocatalyst that is rationally designed, easily prepared, and achieves
high performance for visible-light-driven H2 evolution. Accordingly, a first methylthio-
functionalized porous MOF decorated with Pt co-catalyst for efficient photocatalytic H2
evolution was achieved, which exhibited a high quantum yield (8.90 %) at 420 nm by use
sacrificial triethanolamine. This hybrid material exhibited perfect H2 production rate as high as
3814.0 μmol g-1 h-1, which even is one order of magnitude higher than that of the state-of-the-
art Pt/MOF photocatalyst derived from amino terephthalate. Solid UV−vis spectra were
measured and compared to that of MIL-125 and free H2BDC- (SCH3)2 ligand, as that shown in
Figure 54 B(a). In sharp contrast to that absorption band below 350 nm of MIL-125, intense
131
absorption bands from 350 nm to 550 nm peaking at 420 nm were observed for x%-MIL-125-
(SCH3)2 samples. Such broadened absorptions in visible-light region are also reflected in their
physical crystallization appearance of yellow color. The color variations infer the electronic
band structure around the band gap is significantly modified. Current/time curves under
visiblelight illumination (λ > 400 nm) further demonstrated that the x%- MIL-125-(SCH3)2
samples possessed a high visible-light photocurrent response (~ -5.0 μA cm−2 for 20%-MIL-
125- (SCH3)2 and ~ -4.5 μA/cm-2 for 50%-MIL-125-(SCH3)2) in Na2SO4 which are in sharp
contrast to the negligible photocurrent density of the parent MIL-125 (Figure 54 B(b)).
Moreover, Figure 54 B(c) shows that the capacitance arc for 20%-MIL125-(SCH3)2 is much
smaller than that of MIL-125, suggesting the electron mobility was improved after the ligand
exchanged by BDC-(SCH3)2, thus the electron-hole recombination rate was effectively
prohibited. As shown in Figure 54 B(d), the positive slopes of M−S plots suggest a n-type
semiconductor behavior for both MIL-125 and 20%-MIL-125-(SCH3)2. This work opens a new
way up to obtain more efficient MOF’ photocatalyst with visible-light response, and such type
of thioether-functionalized MOF is worthy of further exploring.

A B

Figure 54. (A) Schematic diagram of SALE process to obtain x%-MIL-125- (SCH3)2 by use
MIL-125 as the parent MOF and H2BDC-(SCH3)2 as the exterior exchange linker. The cage
represents the cavity in MOF structure. (B) (a) UV-vis spectra in comparison and powder colors
(inset) of a: MIL-125, b: 20%-MIL-125-(SCH3)2 and c: 50%-MIL-125-(SCH3)2. (b)
Current/time plots of MIL-125, 20%-MIL-125-(SCH3)2 and 50%-MIL-125- (SCH3)2 measured
at an applied bias potential of 0.24 V vs. RHE with and without visible-light irradiation pulse
of 40 s. (c) Nyquist plots and (d) Mott−Schottky plots of MIL-125 and 20%-MIL-125-(SCH3)2
in 0.1 M Na2SO4Reprinted with permission from [195]. Copyright 2018 Elsevier.

In very recent times, Stylianou and colleagues have been looking into the photocatalytic
hydrogen production performance of amino (NH2) and/or hydroxyl (OH) MIL-125 [529]. In
this regard, the group led by Li was successful in constructing three isostructural UiO-66-based
MOFs decorated with three distinct sulfur-containing groups (X = SCH3, SH, and SO3H).
These MOFs had uniquely unique electronic properties, molecule dimensions, and geometrical
configurations. These UiO-66-X2 derivates exhibited as n-type semiconductors, wherein the
electronic properties and molecule dimensions of the function groups –X played key roles in
determining their band gap (Eg) and photoactive properties. The –SCH3 is shown as the most
efficient functional group and is responsible for that dramatically narrowed Eg of the UiO-66-

132
(SCH3)2 and photocatalytic property. In sharp contrast to that state-of-the-art UiO-66-
NH2 (constructed with 2-aminoterephthalate linkers) having no visible-light induced
photocatalytic activity to split water into H2 even with Pt as co-catalyst, Pt/UiO-66-
(SCH3)2 showed a high efficient H2 generation (3871 μmol/g) from water with
sacrificial ascorbic acid (0.2 M) under λ > 400 nm irradiation [528]. In addition to the
functional groups that were previously mentioned, porphyrin units, which have been shown to
possess excellent visible light-harvesting capability, have also been widely used as ligands to
synthesize MOF materials that can absorb visible light. For instance, Jiang and colleagues were
the first to report the successful preparation of a highly stable out-of-plane (OOP) porphyrin
MOF USTC-8 (In) by employing In (OH)3 as a precursor [200]. In light of previous research,
we can acknowledge that the ligand functionalization method is a flexible and dependable
instrument for enhancing the photocatalytic activity of MOF-based photocatalysts concerning
hydrogen gas generation. Moreover, we can conclude that the ligand functionalization
procedure is an adaptable and reliable way to increase the photocatalytic activity of MOF-based
photocatalysts for hydrogen production.

4.2 Metal doping

It is believed that doping MOFs with metals is an effective strategy for increasing their
activity. The great coordination sites in MOFs provide opportunities for foreign metal
stabilization. This process works as an intermediary to enhance electron transmission,
significantly increasing photocatalytic activity [173]. In addition to the ligand functionalization
method, the metal doping strategy has also been considered a potentially helpful instrument to
optimize the photocatalytic performance of MOF-based photocatalysts [522]. Regarding MOF
materials, there are two approaches to carrying out the metal doping technique. One reason is
that doping metal ions can react with organic ligands with functional groups such as amino,
sulfhydryl, and porphyrin units, each with a lone pair of electrons. The doped metal can operate
as a medium for electron transfer, considerably increasing photocatalytic activity [530], [136].
Lu's group reported one of the most fascinating examples [524]. In their work, they decorated
NH2-MIL-125(Ti), built from TiOx clusters and NH2-containing BDC linkers, with the highly
active Cu II/Cu I as a mixed centre using a simple doping method. Additionally, the metal
clusters might be embellished with various metals. Doping Ti ions into the Zr6 cluster of MOF
structure can greatly boost the photocatalytic activity for H2 generation. Using a post-synthetic
exchange (PSE) approach, Li and colleagues created Ti-substituted NH2-UiO-66(Zr/Ti) [191].
Garcia's group has recently successfully designed a range of UiO-66 MOFs with different metal
clusters (M: Zr, Zr/Ti, Zr/Ce, Zr/Ce/Ti, Ce) [531]. Doping MOF materials with multiple types
of metal ions and doping them with a single metal ion can affect the electron transfer in MOF
materials and thus promote water reaction. For instance, Musho and his colleagues., [532]
constructed functionalized UiO-66-series MOFs using three inorganic metal ions (Zr, Ti, and
Hf) and three functional groups, including an amino group (NH2), a nitro group (NO2), and a
hydrogenated case (H) and then studied the effect that this had on the optical bandgap of
catalytic materials.

4.3 Defect regulation

The inorganic semiconductor materials' properties have typically been altered using
defective engineering techniques to achieve maximum photocatalytic activity for water

133
reduction. These features include electrical and band structures, reactive surface sites,
conductivity, and the concentration of charge carriers [533], [534]. Jiang's group provided the
first example of investigating the influence of MOF flaws on photocatalytic water reduction
performance [525]. When the concentration of defects reached an unacceptable level, the active
sites changed into new recombination zones for photogenerated electrons and holes. To manage
the photocatalytic performance of MOF materials, the defect engineering method, which is also
used for controlling the photocatalytic performance of other inorganic semiconductor
materials, can be applied, according to previous research findings [522]. Zang et al. [535]
investigate activating the proton reduction by opening catalytic sites in the copper-triazoles
framework via defect chemistry. Typically, six nitrogen atoms from six triazole ligands are
associated with the Cu core of MET-Cu, as shown in Figure 55a. The extensive assembly of
the tiny triazole ligand with the copper results in a strong skeletal structure that gives the MET-
Cu framework remarkable stability.

Figure 55. (a) Demonstrates MET-CU and flaws present in MET-CU (b) the photocatalytic H2
production method of MET-Cu-D. Reprinted with permission from [535]. Copyright 2021
Elsevier.

134
The EY molecules absorb enough energy from the radiation to create the single excited state
EY1 *, which then changes into the stable triple keen state EY * through the intersystem
crossover (ISC). Afterward, the electrons of EY− ⋅ are transferred rapidly to the MET-Cu-D
catalyst; simultaneously, the dye molecules return to the ground state (Figure 55b).

4.4 Morphology regulation

Morphological regulation, which includes crystal size, thickness, and facet regulation,
has proven to be one of the most successful strategies for regulating the photocatalytic activity
of the catalysts [536], [537]. For example, two-dimensional MOFs with a shallow thickness
offer significant advantages over bulk MOF materials as illustrated 1) The thinner 2D MOFs
offer the possibility of reducing the charge carrier migration distance. 2) 2D MOFs have an
abundance of reactive surface sites in their structure. 3) It is possible to change the bandgap of
2D MOFs by modifying their thickness, which results in an increase in their light-harvesting
capacity; and 4) Their highly high surface area is advantageous for the diffusion of products or
substrates, but it can also serve as the perfect platform for integrating additional cocatalysts
[522]. A flexible technique for controlling the photocatalytic activity has been thought to be
adjusting the particle size of the photocatalysts.

4.5 Surface or pore decoration of MOFs

Even while functional modification can help to increase the photocatalytic performance
of MOFs to a certain extent, to further improve photocatalytic performance, a more efficient
strategy, such as surface or pore decorating, is required [538]. While doing the surface and
pore decoration, a few scenarios have to be considered, firstly the method by which we do
surface and pore decoration is the incorporation of MOFs with suitable semiconductors like
graphitic carbon nitride (g-C3N4) because of its strong visible-light capacity, outstanding
chemical, and thermal stability, low cost, and high economic viability [173], [539], [540]. By
annealing the g-C3N4 and UiO-66 MOF mixture in an argon medium, Wang et al. [541]
created several hybrids in varying proportions. The schematic illustration of g- C3N4 adhering
to UiO-66 octahedrons through annealing is shown in Figure 56a. Observations made using a
scanning electron microscope (SEM) show that the pure form of UiO-66 is an octahedral
microcrystal with edges that range in length from 200 to 500 nm Figure 56b. Following
annealing with g- C3N4, g- C3N4 nanosheets coat the surface of UiO-66 octahedrons, resulting
in the character of UiO-66 becoming uneven (Figure 56c). As expected, more excellent C3N4
content in starting materials resulted in high C3N4 coverage on UiO-66 octahedrons (See
Figure 56d). C3N4 nanosheets may clearly be seen sticking to the edges of UiO-66
octahedrons. (See Figure 56e.) The energy-dispersive X-ray spectroscopy (EDS) corroborated
the surface C3N4 layer, as illustrated in Figure 56f.

135
Figure 56. (a)The annealing process is illustrated here as a schematic and is used to decorate
g-C3N4 on UiO-66 octahedrons. SEM views of (b) UiO-66, (c) UG-10 and (d) UG-50. (e) The
histogram shows the hydrogen production rate over pristine UiO-66, g- C3N4, and UG-x
samples. (f) The energy-dispersive X-ray spectroscopy (EDS) corroborated the surface C3N4
layer. Reprinted with permission from [541]. Copyright 2015 Elsevier.
Other semiconductors have also been investigated for use with MOFs, such as MoS2 and Cu2O.
MoS2 with a layered structure has received much attention because of its high activity for solar
hydrogen synthesis [371]. For instance, it has been coupled with UiO-66/CdS to function as a
noble-metal-free cocatalyst for HER in visible light. As a low-cost and environmentally
friendly semiconductor, Cu2O has also been used for photocatalytic hydrogen evolution

136
through the post-synthetic encapsulation of the compound into NH2-MIL-125 [542]. Using the
template approach, conventional semiconductor TiO2 hollow nanostructures (TiO2 HNPs) have
recently been created and adorned with ZIF-8 MOFs [543]. This shows that the photocatalytic
performance of this heterogeneous photocatalyst was significantly increased as the shell of
TiO2 HNPs increased the surface area. Metal nanoparticles (MNPs) are another type of great
catalyst as compared to bulk noble metals, as evidenced by their high surface area-to-volume
ratio [544]. Pt NPs have been extensively used in conjunction with MOF for photocatalytic
water splitting. Wen et al. [359] manufactured Pt NPs loaded and RhB-sensitized NH2-MIL-
101(Cr) for HER, dramatically increasing electron-hole separation efficiency. Moreover,
inorganic polyatomic clusters known as polyoxometalates (often abbreviated as POM) are
formed when transition metal atoms and oxygen atoms are joined together. Because it
possesses a high redox potential, diverse structural characteristics, and highly negative charges,
POM is an essential component in the catalytic reactions that take place [545]. It has been
demonstrated that encapsulating POMs within MOFs is a promising strategy for increasing the
catalytic performance as well as the structural stability of MOFs. For instance, POM
(H3PW12O40) and Pt NPs have been encapsulated onto the NH2-MIL-53 [546], which assures
that POM not only stabilized Pt NPs onto the surface of NH2-MIL-53 MOF but also promoted
photocatalytic performance. Simultaneously, a polycationic supramolecular MOF (SMOF-1)
was created, and SMOF-1 that adsorbed anionic Wells-Dawson-type POM demonstrated an
efficient hydrogen production rate [547]. A combination of polyoxometalates (POM) with
SMOF-1 and methanol as a sacrificial agent releases 3.353 mmol h-1·g-1 of hydrogen. Co-based
POM has good catalytic capabilities for water oxidation since it contains earth-rich elements.
As one of the most typical Co-based POM, [Co4(H2O)2(PW9O34)2] 10 − has been embedded
into MIL-101(Cr) by an ion exchange method [548]. As a result, the immobilization of
[Co4(H2O)2(PW9O34)2] 10 − POM in a more appropriate host MIL-100(Fe) MOF, has been
explored and employed in a water oxidation reaction [549]. The outcomes showed that MIL-
100(Fe)/Co-TOF POMs and oxygen yield (9.21003s 1; 72%) were higher than those of MIL-
101(Cr)/Co-POM (7.3103s 1; 66%). Zr-based clusters and a TCPP-H2 linker make up MOF-
545 [550], as illustrated in Figure 57.

137
[(PW9O34)2Co4(H2O)2]10-

[TCPP-H2]4-

P2W18Co4 @ MOF-545
[Zr6(µ3-O)4(µ3-OH)4 (OH)4 (H2O)6]10+
Figure 57. Detailed representation of [Co4(H2O)2(PW9O34)2] 10 − POM, TCPP-H2 linker, Zr-
based unit, and P2W18Co4@MOF-545. Reprinted with permission from [550]. Copyright 2018
Elsevier.

Finally, we may infer that the main principle behind the solutions mentioned above is based on
the flexible architectures of MOF materials. Among the many approaches that have been taken,
ligand functionalization and metal doping are two that have seen extensive use in MOF
materials. However, little attention has been paid to the defect regulation and morphological
regulatory strategies for MOFs in photocatalytic H2 production [548].

5 Stability of metal organic frameworks (MOFs)


As the field develops, greater consideration of the physical properties of metal-organic
frameworks has surfaced. For example, their elastic moduli, stimuli-responsive behaviour,
mesoscale structure and defect content have been studied, and unusual physical properties such
as negative gas adsorption revealed [ 585], [ 586]. The mechanical stabilities of an array of
materials have been researched [616], [587], with recent efforts to predict the limit of pressure-
induced collapse and the use of machine learning to predict limits of stability being particularly
notable [588]. Separately, the chemical stability of MOFs in acidic, basic and hydrolytic media
has been summarized, and provides guidance to their use in industrial settings [589], [590]. A
small number of MOFs have even been studied for their radiolytic stability against gamma
radiation [591]. When analysing thermal stability of MOFs, Healy et al [592] found it helpful
to separate decomposition processes into two qualitative categories, namely ligand-centred
decompositions and node-centred decompositions. They recognize that these processes
probably do not strictly operate independently, as any structural changes to a material will
influence both ligand and node properties. In order to further enhance the stability of the

138
framework, many different approaches, such as the utilization of high-valence metal ions or
nitrogen-donor ligands, were investigated [593]. So far, most stability measurements on MOFs
have been concentrated on their thermal and water stabilities, but further avenues of possible
research are to determine their stabilities towards contaminants such as salts, body fluids, or
corrosive molecules such as H2S or NH3. As future MOFs are developed for more widespread
application, their stability towards more varied chemicals and contaminants should be
determined, and application-built MOFs should be rationally designed to be stable towards
conditions present in their targeted environment.
Assessing the mechanical stability of metal-organic frameworks (MOFs) is critical to
bring these materials to any application. Here, Peyman Z. et al [588] derive the first interactive
map of the structure-mechanical landscape of MOFs by performing a multi-level computational
analysis. They performed high-throughput molecular mechanics calculations to establish
structure-mechanical stability relationships for 3,385 MOF materials with 41 distinct
topologies. Although many MOFs are vulnerable to structural destruction even in an ambient
atmospheric environment due to the lability of coordination bonds between metal ions and
ligands, a growing number of MOFs with exceptional chemical stability have been reported in
recent years. In general, two main types of approaches have been adopted to improve the
chemical stability of MOFs. One is the de novo synthesis of stable unknown MOFs and the
other is to improve the stability of existing MOFs [616]. Given the important roles of metal
ions in the fabrication of stable MOFs, the strategy of introducing two or more types of metal
ions into the metal clusters of MOFs has been developed. Some mixed-metal MOFs have been
found to display enhanced stability as compared to single metal MOFs. There may be several
reasons for this: (1) the formation of stronger coordination bonds compared with pristine bonds;
(2) the enhancement of the inertness of metal clusters; (3) the improvement of surface
hydrophobicity. For instance, to improve the hydrostability of MOF-5, Zn2+ ions were partially
replaced with Ni2+ ions to give the Ni-doped MOF-5 [594, 595]. In some cases, the very open
frameworks with large pore sizes and/or large surface areas are fairly unstable. Introducing
size-matching ligands as brackets into the channels of MOFs via de novo synthesis (i.e. pore
space partition) is a possible strategy to improve the robustness of the above-mentioned MOFs.
Moreover, since the insertion of the stabilizing pillars or these size-matching molecular
building blocks (MBBs) splits the large cage or channel space into smaller segments, this
insertion may remarkably enhance the adsorption performances of small molecules [596].
Moreover, interpenetration (also called framework catenation), which is the interweaving or
entanglement of two or more independent and identical frameworks in MOFs, provides an
alternative strategy to improve the stability of MOFs. The interpenetration of the framework
can not only significantly increase the wall thickness and reduce the pore size, but also prevent
the displacement of the ligand by locking it in place within the framework, resulting in the
formation of more stable structures compared to the non-interpenetrated counterpart [597].
Recently, Bu's group [598] demonstrated that a self-penetrated network was more stable than
an interpenetrated framework (Fig. 58). By employing 5,5′-((thiophene-2,5-
dicarbonyl)bis(azanediyl))diisophthalic acid and 5,5′-([2,2′-bipyridine]-5,5′-
dicarbonyl)bis(azanediyl)diisophthalic acid separately as the organic linkers, NKU-112 with a
two-fold interpenetrated framework and NKU-113 with a self-interpenetrated framework were
prepared. In NKU-113, the coordinated DMF molecules of metal clusters were substituted by
chelating the bipyridine moiety of 5,5′-([2,2′-bipyridine]-5,5′-
dicarbonyl)bis(azanediyl)diisophthalic acid. This greatly increased the framework rigidity of
NKU-113 as compared to NKU-112. After solvent exchange and supercritical drying, NKU-
113 retained its framework and displayed a moderate BET surface area of 1486 m 2 g−1, while
only a negligible amount of N2 was adsorbed in NKU-112 due to its collapsed structure.

139
Fig. 58 Schematic diagram showing the framework and the modification strategy. The blue-
colored balls, silver-colored sticks, yellow-colored sticks, and red-colored cylinders represent
metal SBUs, ligands, coordinate bonds and pores, respectively. Reproduced from [598] with
permission from The Royal Society of Chemistry, copyright 2018.

Given that the rational strategy for preparing stable MOFs via de novo syntheses
remains limited, a large number of strategies for enhancing the stability of existing MOFs have
been developed which includes post-synthetic exchange, post-synthetic modification,
hydrophobic surface treatment, and composite fabrication. Post-synthetic exchange (PSE) can
offer a promising tool to improve the physical and chemical attributes of MOFs via metal-ion
metathesis, linker exchange or the replacement of the counterions of ionic MOFs. Along this
line, the PSE process might increase the bond strength of coordination bonds in the unstable
SBUs of MOFs or tune the water-resistance of MOFs without disturbing the framework
structure. Furthermore, the organic linkers and metal clusters in MOFs provide great
opportunities to a vast range of chemical transformations, which means that the introduction
of particular functional groups via the post-synthetic modification (PSM) approach can alter
the pore environments of MOFs [599]. Although the chemical stability of MOFs can be
significantly improved by the above methods, the porosities and adsorption properties are often
drastically reduced by the occupation of the channels by dangling functional groups. To avoid
this, hydrophobic surface treatment has been developed to be applied to the exterior surface of
MOFs to protect the MOFs from water while preserving their porosity to a large degree. In
general, hydrophobic surface treatment mainly includes surface coating, post-synthetic thermal
annealing, and post-synthetic surface modification [600, 601]. With regard to the experimental

140
stability testing, the combination of multiple means of detection is more reliable to qualify the
maintenance of intrinsic properties of MOFs [602]. Particularly, compared with static stability
analysis, the investigation of the dynamic change course of MOFs under harsh conditions may
be more meaningful.

In the future, the fabrication of very stable MOFs is likely to remain a dream and that would
pave the way to their applications to diverse ends. For example, stable and well-designed
multivariate MOFs or MOF-based materials, which possess the features of each individual part,
can provide extraordinary performance for targeted application. In addition, the precise
location control of customisable functional groups within stable MOFs through stepwise
synthetic techniques will produce great potential or optimal performance for diverse
applications. Moving forward, mechanism analysis related to applications and device-
engineering are also the attractive goals for stable MOFs.

6. Techno socioeconomic and feasible analysis

Techno feasibility is very much crucial for any process as it combines process
modelling and engineering design with economic evaluation and helps to assess the financial
viability [603], [604], [605]. It makes us aware of the cost of the reactor and materials used for
the process, how much the whole setup demands, and the effect of this setup on the economy.
Photocatalytic water splitting is unquestionably a cost-effective and economical method of
producing hydrogen [505], [606]. Here, we're focused on making this technology more
renewable and affordable and how the hydrogen generation efficiency be increased to meet the
energy demand, that is actually what we mean from a techno-feasible analysis [45], [607],
[608]. One crucial question that all scholars working on the subject should be asking because
economics fundamentally drives our energy environment: will hydrogen be obtained at a cost
competitive with fossil fuels if the technical barriers to large-scale implementation of water
splitting are overcome? [609]. Photocatalytic water splitting provides us with energy that
replaces our rare fossil fuels, i.e., oil, gas, and petrol [610], [611]. The energy from water can
be used in various fields without the emission of harmful and poisonous gases [612]. Numerous
review studies have conducted tests for economically viable solar hydrogen production, made
calculations, conducted surveys of the available bench-scale materials, and later computed the
cost of hydrogen production from several large-scale reactors [613]. H2 is a promising energy
carrier for use as a low-cost, wide range of possible energy storage technology [614], [615],
and including H2 in energy strategies is critical to hasten to move away from traditional energy
sources [616]. Dinner and Acar stated that H2 cannot only cut six gigatonnes of CO2 emissions
but also meet 18% of total energy consumption [617], [618]. Advances in the H2 energy sector
suggest that with a manufacturing cost of $2.50/kg, H2 can serve 8% of global energy demand
[619]. Further advances are predicted to reduce production costs to $1.80/kg by 2030 when H2
is expected to meet 15-18% of global energy demand [620]. However, the hydrogen economy
is still not fully achieved [621], [622] because of technological barriers. The cleanliness
standard must be maintained throughout the entire production system to develop a prosperous
hydrogen economy. In conclusion, a well-developed model must be in place to ensure the
cleanliness standard for the whole of the hydrogen-based energy system [623]. It is critical to
carry out a cost analysis for each production method to evaluate their respective practicality. If
hydrogen output surpasses 1000 kg per day, the price of the electrolyzer may be reduced in the

141
following years. The current production scale is limited to 10 kg per day; and if large-scale
hydrogen generation is adopted, the price of the electrolyzer might be reduced to one-quarter
of its current level [624]. Overall, solar-powered technologies hold the most potential for H2
production because they have low health implications and create opportunities for employment
creation. Despite their high prices, solar thermochemical methods and electrolysis cost less
than $10/kg. To compete with other liquid fuels or hydrogen-producing processes, it is essential
to establish the target price for hydrogen [45], [613]. The multi-year research, development,
and demonstration plan for the DOE Fuel Cell Technologies Program called for bringing
hydrogen prices down to $2–$4 per gasoline gallon equivalent (gge) sold at the pump. Given
the lower operating temperature, one gallon of gasoline and one kilogram of hydrogen are
roughly comparable, with a target cost of $2–$4 per kg H2 [625]. According to a prior study,
manufacturing hydrogen using different distributed and centralized systems cost $1.61 per
kilogram, for centralized biomass gasification, $1.33 per kg for natural gas reforming, $4.50
per kg for wind electrolysis, and $2.05 per kg for coal gasification with carbon capture,
excluding the costs of compression, storage, and delivery. Moreover, for the Type 1, Type 2,
Type 3, and Type 4 systems, the baseline levelled production cost of hydrogen was calculated
to be $1.60, $3.20, $10.40, and $4.10 per kg H2 [613]. The study includes a review of the
literature on the Technology Readiness Level (TRL), the Levelised Cost of Hydrogen (LCOH),
various generic criteria, and specific critical metrics. The data discovered regarding STH
efficiency, LCOH, and other general factors are used as input for a Multicriteria Analysis
(MCA) to simulate the outcomes while considering the many stakeholders' preferences [626].
The LCOH is calculated for all methods for the techno-economic analysis, and it is assumed
that they are run off-grid [627], [628]. The results are expressed in $ per kg H2 (we use kg H2
as a proxy for energy) and are obtained via the following equation:
𝑛
𝑙0 + 𝑂𝑀𝑡 + 𝐹𝑡
𝐿𝐶𝑂𝐻 = ∑
(1 + 𝑟)𝑡
𝑡
_______________________
𝐸𝑡
∑𝑛𝑡
(1 + 𝑟)𝑡

where l0 represents the initial investment costs (CAPEX), OMt represents the annual operation
and maintenance costs, Ft represents the yearly fuel costs, r represents the inflation rate
corrected discount rate, Et represents the yearly energy production, and n represents the total
economic life of the energy system. When necessary, data is converted and stated in $/m2 and
corrected for the Consumer Price Index of 2022 to determine the associated cost for the
photoactive area, and sometimes land area, required for each STH technique. The price in 2022
(p2022) is calculated as follows:

𝐶𝑃𝐼2020
𝑝2022 = 𝑝𝑌 ∗
𝐶𝑃𝐼𝑌

where pY is the price in year Y (in $), CPI2022 the dimensionless Consumer Price Index of 2022,
and CPIY is the Consumer Price Index of year Y.

142
PV-electrolysis was the most economical, competitive option with an LCOH of 9.31 $/kg H2.
PC-WS had a LCOH of 18.32 $/kg H2, PEC-WS of 18.98 $/kg H2, DBP of 18.45 $/kg H2, and
IBP of 36.39 $/kg H2 [629].

The amount of hydrogen required to create every month to make the payback at hydrogen
ranges from $1.87 to $4.87 per kilogram. 4.87 $/kg energy is equivalent to 1.20 $/L gasoline.

H2 Prices(/kg) $1 $2 $3 $4 $4.87

Kg/ month 2224.44 1112.22 741.48 556.11 457.22

1 sun system, 2690.75 1345.39 896.93 672.70 553.07

10% efficiency

100 sun system 13.45 6.73 4.48 3.36 2.77

20% efficiency

The area needed to produce the same amount of hydrogen is reduced by the same factor by the
expected doubling in photocatalyst efficiency. This will help reduce the system cost due to the
required material to cover the area. More research is needed to determine how much a
photocatalytic water-splitting system will cost per kilogram of hydrogen.

Moreover, MOFs have great commercial potential; little attention has focused on
optimizing production mechanisms to drive down costs. In these instances, techno-economic
analyses can play a key role in uncovering cost drivers and revealing ways to lower lifetime
production costs. More specifically, they can help assess production mass flows, energy
consumption, and the economic performance of an industrial process; for MOF syntheses, they
can highlight the greatest barriers to commercial viability and the opportunities for further
improvements. However, techno-economic analyses are rarely included in the MOF literature.
On Web of Science, there are 28,067 papers with the keywords “metal-organic framework
synthesis,” and only five of these papers mention techno-economic analyses. Even when
techno-economic analyses are included in the literature, they are oftentimes not sufficiently
robust because they may use inaccurate material cost quotes or do not account for the cost of
labor, inflation, maintenance, electricity, and the cost of capital. In a techno-economic analysis
of industrial-scale MOF synthesis techniques for hydrogen and natural gas storage, DeSantis
et al. [630] modeled the cost breakdowns for producing Mg-MOF-74. For this MOF, they

143
estimated the total production costs from an industrial-scale solvothermal synthesis, a LAG
ball milling synthesis, and an aqueous synthesis technique, which showed that raw material
costs comprised the largest share of total production costs in all three synthesis techniques.
Additionally, they performed a sensitivity analysis in which they varied certain parameters
such as yield, precipitation time, solvent mass ratio, solvent recovery rates, and material costs
to understand the most influential parameters in their model. This sensitivity analysis showed
that the type of synthesis, the solvent mass ratio, and the reaction yield were the most influential
parameters in their cost model. This type of sensitivity analysis is valuable because it identifies
parameters that can be further tuned for greater cost reductions. As seen below, the levelized
cost is equal to the sum of present values for the costs accumulated throughout production
divided by the sum of present values for annual production volumes. This allows us to calculate
a per unit production cost where Ct is the total cost of production in year t, Pt is the total
production volume (kg) of UiO-66-NH2 in year t, N is 10 years, and r is the discount rate.
𝐶𝑡
Levelised cost of MOF(LCOM)= ∑𝑁
𝑡=1 (1+𝑟)𝑡
_______________________
𝑃𝑡
∑𝑁
𝑡=1 (1+𝑟)𝑡

By using the levelized cost methodology, we gain a holistic measurement of the lifetime costs
for producing a kilogram of UiO-66-NH2. Several of the aforementioned techno-economic
analyses listed in the literature review did not calculate the present value of annual production
volumes when they calculated their cost per kilogram of MOF [631]. Inherently, though, a
certain volume of MOFs produced right now has a different value or utility than an equivalent
volume produced in the future, which is why the levelized cost methodology is more accurate.
If an analysis does not calculate the present value of annual production volumes, then the
denominator in the cost function will overestimate production, which results in an artificially
low cost per kilogram of MOF. Additionally, our levelized cost methodology accounts for the
present value of all future capital expenses, variable material costs, maintenance costs,
electricity costs, and labor costs.

7. Current challenges and future perspectives

As the world's population increases to above nine billion people by 2050 and fossil fuel
stocks are depleted, there will be a difficult transition from non-renewable energy sources to
renewable energy sources with zero carbon emissions. In the current situation, photocatalytic
water splitting, often known as "Artificial Photosynthesis," powered by solar energy, emerges
as the most promising approach to solving the world's energy crisis. Sluggish catalysis under
real operating conditions is often limiting. Hydrogen production catalysts based on solid-state
materials are usually stable but exhibit low selectivity and high sensitivity to poisoning. By
contrast, as their chemical properties can be tailored through judicious choices of metal and
ligands, molecular catalysts based on transition metal complexes enjoy good selectivity and
tolerance to varied experimental conditions. Unfortunately, most of the molecular catalysts for
H2 production are not soluble in H2O, which is the most attractive solvent for sustainable
conversion of sunlight to H2. Practical use of molecular catalysts hence will be greatly
facilitated if they are able to utilize H2O as proton source or at least tolerate the presence of
H2O. The various strategies developed for immobilization of molecular catalysts on solid-state
supports represent good solutions to this issue, while also helping the implementation and
144
recycling of the catalytic system [632]. Photocatalytic hydrogen production often suffers from
low efficiency. Most photocatalysts are not very efficient at utilizing the absorbed light energy
for the hydrogen generation process.
Photocatalysts can degrade or become less effective over time due to exposure to harsh reaction
conditions and environmental factors. Developing stable catalysts that can withstand
prolonged use is essential. The cost of materials and processes for photocatalytic hydrogen
production needs to be competitive with other hydrogen production methods to make it
economically viable. This includes the cost of materials, catalysts, and the energy required for
the process. Experiments have provided sufficient evidence to illustrate the potential of MOFs
for applications involving the splitting of water molecules [633]. Due to its regular structural
makeup and synthetic tunability, the integration of MOFs with photocatalytic WS is thought
to increase and expand light absorption, enabling active sites. The OER and HER were
investigated, and MOFs were used as direct catalysts and precursors [125], [118]. They have
demonstrated an intriguing potential in the electrocatalytic and photocatalytic splitting of water
for the creation of hydrogen, with some very fascinating findings. For instance, MOF-based
catalysts have a severe issue with water stability, which can be traced back to the fact that their
ions have a more considerable affinity for water molecules than carboxylate ligands [634],
[635]. During the process of water splitting, one of the other more significant concerns is the
leaching of metallic and organic components from MOFs into the reaction medium [111]. The
reaction mechanisms of MOFs based water-splitting catalysts are still poorly known and
require additional exploration. Recycling MOFs based catalysts is also a substantial challenge
for several reasons, one of which is the nanoscale size of most MOF-based materials.
Furthermore, less photocatalytic stability (e.g., photo corrosion), low production yield and
quantum efficiency, and, most notably, a sluggish photocatalytic activity is perceived as
obstacles in the path of photocatalytic water splitting [636], [637]. One significant difficulty
that limits the use of some MOFs is their good stability in aquatic conditions over a broad pH
range and moisture sensitivity [638], [639]. Furthermore, the organic linkers, particularly
MOFs, are intricate and expensive, making their utilization difficult. To reduce the price of
MOFs based catalysts, developing mixed organic linkers, investigating mixed-organic linkers
and transition metal-based catalysts, and using carbon-based cocatalysts like graphene may be
intriguing strategies in the future. A fundamental barrier to high photocatalytic performance is
the fact that, in addition to being photocatalysts, many MOFs have a limited capacity for light
harvesting by nature. The industrialization and commercialization of photocatalytic hydrogen
evolution are unavoidable problems considering the role of hydrogen energy in future energy
development. Therefore, future research should focus on S-scheme heterojunction
photocatalysts that are highly active, low-cost, reusable, non-toxic and non-hazardous,
recyclable, durable, and with high solar energy conversion efficiency. At present,
computational analysis plays an important role in the theory of photocatalysis. Therefore, DFT
should be vigorously introduced in relevant studies to deeply analyse the interaction between
catalysts. Despite these issues, catalytic water splitting over MOFs-based catalysts has a bright
future because of its unique characteristics, including wide surface area, variable pore volume,
size, 3-dimensional structure, and rich accompanying chemistry and a new set of opportunities
for these MOFs in a range of zones [640], [641]. These materials have unexceptionable thermal
stability compared with inorganic semiconductor photocatalysts. Therefore, the vigorous
development of these materials and the active construction of all-inorganic semiconductor S-
scheme heterojunctions in the field of photocatalysis is essential to improve the stability and
high photocatalytic hydrogen evolution activity of photocatalysts.

145
8. Summary

It is anticipated that photocatalytic solar energy conversion technologies will be crucial in


addressing energy scarcity and environmental degradation issues. This paper contains a
comprehensive assessment of the recent and noteworthy initiatives that have exploited
materials based on MOFs as photocatalysts to speed the creation of hydrogen from splitting
water molecules. These significant achievements demonstrate that MOFs have the potential to
be utilized as novel and efficient photocatalysts for the photolysis of water to produce hydrogen
when exposed to ultraviolet (UV), visible, or even near-infrared (NIR) light. All of the MOFs
based materials discussed above can efficiently convert solar energy to hydrogen energy due
to their high stability and distinctive semiconductor feature, and the majority of them can be
recycled multiple times. Many factors affect the catalytic activity of MOFs, such as their
surface area and porosity, the loading of cocatalysts, the heterostructure between the metals
and MOFs, and the functional group. All these factors are discussed in detail to study their
effect on the performance of hydrogen production. MOFs and their composites retain their
original structures after being changed with nanoparticles or other guests, but MOF-derived
photocatalysts dismantle the original frameworks to produce new composites with improved
photocatalytic activity due to shorter electron transfer distances. However, specific new tactics
and/or difficulties should be considered further to improve the photocatalyst. In this review,
recent advances in using MOFs as photocatalysts have been briefly summarized, with particular
attention paid to several key strategies developed to produce high-efficiency MOF-based
photocatalysts. These strategies include ligand functionalization, MOF defect engineering,
morphology regulation, and pore and surface decoration with other semiconductors. These
strategies overcome the defects in MOFs like low band gap or stability. Our study concludes
with 100 different MOFs, synthesis methods, and process conditions. The synthesis method
which is vastly used is solvo/hydrothermal method and their key findings prove that MOFs are
becoming a hot topic for photocatalytic water splitting. Moreover, the techno feasibility of
MOFs in recent years is also explained. Although research on photocatalytic water splitting to
produce hydrogen using MOFs based materials is still in its early stages, preliminary results
show that they can be effective photocatalysts for photocatalytic HER. MOF-based materials
are expected to successfully transform solar energy into hydrogen energy in the not-too-distant
future, eventually leading to fossil fuel displacement.

Acknowledgment

1. This work was supported by the National Natural Science Foundation of China (NSFC)
[grant numbers 61975148].
2. The authors extend their appreciation to the Deanship of Scientific Research at King Khalid
University for funding this work through large group Research Project under grant number
[RGP2/3/44].

References

146
[1] Hisatomi, T. and Domen, K., 2019. Reaction systems for solar hydrogen production via
water splitting with particulate semiconductor photocatalysts. Nature Catalysis, 2(5),
pp.387-399. https://doi.org/10.1038/s41929-019-0242-6

[2] Shafiee, S. and Topal, E., 2008. An econometrics view of worldwide fossil fuel
consumption and the role of US. Energy policy, 36(2), pp.775-786.
https://doi.org/10.1016/j.enpol.2007.11.002

[3] Balat, M., 2007. Status of fossil energy resources: A global perspective. Energy Sources,
Part B: Economics, Planning, and Policy, 2(1), pp.31-47.
[https://doi.org/10.1080/15567240500400895]

[4] Abas, N., Kalair, A. and Khan, N., 2015. Review of fossil fuels and future energy
technologies. Futures, 69, pp.31-49. [https://doi.org/10.1016/j.futures.2015.03.003].

[5] Liao, C.H., Huang, C.W. and Wu, J.C., 2012. Hydrogen production from semiconductor-
based photocatalysis via water splitting. Catalysts, 2(4), pp.490-516.
https://doi.org/10.3390/catal2040490

[6] Ara, N. and Das, S., 2022. Social Aspects of Green Technology: A Review on
Environmental Protection. http://doi.org/10.36647/978-93-92106-02-6.22

[7] Sharma, R.R., Kaur, T. and Syan, A.S., 2021. Sustainability marketing–the
environmental perspective. In Sustainability Marketing (pp. 79-92). Emerald Publishing
Limited. https://doi.org/10.1108/978-1-80071-244-720211006

[8] Ebhota, W.S. and Jen, T.C., 2020. Fossil fuels environmental challenges and the role of
solar photovoltaic technology advances in fast tracking hybrid renewable energy
system. International Journal of Precision Engineering and Manufacturing-Green
Technology, 7(1), pp.97-117. https://doi.org/10.1007/s40684-019-00101-9

[9] Ong, W.J. and Shak, K.P.Y., 2020. 2D/2D heterostructured photocatalysts: an emerging
platform for artificial photosynthesis. Solar Rrl, 4(8), p.2000132.
https://doi.org/10.1002/solr.202000132

[10] Li, X.B., Xin, Z.K., Xia, S.G., Gao, X.Y., Tung, C.H. and Wu, L.Z., 2020.
Semiconductor nanocrystals for small molecule activation via artificial
photosynthesis. Chemical Society Reviews, 49(24), pp.9028-9056.
https://doi.org/10.1039/D0CS00930J

[11] Matsuoka, M., Kitano, M., Takeuchi, M., Tsujimaru, K., Anpo, M. and Thomas, J.M.,
2007. Photocatalysis for new energy production: Recent advances in photocatalytic
water splitting reactions for hydrogen production. Catalysis Today, 122(1-2), pp.51-61.
https://doi.org/10.1016/j.cattod.2007.01.042

[12] Fujishima, A. and Honda, K., 1972. Electrochemical photolysis of water at a


semiconductor electrode. nature, 238(5358), pp.37-38.
https://doi.org/10.1038/238037a0

[13] Liu, G., Sheng, Y., Ager, J.W., Kraft, M. and Xu, R., 2019. Research advances towards
large-scale solar hydrogen production from water. EnergyChem, 1(2), p.100014.
https://doi.org/10.1016/j.enchem.2019.100014

147
[14] He, K., Campbell, E., Huang, Z., Shen, R., Li, Q., Zhang, S., Zhong, Y.L., Zhang, P. and
Li, X., Metal carbide‐based cocatalysts for photocatalytic solar‐to‐fuel
conversion. Small Structures. https://doi.org/10.1002/sstr.202200104

[15] Lei, W., Zhou, T., Pang, X., Xue, S. and Xu, Q., 2022. Low-dimensional MXenes as
noble metal-free co-catalyst for solar-to-fuel production: Progress and
prospects. Journal of Materials Science & Technology, 114, pp.143-164.
https://doi.org/10.1016/j.jmst.2021.10.029

[16] Thomas, J.M., 2014. Heterogeneous catalysis and the challenges of powering the planet,
securing chemicals for civilised life, and clean efficient utilization of renewable
feedstocks. ChemSusChem, 7(7), pp.1801-1832.
https://doi.org/10.1002/cssc.201301202

[17] Martin, D.J., 2015. Investigation into high efficiency visible light photocatalysts for
water reduction and oxidation. Springer. https://doi.org/10.1007/978-3-319-18488-3

[18] Wang, Y., Suzuki, H., Xie, J., Tomita, O., Martin, D.J., Higashi, M., Kong, D., Abe, R.
and Tang, J., 2018. Mimicking natural photosynthesis: solar to renewable H2 fuel
synthesis by Z-scheme water splitting systems. Chemical reviews, 118(10), pp.5201-
5241. https://doi.org/10.1021/acs.chemrev.7b00286

[19] Ali, S.A. and Ahmad, T., 2022. Chemical strategies in molybdenum based chalcogenides
nanostructures for photocatalysis. International Journal of Hydrogen Energy, 47(68),
pp.29255-29283. https://doi.org/10.1016/j.ijhydene.2022.06.269

[20] Pivovar, B., Rustagi, N. and Satyapal, S., 2018. Hydrogen at scale (H2@ Scale): key to
a clean, economic, and sustainable energy system. The Electrochemical Society
Interface, 27(1), p.47. https://doi.org/10.1149/2.F04181if

[21] Lui, J., Chen, W.H., Tsang, D.C. and You, S., 2020. A critical review on the principles,
applications, and challenges of waste-to-hydrogen technologies. Renewable and
Sustainable Energy Reviews, 134, p.110365. https://doi.org/10.1016/j.rser.2020.110365

[22] Wallace, J.S. and Ward, C.A., 1983. Hydrogen as a fuel. International Journal of
Hydrogen Energy, 8(4), pp.255-268.https://doi.org/10.1016/0360-3199(83)90136-2

[23] Takata, T. and Domen, K., 2019. Particulate photocatalysts for water splitting: recent
advances and future prospects. ACS Energy Letters, 4(2), pp.542-549.
https://doi.org/10.1021/acsenergylett.8b02209

[24] Etacheri, V., Di Valentin, C., Schneider, J., Bahnemann, D. and Pillai, S.C., 2015.
Visible-light activation of TiO2 photocatalysts: Advances in theory and
experiments. Journal of Photochemistry and Photobiology C: Photochemistry
Reviews, 25, pp.1-29. https://doi.org/10.1016/j.jphotochemrev.2015.08.003

[25] Lin, L., Yu, Z. and Wang, X., 2019. Crystalline carbon nitride semiconductors for
photocatalytic water splitting. Angewandte Chemie, 131(19), pp.6225-6236.
https://doi.org/10.1002/ange.201809897

148
[26] Zhang, P., Zhang, J. and Gong, J., 2014. Tantalum-based semiconductors for solar water
splitting. Chemical Society Reviews, 43(13), pp.4395-4422.
https://doi.org/10.1039/C3CS60438A

[27] Ni, M., Leung, M.K., Leung, D.Y. and Sumathy, K., 2007. A review and recent
developments in photocatalytic water-splitting using TiO2 for hydrogen
production. Renewable and Sustainable Energy Reviews, 11(3), pp.401-
425.https://doi.org/10.1016/j.rser.2005.01.009

[28] Takata, T., Pan, C. and Domen, K., 2015. Recent progress in oxynitride photocatalysts
for visible-light-driven water splitting. Science and Technology of Advanced Materials.
https://doi.org/10.1088/1468-6996/16/3/033506

[29] Ali, S.A. and Ahmad, T., 2023. Treasure trove for efficient hydrogen evolution through
water splitting using diverse perovskite photocatalysts. Materials Today Chemistry, 29,
p.101387.https://doi.org/10.1016/j.mtchem.2023.101387

[30] Guo, L., Chen, Y., Su, J., Liu, M. and Liu, Y., 2019. Obstacles of solar-powered
photocatalytic water splitting for hydrogen production: a perspective from energy flow
and mass flow. Energy, 172, pp.1079-1086.
https://doi.org/10.1016/j.energy.2019.02.050

[31] Rahman, M.Z., Kibria, M.G. and Mullins, C.B., 2020. Metal-free photocatalysts for
hydrogen evolution. Chemical Society Reviews, 49(6), pp.1887-1931.
https://doi.org/10.1039/C9CS00313D

[32] Ahmad, H., Kamarudin, S.K., Minggu, L.J. and Kassim, M., 2015. Hydrogen from
photo-catalytic water splitting process: A review. Renewable and Sustainable Energy
Reviews, 43, pp.599-610. https://doi.org/10.1016/j.rser.2014.10.101

[33] Xie, G., Zhang, K., Guo, B., Liu, Q., Fang, L. and Gong, J.R., 2013. Graphene‐based
materials for hydrogen generation from light‐driven water splitting. Advanced
materials, 25(28), pp.3820-3839. https://doi.org/10.1002/adma.201301207

[34] Kim, J.H., Hansora, D., Sharma, P., Jang, J.W. and Lee, J.S., 2019. Toward practical
solar hydrogen production–an artificial photosynthetic leaf-to-farm challenge. Chemical
Society Reviews, 48(7), pp.1908-1971. https://doi.org/10.1039/C8CS00699G

[35] Wang, W., Xu, X., Zhou, W. and Shao, Z., 2017. Recent progress in metal‐organic
frameworks for applications in electrocatalytic and photocatalytic water
splitting. Advanced science, 4(4), p.1600371. https://doi.org/10.1002/advs.201600371

[36] Song, F., Li, W. and Sun, Y., 2017. Metal–organic frameworks and their derivatives for
photocatalytic water splitting. Inorganics, 5(3), p.40.
https://doi.org/10.3390/inorganics5030040

[37] Liu, Y., Huang, D., Cheng, M., Liu, Z., Lai, C., Zhang, C., Zhou, C., Xiong, W., Qin,
L., Shao, B. and Liang, Q., 2020. Metal sulfide/MOF-based composites as visible-light-
driven photocatalysts for enhanced hydrogen production from water
splitting. Coordination Chemistry Reviews, 409, p.213220.
https://doi.org/10.1016/j.ccr.2020.213220

149
[38] Huang, L. and Liu, B., 2016. Synthesis of a novel and stable reduced graphene
oxide/MOF hybrid nanocomposite and photocatalytic performance for the degradation
of dyes. RSC advances, 6(22), pp.17873-17879. https://doi.org/10.1039/C5RA25689E

[39] Hu, E., Yao, Y., Cui, Y. and Qian, G., 2021. Strategies for the enhanced water splitting
activity over metal–organic frameworks-based electrocatalysts and
photocatalysts. Materials Today Nano, 15, p.100124.
https://doi.org/10.1016/j.mtnano.2021.100124

[40] Lu, L., Wu, B., Shi, W. and Cheng, P., 2019. Metal–organic framework-derived
heterojunctions as nanocatalysts for photocatalytic hydrogen production. Inorganic
Chemistry Frontiers, 6(12), pp.3456-3467. https://doi.org/10.1039/C9QI00964G

[41] Nemiwal, M., Gosu, V., Zhang, T.C. and Kumar, D., 2021. Metal organic frameworks
as electrocatalysts: Hydrogen evolution reactions and overall water
splitting. International Journal of Hydrogen Energy, 46(17), pp.10216-10238.
https://doi.org/10.1016/j.ijhydene.2020.12.146

[42] Ding, Q., Gou, L., Wei, D., Xu, D., Fan, W. and Shi, W., 2021. Metal-organic framework
derived Co3O4/TiO2 heterostructure nanoarrays for promote photoelectrochemical water
splitting. International Journal of Hydrogen Energy, 46(49), pp.24965-24976.
https://doi.org/10.1016/j.ijhydene.2021.05.065

[43] Wu, T., Liu, X., Liu, Y., Cheng, M., Liu, Z., Zeng, G., Shao, B., Liang, Q., Zhang, W.
and He, Q., 2020. Application of QD-MOF composites for photocatalysis: Energy
production and environmental remediation. Coordination Chemistry Reviews, 403,
p.213097. https://doi.org/10.1016/j.ccr.2019.213097

[44] Baykara, S.Z., 2018. Hydrogen: A brief overview on its sources, production and
environmental impact. International Journal of Hydrogen Energy, 43(23), pp.10605-
10614. https://doi.org/10.1016/j.ijhydene.2018.02.022

[45] Ball, M. and Wietschel, M., 2009. The future of hydrogen–opportunities and
challenges. International journal of hydrogen energy, 34(2), pp.615-627.
https://doi.org/10.1016/j.ijhydene.2008.11.014

[46] Dunn, S., 2002. Hydrogen futures: toward a sustainable energy system. International
journal of hydrogen energy, 27(3), pp.235-264. https://doi.org/10.1016/S0360-
3199(01)00131-8

[47] T-raissi, A. and Block, D.L., 2004. Hydrogen: automotive fuel of the future. IEEE
Power and Energy Magazine, 2(6), pp.40-45.
https://doi.org/10.1109/MPAE.2004.1359020

[48] Møller, K.T., Jensen, T.R., Akiba, E. and Li, H.W., 2017. Hydrogen-A sustainable
energy carrier. Progress in Natural Science: Materials International, 27(1), pp.34-40.
https://doi.org/10.1016/j.pnsc.2016.12.014

150
[49] Navarro, R.M., Pena, M.A. and Fierro, J.L.G., 2007. Hydrogen production reactions
from carbon feedstocks: fossil fuels and biomass. Chemical reviews, 107(10), pp.3952-
3991. https://doi.org/10.1021/cr0501994

[50] Nazir, H., Muthuswamy, N., Louis, C., Jose, S., Prakash, J., Buan, M.E., Flox, C.,
Chavan, S., Shi, X., Kauranen, P. and Kallio, T., 2020. Is the H2 economy realizable in
the foreseeable future? Part II: H2 storage, transportation, and distribution. International
journal of hydrogen energy, 45(41), pp.20693-20708.
https://doi.org/10.1016/j.ijhydene.2020.05.241

[51] Marbán, G. and Valdés-Solís, T., 2007. Towards the hydrogen economy?. International
journal of hydrogen energy, 32(12), pp.1625-1637.
https://doi.org/10.1016/j.tej.2005.06.003

[52] Puga, A.V., 2016. Photocatalytic production of hydrogen from biomass-derived


feedstocks. Coordination Chemistry Reviews, 315, pp.1-66.
https://doi.org/10.1016/j.ccr.2015.12.009

[53] Nikolaidis, P. and Poullikkas, A., 2017. A comparative overview of hydrogen


production processes. Renewable and sustainable energy reviews, 67, pp.597-611.
https://doi.org/10.1016/j.rser.2016.09.044

[54] Steinberg, M. and Cheng, H.C., 1989. Modern and prospective technologies for
hydrogen production from fossil fuels. International Journal of Hydrogen
Energy, 14(11), pp.797-820. https://doi.org/10.1016/0360-3199(89)90018-9

[55] Acar, C. and Dincer, I., 2015. Impact assessment and efficiency evaluation of hydrogen
production methods. International journal of energy research, 39(13), pp.1757-1768.
https://doi.org/10.1002/er.3302

[56] Rafique, M., Mubashar, R., Irshad, M., Gillani, S.S.A., Tahir, M.B., Khalid, N.R.,
Yasmin, A. and Shehzad, M.A., 2020. A comprehensive study on methods and materials
for photocatalytic water splitting and hydrogen production as a renewable energy
resource. Journal of Inorganic and Organometallic Polymers and Materials, 30(10),
pp.3837-3861. https://doi.org/10.1007/s10904-020-01611-9

[57] Yuan, L., Han, C., Yang, M.Q. and Xu, Y.J., 2016. Photocatalytic water splitting for
solar hydrogen generation: fundamentals and recent advancements. International
Reviews in Physical Chemistry, 35(1), pp.1-36.
https://doi.org/10.1080/0144235X.2015.1127027

[58] Tee, S.Y., Win, K.Y., Teo, W.S., Koh, L.D., Liu, S., Teng, C.P. and Han, M.Y., 2017.
Recent progress in energy‐driven water splitting. Advanced science, 4(5), p.1600337.
https://doi.org/10.1002/advs.201600337

[59] Ishaq, H. and Dincer, I., 2021. Comparative assessment of renewable energy-based
hydrogen production methods. Renewable and Sustainable Energy Reviews, 135,
p.110192.https://doi.org/10.1016/j.rser.2020.110192

[60] Safari, F. and Dincer, I., 2020. A review and comparative evaluation of thermochemical
water splitting cycles for hydrogen production. Energy Conversion and
Management, 205, p.112182.https://doi.org/10.1016/j.enconman.2019.112182
151
[61] O'keefe, D.R., Norman, J.H. and Williamson, D.G., 1980. Catalysis research in
thermochemical water-splitting processes. Catalysis Reviews Science and
Engineering, 22(3), pp.325-369. https://doi.org/10.1080/03602458008067537

[62] Mehrpooya, M. and Habibi, R., 2020. A review on hydrogen production thermochemical
water-splitting cycles. Journal of Cleaner Production, 275, p.123836.
https://doi.org/10.1016/j.jclepro.2020.123836

[63] Li, X., Sun, X., Song, Q., Yang, Z., Wang, H. and Duan, Y., 2022. A critical review on
integrated system design of solar thermochemical water-splitting cycle for hydrogen
production. International Journal of Hydrogen Energy.
https://doi.org/10.1016/j.ijhydene.2022.07.249

[64] Oruc, O. and Dincer, I., 2021. Assessing the potential of thermo-chemical water splitting
cycles: A bridge towards clean and sustainable hydrogen generation. Fuel, 286,
p.119325. https://doi.org/10.1016/j.fuel.2020.119325

[65] Ni, M., Leung, M.K., Sumathy, K. and Leung, D.Y., 2006. Potential of renewable
hydrogen production for energy supply in Hong Kong. International journal of
hydrogen energy, 31(10), pp.1401-1412. https://doi.org/10.1002/er.1372

[66] Osterloh, F.E. and Parkinson, B.A., 2011. Recent developments in solar water-splitting
photocatalysis. MRS bulletin, 36(1), pp.17-22. https://doi.org/10.1557/mrs.2010.5

[67] Peter, L.M. and Upul Wijayantha, K.G., 2014. Photoelectrochemical water splitting at
semiconductor electrodes: fundamental problems and new
perspectives. ChemPhysChem, 15(10), pp.1983-1995.
https://doi.org/10.1002/cphc.201402024

[68] Zhang, K., Ma, M., Li, P., Wang, D.H. and Park, J.H., 2016. Water splitting progress in
tandem devices: moving photolysis beyond electrolysis. Advanced Energy
Materials, 6(15), p.1600602. https://doi.org/10.1002/aenm.201600602

[69] Yaqoob, L., Noor, T., Iqbal, N., Nasir, H., Sohail, M., Zaman, N. and Usman, M., 2020.
Nanocomposites of cobalt benzene tricarboxylic acid MOF with rGO: an efficient and
robust electrocatalyst for oxygen evolution reaction (OER). Renewable Energy, 156,
pp.1040-1054. https://doi.org/10.1016/j.renene.2020.04.131

[70] Yan, M., Tang, R.L., Liu, G.X., Huai, L., Liu, W. and Guo, S.P., 2022. Pb 5
(GeO4)(Ge2O7) and Pb3. 32Ca1. 68 (GeO4)(Ge2O7): Two Nonlinear Optical Germanates
Induced by Diverse PbO x Polyhedra. Inorganic Chemistry, 61(34), pp.13637-13643.
https://doi.org/10.1021/acs.inorgchem.2c02567

[71] Matheu, R., Garrido-Barros, P., Gil-Sepulcre, M., Ertem, M.Z., Sala, X., Gimbert-
Suriñach, C. and Llobet, A., 2019. The development of molecular water oxidation
catalysts. Nature Reviews Chemistry, 3(5), pp.331-341. https://doi.org/10.1038/s41570-
019-0096-0

152
[72] Wang, X., Dong, A., Zhu, Z., Chai, L., Ding, J., Zhong, L., Li, T.T., Hu, Y., Qian, J. and
Huang, S., 2020. Surfactant‐Mediated Morphological Evolution of MnCo Prussian Blue
Structures. Small, 16(43), p.2004614. https://doi.org/10.1002/smll.202004614

[73] Fang, B., Xing, Z., Sun, D., Li, Z. and Zhou, W., 2021. Hollow semiconductor
photocatalysts for solar energy conversion. Advanced Powder Materials.
https://doi.org/10.1016/j.apmate.2021.11.008

[74] Fang, B., Xing, Z., Sun, D., Li, Z. and Zhou, W., 2021. Hollow semiconductor
photocatalysts for solar energy conversion. Advanced Powder Materials.
https://doi.org/10.1021/acs.chemrev.9b00201

[75] Sun, Y., He, J., Yang, G., Sun, G. and Sage, V., 2019. A review of the enhancement of
bio-hydrogen generation by chemicals addition. Catalysts, 9(4),
p.353.https://doi.org/10.3390/catal9040353

[76] Boddy, P.J., 1968. Oxygen evolution on semiconducting TiO2. Journal of The
Electrochemical Society, 115(2), p.199.

[77] Chandra, M., Bhunia, K. and Pradhan, D., 2018. Controlled synthesis of CuS/TiO2
heterostructured nanocomposites for enhanced photocatalytic hydrogen generation
through water splitting. Inorganic chemistry, 57(8), pp.4524-4533.
https://doi.org/10.1021/acs.inorgchem.8b00283

[78] Ismael, M., 2020. A review and recent advances in solar-to-hydrogen energy conversion
based on photocatalytic water splitting over doped-TiO2 nanoparticles. Solar
Energy, 211, pp.522-546. https://doi.org/10.1016/j.solener.2020.09.073

[79] Horiuchi, Y., Toyao, T., Takeuchi, M., Matsuoka, M. and Anpo, M., 2013. Recent
advances in visible-light-responsive photocatalysts for hydrogen production and solar
energy conversion–from semiconducting TiO2 to MOF/PCP photocatalysts. Physical
Chemistry Chemical Physics, 15(32), pp.13243-13253.
https://doi.org/10.1039/C3CP51427G

[80] Chen, S., Takata, T. and Domen, K., 2017. Particulate photocatalysts for overall water
splitting. Nature Reviews Materials, 2(10), pp.1-17.
https://doi.org/10.1038/natrevmats.2017.50

[81] Navarro Yerga, R.M., Álvarez Galván, M.C., Del Valle, F., Villoria de la Mano, J.A.
and Fierro, J.L., 2009. Water splitting on semiconductor catalysts under visible‐light
irradiation. ChemSusChem: Chemistry & Sustainability Energy & Materials, 2(6),
pp.471-485. https://doi.org/10.1002/cssc.200900018

[82] Acar, C., Dincer, I. and Naterer, G.F., 2016. Review of photocatalytic water‐splitting
methods for sustainable hydrogen production. International Journal of Energy
Research, 40(11), pp.1449-1473. https://doi.org/10.1002/er.3549

[83] Maeda, K., 2011. Photocatalytic water splitting using semiconductor particles: history
and recent developments. Journal of Photochemistry and Photobiology C:
Photochemistry Reviews, 12(4), pp.237-268.
https://doi.org/10.1016/j.jphotochemrev.2011.07.001

153
[84] Maeda, K. and Domen, K., 2010. Photocatalytic water splitting: recent progress and
future challenges. The Journal of Physical Chemistry Letters, 1(18), pp.2655-
2661.https://doi.org/10.1021/jz1007966

[85] Scholz, J.T. and Stiftel, B., 2010. Introduction: the challenges of adaptive governance.
In Adaptive Governance and Water Conflict (pp. 15-26). Routledge.
https://doi.org/10.1016/B978-0-12-385469-8.00048-4

[86] Zhang, Z., Jiang, X., Liu, B., Guo, L., Lu, N., Wang, L., Huang, J., Liu, K. and Dong,
B., 2018. IR‐Driven Ultrafast Transfer of Plasmonic Hot Electrons in Nonmetallic
Branched Heterostructures for Enhanced H2 Generation. Advanced Materials, 30(9),
p.1705221.https://doi.org/10.1002/adma.201705221

[87] Peng, J., Shen, J., Yu, X., Tang, H. and Liu, Q., 2021. Construction of LSPR-enhanced
0D/2D CdS/MoO3− x S-scheme heterojunctions for visible-light-driven photocatalytic
H2 evolution. Chinese Journal of Catalysis, 42(1), pp.87-96.
https://doi.org/10.1016/S1872-2067(20)63595-1

[88] He, X., Liu, Q., Xu, D., Wang, L. and Tang, H., 2022. Plasmonic TiN nanobelts assisted
broad spectrum photocatalytic H2 generation. Journal of Materials Science &
Technology, 116, pp.1-10. https://doi.org/10.1016/j.jmst.2021.10.033

[89] Takanabe, K., 2017. Photocatalytic water splitting: quantitative approaches toward
photocatalyst by design. Acs Catalysis, 7(11), pp.8006-8022.
https://doi.org/10.1021/acscatal.7b02662

[90] Wang, Z., Li, C. and Domen, K., 2019. Recent developments in heterogeneous
photocatalysts for solar-driven overall water splitting. Chemical Society Reviews, 48(7),
pp.2109-2125. https://doi.org/10.1039/C8CS00542G

[91] Li, H., Xiao, J., Vequizo, J.J.M., Hisatomi, T., Nakabayashi, M., Pan, Z., Shibata, N.,
Yamakata, A., Takata, T. and Domen, K., 2022. One-Step Excitation Overall Water
Splitting over a Modified Mg-Doped BaTaO2N Photocatalyst. ACS Catalysis, 12(16),
pp.10179-10185. https://doi.org/10.1021/acscatal.2c02394

[92] Maeda, K., Takata, T., Hara, M., Saito, N., Inoue, Y., Kobayashi, H. and Domen, K.,
2005. GaN: ZnO solid solution as a photocatalyst for visible-light-driven overall water
splitting. Journal of the American Chemical Society, 127(23), pp.8286-
8287.https://doi.org/10.1021/ja0518777

[93] Higashi, M., Abe, R., Teramura, K., Takata, T., Ohtani, B. and Domen, K., 2008. Two
step water splitting into H2 and O2 under visible light by ATaO2N (A= Ca, Sr, Ba) and
WO3 with IO3-/I-shuttle redox mediator. Chemical Physics Letters, 452(1-3), pp.120-
123. https://doi.org/10.1016/j.cplett.2007.12.021

[94] Yu, J., Qi, L. and Jaroniec, M., 2010. Hydrogen production by photocatalytic water
splitting over Pt/TiO2 nanosheets with exposed (001) facets. The Journal of Physical
Chemistry C, 114(30), pp.13118-13125. https://doi.org/10.1627/jpi.56.280

[95] Maeda, K., 2013. Z-scheme water splitting using two different semiconductor
photocatalysts. ACS catalysis, 3(7), pp.1486-1503. https://doi.org/10.1021/cs4002089

154
[96] Ng, B.J., Putri, L.K., Kong, X.Y., Teh, Y.W., Pasbakhsh, P. and Chai, S.P., 2020. Z‐
scheme photocatalytic systems for solar water splitting. Advanced Science, 7(7),
p.1903171. https://doi.org/10.1002/advs.201903171

[97] Abdul Nasir, J., Munir, A., Ahmad, N., Haq, T.U., Khan, Z. and Rehman, Z., 2021.
Photocatalytic Z‐Scheme Overall Water Splitting: Recent Advances in Theory and
Experiments. Advanced Materials, 33(52), p.2105195.
https://doi.org/10.1002/adma.202105195

[98] Maeda, K., Takata, T., Hara, M., Saito, N., Inoue, Y., Kobayashi, H. and Domen, K.,
2005. GaN: ZnO solid solution as a photocatalyst for visible-light-driven overall water
splitting. Journal of the American Chemical Society, 127(23), pp.8286-
8287.https://doi.org/10.1016/S1010-6030(02)00070-9

[99] Yan, H., Wang, X., Yao, M. and Yao, X., 2013. Band structure design of semiconductors
for enhanced photocatalytic activity: The case of TiO2. Progress in Natural Science:
Materials International, 23(4), pp.402-407. https://doi.org/10.1016/j.pnsc.2013.06.002

[100] Niu, P. and Li, L., 2021. Photocatalytic overall water splitting of carbon nitride by band-
structure modulation. Matter, 4(6), pp.1765-1767.
https://doi.org/10.1016/j.matt.2021.04.022

[101] Zhang, C., Xie, C., Gao, Y., Tao, X., Ding, C., Fan, F. and Jiang, H.L., 2022. Charge
separation by creating band bending in metal–organic frameworks for improved
photocatalytic hydrogen evolution. Angewandte Chemie International Edition, 61(28),
p.e202204108.https://doi.org/10.1002/anie.202204108

[102] Jang, J.S., Kim, H.G. and Lee, J.S., 2012. Heterojunction semiconductors: A strategy to
develop efficient photocatalytic materials for visible light water splitting. Catalysis
today, 185(1), pp.270-277. https://doi.org/10.1016/j.cattod.2011.07.008

[103] Mehtab, A., Alshehri, S.M. and Ahmad, T., 2022. Photocatalytic and
photoelectrocatalytic water splitting by porous g-C3N4 nanosheets for hydrogen
generation. ACS Applied Nano Materials, 5(9), pp.12656-12665.
https://doi.org/10.1021/acsanm.2c02460

[104] Bhunia, K., Chandra, M., Sharma, S.K., Pradhan, D. and Kim, S.J., 2023. A critical
review on transition metal phosphide based catalyst for electrochemical hydrogen
evolution reaction: Gibbs free energy, composition, stability, and true identity of active
site. Coordination Chemistry Reviews, 478,
p.214956.https://doi.org/10.1016/j.ccr.2022.214956

[105] Fan, W., Chang, H., Zhong, J., Lu, J., Ma, G., Zhang, H., Jiang, Z. and Yin, G., 2023.
Facile synthesis of ZnCdS quantum dots via a novel photoetching MOF strategy for
boosting photocatalytic hydrogen evolution. Separation and Purification Technology,
p.125258. https://doi.org/10.1016/j.seppur.2023.125258

[106] Pramoda, K. and Rao, C.N.R., 2023. 2D transition metal-based phospho-chalcogenides


and their applications in photocatalytic and electrocatalytic hydrogen evolution
reactions. Journal of Materials Chemistry A, 11(32), pp.16933-16962.
https://doi.org/10.1039/D3TA01629C

155
[107] Remiro-Buenamanana, S., Cabrero-Antonino, M., Martinez-Guanter, M., Alvaro, M.,
Navalon, S. and Garcia, H., 2019. Influence of co-catalysts on the photocatalytic activity
of MIL-125 (Ti)-NH2 in the overall water splitting. Applied Catalysis B:
Environmental, 254, pp.677-684. https://doi.org/10.1016/j.apcatb.2019.05.027

[108] Zhang, J., Bai, T., Huang, H., Yu, M.H., Fan, X., Chang, Z. and Bu, X.H., 2020. Metal–
organic‐framework‐based photocatalysts optimized by spatially separated cocatalysts
for overall water splitting. Advanced Materials, 32(49), p.2004747.
https://doi.org/10.1002/adma.202004747

[109] Liu, X., Inagaki, S. and Gong, J., 2016. Heterogeneous molecular systems for
photocatalytic CO2 reduction with water oxidation. Angewandte Chemie International
Edition, 55(48), pp.14924-14950.https://doi.org/10.1002/anie.201600395

[110] Hu, H., Wang, Z., Cao, L., Zeng, L., Zhang, C., Lin, W. and Wang, C., 2021. Metal–
organic frameworks embedded in a liposome facilitate overall photocatalytic water
splitting. Nature Chemistry, 13(4), pp.358-366.https://doi.org/10.1038/s41557-020-
00635-5

[111] Yoon, M., Srirambalaji, R. and Kim, K., 2012. Homochiral metal–organic frameworks
for asymmetric heterogeneous catalysis. Chemical reviews, 112(2), pp.1196-1231.
https://doi.org/10.1021/cr2003147

[112] Pan, Y., Abazari, R., Yao, J. and Gao, J., 2021. Recent progress in 2D metal-organic
framework photocatalysts: Synthesis, photocatalytic mechanism and
applications. Journal of Physics: Energy, 3(3), p.032010.DOI 10.1088/2515-
7655/abf721

[113] Clough, A.J., Yoo, J.W., Mecklenburg, M.H. and Marinescu, S.C., 2015. Two-
dimensional metal–organic surfaces for efficient hydrogen evolution from
water. Journal of the American Chemical Society, 137(1), pp.118-121.
https://doi.org/10.1021/ja5116937

[114] Batten, S.R., Champness, N.R., Chen, X.M., Garcia-Martinez, J., Kitagawa, S.,
Öhrström, L., O'Keeffe, M., Suh, M.P. and Reedijk, J., 2012. Coordination polymers,

156
metal–organic frameworks and the need for terminology
guidelines. CrystEngComm, 14(9), pp.3001-3004. https://doi.org/10.1039/C2CE06488J

[115] Meek, S.T., Greathouse, J.A. and Allendorf, M.D., 2011. Metal‐organic frameworks: A
rapidly growing class of versatile nanoporous materials. Advanced materials, 23(2),
pp.249-267. https://doi.org/10.1002/adma.201002854

[116] Kortlever, R., Shen, J., Schouten, K.J.P., Calle-Vallejo, F. and Koper, M.T., 2015.
Catalysts and reaction pathways for the electrochemical reduction of carbon
dioxide. The journal of physical chemistry letters, 6(20), pp.4073-4082.

https://doi.org/10.1021/acs.jpclett.5b01559

[117] Xu, C., Liu, H., Li, D., Su, J.H. and Jiang, H.L., 2018. Direct evidence of charge
separation in a metal–organic framework: efficient and selective photocatalytic
oxidative coupling of amines via charge and energy transfer. Chemical science, 9(12),
pp.3152-3158. https://doi.org/10.1039/C7SC05296K

[118] Li, D., Xu, H.Q., Jiao, L. and Jiang, H.L., 2019. Metal-organic frameworks for catalysis:
State of the art, challenges, and opportunities. EnergyChem, 1(1), p.100005.
https://doi.org/10.1016/j.enchem.2019.100005

[119] Hao, X.L., Ma, Y.Y., Zang, H.Y., Wang, Y.H., Li, Y.G. and Wang, E.B., 2015. A
polyoxometalate‐encapsulating cationic metal–organic framework as a heterogeneous
catalyst for desulfurization. Chemistry–A European Journal, 21(9), pp.3778-3784.
https://doi.org/10.1002/chem.201405825

[120] Wu, H.B., Xia, B.Y., Yu, L., Yu, X.Y. and Lou, X.W.D., 2015. Porous molybdenum
carbide nano-octahedrons synthesized via confined carburization in metal-organic
frameworks for efficient hydrogen production. Nature communications, 6(1), pp.1-8.
https://doi.org/10.1038/ncomms7512

[121] Chen, Y.Z., Cai, G., Wang, Y., Xu, Q., Yu, S.H. and Jiang, H.L., 2016. Palladium
nanoparticles stabilized with N-doped porous carbons derived from metal–organic
frameworks for selective catalysis in biofuel upgrade: the role of catalyst
wettability. Green chemistry, 18(5), pp.1212-1217.
https://doi.org/10.1039/C5GC02530C

157
[122] Liu, Y., Su, Y., Quan, X., Fan, X., Chen, S., Yu, H., Zhao, H., Zhang, Y. and Zhao, J.,
2018. Facile ammonia synthesis from electrocatalytic N2 reduction under ambient
conditions on N-doped porous carbon. ACS Catalysis, 8(2), pp.1186-1191.
https://doi.org/10.1021/acscatal.7b02165

[123] Xia, T., Lin, Y., Li, W. and Ju, M., 2021. Photocatalytic degradation of organic
pollutants by MOFs based materials: A review. Chinese Chemical Letters, 32(10),
pp.2975-2984. https://doi.org/10.1016/j.cclet.2021.02.058

[124] Kataoka, Y., Sato, K., Miyazaki, Y., Masuda, K., Tanaka, H., Naito, S. and Mori, W.,
2009. Photocatalytic hydrogen production from water using porous material [Ru 2 (p-
BDC)2]n. Energy & Environmental Science, 2(4), pp.397-400.
https://doi.org/10.1039/B814539C

[125] Zhao, S.N., Song, X.Z., Song, S.Y. and Zhang, H.J., 2017. Highly efficient
heterogeneous catalytic materials derived from metal-organic framework
supports/precursors. Coordination Chemistry Reviews, 337, pp.80-96.
https://doi.org/10.1016/j.ccr.2017.02.010

[126] Dang, S., Zhu, Q.L. and Xu, Q., 2017. Nanomaterials derived from metal–organic
frameworks. Nature Reviews Materials, 3(1), pp.1-14.
https://doi.org/10.1038/natrevmats.2017.75

[127] Zaman, N., Iqbal, N. and Noor, T., 2022. Advances and challenges of MOF derived
carbon-based electrocatalysts and photocatalyst for water splitting: A review. Arabian
Journal of Chemistry, p.103906. https://doi.org/10.1016/j.arabjc.2022.103906

[128] Kang, Y.S., Lu, Y., Chen, K., Zhao, Y., Wang, P. and Sun, W.Y., 2019. Metal–organic
frameworks with catalytic centers: From synthesis to catalytic application. Coordination
chemistry reviews, 378, pp.262-280. https://doi.org/10.1016/j.ccr.2018.02.009

[129] Ali, M., Pervaiz, E., Noor, T., Rabi, O., Zahra, R. and Yang, M., 2021. Recent
advancements in MOF‐based catalysts for applications in electrochemical and
photoelectrochemical water splitting: A review. International Journal of Energy
Research, 45(2), pp.1190-1226. https://doi.org/10.1002/er.5807

[130] Sun, D., Ye, L., Sun, F., García, H. and Li, Z., 2017. From mixed-metal MOFs to carbon-
coated Core–Shell metal alloy@ metal oxide solid solutions: transformation of Co/Ni-
MOF-74 to Co x Ni1–x@ Co y Ni1–y O@ C for the oxygen evolution reaction. Inorganic
chemistry, 56(9), pp.5203-5209. https://doi.org/10.1021/acs.inorgchem.7b00333

[131] Furukawa, S., Reboul, J., Diring, S., Sumida, K. and Kitagawa, S., 2014. Structuring of
metal–organic frameworks at the mesoscopic/macroscopic scale. Chemical Society
Reviews, 43(16), pp.5700-5734. https://doi.org/10.1039/C4CS00106K

[132] Sharmin, E. and Zafar, F., 2016. Introductory chapter: metal organic frameworks
(MOFs). In Metal-organic frameworks. IntechOpen. DOI: 10.5772/64797

158
[133] Guerrero, V.V., Yoo, Y., McCarthy, M.C. and Jeong, H.K., 2010. HKUST-1 membranes
on porous supports using secondary growth. Journal of Materials Chemistry, 20(19),
pp.3938-3943. https://doi.org/10.1039/B924536G

[134] Vaid, T.P., Kelley, S.P. and Rogers, R.D., 2017. Structure-directing effects of ionic
liquids in the ionothermal synthesis of metal–organic frameworks. IUCrJ, 4(4), pp.380-
392. http://dx.doi.org/10.1107/S2052252517008326

[135] Freund, R., Zaremba, O., Arnauts, G., Ameloot, R., Skorupskii, G., Dincă, M.,
Bavykina, A., Gascon, J., Ejsmont, A., Goscianska, J. and Kalmutzki, M., 2021. The
current status of MOF and COF applications. Angewandte Chemie International
Edition, 60(45), pp.23975-24001. https://doi.org/10.1002/anie.202106259

[136] Masoomi, M.Y., Morsali, A., Dhakshinamoorthy, A. and Garcia, H., 2019. Mixed‐metal
MOFs: unique opportunities in metal–organic framework (MOF) functionality and
design. Angewandte Chemie, 131(43), pp.15330-15347.
https://doi.org/10.1002/ange.201902229

[137] Brandenberger, S., Kröcher, O., Tissler, A. and Althoff, R., 2008. The state of the art in
selective catalytic reduction of NOx by ammonia using metal‐exchanged zeolite
catalysts. Catalysis Reviews, 50(4), pp.492-531.
https://doi.org/10.1080/01614940802480122

[138] Zhang, C., Wang, W., Duan, A., Zeng, G., Huang, D., Lai, C., Tan, X., Cheng, M.,
Wang, R., Zhou, C. and Xiong, W., 2019. Adsorption behavior of engineered carbons
and carbon nanomaterials for metal endocrine disruptors: experiments and theoretical
calculation. Chemosphere, 222, pp.184-194.
https://doi.org/10.1016/j.chemosphere.2019.01.128

[139] Wang, H., Zeng, Z., Xu, P., Li, L., Zeng, G., Xiao, R., Tang, Z., Huang, D., Tang, L.,
Lai, C. and Jiang, D., 2019. Recent progress in covalent organic framework thin films:
fabrications, applications and perspectives. Chemical Society Reviews, 48(2), pp.488-
516. https://doi.org/10.1039/C8CS00376A

[140] Kalhorizadeh, T., Dahrazma, B., Zarghami, R., Mirzababaei, S., Kirillov, A.M. and
Abazari, R., 2022. Quick removal of metronidazole from aqueous solutions using metal–
organic frameworks. New Journal of Chemistry, 46(19), pp.9440-9450.
https://doi.org/10.1039/D1NJ06107K

[141] Shi, Y., Yang, A.F., Cao, C.S. and Zhao, B., 2019. Applications of MOFs: Recent
advances in photocatalytic hydrogen production from water. Coordination Chemistry
Reviews, 390, pp.50-75. https://doi.org/10.1016/j.ccr.2019.03.012

[142] Xiao, Y., Guo, X., Yang, N. and Zhang, F., 2021. Heterostructured MOFs photocatalysts
for water splitting to produce hydrogen. Journal of Energy Chemistry, 58, pp.508-522.
https://doi.org/10.1016/j.jechem.2020.10.008

[143] Qu, Y. and Duan, X., 2013. Progress, challenge and perspective of heterogeneous
photocatalysts. Chemical Society Reviews, 42(7), pp.2568-2580.
https://doi.org/10.1039/C2CS35355E
159
[144] Li, X., Wang, Z. and Wang, L., 2021. Metal–Organic Framework‐Based Materials for
Solar Water Splitting. Small Science, 1(5), p.2000074.
https://doi.org/10.1002/smsc.202000074

[145] Abodunrin, T.O., Oladipo, A.C., Oladeji, S.O., Bankole, D.T., Egharevba, G.O. and
Bello, O.S., 2022. MOF-based photocatalysts for hydrogen generation by water splitting.
In Metal-Organic Framework-Based Nanomaterials for Energy Conversion and
Storage (pp. 709-726). Elsevier. https://doi.org/10.1016/B978-0-323-91179-5.00028-0

[146] Zhang, X., Tong, S., Huang, D., Liu, Z., Shao, B., Liang, Q., Wu, T., Pan, Y., Huang,
J., Liu, Y. and Cheng, M., 2021. Recent advances of Zr based metal organic frameworks
photocatalysis: Energy production and environmental remediation. Coordination
Chemistry Reviews, 448, p.214177. https://doi.org/10.1016/j.ccr.2021.214177

[147] Zheng, S., Li, X., Yan, B., Hu, Q., Xu, Y., Xiao, X., Xue, H. and Pang, H., 2017.
Transition‐metal (Fe, Co, Ni) based metal‐organic frameworks for electrochemical
energy storage. Advanced Energy Materials, 7(18), p.1602733.
https://doi.org/10.1002/aenm.201602733

[148] Patial, S., Raizada, P., Hasija, V., Singh, P., Thakur, V.K. and Nguyen, V.H., 2021.
Recent advances in photocatalytic multivariate metal organic frameworks-based
nanostructures toward renewable energy and the removal of environmental
pollutants. Materials Today Energy, 19, p.100589.
https://doi.org/10.1016/j.mtener.2020.100589

[149] Gautam, S., Agrawal, H., Thakur, M., Akbari, A., Sharda, H., Kaur, R. and Amini, M.,
2020. Metal oxides and metal organic frameworks for the photocatalytic degradation: A
review. Journal of Environmental Chemical Engineering, 8(3), p.103726.
https://doi.org/10.1016/j.jece.2020.103726

[150] Younis, S.A., Kwon, E.E., Qasim, M., Kim, K.H., Kim, T., Kukkar, D., Dou, X. and Ali,
I., 2020. Metal-organic framework as a photocatalyst: Progress in modulation strategies
and environmental/energy applications. Progress in Energy and Combustion
Science, 81, p.100870. https://doi.org/10.1016/j.pecs.2020.100870

[151] Kataoka, Y., Sato, K., Miyazaki, Y., Masuda, K., Tanaka, H., Naito, S. and Mori, W.,
2009. Photocatalytic hydrogen production from water using porous material [Ru2 (p-
BDC)2] n. Energy & Environmental Science, 2(4), pp.397-400. DOI: 10.1039/b814539c

[152] Tasleem, S., Tahir, M. and Khalifa, W.A., 2021. Current trends in structural
development and modification strategies for metal-organic frameworks (MOFs) towards
photocatalytic H2 production: a review. International Journal of Hydrogen
Energy, 46(27), pp.14148-14189. https://doi.org/10.1016/j.ijhydene.2021.01.162

[153] Iftekhar, S., Heidari, G., Amanat, N., Zare, E.N., Asif, M.B., Hassanpour, M., Lehto,
V.P. and Sillanpaa, M., 2022. Porous materials for the recovery of rare earth elements,
platinum group metals, and other valuable metals: a review. Environmental Chemistry
Letters, pp.1-50. https://doi.org/10.1007/s10311-022-01486-x

160
[154] Azharuddin, M., Zhu, G.H., Das, D., Ozgur, E., Uzun, L., Turner, A.P. and Patra, H.K.,
2019. A repertoire of biomedical applications of noble metal nanoparticles. Chemical
Communications, 55(49), pp.6964-6996. https://doi.org/10.1039/C9CC01741K

[155] Huang, W., Meng, H., Gao, Y., Wang, J., Yang, C., Liu, D., Liu, J., Guo, C., Yang, B.
and Cao, W., 2019. Metallic tungsten carbide nanoparticles as a near-infrared-driven
photocatalyst. Journal of Materials Chemistry A, 7(31), pp.18538-18546.
https://doi.org/10.1039/C9TA03151K

[156] Rui, K., Zhao, G., Lao, M., Cui, P., Zheng, X., Zheng, X., Zhu, J., Huang, W., Dou, S.X.
and Sun, W., 2019. Direct hybridization of noble metal nanostructures on 2D metal–
organic framework nanosheets to catalyze hydrogen evolution. Nano Letters, 19(12),
pp.8447-8453. https://doi.org/10.1021/acs.nanolett.9b02729

[157] Zhang, F., Zhu, Y., Lin, Q., Zhang, L., Zhang, X. and Wang, H., 2021. Noble-metal
single-atoms in thermocatalysis, electrocatalysis, and photocatalysis. Energy &
Environmental Science, 14(5), pp.2954-3009. https://doi.org/10.1039/D1EE00247C

[158] Gao, F., Zhang, Y., Wu, Z., You, H. and Du, Y., 2021. Universal strategies to multi-
dimensional noble-metal-based catalysts for electrocatalysis. Coordination Chemistry
Reviews, 436, p.213825. https://doi.org/10.1016/j.ccr.2021.213825

[159] Dong, K., Lei, Y., Zhao, H., Liang, J., Ding, P., Liu, Q., Xu, Z., Lu, S., Li, Q. and Sun,
X., 2020. Noble-metal-free electrocatalysts toward H2 O2 production. Journal of
Materials Chemistry A, 8(44), pp.23123-23141. https://doi.org/10.1039/D0TA08894C

[160] Reddy, D.A., Kim, Y., Gopannagari, M., Kumar, D.P. and Kim, T.K., 2021. Recent
advances in metal–organic framework-based photocatalysts for hydrogen
production. Sustainable Energy & Fuels, 5(6), pp.1597-1618.
https://doi.org/10.1039/C9SE00749K

[161] Lan, G., Zhu, Y.Y., Veroneau, S.S., Xu, Z., Micheroni, D. and Lin, W., 2018. Electron
injection from photoexcited metal–organic framework ligands to Ru2 secondary building
units for visible-light-driven hydrogen evolution. Journal of the American Chemical
Society, 140(16), pp.5326-5329. http://dx.doi.org/10.1021/jacs.8b01601

[162] Maeda, K., Teramura, K., Lu, D., Saito, N., Inoue, Y. and Domen, K., 2006. Noble‐
metal/Cr2O3 core/shell nanoparticles as a cocatalyst for photocatalytic overall water
splitting. Angewandte Chemie, 118(46), pp.7970-7973.
https://doi.org/10.1002/ange.200602473

[163] Liu, M., Hof, F., Moro, M., Valenti, G., Paolucci, F. and Pénicaud, A., 2020. Carbon
supported noble metal nanoparticles as efficient catalysts for electrochemical water
splitting. Nanoscale, 12(39), pp.20165-20170. https://doi.org/10.1039/D0NR05659F

[164] Sakamoto, N., Ohtsuka, H., Ikeda, T., Maeda, K., Lu, D., Kanehara, M., Teramura, K.,
Teranishi, T. and Domen, K., 2009. Highly dispersed noble-metal/chromia (core/shell)
nanoparticles as efficient hydrogen evolution promoters for photocatalytic overall water
splitting under visible light. Nanoscale, 1(1), pp.106-109.
https://doi.org/10.1039/B9NR00186G
161
[165] Guo, R., Lai, X., Huang, J., Du, X., Yan, Y., Sun, Y., Zou, G. and Xiong, J., 2018.
Phosphate‐based electrocatalysts for water splitting: recent
progress. ChemElectroChem, 5(24), pp.3822-3834.
https://doi.org/10.1002/celc.201800996

[166] Tao, Z., Wang, T., Wang, X., Zheng, J. and Li, X., 2016. MOF-derived noble metal free
catalysts for electrochemical water splitting. ACS applied materials & interfaces, 8(51),
pp.35390-35397. https://doi.org/10.1021/acsami.6b13411

[167] Shi, D., Zheng, R., Liu, C.S., Chen, D.M., Zhao, J. and Du, M., 2019. Dual-
functionalized mixed Keggin-and Lindqvist-type Cu24-based POM@ MOF for visible-
light-driven H2 and O2 evolution. Inorganic Chemistry, 58(11), pp.7229-7235.
https://doi.org/10.1021/acs.inorgchem.9b00206

[168] Yang, H., Zhao, M., Zhang, J., Ma, J., Wu, P., Liu, W. and Wen, L., 2019. A noble-
metal-free photocatalyst system obtained using BODIPY-based MOFs for highly
efficient visible-light-driven H 2 evolution. Journal of Materials Chemistry A, 7(36),
pp.20742-20749. https://doi.org/10.1039/C9TA06989E

[169] Luo, S., Liu, X., Wei, X., Fu, M., Lu, P., Li, X., Jia, Y., Ren, Q. and He, Y., 2020. Noble-
metal-free cobaloxime coupled with metal-organic frameworks NH2-MIL-125: a novel
bifunctional photocatalyst for photocatalytic NO removal and H2 evolution under visible
light irradiation. Journal of Hazardous Materials, 399, p.122824.
https://doi.org/10.1016/j.jhazmat.2020.122824

[170] Singh, A.K., Gonuguntla, S., Mahajan, B. and Pal, U., 2019. Noble metal-free integrated
UiO-66-PANI-Co 3 O 4 catalyst for visible-light-induced H 2 production. Chemical
Communications, 55(96), pp.14494-14497. https://doi.org/10.1039/C9CC07414G

[171] Hu, Y., Abazari, R., Sanati, S., Nadafan, M., Carpenter-Warren, C.L., Slawin, A.M.,
Zhou, Y. and Kirillov, A.M., 2023. A Dual-Purpose Ce (III)–Organic Framework with
Amine Groups and Open Metal Sites: Third-Order Nonlinear Optical Activity and
Catalytic CO2 Fixation. ACS Applied Materials & Interfaces, 15(31), pp.37300-
37311.https://doi.org/10.1021/acsami.3c04506

[172] Wang, J., Abazari, R., Sanati, S., Ejsmont, A., Goscianska, J., Zhou, Y. and Dubal, D.P.,
2023. Water‐Stable Fluorous Metal–Organic Frameworks with Open Metal Sites and
Amine Groups for Efficient Urea Electrocatalytic Oxidation. Small,
p.2300673.https://doi.org/10.1002/smll.202300673

[173] Luo, H., Zeng, Z., Zeng, G., Zhang, C., Xiao, R., Huang, D., Lai, C., Cheng, M., Wang,
W., Xiong, W. and Yang, Y., 2020. Recent progress on metal-organic frameworks
based-and derived-photocatalysts for water splitting. Chemical Engineering
Journal, 383, p.123196. https://doi.org/10.1016/j.cej.2019.123196

[174] Dong, X.Y., Zhang, M., Pei, R.B., Wang, Q., Wei, D.H., Zang, S.Q., Fan, Y.T. and Mak,
T.C., 2016. A crystalline copper (II) coordination polymer for the efficient visible‐light‐
driven generation of hydrogen. Angewandte Chemie International Edition, 55(6),
pp.2073-2077. https://doi.org/10.1002/anie.201509744

162
[175] Zhao, J., Wang, Y., Zhou, J., Qi, P., Li, S., Zhang, K., Feng, X., Wang, B. and Hu, C.,
2016. A copper (II)-based MOF film for highly efficient visible-light-driven hydrogen
production. Journal of Materials Chemistry A, 4(19), pp.7174-
7177.https://doi.org/10.1039/C6TA00431H

[176] Song, Y., Xin, X., Guo, S., Zhang, Y., Yang, L., Wang, B. and Li, X., 2020. One-step
MOFs-assisted synthesis of intimate contact MoP-Cu3P hybrids for photocatalytic water
splitting. Chemical Engineering Journal, 384, p.123337.
https://doi.org/10.1016/j.cej.2019.123337

[177] Li, L., Zhao, Y., Wang, Q., Liu, Z.Y., Wang, X.G., Yang, E.C. and Zhao, X.J., 2021.
Boosting photocatalytic hydrogen production activity by a microporous Cu II-MOF
nanoribbon decorated with Pt nanoparticles. Inorganic Chemistry Frontiers, 8(14),
pp.3556-3565. https://doi.org/10.1039/D1QI00516B

[178] Ali, M., Pervaiz, E., Sikandar, U. and Khan, Y., 2021. A review on the recent
developments in zirconium and carbon-based catalysts for photoelectrochemical water-
splitting. International Journal of Hydrogen Energy, 46(35), pp.18257-18283.
https://doi.org/10.1016/j.ijhydene.2021.02.202

[179] Hou, W., Chen, M., Chen, C., Wang, Y. and Xu, Y., 2021. Increased production of H2
under visible light by packing CdS in a Ti, Zr-Based metal organic framework. Journal
of Colloid and Interface Science, 604, pp.310-318.
https://doi.org/10.1016/j.jcis.2021.06.150

[180] Cavka, J.H., Jakobsen, S., Olsbye, U., Guillou, N., Lamberti, C., Bordiga, S. and
Lillerud, K.P., 2008. A new zirconium inorganic building brick forming metal organic
frameworks with exceptional stability. Journal of the American Chemical
Society, 130(42), pp.13850-13851. https://doi.org/10.1021/ja8057953

[181] Kim, M. and Cohen, S.M., 2012. Discovery, development, and functionalization of Zr
(iv)-based metal–organic frameworks. CrystEngComm, 14(12), pp.4096-4104.
https://doi.org/10.1039/C2CE06491J

[182] Kandiah, M., Nilsen, M.H., Usseglio, S., Jakobsen, S., Olsbye, U., Tilset, M., Larabi,
C., Quadrelli, E.A., Bonino, F. and Lillerud, K.P., 2010. Synthesis and stability of tagged
UiO-66 Zr-MOFs. Chemistry of Materials, 22(24), pp.6632-6640.
https://doi.org/10.1021/cm102601v

[183] Gomes Silva, C., Luz, I., Llabres i Xamena, F.X., Corma, A. and García, H., 2010.
Water stable Zr–benzenedicarboxylate metal–organic frameworks as photocatalysts for
hydrogen generation. Chemistry–A European Journal, 16(36), pp.11133-11138.
https://doi.org/10.1002/chem.200903526

[184] Hou, C.C., Li, T.T., Cao, S., Chen, Y. and Fu, W.F., 2015. Incorporation of a [Ru
(dcbpy)(bpy) 2] 2+ photosensitizer and a Pt (dcbpy) Cl 2 catalyst into metal–organic
frameworks for photocatalytic hydrogen evolution from aqueous solution. Journal of
Materials Chemistry A, 3(19), pp.10386-10394. https://doi.org/10.1039/C5TA01135C

[185] Shen, L., Liang, S., Wu, W., Liang, R. and Wu, L., 2013. CdS-decorated UiO–66 (NH
2) nanocomposites fabricated by a facile photodeposition process: an efficient and stable

163
visible-light-driven photocatalyst for selective oxidation of alcohols. Journal of
Materials Chemistry A, 1(37), pp.11473-11482. https://doi.org/10.1039/C3TA12645E

[186] Sun, D., Fu, Y., Liu, W., Ye, L., Wang, D., Yang, L., Fu, X. and Li, Z., 2013. Studies
on photocatalytic CO2 reduction over NH2‐Uio‐66 (Zr) and its derivatives: towards a
better understanding of photocatalysis on metal–organic frameworks. Chemistry–A
European Journal, 19(42), pp.14279-14285. https://doi.org/10.1002/chem.201301728

[187] Lin, R., Shen, L., Ren, Z., Wu, W., Tan, Y., Fu, H., Zhang, J. and Wu, L., 2014.
Enhanced photocatalytic hydrogen production activity via dual modification of MOF
and reduced graphene oxide on CdS. Chemical Communications, 50(62), pp.8533-
8535. https://doi.org/10.1039/C4CC01776E

[188] He, J., Wang, J., Chen, Y., Zhang, J., Duan, D., Wang, Y. and Yan, Z., 2014. A dye-
sensitized Pt@ UiO-66 (Zr) metal–organic framework for visible-light photocatalytic
hydrogen production. Chemical communications, 50(53), pp.7063-7066.
https://doi.org/10.1039/C4CC01086H

[189] Sabo, M., Böhlmann, W. and Kaskel, S., 2006. Titanium terephthalate (TT-1) hybrid
materials with high specific surface area. Journal of Materials Chemistry, 16(24),
pp.2354-2357. https://doi.org/10.1039/B601043A

[190] Assi, H., Pardo Pérez, L.C., Mouchaham, G., Ragon, F., Nasalevich, M., Guillou, N.,
Martineau, C., Chevreau, H., Kapteijn, F., Gascon, J. and Fertey, P., 2016. Investigating
the case of titanium (IV) carboxyphenolate photoactive coordination
polymers. Inorganic chemistry, 55(15), pp.7192-7199.
https://doi.org/10.1021/acs.inorgchem.6b01060

[191] Sun, D., Liu, W., Qiu, M., Zhang, Y. and Li, Z., 2015. Introduction of a mediator for
enhancing photocatalytic performance via post-synthetic metal exchange in metal–
organic frameworks (MOFs). Chemical communications, 51(11), pp.2056-2059.
https://doi.org/10.1039/C4CC09407G

[192] Yuan, S., Qin, J.S., Xu, H.Q., Su, J., Rossi, D., Chen, Y., Zhang, L., Lollar, C., Wang,
Q., Jiang, H.L. and Son, D.H., 2018. [Ti8Zr2O12 (COO) 16] Cluster: an ideal inorganic
building unit for photoactive metal–organic frameworks. ACS central science, 4(1),
pp.105-111. https://doi.org/10.1021/acscentsci.7b00497

[193] Nasalevich, M.A., Goesten, M.G., Savenije, T.J., Kapteijn, F. and Gascon, J., 2013.
Enhancing optical absorption of metal–organic frameworks for improved visible light
photocatalysis. Chemical Communications, 49(90), pp.10575-10577.
https://doi.org/10.1039/C3CC46398B

[194] Guo, F., Guo, J.H., Wang, P., Kang, Y.S., Liu, Y., Zhao, J. and Sun, W.Y., 2019. Facet-
dependent photocatalytic hydrogen production of metal–organic framework NH 2-MIL-
125 (Ti). Chemical science, 10(18), pp.4834-4838.
https://doi.org/10.1039/C8SC05060K

[195] Han, S.Y., Pan, D.L., Chen, H., Bu, X.B., Gao, Y.X., Gao, H., Tian, Y., Li, G.S., Wang,
G., Cao, S.L. and Wan, C.Q., 2018. A Methylthio‐Functionalized‐MOF Photocatalyst

164
with High Performance for Visible‐Light‐Driven H2 Evolution. Angewandte Chemie
International Edition, 57(31), pp.9864-9869. https://doi.org/10.1002/anie.201806077

[196] Yuan, S., Liu, T.F., Feng, D., Tian, J., Wang, K., Qin, J., Zhang, Q., Chen, Y.P., Bosch,
M., Zou, L. and Teat, S.J., 2015. A single crystalline porphyrinic titanium metal–
organic framework. Chemical science, 6(7), pp.3926-3930.
https://doi.org/10.1039/C5SC00916B

[197] Wang, J., Liu, Y., Li, Y., Xia, L., Jiang, M. and Wu, P., 2018. Highly efficient visible-
light-driven H2 production via an Eosin Y-based metal–organic framework. Inorganic
Chemistry, 57(13), pp.7495-7498. https://doi.org/10.1021/acs.inorgchem.8b00718

[198] Xiao, Y., Qi, Y., Wang, X., Wang, X., Zhang, F. and Li, C., 2018. Visible‐light‐
responsive 2D cadmium–organic framework single crystals with dual functions of water
reduction and oxidation. Advanced Materials, 30(44), p.1803401.
https://doi.org/10.1002/adma.201803401

[199] Song, T., Zhang, P., Zeng, J., Wang, T., Ali, A. and Zeng, H., 2017. Tunable conduction
band energy and metal-to-ligand charge transfer for wide-spectrum photocatalytic H2
evolution and stability from isostructural metal-organic frameworks. International
Journal of Hydrogen Energy, 42(43), pp.26605-26616.
https://doi.org/10.1016/j.ijhydene.2017.09.081

[200] Leng, F., Liu, H., Ding, M., Lin, Q.P. and Jiang, H.L., 2018. Boosting photocatalytic
hydrogen production of porphyrinic MOFs: the metal location in metalloporphyrin
matters. ACS Catalysis, 8(5), pp.4583-4590. https://doi.org/10.1021/acscatal.8b00764

[201] Lu, C., Xiong, D., Chen, C., Wang, J., Kong, Y., Liu, T., Ying, S. and Yi, F.Y., 2022.
Indium-Based Metal–Organic Framework for Efficient Photocatalytic Hydrogen
Evolution. Inorganic Chemistry, 61(5), pp.2587-
2594.https://doi.org/10.1021/acs.inorgchem.1c03628

[202] Guo, W., Chen, Z., Yang, C., Neumann, T., Kübel, C., Wenzel, W., Welle, A., Pfleging,
W., Shekhah, O., Wöll, C. and Redel, E., 2016. Bi 2 O 3 nanoparticles encapsulated in
surface mounted metal–organic framework thin films. Nanoscale, 8(12), pp.6468-6472.
https://doi.org/10.1039/C6NR00532B

[203] Zhao, X., Zhang, Y., Wen, P., Xu, G., Ma, D. and Qiu, P., 2018. NH2-MIL-125 (Ti)/TiO2
composites as superior visible-light photocatalysts for selective oxidation of
cyclohexane. Molecular Catalysis, 452, pp.175-183.
https://doi.org/10.1016/j.mcat.2018.04.004

[204] Panda, R., Rahut, S. and Basu, J.K., 2016. Preparation of a Fe 2 O 3/MIL-53 (Fe)
composite by partial thermal decomposition of MIL-53 (Fe) nanorods and their
photocatalytic activity. RSC advances, 6(84), pp.80981-80985.
https://doi.org/10.1039/C6RA15792K

[205] Liu, Q., Low, Z.X., Li, L., Razmjou, A., Wang, K., Yao, J. and Wang, H., 2013. ZIF-
8/Zn 2 GeO4 nanorods with an enhanced CO 2 adsorption property in an aqueous medium

165
for photocatalytic synthesis of liquid fuel. Journal of Materials Chemistry A, 1(38),
pp.11563-11569. https://doi.org/10.1039/C3TA12433A

[206] Subudhi, S., Mansingh, S., Swain, G., Behera, A., Rath, D. and Parida, K., 2019. HPW-
anchored UiO-66 metal–organic framework: a promising photocatalyst effective toward
tetracycline hydrochloride degradation and H2 evolution via Z-scheme charge
dynamics. Inorganic Chemistry, 58(8), pp.4921-4934.
https://doi.org/10.1021/acs.inorgchem.8b03544

[207] Zeng, X., Huang, L., Wang, C., Wang, J., Li, J. and Luo, X., 2016. Sonocrystallization
of ZIF-8 on electrostatic spinning TiO2 nanofibers surface with enhanced photocatalysis
property through synergistic effect. ACS Applied Materials & Interfaces, 8(31),
pp.20274-20282. https://doi.org/10.1021/acsami.6b05746

[208] Hu, P., Morabito, J.V. and Tsung, C.K., 2013. ACS Catal. 2014, 4, 4409.(d) Furukawa,
H.; Cordova, KE; O’Keeffe, M.; Yaghi, OM. Science, 341, p.1230444.
https://doi.org/10.1126/science.1230444

[209] He, L., Li, L., Wang, T., Gao, H., Li, G., Wu, X., Su, Z. and Wang, C., 2014. Fabrication
of Au/ZnO nanoparticles derived from ZIF-8 with visible light photocatalytic hydrogen
production and degradation dye activities. Dalton Transactions, 43(45), pp.16981-
16985. https://doi.org/10.1039/C4DT02557A

[210] Yan, B., Zhang, L., Tang, Z., Al-Mamun, M., Zhao, H. and Su, X., 2017. Palladium-
decorated hierarchical titania constructed from the metal-organic frameworks NH2-
MIL-125 (Ti) as a robust photocatalyst for hydrogen evolution. Applied Catalysis B:
Environmental, 218, pp.743-750. http://dx.doi.org/doi:10.1016/j.apcatb.2017.07.020

[211] Zhao, C.W., Li, Y.A., Wang, X.R., Chen, G.J., Liu, Q.K., Ma, J.P. and Dong, Y.B.,
2015. Fabrication of Cd (ii)-MOF-based ternary photocatalytic composite materials for
H2 production via a gel-to-crystal approach. Chemical Communications, 51(88),
pp.15906-15909. https://doi.org/10.1039/C5CC06291H

[212] Hao, X., Jin, Z., Yang, H., Lu, G. and Bi, Y., 2017. Peculiar synergetic effect of MoS 2
quantum dots and graphene on Metal-Organic Frameworks for photocatalytic hydrogen
evolution. Applied Catalysis B: Environmental, 210, pp.45-56.
https://doi.org/10.1016/j.apcatb.2017.03.057

[213] Zhang, C.F., Qiu, L.G., Ke, F., Zhu, Y.J., Yuan, Y.P., Xu, G.S. and Jiang, X., 2013. A
novel magnetic recyclable photocatalyst based on a core–shell metal–organic framework
Fe3O4@ MIL-100 (Fe) for the decolorization of methylene blue dye. Journal of
Materials Chemistry A, 1(45), pp.14329-14334. https://doi.org/10.1039/C3TA13030D

[214] Kim, J., Hwang, D.W., Kim, H.G., Bae, S.W., Lee, J.S., Li, W. and Oh, S.H., 2005.
Highly efficient overall water splitting through optimization of preparation and
operation conditions of layered perovskite photocatalysts. Topics in Catalysis, 35(3),
pp.295-303. https://doi.org/10.1007/s11244-005-3837-x

[215] Mizoguchi, H., Ueda, K., Orita, M., Moon, S.C., Kajihara, K., Hirano, M. and Hosono,
H., 2002. Decomposition of water by a CaTiO3 photocatalyst under UV light
166
irradiation. Materials Research Bulletin, 37(15), pp.2401-2406.
https://doi.org/10.1016/S0025-5408(02)00974-1

[216] Kondo, J., 1998. Cu 2 O as a photocatalyst for overall water splitting under visible light
irradiation. Chemical communications, (3), pp.357-358.
https://doi.org/10.1039/A707440I

[217] Dan-Hardi, M., Serre, C., Frot, T., Rozes, L., Maurin, G., Sanchez, C. and Férey, G.,
2009. A new photoactive crystalline highly porous titanium (IV) dicarboxylate. Journal
of the American Chemical Society, 131(31), pp.10857-10859.
https://doi.org/10.1021/ja903726m

[218] Horiuchi, Y., Toyao, T., Saito, M., Mochizuki, K., Iwata, M., Higashimura, H., Anpo,
M. and Matsuoka, M., 2012. Visible-light-promoted photocatalytic hydrogen
production by using an amino-functionalized Ti (IV) metal–organic framework. The
Journal of Physical Chemistry C, 116(39), pp.20848-20853.
https://doi.org/10.1021/jp3046005

[219] Joshi, U.A., Palasyuk, A.M. and Maggard, P.A., 2011. Photoelectrochemical
investigation and electronic structure of ap-type CuNbO3 photocathode. The Journal of
Physical Chemistry C, 115(27), pp.13534-13539. https://doi.org/10.1021/jp204631a

[220] Liu, N., Shang, Q., Gao, K., Cheng, Q. and Pan, Z., 2020. Construction of ZnO/ZIF-9
heterojunction photocatalyst: enhanced photocatalytic performance and mechanistic
insight. New Journal of Chemistry, 44(16), pp.6384-6393.
https://doi.org/10.1039/D0NJ00510J

[221] Jin, M., Qian, X., Gao, J., Chen, J., Hensley, D.K., Ho, H.C., Percoco, R.J., Ritzi, C.M.
and Yue, Y., 2019. Solvent-free synthesis of CuO/HKUST-1 composite and its
photocatalytic application. Inorganic Chemistry, 58(13), pp.8332-8338.
https://doi.org/10.1021/acs.inorgchem.9b00362

[222] Subudhi, S., Tripathy, S.P. and Parida, K., 2021. Metal oxide integrated metal organic
frameworks (MO@ MOF): rational design, fabrication strategy, characterization and
emerging photocatalytic applications. Inorganic Chemistry Frontiers, 8(6), pp.1619-
1636. https://doi.org/10.1039/D0QI01117G

[223] Zong, X. and Li, C., 2018. Photocatalytic water splitting on metal oxide-based
semiconductor photocatalysts. In Metal oxides in heterogeneous catalysis (pp. 355-
399). Elsevier. https://doi.org/10.1016/B978-0-12-811631-9.00007-7

[224] He, J., Yan, Z., Wang, J., Xie, J., Jiang, L., Shi, Y., Yuan, F., Yu, F. and Sun, Y., 2013.
Significantly enhanced photocatalytic hydrogen evolution under visible light over CdS
embedded on metal–organic frameworks. Chemical communications, 49(60), pp.6761-
6763. https://doi.org/10.1039/C3CC43218A

[225] Hussain, M.Z., Yang, Z., Huang, Z., Jia, Q., Zhu, Y. and Xia, Y., 2021. Recent advances
in metal–organic frameworks derived nanocomposites for photocatalytic applications in
energy and environment. Advanced Science, 8(14), p.2100625.
https://doi.org/10.1002/advs.202100625

167
[226] Xiao, J.D. and Jiang, H.L., 2017. Thermally Stable Metal‐Organic Framework‐
Templated Synthesis of Hierarchically Porous Metal Sulfides: Enhanced Photocatalytic
Hydrogen Production. Small, 13(28), p.1700632.
https://doi.org/10.1002/smll.201700632

[227] Su, Y., Ao, D., Liu, H. and Wang, Y., 2017. MOF-derived yolk–shell CdS microcubes
with enhanced visible-light photocatalytic activity and stability for hydrogen
evolution. Journal of Materials Chemistry A, 5(18), pp.8680-8689.
https://doi.org/10.1039/C7TA00855D

[228] Chen, W., Fang, J., Zhang, Y., Chen, G., Zhao, S., Zhang, C., Xu, R., Bao, J., Zhou, Y.
and Xiang, X., 2018. CdS nanosphere-decorated hollow polyhedral ZCO derived from
a metal–organic framework (MOF) for effective photocatalytic water
evolution. Nanoscale, 10(9), pp.4463-4474. https://doi.org/10.1039/C7NR08943K

[229] Tang, X., Zhao, J.H., Li, Y.H., Zhou, Z.J., Li, K., Liu, F.T. and Lan, Y.Q., 2017. Co-
Doped Zn 1− x Cd x S nanocrystals from metal–organic framework precursors: porous
microstructure and efficient photocatalytic hydrogen evolution. Dalton
Transactions, 46(32), pp.10553-10557. https://doi.org/10.1039/C7DT01970J

[230] Kumar, D.P., Park, H., Kim, E.H., Hong, S., Gopannagari, M., Reddy, D.A. and Kim,
T.K., 2018. Noble metal-free metal-organic framework-derived onion slice-type hollow
cobalt sulfide nanostructures: Enhanced activity of CdS for improving photocatalytic
hydrogen production. Applied Catalysis B: Environmental, 224, pp.230-238.
https://doi.org/10.1016/j.apcatb.2017.10.051

[231] Zhao, X., Yang, H., Jing, P., Shi, W., Yang, G. and Cheng, P., 2017. A Metal‐Organic
Framework Approach toward Highly Nitrogen‐Doped Graphitic Carbon as a Metal‐Free
Photocatalyst for Hydrogen Evolution. Small, 13(9), p.1603279.
https://doi.org/10.1002/smll.201603279

[232] Zhao, X., Feng, J., Liu, J., Lu, J., Shi, W., Yang, G., Wang, G., Feng, P. and Cheng, P.,
2018. Metal–Organic Framework‐Derived ZnO/ZnS Heteronanostructures for Efficient
Visible‐Light‐Driven Photocatalytic Hydrogen Production. Advanced Science, 5(4),
p.1700590. https://doi.org/10.1002/advs.201700590

[233] Chen, L., Wang, H.F., Li, C. and Xu, Q., 2020. Bimetallic metal–organic frameworks
and their derivatives. Chemical science, 11(21), pp.5369-5403.
https://doi.org/10.1039/D0SC01432J

[234] Abednatanzi, S., Derakhshandeh, P.G., Depauw, H., Coudert, F.X., Vrielinck, H., Van
Der Voort, P. and Leus, K., 2019. Mixed-metal metal–organic frameworks. Chemical
Society Reviews, 48(9), pp.2535-2565. https://doi.org/10.1039/C8CS00337H

[235] Jaryal, R., Kumar, R. and Khullar, S., 2022. Mixed metal-metal organic frameworks
(MM-MOFs) and their use as efficient photocatalysts for hydrogen evolution from water
splitting reactions. Coordination Chemistry Reviews, 464, p.214542.
https://doi.org/10.1016/j.ccr.2022.214542

168
[236] Dekrafft, K.E., Wang, C. and Lin, W., 2012. Metal‐organic framework templated
synthesis of Fe2O3/TiO2 nanocomposite for hydrogen production. Advanced
materials, 24(15), pp.2014-2018. https://doi.org/10.1002/adma.201200330

[237] Zhang, M., Huang, Y.L., Wang, J.W. and Lu, T.B., 2016. A facile method for the
synthesis of a porous cobalt oxide–carbon hybrid as a highly efficient water oxidation
catalyst. Journal of Materials Chemistry A, 4(5), pp.1819-1827.
https://doi.org/10.1039/C5TA07813J

[238] Chen, J., Abazari, R., Adegoke, K.A., Maxakato, N.W., Bello, O.S., Tahir, M., Tasleem,
S., Sanati, S., Kirillov, A.M. and Zhou, Y., 2022. Metal–organic frameworks and derived
materials as photocatalysts for water splitting and carbon dioxide
reduction. Coordination Chemistry Reviews, 469, p.214664.
https://doi.org/10.1016/j.ccr.2022.214664

[239] Guo, J., Wan, Y., Zhu, Y., Zhao, M. and Tang, Z., 2021. Advanced photocatalysts based
on metal nanoparticle/metal-organic framework composites. Nano Research, 14(7),
pp.2037-2052. https://doi.org/10.1007/s12274-020-3182-1

[240] Yang, S., Li, X., Zeng, G., Cheng, M., Huang, D., Liu, Y., Zhou, C., Xiong, W., Yang,
Y., Wang, W. and Zhang, G., 2021. Materials Institute Lavoisier (MIL) based materials
for photocatalytic applications. Coordination Chemistry Reviews, 438, p.213874.
https://doi.org/10.1016/j.ccr.2021.213874

[241] Hussain, I., Jalil, A.A., Hamid, M.Y.S. and Hassan, N.S., 2021. Recent advances in
catalytic systems in the prism of physicochemical properties to remediate toxic CO
pollutants: A state-of-the-art review. Chemosphere, 277, p.130285.
https://doi.org/10.1016/j.chemosphere.2021.130285

[242] Jiang, Q. and Zhao, M., 2005. SIZE DEPENDENT THERMODYNAMIC AND
KINETIC CHARACTERIZATIONS OF SEMICONDUCTORS. Focus on
Semiconductor Research, p.1. http://dx.doi.org/10.1016/j.surfrep.2008.07.001

[243] Pour, A.N., Housaindokht, M.R., Irani, M. and Shahri, S.M.K., 2014. Size-dependent
studies of Fischer–Tropsch synthesis on iron based catalyst: New kinetic
model. Fuel, 116, pp.787-793. https://doi.org/10.1016/j.fuel.2013.08.080

[244] Zhou, X., Xu, W., Liu, G., Panda, D. and Chen, P., 2010. Size-dependent catalytic
activity and dynamics of gold nanoparticles at the single-molecule level. Journal of the
American Chemical Society, 132(1), pp.138-146. https://doi.org/10.1021/ja904307n

[245] Hamilton, J.F. and Baetzold, R.C., 1979. Catalysis by Small Metal Clusters: Size-
dependent catalytic activity can be correlated with changes in physical
properties. Science, 205(4412), pp.1213-1220.
https://doi.org/10.1126/science.205.4412.1213

[246] Cao, S., Tao, F.F., Tang, Y., Li, Y. and Yu, J., 2016. Size-and shape-dependent catalytic
performances of oxidation and reduction reactions on nanocatalysts. Chemical Society
Reviews, 45(17), pp.4747-4765. https://doi.org/10.1039/C6CS00094K

169
[247] Dong, C., Lian, C., Hu, S., Deng, Z., Gong, J., Li, M., Liu, H., Xing, M. and Zhang, J.,
2018. Size-dependent activity and selectivity of carbon dioxide photocatalytic reduction
over platinum nanoparticles. Nature communications, 9(1), pp.1-11.
https://doi.org/10.1038/s41467-018-03666-2

[248] Pasternak, S. and Paz, Y., 2013. On the similarity and dissimilarity between
photocatalytic water splitting and photocatalytic degradation of
pollutants. ChemPhysChem, 14(10), pp.2059-2070.
https://doi.org/10.1002/cphc.201300247

[249] Ishaq, T., Yousaf, M., Bhatti, I.A., Batool, A., Asghar, M.A., Mohsin, M. and Ahmad,
M., 2021. A perspective on possible amendments in semiconductors for enhanced
photocatalytic hydrogen generation by water splitting. international journal of hydrogen
energy, 46(79), pp.39036-39057. https://doi.org/10.1016/j.ijhydene.2021.09.165

[250] Gan, X., Lei, D. and Wong, K.Y., 2018. Two-dimensional layered nanomaterials for
visible-light-driven photocatalytic water splitting. Materials today energy, 10, pp.352-
367. https://doi.org/10.1016/j.mtener.2018.10.015

[251] Wang, T. and Gong, J., 2015. Single‐crystal semiconductors with narrow band gaps for
solar water splitting. Angewandte Chemie International Edition, 54(37), pp.10718-
10732. https://doi.org/10.1002/anie.201503346

[252] Zuo, F., Wang, L. and Feng, P., 2014. Self-doped Ti3+@ TiO2 visible light
photocatalyst: Influence of synthetic parameters on the H2 production
activity. International journal of hydrogen energy, 39(2), pp.711-717.
https://doi.org/10.1016/j.ijhydene.2013.10.120

[253] Wu, N., Wang, J., Tafen, D.N., Wang, H., Zheng, J.G., Lewis, J.P., Liu, X., Leonard,
S.S. and Manivannan, A., 2010. Shape-enhanced photocatalytic activity of single-
crystalline anatase TiO2 (101) nanobelts. Journal of the American Chemical
Society, 132(19), pp.6679-6685. https://doi.org/10.1021/ja909456f

[254] Yu, J., Xiang, Q., Ran, J. and Mann, S., 2010. One-step hydrothermal fabrication and
photocatalytic activity of surface-fluorinated TiO2 hollow microspheres and tabular
anatase single micro-crystals with high-energy facets. CrystEngComm, 12(3), pp.872-
879. https://doi.org/10.1039/B914385H

[255] Feng, X., Wang, H., Bo, X. and Guo, L., 2019. Bimetal–organic framework-derived
porous rodlike Cobalt/Nickel nitride for All-pH value electrochemical hydrogen
evolution. ACS applied materials & interfaces, 11(8), pp.8018-8024.
https://doi.org/10.1021/acsami.8b21369

[256] Guan, J., Pal, T., Kamiya, K., Fukui, N., Maeda, H., Sato, T., Suzuki, H., Tomita, O.,
Nishihara, H., Abe, R. and Sakamoto, R., 2022. Two-Dimensional Metal–Organic
Framework Acts as a Hydrogen Evolution Cocatalyst for Overall Photocatalytic Water
Splitting. ACS Catalysis, 12(7), pp.3881-3889.
https://doi.org/10.1021/acscatal.1c05889

170
[257] Liu, L., Du, S., Guo, X., Xiao, Y., Yin, Z., Yang, N., Bao, Y., Zhu, X., Jin, S., Feng, Z.
and Zhang, F., 2022. Water-Stable Nickel Metal–Organic Framework Nanobelts for
Cocatalyst-Free Photocatalytic Water Splitting to Produce Hydrogen. Journal of the
American Chemical Society, 144(6), pp.2747-2754.
https://doi.org/10.1021/jacs.1c12179

[258] Yang, X., Singh, D. and Ahuja, R., 2020. Recent advancements and future prospects in
ultrathin 2D semiconductor-based photocatalysts for water splitting. Catalysts, 10(10),
p.1111. https://doi.org/10.3390/catal10101111

[259] Tan, J.B. and Li, G.R., 2020. Recent progress on metal–organic frameworks and their
derived materials for electrocatalytic water splitting. Journal of Materials Chemistry
A, 8(29), pp.14326-14355. https://doi.org/10.1039/D0TA04016A

[260] Jiao, L., Wang, Y., Jiang, H.L. and Xu, Q., 2018. Metal–organic frameworks as
platforms for catalytic applications. Advanced Materials, 30(37), p.1703663.
https://doi.org/10.1002/adma.201703663

[261] Zhu, B., Zou, R. and Xu, Q., 2018. Metal–organic framework based catalysts for
hydrogen evolution. Advanced Energy Materials, 8(24), p.1801193.
https://doi.org/10.1002/aenm.201801193

[262] Pullen, S. and Ott, S., 2016. Photochemical Hydrogen Production with Metal–Organic
Frameworks. Topics in Catalysis, 59(19), pp.1712-1721.
https://doi.org/10.1007/s11244-016-0690-z

[263] Wen, M., Mori, K., Kuwahara, Y., An, T. and Yamashita, H., 2017. Design and
architecture of metal organic frameworks for visible light enhanced hydrogen
production. Applied Catalysis B: Environmental, 218, pp.555-
569.https://doi.org/10.1016/j.apcatb.2017.06.082

[264] Zhai, X. and Chen, F., 2019. Path planning of a type of porous structures for additive
manufacturing. Computer-Aided Design, 115, pp.218-230.
https://doi.org/10.1016/j.cad.2019.06.002

[265] Liu, C. and Liu, G., 2021. Characterization of pore structure parameters of foam concrete
by 3D reconstruction and image analysis. Construction and Building Materials, 267,
p.120958. https://doi.org/10.1016/j.conbuildmat.2020.120958

[266] Salcedo‐Abraira, P., Babaryk, A.A., Montero‐Lanzuela, E., Contreras‐Almengor, O.R.,


Cabrero‐Antonino, M., Grape, E.S., Willhammar, T., Navalón, S., Elkäim, E., García,
H. and Horcajada, P., 2021. A Novel Porous Ti‐Squarate as Efficient Photocatalyst in
the Overall Water Splitting Reaction under Simulated Sunlight Irradiation. Advanced
Materials, 33(52), p.2106627. https://doi.org/10.1002/adma.202106627

171
[267] Lan, M., Guo, R.M., Dou, Y., Zhou, J., Zhou, A. and Li, J.R., 2017. Fabrication of
porous Pt-doping heterojunctions by using bimetallic MOF template for photocatalytic
hydrogen generation. Nano Energy, 33, pp.238-246.
https://doi.org/10.1016/j.nanoen.2017.01.046
[268] Sharma, K., Hasija, V., Patial, S., Singh, P., Nguyen, V.H., Nadda, A.K., Thakur, S.,
Nguyen-Tri, P., Nguyen, C.C., Kim, S.Y. and Van Le, Q., 2022. Recent progress on
MXenes and MOFs hybrids: structure, synthetic strategies and catalytic water
splitting. International Journal of Hydrogen Energy.
https://doi.org/10.1016/j.ijhydene.2022.01.004

[269] Xiao, M., Wang, Z., Lyu, M., Luo, B., Wang, S., Liu, G., Cheng, H.M. and Wang, L.,
2019. Hollow nanostructures for photocatalysis: advantages and challenges. Advanced
Materials, 31(38), p.1801369. https://doi.org/10.1002/adma.201801369

[270] Zhu, Q.L. and Xu, Q., 2016. Immobilization of ultrafine metal nanoparticles to high-
surface-area materials and their catalytic applications. Chem, 1(2), pp.220-245.
https://doi.org/10.1016/j.chempr.2016.07.005

[271] Wang, Y., Zhao, S., Zhu, Y., Qiu, R., Gengenbach, T., Liu, Y., Zu, L., Mao, H., Wang,
H., Tang, J. and Zhao, D., 2020. Three-dimensional hierarchical porous nanotubes
derived from metal-organic frameworks for highly efficient overall water
splitting. Iscience, 23(1), p.100761. https://doi.org/10.1016/j.isci.2019.100761

[272] Zhang, Y., Liu, J., Li, S.L., Su, Z.M. and Lan, Y.Q., 2019. Polyoxometalate-based
materials for sustainable and clean energy conversion and storage. EnergyChem, 1(3),
p.100021. https://doi.org/10.1016/j.enchem.2019.100021

[273] Jin, Z., Wang, Z., Yuan, H. and Han, F., 2019. Inserting MOF into flaky CdS
photocatalyst forming special structure and active sites for efficient hydrogen
production. International Journal of Hydrogen Energy, 44(36), pp.19640-19649.
https://doi.org/10.1016/j.ijhydene.2019.05.233

[274] Xu, G.R., An, Z.H., Xu, K., Liu, Q., Das, R. and Zhao, H.L., 2021. Metal organic
framework (MOF)-based micro/nanoscaled materials for heavy metal ions removal: The
cutting-edge study on designs, synthesis, and applications. Coordination Chemistry
Reviews, 427, p.213554. https://doi.org/10.1016/j.ccr.2020.213554

[275] Li, T. and Jin, Z., 2022. Unique ternary Ni-MOF-74/Ni2P/MoSx composite for efficient
photocatalytic hydrogen production: Role of Ni2P for accelerating separation of
photogenerated carriers. Journal of Colloid and Interface Science, 605, pp.385-397.
https://doi.org/10.1016/j.jcis.2021.07.098

[276] Li, T., Guo, X., Zhang, L., Yan, T. and Jin, Z., 2021. 2D CoP supported 0D WO3
constructed S-scheme for efficient photocatalytic hydrogen evolution. International
Journal of Hydrogen Energy, 46(39), pp.20560-20572.
https://doi.org/10.1016/j.ijhydene.2021.03.169

172
[277] Dhakshinamoorthy, A., Li, Z. and Garcia, H., 2018. Catalysis and photocatalysis by
metal organic frameworks. Chemical Society Reviews, 47(22), pp.8134-8172.
https://doi.org/10.1039/C8CS00256H

[278] Zhang, F.M., Sheng, J.L., Yang, Z.D., Sun, X.J., Tang, H.L., Lu, M., Dong, H., Shen,
F.C., Liu, J. and Lan, Y.Q., 2018. Rational design of MOF/COF hybrid materials for
photocatalytic H2 evolution in the presence of sacrificial electron donors. Angewandte
Chemie International Edition, 57(37), pp.12106-12110.
https://doi.org/10.1002/anie.201806862

[279] Yang, J., Wang, D., Han, H. and Li, C.A.N., 2013. Roles of cocatalysts in photocatalysis
and photoelectrocatalysis. Accounts of chemical research, 46(8), pp.1900-1909.
https://doi.org/10.1021/ar300227e

[280] Trasatti, S., 1972. Work function, electronegativity, and electrochemical behaviour of
metals: III. Electrolytic hydrogen evolution in acid solutions. Journal of
Electroanalytical Chemistry and Interfacial Electrochemistry, 39(1), pp.163-184.
https://doi.org/10.1016/S0022-0728(72)80485-6

[281] Hejazi, S., Mohajernia, S., Osuagwu, B., Zoppellaro, G., Andryskova, P., Tomanec, O.,
Kment, S., Zbořil, R. and Schmuki, P., 2020. On the controlled loading of single
platinum atoms as a Co‐catalyst on TiO2 anatase for optimized photocatalytic H2
generation. Advanced Materials, 32(16), p.1908505.
https://doi.org/10.1002/adma.201908505

[282] Li, F., Huang, H., Li, G. and Leung, D.Y., 2019. TiO2 nanotube arrays modified with
nanoparticles of platinum group metals (Pt, Pd, Ru): enhancement on
photoelectrochemical performance. Journal of Nanoparticle Research, 21(2), pp.1-13.
https://doi.org/10.1007/s11051-018-4443-8

[283] Jiang, Z., Zhang, Z., Shangguan, W., Isaacs, M.A., Durndell, L.J., Parlett, C.M. and Lee,
A.F., 2016. Photodeposition as a facile route to tunable Pt photocatalysts for hydrogen
production: on the role of methanol. Catalysis Science & Technology, 6(1), pp.81-88.
https://doi.org/10.1039/C5CY01364J

[284] Feng, R., Wan, K., Sui, X., Zhao, N., Li, H., Lei, W., Yu, J., Liu, X., Shi, X., Zhai, M.
and Liu, G., 2021. Anchoring single Pt atoms and black phosphorene dual co-catalysts
on CdS nanospheres to boost visible-light photocatalytic H2 evolution. Nano Today, 37,
p.101080. https://doi.org/10.1016/j.nantod.2021.101080

[285] Xiao, N., Li, S., Li, X., Ge, L., Gao, Y. and Li, N., 2020. The roles and mechanism of
cocatalysts in photocatalytic water splitting to produce hydrogen. Chinese Journal of
Catalysis, 41(4), pp.642-671. https://doi.org/10.1016/S1872-2067(19)63469-8

[286] Yu, G., Qian, J., Zhang, P., Zhang, B., Zhang, W., Yan, W. and Liu, G., 2019. Collective
excitation of plasmon-coupled Au-nanochain boosts photocatalytic hydrogen evolution
of semiconductor. Nature communications, 10(1), pp.1-8.
https://doi.org/10.1038/s41467-019-12853-8
173
[287] Zong, X., Wu, G., Yan, H., Ma, G., Shi, J., Wen, F., Wang, L. and Li, C., 2010.
Photocatalytic H2 evolution on MoS2/CdS catalysts under visible light irradiation. The
Journal of Physical Chemistry C, 114(4), pp.1963-1968.
https://doi.org/10.1021/jp904350e

[288] Xiao, J.D., Han, L., Luo, J., Yu, S.H. and Jiang, H.L., 2018. Integration of Plasmonic
Effects and Schottky Junctions into Metal–Organic Framework Composites: Steering
Charge Flow for Enhanced Visible‐Light Photocatalysis. Angewandte Chemie
International Edition, 57(4), pp.1103-1107. https://doi.org/10.1002/anie.201711725

[289] Khan, M.R., Chuan, T.W., Yousuf, A., Chowdhury, M.N.K. and Cheng, C.K., 2015.
Schottky barrier and surface plasmonic resonance phenomena towards the
photocatalytic reaction: study of their mechanisms to enhance photocatalytic
activity. Catalysis Science & Technology, 5(5), pp.2522-2531.
https://doi.org/10.1039/C4CY01545B

[290] Sanzone, G., Zimbone, M., Cacciato, G., Ruffino, F., Carles, R., Privitera, V. and
Grimaldi, M.G., 2018. Ag/TiO2 nanocomposite for visible light-driven
photocatalysis. Superlattices and Microstructures, 123, pp.394-402.
https://doi.org/10.1016/j.spmi.2018.09.028

[291] Wu, Y., Wang, P., Zhu, X., Zhang, Q., Wang, Z., Liu, Y., Zou, G., Dai, Y., Whangbo,
M.H. and Huang, B., 2018. Composite of CH3NH3PbI3 with reduced graphene oxide as
a highly efficient and stable visible‐light photocatalyst for hydrogen evolution in
aqueous HI solution. Advanced Materials, 30(7), p.1704342.
https://doi.org/10.1002/adma.201704342

[292] Wang, X., Wang, H., Zhang, H., Yu, W., Wang, X., Zhao, Y., Zong, X. and Li, C.,
2018. Dynamic interaction between methylammonium lead iodide and TiO2
nanocrystals leads to enhanced photocatalytic H2 evolution from HI splitting. ACS
Energy Letters, 3(5), pp.1159-1164. https://doi.org/10.1021/acsenergylett.8b00488

[293] Batool, S., Nandan, S.P., Myakala, S.N., Rajagopal, A., Schubert, J.S., Ayala, P.,
Naghdi, S., Saito, H., Bernardi, J., Streb, C. and Cherevan, A., 2022. Surface Anchoring
and Active Sites of [Mo3S13]2–Clusters as Co-Catalysts for Photocatalytic Hydrogen
Evolution. ACS Catalysis, 12, pp.6641-6650. https://doi.org/10.1021/acscatal.2c00972

[294] Li, R., Zhao, Y. and Li, C., 2017. Spatial distribution of active sites on a ferroelectric
PbTiO3 photocatalyst for photocatalytic hydrogen production. Faraday
Discussions, 198, pp.463-472. https://doi.org/10.1039/C6FD00199H

[295] He, K., Xie, J., Yang, Z., Shen, R., Fang, Y., Ma, S., Chen, X. and Li, X., 2017. Earth-
abundant WC nanoparticles as an active noble-metal-free co-catalyst for the highly
boosted photocatalytic H2 production over gC3N4 nanosheets under visible
light. Catalysis Science & Technology, 7(5), pp.1193-1202.
https://doi.org/10.1039/C7CY00029D

[296] Wang, H., Zhang, L., Chen, Z., Hu, J., Li, S., Wang, Z., Liu, J. and Wang, X., 2014.
Semiconductor heterojunction photocatalysts: design, construction, and photocatalytic
performances. Chemical Society Reviews, 43(15), pp.5234-5244.
https://doi.org/10.1039/C4CS00126E

174
[297] Li, H., Zhou, Y., Tu, W., Ye, J. and Zou, Z., 2015. State‐of‐the‐art progress in diverse
heterostructured photocatalysts toward promoting photocatalytic
performance. Advanced Functional Materials, 25(7), pp.998-1013.
https://doi.org/10.1002/adfm.201401636

[298] Zhu, Y., Wang, T., Xu, T., Li, Y. and Wang, C., 2019. Size effect of Pt co-catalyst on
photocatalytic efficiency of g-C3N4 for hydrogen evolution. Applied Surface
Science, 464, pp.36-42. https://doi.org/10.1016/j.apsusc.2018.09.061

[299] Qu, R., Zhang, W., Liu, N., Zhang, Q., Liu, Y., Li, X., Wei, Y. and Feng, L., 2018.
Antioil Ag3PO4 nanoparticle/polydopamine/Al2O3 sandwich structure for complex
wastewater treatment: dynamic catalysis under natural light. ACS Sustainable
Chemistry & Engineering, 6(6), pp.8019-8028.
https://doi.org/10.1021/acssuschemeng.8b01469

[300] Al-Azri, Z.H., Chen, W.T., Chan, A., Jovic, V., Ina, T., Idriss, H. and Waterhouse, G.I.,
2015. The roles of metal co-catalysts and reaction media in photocatalytic hydrogen
production: Performance evaluation of M/TiO2 photocatalysts (M= Pd, Pt, Au) in
different alcohol–water mixtures. Journal of catalysis, 329, pp.355-367.
https://doi.org/10.1016/j.jcat.2015.06.005

[301] Mo, Z., Xu, H., She, X., Song, Y., Yan, P., Yi, J., Zhu, X., Lei, Y., Yuan, S. and Li, H.,
2019. Constructing Pd/2D-C3N4 composites for efficient photocatalytic H2 evolution
through nonplasmon-induced bound electrons. Applied Surface Science, 467, pp.151-
157. https://doi.org/10.1016/j.apsusc.2018.10.115

[302] Chen, T., Quan, W., Yu, L., Hong, Y., Song, C., Fan, M., Xiao, L., Gu, W. and Shi, W.,
2016. One-step synthesis and visible-light-driven H2 production from water splitting of
Ag quantum dots/g-C3N4 photocatalysts. Journal of Alloys and Compounds, 686,
pp.628-634. https://doi.org/10.1016/j.jallcom.2016.06.076

[303] Ran, J., Zhang, J., Yu, J., Jaroniec, M. and Qiao, S.Z., 2014. Earth-abundant cocatalysts
for semiconductor-based photocatalytic water splitting. Chemical Society
Reviews, 43(22), pp.7787-7812. https://doi.org/10.1039/C3CS60425J

[304] Li, Q., Li, X., Wageh, S., Al‐Ghamdi, A.A. and Yu, J., 2015. CdS/graphene
nanocomposite photocatalysts. Advanced Energy Materials, 5(14), p.1500010.
https://doi.org/10.1002/aenm.201500010

[305] Yi, S.S., Zhang, X.B., Wulan, B.R., Yan, J.M. and Jiang, Q., 2018. Non-noble metals
applied to solar water splitting. Energy & Environmental Science, 11(11), pp.3128-
3156. https://doi.org/10.1039/C8EE02096E

[306] Zou, X. and Zhang, Y., 2015. Noble metal-free hydrogen evolution catalysts for water
splitting. Chemical Society Reviews, 44(15), pp.5148-5180.
https://doi.org/10.1039/C4CS00448E

[307] Irfan, R.M., Tahir, M.H., Maqsood, M., Lin, Y., Bashir, T., Iqbal, S., Zhao, J., Gao, L.
and Haroon, M., 2020. CoSe as non-noble-metal cocatalyst integrated with
heterojunction photosensitizer for inexpensive H2 production under visible
light. Journal of Catalysis, 390, pp.196-205. https://doi.org/10.1016/j.jcat.2020.07.034

175
[308] Lin, H., Ma, X., Wang, H., Wang, B., Li, S.X., Yi, X., Li, Y.Y. and Wang, L., 2021.
Promoted Interfacial Charge Transport and Separation of Size-Uniform Zn, Ni-Doped
CdS-1T/2H O-MoS2 Nanoassemblies for Efficient Visible-Light Photocatalytic Water
Splitting. Crystal Growth & Design, 21(2), pp.1278-1289.
https://doi.org/10.1021/acs.cgd.0c01546

[309] Eder, M., Courtois, C., Petzoldt, P., Mackewicz, S., Tschurl, M. and Heiz, U., 2022. Size
and coverage effects of Ni and Pt co-catalysts in the photocatalytic hydrogen evolution
from methanol on TiO2 (110). ACS Catalysis, 12(15), pp.9579-9588.
https://doi.org/10.1021/acscatal.2c02230

[310] Zhang, Q., Xiao, Y., Li, Y., Zhao, K., Deng, H., Lou, Y., Chen, J. and Cheng, L., 2020.
NiS‐decorated ZnO/ZnS nanorod heterostructures for enhanced photocatalytic
hydrogen production: insight into the role of NiS. Solar RRL, 4(4), p.1900568.
https://doi.org/10.1002/solr.201900568

[311] Zhang, W., Wang, Y. and Xu, R., 2020. Development of metal sulfide–based
photocatalysts for hydrogen evolution under visible light. In Current Developments in
Photocatalysis and Photocatalytic Materials (pp. 369-384). Elsevier.
https://doi.org/10.1016/B978-0-12-819000-5.00023-0

[312] Peng, S., Yang, Y., Tan, J., Gan, C. and Li, Y., 2018. In situ loading of Ni2P on Cd0.
5Zn0. 5S with red phosphorus for enhanced visible light photocatalytic H2
evolution. Applied Surface Science, 447, pp.822-828.
https://doi.org/10.1016/j.apsusc.2018.04.050

[313] Xu, H., Li, S., Ge, L., Han, C., Gao, Y. and Dai, D., 2017. In-situ synthesis of novel
plate-like Co (OH)2 co-catalyst decorated TiO2 nanosheets with efficient photocatalytic
H2 evolution activity. International Journal of Hydrogen Energy, 42(36), pp.22877-
22886. https://doi.org/10.1016/j.ijhydene.2017.07.186

[314] Bagal, I.V., Chodankar, N.R., Hassan, M.A., Waseem, A., Johar, M.A., Kim, D.H. and
Ryu, S.W., 2019. Cu2O as an emerging photocathode for solar water splitting-A status
review. International Journal of Hydrogen Energy, 44(39), pp.21351-21378.
https://doi.org/10.1016/j.ijhydene.2019.06.184

[315] Wang, Z., Huang, J., Mao, J., Guo, Q., Chen, Z. and Lai, Y., 2020. Metal–organic
frameworks and their derivatives with graphene composites: Preparation and
applications in electrocatalysis and photocatalysis. Journal of Materials Chemistry
A, 8(6), pp.2934-2961. https://doi.org/10.1039/C9TA12776C

[316] Li, B., Wen, H.M., Cui, Y., Zhou, W., Qian, G. and Chen, B., 2016. Emerging
multifunctional metal–organic framework materials. Advanced Materials, 28(40),
pp.8819-8860. https://doi.org/10.1002/adma.201601133

[317] Chen, Y.Z., Zhang, R., Jiao, L. and Jiang, H.L., 2018. Metal–organic framework-
derived porous materials for catalysis. Coordination Chemistry Reviews, 362, pp.1-23.
https://doi.org/10.1016/j.ccr.2018.02.008

[318] Xu, H.Q., Yang, S., Ma, X., Huang, J. and Jiang, H.L., 2018. Unveiling charge-
separation dynamics in CdS/metal–organic framework composites for enhanced

176
photocatalysis. Acs Catalysis, 8(12), pp.11615-11621.
https://doi.org/10.1021/acscatal.8b03233

[319] Wang, C.C., Yi, X.H. and Wang, P., 2019. Powerful combination of MOFs and C 3N4
for enhanced photocatalytic performance. Applied Catalysis B: Environmental, 247,
pp.24-48. https://doi.org/10.1016/j.apcatb.2019.01.091

[320] Domen, K., Naito, S., Onishi, T., Tamaru, K. and Soma, M., 1982. Study of the
photocatalytic decomposition of water vapor over a nickel (II) oxide-strontium titanate
(SrTiO3) catalyst. The Journal of Physical Chemistry, 86(18), pp.3657-3661.

[321] Ma, B., Guan, P.Y., Li, Q.Y., Zhang, M. and Zang, S.Q., 2016. MOF-derived flower-
like MoS2@ TiO2 nanohybrids with enhanced activity for hydrogen evolution. ACS
applied materials & interfaces, 8(40), pp.26794-26800.
https://doi.org/10.1021/acsami.6b08740

[322] Zhang, Y., Jin, Z. and Wang, G., 2018. Charge transfer behaviors over MOF-5@ g-C3N4
with NixMo1− xS2 modification. International Journal of Hydrogen Energy, 43(21),
pp.9914-9923. https://doi.org/10.1016/j.ijhydene.2018.04.071

[323] Dhiman, P., Rana, G., Kumar, A., Sharma, G., Vo, D.V.N. and Naushad, M., 2022. ZnO-
based heterostructures as photocatalysts for hydrogen generation and depollution: a
review. Environmental Chemistry Letters, pp.1-35. https://doi.org/10.1007/s10311-021-
01361-1

[324] Reddy, C.V., Reddy, K.R., Harish, V.A., Shim, J., Shankar, M.V., Shetti, N.P. and
Aminabhavi, T.M., 2020. Metal-organic frameworks (MOFs)-based efficient
heterogeneous photocatalysts: synthesis, properties and its applications in photocatalytic
hydrogen generation, CO2 reduction and photodegradation of organic
dyes. International Journal of Hydrogen Energy, 45(13), pp.7656-7679.
https://doi.org/10.1016/j.ijhydene.2019.02.144

[325] Chen, S., Qi, Y., Li, C., Domen, K. and Zhang, F., 2018. Surface strategies for
particulate photocatalysts toward artificial photosynthesis. Joule, 2(11), pp.2260-2288.
https://doi.org/10.1016/j.joule.2018.07.030

[326] Niu, F., Wang, D., Li, F., Liu, Y., Shen, S. and Meyer, T.J., 2020. Hybrid
photoelectrochemical water splitting systems: from interface design to system
assembly. Advanced Energy Materials, 10(11), p.1900399.
https://doi.org/10.1002/aenm.201900399

[327] Wang, Y., Wang, Q., Zhan, X., Wang, F., Safdar, M. and He, J., 2013. Visible light
driven type II heterostructures and their enhanced photocatalysis properties: a
review. Nanoscale, 5(18), pp.8326-8339. https://doi.org/10.1039/C3NR01577G

177
[328] Zhong, Y., Peng, C., He, Z., Chen, D., Jia, H., Zhang, J., Ding, H. and Wu, X., 2021.
Interface engineering of heterojunction photocatalysts based on 1D
nanomaterials. Catalysis Science & Technology, 11(1), pp.27-42.
https://doi.org/10.1039/D0CY01847C

[329] Wang, Y., Wu, Q., Li, Y., Liu, L., Geng, Z., Li, Y., Chen, J., Bai, W., Jiang, G. and
Zhao, Z., 2018. Controlled fabrication of TiO2/C3N4 core–shell nanowire arrays: a
visible-light-responsive and environmental-friendly electrode for photoelectrocatalytic
degradation of bisphenol A. Journal of Materials Science, 53(15), pp.11015-11026.
https://doi.org/10.1007/s10853-018-2368-3

[330] Zhang, J., Wang, Y., Jin, J., Zhang, J., Lin, Z., Huang, F. and Yu, J., 2013. Efficient
visible-light photocatalytic hydrogen evolution and enhanced photostability of
core/shell CdS/g-C3N4 nanowires. ACS applied materials & interfaces, 5(20),
pp.10317-10324. https://doi.org/10.1021/am403327g

[331] Butanovs, E., Vlassov, S., Kuzmin, A., Piskunov, S., Butikova, J. and Polyakov, B.,
2018. Fast-response single-nanowire photodetector based on ZnO/WS2 core/shell
heterostructures. ACS applied materials & interfaces, 10(16), pp.13869-13876.
https://doi.org/10.1021/acsami.8b02241

[332] Tripathy, S.P., Subudhi, S., Das, S., Ghosh, M.K., Das, M., Acharya, R., Acharya, R.
and Parida, K., 2022. Hydrolytically stable citrate capped Fe3O4@ UiO-66-NH2 MOF:
A hetero-structure composite with enhanced activity towards Cr (VI) adsorption and
photocatalytic H2 evolution. Journal of Colloid and Interface Science, 606, pp.353-
366.https://doi.org/10.1016/j.jcis.2021.08.031

[333] Meng, X.B., Sheng, J.L., Tang, H.L., Sun, X.J., Dong, H. and Zhang, F.M., 2019.
Metal-organic framework as nanoreactors to co-incorporate carbon nanodots and CdS
quantum dots into the pores for improved H2 evolution without noble-metal
cocatalyst. Applied Catalysis B: Environmental, 244, pp.340-346.
https://doi.org/10.1016/j.apcatb.2018.11.018
[334] Zhao, C., Zhang, Y., Jiang, H., Chen, J., Liu, Y., Liang, Q., Zhou, M., Li, Z. and Zhou,
Y., 2019. Combined effects of octahedron NH2-UiO-66 and flowerlike ZnIn2S4
microspheres for photocatalytic dye degradation and hydrogen evolution under visible
light. The Journal of Physical Chemistry C, 123(29), pp.18037-18049.
https://doi.org/10.1021/acs.jpcc.9b03807

[335] Chen, Y., Yang, D., Shi, B., Dai, W., Ren, H., An, K., Zhou, Z., Zhao, Z., Wang, W.
and Jiang, Z., 2020. In situ construction of hydrazone-linked COF-based core–shell
hetero-frameworks for enhanced photocatalytic hydrogen evolution. Journal of
materials chemistry A, 8(16), pp.7724-7732. https://doi.org/10.1039/D0TA00901F

[336] Chen, W., Liu, T.Y., Huang, T., Liu, X.H., Duan, G.R., Yang, X.J. and Chen, S.M.,
2015. A novel yet simple strategy to fabricate visible light responsive C, N-TiO 2/gC 3
N4 heterostructures with significantly enhanced photocatalytic hydrogen
generation. Rsc Advances, 5(122), pp.101214-101220.
https://doi.org/10.1039/C5RA18302B

[337] Sohail, M., Kim, H. and Kim, T.W., 2019. Enhanced photocatalytic performance of a
Ti-based metal-organic framework for hydrogen production: Hybridization with ZnCr-

178
LDH nanosheets. Scientific reports, 9(1), pp.1-11. https://doi.org/10.1038/s41598-019-
44008-6

[338] Lin, X.Y., Li, Y.H., Qi, M.Y., Tang, Z.R., Jiang, H.L. and Xu, Y.J., 2020. A unique
coordination-driven route for the precise nanoassembly of metal sulfides on metal–
organic frameworks. Nanoscale Horizons, 5(4), pp.714-719.
https://doi.org/10.1039/C9NH00769E

[339] Mehtab, A., Banerjee, S., Mao, Y. and Ahmad, T., 2022. Type-II CuFe2O4/graphitic
carbon nitride heterojunctions for high-efficiency photocatalytic and electrocatalytic
hydrogen generation. ACS applied materials & interfaces, 14(39), pp.44317-44329.
https://doi.org/10.1021/acsami.2c11140

[340] Ali, S.A., Sadiq, I. and Ahmad, T., 2023. Oxide based heterostructured photocatalysts
for CO2 reduction and hydrogen generation. ChemistrySelect, 8(8), p.e202203176.
https://doi.org/10.1002/slct.202203176

[341] Zhao, M., Huang, Y., Peng, Y., Huang, Z., Ma, Q. and Zhang, H., 2018. Two-
dimensional metal–organic framework nanosheets: synthesis and
applications. Chemical Society Reviews, 47(16), pp.6267-6295.
https://doi.org/10.1039/C8CS00268A

[342] Su, Q., Li, Y., Hu, R., Song, F., Liu, S., Guo, C., Zhu, S., Liu, W. and Pan, J., 2020.
Heterojunction photocatalysts based on 2D materials: the role of
configuration. Advanced Sustainable Systems, 4(9), p.2000130.
https://doi.org/10.1002/adsu.202000130

[343] Bard, A.J., 1979. Photo electrochemistry and heterogeneous photo-catalysis at


semiconductors. Journal of Photochemistry, 10(1), pp.59-75.
https://doi.org/10.1016/0047-2670(79)80037-4

[344] Tahir, M., Tasleem, S. and Tahir, B., 2020. Recent development in band engineering
of binary semiconductor materials for solar driven photocatalytic hydrogen
production. International Journal of Hydrogen Energy, 45(32), pp.15985-16038.
https://doi.org/10.1016/j.ijhydene.2020.04.071

[345] Abe, R., Sayama, K. and Sugihara, H., 2005. Development of new photocatalytic water
splitting into H2 and O2 using two different semiconductor photocatalysts and a shuttle
redox mediator IO3-/I-. The journal of physical chemistry B, 109(33), pp.16052-16061.
https://doi.org/10.1021/jp052848l

[346] Li, H., Tu, W., Zhou, Y. and Zou, Z., 2016. Z‐Scheme photocatalytic systems for
promoting photocatalytic performance: recent progress and future
challenges. Advanced science, 3(11), p.1500389.
https://doi.org/10.1002/advs.201500389

[347] Chandran, R.B., Breen, S., Shao, Y., Ardo, S. and Weber, A.Z., 2018. Evaluating
particle-suspension reactor designs for Z-scheme solar water splitting via transport and
kinetic modeling. Energy & Environmental Science, 11(1), pp.115-135.
https://doi.org/10.1039/C7EE01360D
179
[348] Maeda, K., Higashi, M., Lu, D., Abe, R. and Domen, K., 2010. Efficient nonsacrificial
water splitting through two-step photoexcitation by visible light using a modified
oxynitride as a hydrogen evolution photocatalyst. Journal of the American Chemical
Society, 132(16), pp.5858-5868. https://doi.org/10.1021/ja1009025

[349] Zhou, G., Wu, M.F., Xing, Q.J., Li, F., Liu, H., Luo, X.B., Zou, J.P., Luo, J.M. and
Zhang, A.Q., 2018. Synthesis and characterizations of metal-free
Semiconductor/MOFs with good stability and high photocatalytic activity for H2
evolution: A novel Z-Scheme heterostructured photocatalyst formed by covalent
bonds. Applied Catalysis B: Environmental, 220, pp.607-614.
https://doi.org/10.1016/j.apcatb.2017.08.086

[350] Liu, M., Qiao, L.Z., Dong, B.B., Guo, S., Yao, S., Li, C., Zhang, Z.M. and Lu, T.B.,
2020. Photocatalytic coproduction of H2 and industrial chemical over MOF-derived
direct Z-scheme heterostructure. Applied Catalysis B: Environmental, 273, p.119066.
https://doi.org/10.1016/j.apcatb.2020.119066

[351] Wu, Q., Zhang, C., Sun, K. and Jiang, H.L., 2020. Microwave-assisted synthesis and
photocatalytic performance of a soluble porphyrinic MOF. Acta Chimica Sinica, 78(7),
p.688. DOI: 10.6023/A20050141

[352] Cai, X., Xie, Z., Li, D., Kassymova, M., Zang, S.Q. and Jiang, H.L., 2020. Nano-sized
metal-organic frameworks: Synthesis and applications. Coordination Chemistry
Reviews, 417, p.213366. https://doi.org/10.1016/j.ccr.2020.213366

[353] Guillerm, V. and Eddaoudi, M., 2021. The importance of highly connected building
units in reticular chemistry: Thoughtful design of metal–organic frameworks. Accounts
of Chemical Research, 54(17), pp.3298-3312.
https://doi.org/10.1021/acs.accounts.1c00214

[354] Lu, W., Wei, Z., Gu, Z.Y., Liu, T.F., Park, J., Park, J., Tian, J., Zhang, M., Zhang, Q.,
Gentle III, T. and Bosch, M., 2014. Tuning the structure and function of metal–organic
frameworks via linker design. Chemical Society Reviews, 43(16), pp.5561-5593.
https://doi.org/10.1039/C4CS00003J

[355] Furukawa, H., Müller, U. and Yaghi, O.M., 2015. “Heterogeneity within order” in
metal–organic frameworks. Angewandte Chemie International Edition, 54(11),
pp.3417-3430. https://doi.org/10.1002/anie.201410252

[356] Qin, L. and Zheng, H.G., 2017. Structures and applications of metal–organic
frameworks featuring metal clusters. CrystEngComm, 19(5), pp.745-757.
https://doi.org/10.1039/C6CE02293F

[357] Yang, H., Wang, J., Ma, J., Yang, H., Zhang, J., Lv, K., Wen, L. and Peng, T., 2019. A
novel BODIPY-based MOF photocatalyst for efficient visible-light-driven hydrogen
evolution. Journal of Materials Chemistry A, 7(17), pp.10439-10445.
https://doi.org/10.1039/C9TA02357G

[358] Seo, Y.K., Yoon, J.W., Lee, J.S., Hwang, Y.K., Jun, C.H., Chang, J.S., Wuttke, S.,
Bazin, P., Vimont, A., Daturi, M. and Bourrelly, S., 2012. Energy‐efficient
180
dehumidification over hierachically porous metal–organic frameworks as advanced
water adsorbents. Advanced Materials, 24(6), pp.806-810.
https://doi.org/10.1002/adma.201104084

[359] Wen, M., Mori, K., Kamegawa, T. and Yamashita, H., 2014. Amine-functionalized
MIL-101 (Cr) with imbedded platinum nanoparticles as a durable photocatalyst for
hydrogen production from water. Chemical Communications, 50(79), pp.11645-11648.
https://doi.org/10.1039/C4CC02994A

[360] Wang, X., Zhang, X., Zhou, W., Liu, L., Ye, J. and Wang, D., 2019. An ultrathin
porphyrin-based metal-organic framework for efficient photocatalytic hydrogen
evolution under visible light. Nano Energy, 62, pp.250-258.
https://doi.org/10.1016/j.nanoen.2019.05.023

[361] Huang, C.W., Nguyen, V.H., Zhou, S.R., Hsu, S.Y., Tan, J.X. and Wu, K.C.W., 2020.
Metal–organic frameworks: preparation and applications in highly efficient
heterogeneous photocatalysis. Sustainable Energy & Fuels, 4(2), pp.504-521.
https://doi.org/10.1039/C9SE00972H

[362] He, T., Chen, S., Ni, B., Gong, Y., Wu, Z., Song, L., Gu, L., Hu, W. and Wang, X.,
2018. Zirconium–porphyrin‐based metal–organic framework hollow nanotubes for
immobilization of noble‐metal single atoms. Angewandte Chemie, 130(13), pp.3551-
3556. https://doi.org/10.1002/ange.201800817

[363] Dooris, E., McAnally, C.A., Cussen, E.J., Kennedy, A.R. and Fletcher, A.J., 2016. A
Family of nitrogen-enriched metal organic frameworks with CCS
potential. Crystals, 6(1), p.14. https://doi.org/10.3390/cryst6010014

[364] Cirujano, F.G. and Dhakshinamoorthy, A., 2021. Engineering of active sites in metal–
organic frameworks for biodiesel production. Advanced Sustainable Systems, 5(8),
p.2100101. https://doi.org/10.1002/adsu.202100101

[365] Roberts, J.M., Fini, B.M., Sarjeant, A.A., Farha, O.K., Hupp, J.T. and Scheidt, K.A.,
2012. Urea metal–organic frameworks as effective and size-selective hydrogen-bond
catalysts. Journal of the American Chemical Society, 134(7), pp.3334-3337.
https://doi.org/10.1021/ja2108118

[366] Mylonas‐Margaritis, I., Mayans, J., Efthymiou, C.G., McArdle, P. and


Papatriantafyllopoulou, C., 2022. Mixed‐Ligand Metal‐Organic Frameworks:
Synthesis and Characterization of New MOFs Containing Pyridine‐2, 6‐dimethanolate
and Benzene‐1, 4‐dicarboxylate Ligands. European Journal of Inorganic
Chemistry, 2022(19), p.e202200140.https://doi.org/10.1002/ejic.202200140

[367] Deng, X., Zheng, S.L., Zhong, Y.H., Hu, J., Chung, L.H. and He, J., 2022. Conductive
MOFs based on Thiol-functionalized Linkers: Challenges, Opportunities, and Recent
Advances. Coordination Chemistry Reviews, 450, p.214235.
https://doi.org/10.1016/j.ccr.2021.214235

181
[368] Yee, K.K., Reimer, N., Liu, J., Cheng, S.Y., Yiu, S.M., Weber, J., Stock, N. and Xu,
Z., 2013. Effective mercury sorption by thiol-laced metal–organic frameworks: in
strong acid and the vapor phase. Journal of the American Chemical Society, 135(21),
pp.7795-7798. https://doi.org/10.1021/ja400212k

[369] Wu, L.Y., Mu, Y.F., Guo, X.X., Zhang, W., Zhang, Z.M., Zhang, M. and Lu, T.B.,
2019. Encapsulating perovskite quantum dots in iron‐based metal–organic frameworks
(MOFs) for efficient photocatalytic CO2 reduction. Angewandte Chemie International
Edition, 58(28), pp.9491-9495. https://doi.org/10.1002/anie.201904537

[370] Saha, S., Das, G., Thote, J. and Banerjee, R., 2014. Photocatalytic metal–organic
framework from CdS quantum dot incubated luminescent metallohydrogel. Journal of
the American Chemical Society, 136(42), pp.14845-14851.
https://doi.org/10.1021/ja509019k

[371] Shen, L., Luo, M., Liu, Y., Liang, R., Jing, F. and Wu, L., 2015. Noble-metal-free MoS2
co-catalyst decorated UiO-66/CdS hybrids for efficient photocatalytic H2
production. Applied Catalysis B: Environmental, 166, pp.445-453.
https://doi.org/10.1016/j.apcatb.2014.11.056

[372] Low, J.J., Benin, A.I., Jakubczak, P., Abrahamian, J.F., Faheem, S.A. and Willis, R.R.,
2009. Virtual high throughput screening confirmed experimentally: porous
coordination polymer hydration. Journal of the American Chemical Society, 131(43),
pp.15834-15842. https://doi.org/10.1021/ja9061344

[373] Shen, L., Liang, R., Luo, M., Jing, F. and Wu, L., 2015. Electronic effects of ligand
substitution on metal–organic framework photocatalysts: the case study of UiO-
66. Physical chemistry chemical physics, 17(1), pp.117-121.
https://doi.org/10.1039/C4CP04162C

[374] Wu, H., Yildirim, T. and Zhou, W., 2013. Exceptional mechanical stability of highly
porous zirconium metal–organic framework UiO-66 and its important
implications. The journal of physical chemistry letters, 4(6), pp.925-930.
https://doi.org/10.1021/jz4002345

[375] Wang, J., Cherevan, A.S., Hannecart, C., Naghdi, S., Nandan, S.P., Gupta, T. and Eder,
D., 2021. Ti-based MOFs: New insights on the impact of ligand composition and hole
scavengers on stability, charge separation and photocatalytic hydrogen
evolution. Applied Catalysis B: Environmental, 283, p.119626.
https://doi.org/10.1016/j.apcatb.2020.119626

[376] Wang, J.L., Wang, C. and Lin, W., 2012. ACS Catal. 2012, 2, 2630. Fu, D. Sun, Y.
Chen, R. Huang, Z. Ding, X. Fu, Z. Li, Angew. Chem., Int. Ed, 51, p.3364.
https://doi.org/10.1002/anie.201108357

[377] JG, N.M.H.C.S., 2016. Svane K. van der Linden B. Veber SL Fedin MV Houtepen AJ
van der Veen MA Kapteijn F. Walsh A. Gascon J. Sci. Rep, 6,
p.23676.https://doi.org/10.1038/srep23676

[378] Farha, O.K., Eryazici, I., Jeong, N.C., Hauser, B.G., Wilmer, C.E., Sarjeant, A.A.,
Snurr, R.Q., Nguyen, S.T., Yazaydın, A.O. and Hupp, J.T., 2012. Metal–organic

182
framework materials with ultrahigh surface areas: is the sky the limit?. Journal of the
American Chemical Society, 134(36), pp.15016-15021.
https://doi.org/10.1021/ja3055639

[379] Morshedy, A.S., Abd El Salam, H.M., El Naggar, A.M. and Zaki, T., 2020. Hydrogen
Production and In Situ Storage through Process of Water Splitting Using Mono/Binary
Metal–Organic Framework (MOF) Structures as New Chief Photocatalysts. Energy &
Fuels, 34(9), pp.11660-11669. https://doi.org/10.1021/acs.energyfuels.0c01559

[380] El Naggar, A.M., Nassar, I.M. and Gobara, H.M., 2013. Enhanced hydrogen production
from water via a photo-catalyzed reaction using chalcogenide d-element nanoparticles
induced by UV light. Nanoscale, 5(20), pp.9994-9999.
https://doi.org/10.1039/C3NR02640J

[381] Al Amery, N., Abid, H.R., Al-Saadi, S., Wang, S. and Liu, S., 2020. Facile directions
for synthesis, modification and activation of MOFs. Materials Today Chemistry, 17,
p.100343. https://doi.org/10.1016/j.mtchem.2020.100343

[382] Howarth, A.J., Peters, A.W., Vermeulen, N.A., Wang, T.C., Hupp, J.T. and Farha,
O.K., 2017. Best practices for the synthesis, activation, and characterization of metal–
organic frameworks. Chemistry of Materials, 29(1), pp.26-39.
https://doi.org/10.1021/acs.chemmater.6b02626

[383] Chen, L., Song, Z., Wang, X., Prikhodko, S.V., Hu, J., Kodambaka, S. and Richards,
R., 2009. Three-dimensional morphology control during wet chemical synthesis of
porous chromium oxide spheres. ACS applied materials & interfaces, 1(9), pp.1931-
1937. https://doi.org/10.1021/am900334q

[384] Meng, X. and Xiao, F.S., 2014. Green routes for synthesis of zeolites. Chemical
reviews, 114(2), pp.1521-1543. https://doi.org/10.1021/cr4001513

[385] Kumar, S., Jain, S., Nehra, M., Dilbaghi, N., Marrazza, G. and Kim, K.H., 2020. Green
synthesis of metal–organic frameworks: A state-of-the-art review of potential
environmental and medical applications. Coordination Chemistry Reviews, 420,
p.213407. https://doi.org/10.1016/j.ccr.2020.213407

[386] DEYC, K., 2014. Crystallinemetal-organicframeworks (MOFs): synthesis,


structureandfunction. ActaCryst, 70, pp.3-10.
https://doi.org/10.1107/S2052520613029557

[387] Murinzi, T.W., Hosten, E. and Watkins, G.M., 2017. Synthesis and characterization of
a cobalt-2, 6-pyridinedicarboxylate MOF with potential application in electrochemical
sensing. Polyhedron, 137, pp.188-196. https://doi.org/10.1016/j.poly.2017.08.030

[388] Bilal, M., Adeel, M., Rasheed, T. and Iqbal, H.M., 2019. Multifunctional metal–organic
frameworks-based biocatalytic platforms: recent developments and future
prospects. Journal of Materials Research and Technology, 8(2), pp.2359-2371.
https://doi.org/10.1016/j.jmrt.2018.12.001

[389] Crane, J.L., Anderson, K.E. and Conway, S.G., 2015. Hydrothermal synthesis and
characterization of a metal–organic framework by Thermogravimetric analysis, powder
X-ray diffraction, and infrared spectroscopy: an integrative inorganic chemistry
183
experiment. Journal of Chemical Education, 92(2), pp.373-377.
https://doi.org/10.1021/ed5000839

[390] Walton, R.I., 2002. Subcritical solvothermal synthesis of condensed inorganic


materials. Chemical Society Reviews, 31(4), pp.230-238.
https://doi.org/10.1039/B105762F

[391] Karagiaridi, O., Bury, W., Sarjeant, A.A., Hupp, J.T. and Farha, O.K., 2014. Synthesis
and characterization of functionalized metal-organic frameworks. JoVE (Journal of
Visualized Experiments), (91), p.e52094. https://doi.org/10.3791/52094

[392] Sun, Y. and Zhou, H.C., 2015. Recent progress in the synthesis of metal–organic
frameworks. Science and technology of advanced materials.
http://dx.doi.org/10.1088/1468-6996/16/5/054202

[393] Paz, F.A.A. and Klinowski, J., 2004. Hydrothermal synthesis and structural
characterization of a novel cadmium-organic framework. Journal of Solid State
Chemistry, 177(10), pp.3423-3432. https://doi.org/10.1016/j.jssc.2004.05.022

[394] Schlesinger, M., Schulze, S., Hietschold, M. and Mehring, M., 2010. Evaluation of
synthetic methods for microporous metal–organic frameworks exemplified by the
competitive formation of [Cu2 (btc)3 (H2O) 3] and [Cu2 (btc)(OH)(H2O)]. Microporous
and Mesoporous Materials, 132(1-2), pp.121-127.
https://doi.org/10.1016/j.micromeso.2010.02.008

[395] Hausdorf, S., Baitalow, F., Seidel, J. and Mertens, F.O., 2007. Gaseous species as
reaction tracers in the solvothermal synthesis of the zinc oxide terephthalate MOF-
5. The Journal of Physical Chemistry A, 111(20), pp.4259-4266.
https://doi.org/10.1021/jp0708291

[396] Han, Z.B., Lu, R.Y., Liang, Y.F., Zhou, Y.L., Chen, Q. and Zeng, M.H., 2012. Mn (II)-
based porous metal–organic framework showing metamagnetic properties and high
hydrogen adsorption at low pressure. Inorganic Chemistry, 51(1), pp.674-679.
https://doi.org/10.1021/ic2021929

[397] Serre, C., Millange, F., Surblé, S. and Férey, G., 2004. A route to the synthesis of
trivalent transition‐metal porous carboxylates with trimeric secondary building
units. Angewandte Chemie International Edition, 43(46), pp.6285-6289.
https://doi.org/10.1002/anie.200454250

[398] Surblé, S., Serre, C., Mellot-Draznieks, C., Millange, F. and Férey, G., 2006. A new
isoreticular class of metal-organic-frameworks with the MIL-88 topology. Chemical
communications, (3), pp.284-286. https://doi.org/10.1039/B512169H

[399] Guillerm, V., Gross, S., Serre, C., Devic, T., Bauer, M. and Férey, G., 2010. A
zirconium methacrylate oxocluster as precursor for the low-temperature synthesis of
porous zirconium (IV) dicarboxylates. Chemical communications, 46(5), pp.767-769.
https://doi.org/10.1039/B914919H

[400] Hashemi, L., Morsali, A., Yilmaz, V.T., Büyükgüngor, O., Khavasi, H.R., Ashouri, F.
and Bagherzadeh, M., 2014. Sonochemical syntheses of two nano-sized lead (II) metal–
organic frameworks; application for catalysis and preparation of lead (II) oxide
184
nanoparticles. Journal of Molecular Structure, 1072, pp.260-266.
http://dx.doi.org/10.1016/j.molstruc.2014.05.021

[401] Yu, K., Lee, Y.R., Seo, J.Y., Baek, K.Y., Chung, Y.M. and Ahn, W.S., 2021.
Sonochemical synthesis of Zr-based porphyrinic MOF-525 and MOF-545:
Enhancement in catalytic and adsorption properties. Microporous and Mesoporous
Materials, 316, p.110985. https://doi.org/10.1016/j.micromeso.2021.110985

[402] Chatel, G. and Colmenares, J.C., 2017. Sonochemistry: from basic principles to
innovative applications. Topics in Current Chemistry, 375(1), pp.1-4.
https://doi.org/10.1007/s41061-016-0096-1

[403] Suslick, K.S., Choe, S.B., Cichowlas, A.A. and Grinstaff, M.W., 1991. Sonochemical
synthesis of amorphous iron. nature, 353(6343), pp.414-416.
https://doi.org/10.1038/353414a0

[404] Son, W.J., Kim, J., Kim, J. and Ahn, W.S., 2008. Sonochemical synthesis of MOF-
5. Chemical Communications, (47), pp.6336-6338. https://doi.org/10.1039/B814740J

[405] Gedanken, A., 2004. Using sonochemistry for the fabrication of


nanomaterials. Ultrasonics sonochemistry, 11(2), pp.47-
55.https://doi.org/10.1016/j.ultsonch.2004.01.037

[406] Haque, E., Khan, N.A., Park, J.H. and Jhung, S.H., 2010. Synthesis of a metal–organic
framework material, iron terephthalate, by ultrasound, microwave, and conventional
electric heating: a kinetic study. Chemistry–A European Journal, 16(3), pp.1046-1052.
https://doi.org/10.1002/chem.200902382

[407] Jung, D.W., Yang, D.A., Kim, J., Kim, J. and Ahn, W.S., 2010. Facile synthesis of
MOF-177 by a sonochemical method using 1-methyl-2-pyrrolidinone as a
solvent. Dalton Transactions, 39(11), pp.2883-2887.
https://doi.org/10.1039/B925088C

[408] Vaitsis, C., Sourkouni, G. and Argirusis, C., 2020. Sonochemical synthesis of MOFs.
In Metal-Organic Frameworks for Biomedical Applications (pp. 223-244). Woodhead
Publishing. https://doi.org/10.1016/B978-0-12-816984-1.00013-5

[409] Lee, Y.R., Kim, J. and Ahn, W.S., 2013. Synthesis of metal-organic frameworks: A
mini review. Korean Journal of Chemical Engineering, 30(9), pp.1667-1680.
https://doi.org/10.1007/s11814-013-0140-6

[410] Azad, F.N., Ghaedi, M., Dashtian, K., Hajati, S. and Pezeshkpour, V., 2016.
Ultrasonically assisted hydrothermal synthesis of activated carbon–HKUST-1-MOF
hybrid for efficient simultaneous ultrasound-assisted removal of ternary organic dyes

185
and antibacterial investigation: Taguchi optimization. Ultrasonics sonochemistry, 31,
pp.383-393. https://doi.org/10.1016/j.ultsonch.2016.01.024

[411] Li, Z.Q., Qiu, L.G., Xu, T., Wu, Y., Wang, W., Wu, Z.Y. and Jiang, X., 2009. Ultrasonic
synthesis of the microporous metal–organic framework Cu3 (BTC)2 at ambient
temperature and pressure: an efficient and environmentally friendly method. Materials
Letters, 63(1), pp.78-80. https://doi.org/10.1016/j.matlet.2008.09.010

[412] Stawowy, M., Róziewicz, M., Szczepańska, E., Silvestre-Albero, J., Zawadzki, M.,
Musioł, M., Łuzny, R., Kaczmarczyk, J., Trawczyński, J. and Łamacz, A., 2019. The
impact of synthesis method on the properties and CO2 sorption capacity of UiO-66
(Ce). Catalysts, 9(4), p.309. https://doi.org/10.3390/catal9040309

[413] Lagashetty, A., Havanoor, V., Basavaraja, S., Balaji, S.D. and Venkataraman, A., 2007.
Microwave-assisted route for synthesis of nanosized metal oxides. Science and
Technology of Advanced Materials, 8(6), p.484.
https://doi.org/10.1016/j.stam.2007.07.001

[414] Gangu, K.K., Maddila, S., Mukkamala, S.B. and Jonnalagadda, S.B., 2016. A review
on contemporary metal–organic framework materials. Inorganica Chimica Acta, 446,
pp.61-74. http://dx.doi.org/10.1016/j.ica.2016.02.062

[415] Horcajada, P., Chalati, T., Serre, C., Gillet, B., Sebrie, C., Baati, T., Eubank, J.F.,
Heurtaux, D., Clayette, P., Kreuz, C. and Chang, J.S., 2010. Porous metal–organic-
framework nanoscale carriers as a potential platform for drug delivery and
imaging. Nature materials, 9(2), pp.172-178. https://doi.org/10.1038/nmat2608

[416] Jhung, S.H., Lee, J.H., Forster, P.M., Férey, G., Cheetham, A.K. and Chang, J.S., 2006.
Microwave Synthesis of Hybrid Inorganic–Organic Porous Materials: Phase‐Selective
and Rapid Crystallization. Chemistry–A European Journal, 12(30), pp.7899-7905.
https://doi.org/10.1002/chem.200600270

[417] Isaeva, V.I. and Kustov, L.M., 2016. Microwave activation as an alternative production
of metal-organic frameworks. Russian Chemical Bulletin, 65(9), pp.2103-2114.
https://doi.org/10.1007/s11172-016-1559-9

[418] Dai, X., Jia, X., Zhao, P., Wang, T., Wang, J., Huang, P., He, L. and Hou, X., 2016. A
combined experimental/computational study on metal-organic framework MIL-101
(Cr) as a SPE sorbent for the determination of sulphonamides in environmental water
samples coupling with UPLC-MS/MS. Talanta, 154, pp.581-588.
https://doi.org/10.1016/j.talanta.2016.03.042

[419] Liang, W. and D'Alessandro, D.M., 2013. Microwave-assisted solvothermal synthesis


of zirconium oxide based metal–organic frameworks. Chemical
Communications, 49(35), pp.3706-3708. https://doi.org/10.1039/C3CC40368H

[420] Sabouni, R., Kazemian, H. and Rohani, S., 2012. Microwave Synthesis of the CPM‐5
Metal Organic Framework. Chemical engineering & technology, 35(6), pp.1085-1092.
https://doi.org/10.1002/ceat.201100626

186
[421] Klinowski, J., Paz, F.A.A., Silva, P. and Rocha, J., 2011. Microwave-assisted synthesis
of metal–organic frameworks. Dalton Transactions, 40(2), pp.321-330.
https://doi.org/10.1039/C0DT00708K

[422] Vakili, R., Xu, S., Al-Janabi, N., Gorgojo, P., Holmes, S.M. and Fan, X., 2018.
Microwave-assisted synthesis of zirconium-based metal organic frameworks (MOFs):
Optimization and gas adsorption. Microporous and Mesoporous Materials, 260, pp.45-
53. https://doi.org/10.1016/j.micromeso.2017.10.028

[423] Pirzadeh, K., Ghoreyshi, A.A., Rahimnejad, M. and Mohammadi, M., 2018.
Electrochemical synthesis, characterization and application of a microstructure Cu3
(BTC)2 metal organic framework for CO2 and CH4 separation. Korean Journal of
Chemical Engineering, 35(4), pp.974-983. https://doi.org/10.1007/s11814-017-0340-6

[424] Zhao, S., Wang, Y., Dong, J., He, C.T., Yin, H., An, P., Zhao, K., Zhang, X., Gao, C.,
Zhang, L. and Lv, J., 2016. Ultrathin metal–organic framework nanosheets for
electrocatalytic oxygen evolution. Nature Energy, 1(12), pp.1-10.
https://doi.org/10.1038/nenergy.2016.184

[425] Ramimoghadam, D., Bagheri, S. and Abd Hamid, S.B., 2014. Progress in
electrochemical synthesis of magnetic iron oxide nanoparticles. Journal of Magnetism
and Magnetic Materials, 368, pp.207-229.
https://doi.org/10.1016/j.jmmm.2014.05.015

[426] Martinez Joaristi, A., Juan-Alcañiz, J., Serra-Crespo, P., Kapteijn, F. and Gascon, J.,
2012. Electrochemical synthesis of some archetypical Zn2+, Cu2+, and Al3+ metal
organic frameworks. Crystal Growth & Design, 12(7), pp.3489-3498.
https://doi.org/10.1021/cg300552w

[427] Li, M. and Dincă, M., 2015. On the mechanism of MOF-5 formation under cathodic
bias. Chemistry of Materials, 27(9), pp.3203-3206.
https://doi.org/10.1021/acs.chemmater.5b00899

[428] Varsha, M.V. and Nageswaran, G., 2020. Direct electrochemical synthesis of metal
organic frameworks. Journal of the Electrochemical Society, 167(15), p.155527.
DOI 10.1149/1945-7111/abc6c6

[429] Al‐Kutubi, H., Gascon, J., Sudhölter, E.J. and Rassaei, L., 2015. Electrosynthesis of
metal–organic frameworks: challenges and opportunities. ChemElectroChem, 2(4),
pp.462-474. https://doi.org/10.1002/celc.201402429

[430] Campagnol, N., Van Assche, T., Boudewijns, T., Denayer, J., Binnemans, K., De Vos,
D. and Fransaer, J., 2013. High pressure, high temperature electrochemical synthesis of
metal–organic frameworks: films of MIL-100 (Fe) and HKUST-1 in different
morphologies. Journal of Materials Chemistry A, 1(19), pp.5827-5830.
https://doi.org/10.1039/C3TA10419B

187
[431] Wu, W., Decker, G.E., Weaver, A.E., Arnoff, A.I., Bloch, E.D. and Rosenthal, J., 2021.
Facile and Rapid Room-Temperature Electrosynthesis and Controlled Surface Growth
of Fe-MIL-101 and Fe-MIL-101-NH2. ACS central science, 7(8), pp.1427-1433.
https://doi.org/10.1021/acscentsci.1c00686

[432] Stassen, I., Styles, M., Van Assche, T., Campagnol, N., Fransaer, J., Denayer, J., Tan,
J.C., Falcaro, P., De Vos, D. and Ameloot, R., 2015. Electrochemical film deposition
of the zirconium metal–organic framework UiO-66 and application in a miniaturized
sorbent trap. Chemistry of materials, 27(5), pp.1801-1807.
https://doi.org/10.1021/cm504806p

[433] Ameloot, R., Stappers, L., Fransaer, J., Alaerts, L., Sels, B.F. and De Vos, D.E., 2009.
Patterned growth of metal-organic framework coatings by electrochemical
synthesis. Chemistry of Materials, 21(13), pp.2580-2582.
https://doi.org/10.1021/cm900069f

[434] Wei, J.Z., Gong, F.X., Sun, X.J., Li, Y., Zhang, T., Zhao, X.J. and Zhang, F.M., 2019.
Rapid and low-cost electrochemical synthesis of UiO-66-NH2 with enhanced
fluorescence detection performance. Inorganic chemistry, 58(10), pp.6742-6747.
https://doi.org/10.1021/acs.inorgchem.9b00157

[435] Jabarian, S. and Ghaffarinejad, A., 2019. Electrochemical synthesis of NiBTC metal
organic framework thin layer on nickel foam: An efficient electrocatalyst for the
hydrogen evolution reaction. Journal of Inorganic and Organometallic Polymers and
Materials, 29(5), pp.1565-1574. https://doi.org/10.1007/s10904-019-01120-4

[436] Hafizovic, J., Bjørgen, M., Olsbye, U., Dietzel, P.D., Bordiga, S., Prestipino, C.,
Lamberti, C. and Lillerud, K.P., 2007. The inconsistency in adsorption properties and
powder XRD data of MOF-5 is rationalized by framework interpenetration and the
presence of organic and inorganic species in the nanocavities. Journal of the American
Chemical Society, 129(12), pp.3612-3620. https://doi.org/10.1021/ja0675447

[437] Ameloot, R., Gobechiya, E., Uji‐i, H., Martens, J.A., Hofkens, J., Alaerts, L., Sels, B.F.
and De Vos, D.E., 2010. Direct patterning of oriented metal–organic framework
crystals via control over crystallization kinetics in clear precursor solutions. Advanced
materials, 22(24), pp.2685-2688. https://doi.org/10.1002/adma.200903867

[438] Tranchemontagne, D.J., Hunt, J.R. and Yaghi, O.M., 2008. Room temperature
synthesis of metal-organic frameworks: MOF-5, MOF-74, MOF-177, MOF-199, and
IRMOF-0. Tetrahedron, 64(36), pp.8553-8557.
https://doi.org/10.1016/j.tet.2008.06.036

[439] Seetharaj, R., Vandana, P.V., Arya, P. and Mathew, S., 2019. Dependence of solvents,
pH, molar ratio and temperature in tuning metal organic framework
architecture. Arabian journal of chemistry, 12(3), pp.295-315.
https://doi.org/10.1016/j.arabjc.2016.01.003

[440] Michaelides, A. and Skoulika, S., 2005. Reactive microporous rare-earth coordination
polymers that exhibit single-crystal-to-single-crystal dehydration and

188
rehydration. Crystal growth & design, 5(2), pp.529-533.
https://doi.org/10.1021/cg049725v

[441] Vehring, R., 2008. Pharmaceutical particle engineering via spray


drying. Pharmaceutical research, 25(5), pp.999-1022. https://doi.org/10.1007/s11095-
007-9475-1

[442] Carné-Sánchez, A., Imaz, I., Cano-Sarabia, M. and Maspoch, D., 2013. A spray-drying
strategy for synthesis of nanoscale metal–organic frameworks and their assembly into
hollow superstructures. Nature Chemistry, 5(3), pp.203-211.
https://doi.org/10.1038/nchem.1569

[443] Gholampour, N., Chaemchuen, S., Hu, Z.Y., Mousavi, B., Van Tendeloo, G. and
Verpoort, F., 2017. Simultaneous creation of metal nanoparticles in metal organic
frameworks via spray drying technique. Chemical Engineering Journal, 322, pp.702-
709. https://doi.org/10.1016/j.cej.2017.04.085

[444] Tanaka, S. and Miyashita, R., 2017. Aqueous-system-enabled spray-drying technique


for the synthesis of hollow polycrystalline ZIF-8 MOF particles. ACS omega, 2(10),
pp.6437-6445. https://doi.org/10.1021/acsomega.7b01325

[445] Avci-Camur, C., Troyano, J., Pérez-Carvajal, J., Legrand, A., Farrusseng, D., Imaz, I.
and Maspoch, D., 2018. Aqueous production of spherical Zr-MOF beads via
continuous-flow spray-drying. Green chemistry, 20(4), pp.873-878.
https://doi.org/10.1039/C7GC03132G

[446] Guillerm, V., Garzón‐Tovar, L., Yazdi, A., Imaz, I., Juanhuix, J. and Maspoch, D.,
2017. Continuous One‐Step Synthesis of Porous M‐XF6‐Based Metal‐Organic and
Hydrogen‐Bonded Frameworks. Chemistry–A European Journal, 23(28), pp.6829-
6835. https://doi.org/10.1002/chem.201605507

[447] Crawford, D., Casaban, J., Haydon, R., Giri, N., McNally, T. and James, S.L., 2015.
Synthesis by extrusion: continuous, large-scale preparation of MOFs using little or no
solvent. Chemical science, 6(3), pp.1645-1649. https://doi.org/10.1039/C4SC03217A

[448] Friščić, T., 2010. New opportunities for materials synthesis using
mechanochemistry. Journal of Materials Chemistry, 20(36), pp.7599-7605.
https://doi.org/10.1039/C0JM00872A

[449] Friščić, T., Halasz, I., Štrukil, V., Eckert-Maksić, M. and Dinnebier, R.E., 2012. Clean
and efficient synthesis using mechanochemistry: coordination polymers, metal-organic
frameworks and metallodrugs. Croatica chemica acta, 85(3), pp.367-378.
http://dx.doi.org/10.5562/cca2014

[450] Klimakow, M., Klobes, P., Thünemann, A.F., Rademann, K. and Emmerling, F., 2010.
Mechanochemical synthesis of metal− organic frameworks: a fast and facile approach
toward quantitative yields and high specific surface areas. Chemistry of
Materials, 22(18), pp.5216-5221. https://doi.org/10.1021/cm1012119

[451] Braga, D., Curzi, M., Johansson, A., Polito, M., Rubini, K. and Grepioni, F., 2006.
Simple and quantitative mechanochemical preparation of a porous crystalline material

189
based on a 1D coordination network for uptake of small molecules. Angewandte
Chemie, 118(1), pp.148-152. https://doi.org/10.1002/ange.200502597

[452] Pichon, A., Lazuen-Garay, A. and James, S.L., 2006. Solvent-free synthesis of a
microporous metal–organic framework. CrystEngComm, 8(3), pp.211-214.
https://doi.org/10.1039/B513750K

[453] Braga, D., Giaffreda, S.L., Grepioni, F., Pettersen, A., Maini, L., Curzi, M. and Polito,
M., 2006. Mechanochemical preparation of molecular and supramolecular
organometallic materials and coordination networks. Dalton Transactions, (10),
pp.1249-1263. https://doi.org/10.1039/B516165G

[454] Beldon, P.J., Fábián, L., Stein, R.S., Thirumurugan, A., Cheetham, A.K. and Friščić,
T., 2010. Rapid room‐temperature synthesis of zeolitic imidazolate frameworks by
using mechanochemistry. Angewandte Chemie, 122(50), pp.9834-9837.
https://doi.org/10.1002/ange.201005547

[455] Friščić, T. and Fábián, L., 2009. Mechanochemical conversion of a metal oxide into
coordination polymers and porous frameworks using liquid-assisted grinding
(LAG). CrystEngComm, 11(5), pp.743-745. https://doi.org/10.1039/B822934C

[456] Friščić, T., Reid, D.G., Halasz, I., Stein, R.S., Dinnebier, R.E. and Duer, M.J., 2010.
Ion‐and liquid‐assisted grinding: improved mechanochemical synthesis of metal–
organic frameworks reveals salt inclusion and anion templating. Angewandte
Chemie, 122(4), pp.724-727. https://doi.org/10.1002/ange.200906583

[457] Tanaka, S., Kida, K., Nagaoka, T., Ota, T. and Miyake, Y., 2013. Mechanochemical
dry conversion of zinc oxide to zeolitic imidazolate framework. Chemical
Communications, 49(72), pp.7884-7886. https://doi.org/10.1039/C3CC43028F

[458] Julien, P.A., Užarević, K., Katsenis, A.D., Kimber, S.A., Wang, T., Farha, O.K., Zhang,
Y., Casaban, J., Germann, L.S., Etter, M. and Dinnebier, R.E., 2016. In situ monitoring
and mechanism of the mechanochemical formation of a microporous MOF-74
framework. Journal of the American Chemical Society, 138(9), pp.2929-2932.
https://doi.org/10.1021/jacs.5b13038

[459] Guo, Z., Zhang, Q., Cheng, Z., Liu, Q., Zuo, J., Jin, B. and Peng, R., 2020. Air-flow
impacting synthesis of metal organic frameworks: a continuous, highly efficient, large-
scale mechanochemical synthetic method. ACS Sustainable Chemistry &
Engineering, 8(10), pp.4037-4043. https://doi.org/10.1021/acssuschemeng.9b05395

[460] Huang, Y.H., Lo, W.S., Kuo, Y.W., Chen, W.J., Lin, C.H. and Shieh, F.K., 2017. Green
and rapid synthesis of zirconium metal–organic frameworks via mechanochemistry:
UiO-66 analog nanocrystals obtained in one hundred seconds. Chemical
Communications, 53(43), pp.5818-5821. https://doi.org/10.1039/C7CC03105J

[461] Zhang, R., Tao, C.A., Chen, R., Wu, L., Zou, X. and Wang, J., 2018. Ultrafast synthesis
of Ni-MOF in one minute by ball milling. Nanomaterials, 8(12), p.1067.
https://doi.org/10.3390/nano8121067

[462] Wang, T., Zhai, Y., Zhu, Y., Li, C. and Zeng, G., 2018. A review of the hydrothermal
carbonization of biomass waste for hydrochar formation: Process conditions,
190
fundamentals, and physicochemical properties. Renewable and Sustainable Energy
Reviews, 90, pp.223-247. https://doi.org/10.1016/j.rser.2018.03.071

[463] Campanati, M., Fornasari, G. and Vaccari, A., 2003. Fundamentals in the preparation
of heterogeneous catalysts. Catalysis today, 77(4), pp.299-314.
https://doi.org/10.1016/S0920-5861(02)00375-9

[464] Kinsinger, N.M., Dudchenko, A., Wong, A. and Kisailus, D., 2013. Synergistic effect
of pH and phase in a nanocrystalline titania photocatalyst. ACS Applied Materials &
Interfaces, 5(13), pp.6247-6254. https://doi.org/10.1021/am401247h

[465] Jassby, D., Farner Budarz, J. and Wiesner, M., 2012. Impact of aggregate size and
structure on the photocatalytic properties of TiO2 and ZnO
nanoparticles. Environmental science & technology, 46(13), pp.6934-6941.
https://doi.org/10.1021/es202009h

[466] Grzechulska, J. and Morawski, A.W., 2002. Photocatalytic decomposition of azo-dye


acid black 1 in water over modified titanium dioxide. Applied Catalysis B:
Environmental, 36(1), pp.45-51. https://doi.org/10.1016/S0926-3373(01)00275-2

[467] Tschirch, J., Dillert, R., Bahnemann, D., Proft, B., Biedermann, A. and Goer, B., 2008.
Photodegradation of methylene blue in water, a standard method to determine the
activity of photocatalytic coatings? Research on Chemical Intermediates, 34(4),
pp.381-392. https://doi.org/10.1163/156856708784040588

[468] Maeda, K., Teramura, K., Saito, N., Inoue, Y., Kobayashi, H. and Domen, K., 2006.
Overall water splitting using (oxy) nitride photocatalysts. Pure and applied
chemistry, 78(12), pp.2267-2276. https://doi.org/10.1351/pac200678122267

[469] Zhang, G., Zhang, W., Minakata, D., Chen, Y., Crittenden, J. and Wang, P., 2013. The
pH effects on H2 evolution kinetics for visible light water splitting over the Ru/(CuAg)
0.15 In0.3Zn1.4S2 photocatalyst. international journal of hydrogen energy, 38(27),
pp.11727-11736. https://doi.org/10.1016/j.ijhydene.2013.06.140

[470] Yang, H., Jin, Z., Fan, K., Liu, D. and Lu, G., 2017. The roles of Ni nanoparticles over
CdS nanorods for improved photocatalytic stability and activity. Superlattices and
Microstructures, 111, pp.687-695. https://doi.org/10.1016/j.spmi.2017.07.025

[471] Choi, J., Ryu, S.Y., Balcerski, W., Lee, T.K. and Hoffmann, M.R., 2008. Photocatalytic
production of hydrogen on Ni/NiO/KNbO3/CdS nanocomposites using visible
light. Journal of Materials Chemistry, 18(20), pp.2371-2378.
https://doi.org/10.1039/B718535A

[472] Zhao, X., Liu, D., Huang, H., Zhang, W., Yang, Q. and Zhong, C., 2014. The stability
and defluoridation performance of MOFs in fluoride solutions. Microporous and
mesoporous materials, 185, pp.72-78.
https://doi.org/10.1016/j.micromeso.2013.11.002

[473] Zhou, T., Du, Y., Borgna, A., Hong, J., Wang, Y., Han, J., Zhang, W. and Xu, R., 2013.
Post-synthesis modification of a metal–organic framework to construct a bifunctional
photocatalyst for hydrogen production. Energy & Environmental Science, 6(11),
pp.3229-3234. https://doi.org/10.1039/C3EE41548A
191
[474] Shamzhy, M.V., Opanasenko, M.V., Garcia, H. and Čejka, J., 2015. Annulation of
phenols with methylbutenol over MOFs: The role of catalyst structure and acid strength
in producing 2, 2-dimethylbenzopyran derivatives. Microporous and Mesoporous
Materials, 202, pp.297-302. https://doi.org/10.1016/j.micromeso.2014.10.003

[475] Feng, Y., Chen, C., Liu, Z., Fei, B., Lin, P., Li, Q., Sun, S. and Du, S., 2015. Application
of a Ni mercaptopyrimidine MOF as highly efficient catalyst for sunlight-driven
hydrogen generation. Journal of Materials Chemistry A, 3(13), pp.7163-7169.
https://doi.org/10.1039/C5TA00136F

[476] Han, Z., McNamara, W.R., Eum, M.S., Holland, P.L. and Eisenberg, R., 2012. A nickel
thiolate catalyst for the long‐lived photocatalytic production of hydrogen in a noble‐
metal‐free system. Angewandte Chemie, 124(7), pp.1699-1702.
https://doi.org/10.1002/anie.201107329

[477] Ejsmont, A., Jankowska, A. and Goscianska, J., 2022. Insight into the Photocatalytic
Activity of Cobalt-Based Metal–Organic Frameworks and Their
Composites. Catalysts, 12(2), p.110. https://doi.org/10.3390/catal12020110

[478] Jin, Z. and Yang, H., 2017. Exploration of Zr–metal–organic framework as efficient
photocatalyst for hydrogen production. Nanoscale research letters, 12, pp.1-10.
https://doi.org/10.1186/s11671-017-2311-6

[479] Min, S. and Lu, G., 2011. Dye-sensitized reduced graphene oxide photocatalysts for
highly efficient visible-light-driven water reduction. The Journal of Physical Chemistry
C, 115(28), pp.13938-13945. https://doi.org/10.1021/jp203750z

[480] Chen, X., Liu, L., Yu, P.Y. and Mao, S.S., 2011. Increasing solar absorption for
photocatalysis with black hydrogenated titanium dioxide
nanocrystals. Science, 331(6018), pp.746-750.
https://doi.org/10.1126/science.1200448

[481] Tao, X., Zhao, Y., Mu, L., Wang, S., Li, R. and Li, C., 2018. Bismuth tantalum
oxyhalogen: a promising candidate photocatalyst for solar water splitting. Advanced
Energy Materials, 8(1), p.1701392. https://doi.org/10.1002/aenm.201701392

[482] Cao, S., Chen, Y., Wang, C.J., He, P. and Fu, W.F., 2014. Highly efficient
photocatalytic hydrogen evolution by nickel phosphide nanoparticles from aqueous

192
solution. Chemical Communications, 50(72), pp.10427-10429.
https://doi.org/10.1039/C4CC05026F

[483] Rather, R.A., Singh, S. and Pal, B., 2017. Visible and direct sunlight induced H2
production from water by plasmonic Ag-TiO2 nanorods hybrid interface. Solar Energy
Materials and Solar Cells, 160, pp.463-469.
https://doi.org/10.1016/j.solmat.2016.11.017

[484] Cao, S. and Piao, L., 2020. Considerations for a more accurate evaluation method for
photocatalytic water splitting. Angewandte Chemie International Edition, 59(42),
pp.18312-18320. https://doi.org/10.1002/anie.202009633

[485] Navarro, R.M., Del Valle, F., De La Mano, J.V., Álvarez-Galván, M.C. and Fierro,
J.L.G., 2009. Photocatalytic water splitting under visible light: concept and catalysts
development. Advances in chemical engineering, 36, pp.111-143.
https://doi.org/10.1016/S0065-2377(09)00404-9

[486] Wang, Z., Hisatomi, T., Li, R., Sayama, K., Liu, G., Domen, K., Li, C. and Wang, L.,
2021. Efficiency accreditation and testing protocols for particulate photocatalysts
toward solar fuel production. Joule, 5(2), pp.344-359.
https://doi.org/10.1016/j.joule.2021.01.001

[487] Sarina, S., Waclawik, E.R. and Zhu, H., 2013. Photocatalysis on supported gold and
silver nanoparticles under ultraviolet and visible light irradiation. Green
Chemistry, 15(7), pp.1814-1833. https://doi.org/10.1039/C3GC40450A

[488] Fajrina, N. and Tahir, M., 2019. A critical review in strategies to improve photocatalytic
water splitting towards hydrogen production. International Journal of Hydrogen
Energy, 44(2), pp.540-577. https://doi.org/10.1016/j.ijhydene.2018.10.200

[489] Al-Hamdi, A.M., Rinner, U. and Sillanpää, M., 2017. Tin dioxide as a photocatalyst for
water treatment: a review. Process Safety and Environmental Protection, 107, pp.190-
205. https://doi.org/10.1016/j.psep.2017.01.022

[490] Tahir, Z.R. and Asim, M., 2018. Surface measured solar radiation data and solar energy
resource assessment of Pakistan: A review. Renewable and Sustainable Energy
Reviews, 81, pp.2839-2861. https://doi.org/10.1007/978-1-4471-1742-1_2

[491] Hisatomi, T., Minegishi, T. and Domen, K., 2012. Kinetic assessment and numerical
modeling of photocatalytic water splitting toward efficient solar hydrogen
production. Bulletin of the Chemical Society of Japan, 85(6), pp.647-655.
https://doi.org/10.1246/bcsj.20120058

193
[492] Ye, S., Wang, R., Wu, M.Z. and Yuan, Y.P., 2015. A review on g-C3N4 for
photocatalytic water splitting and CO2 reduction. Applied Surface Science, 358, pp.15-
27. http://dx.doi.org/doi:10.1016/j.apsusc.2015.08.173

[493] Martín-Sómer, M., Pablos, C., van Grieken, R. and Marugán, J., 2017. Influence of
light distribution on the performance of photocatalytic reactors: LED vs mercury
lamps. Applied Catalysis B: Environmental, 215, pp.1-7.
https://doi.org/10.1016/j.apcatb.2017.05.048

[494] Pareek, V.K. and Adesina, A.A., 2004. Light intensity distribution in a photocatalytic
reactor using finite volume. AIChE journal, 50(6), pp.1273-1288.
https://doi.org/10.1002/aic.10107

[495] Martindale, B.C., Hutton, G.A., Caputo, C.A. and Reisner, E., 2015. Solar hydrogen
production using carbon quantum dots and a molecular nickel catalyst. Journal of the
American Chemical Society, 137(18), pp.6018-6025.
https://doi.org/10.1021/jacs.5b01650

[496] Jo, W.K. and Tayade, R.J., 2014. New generation energy-efficient light source for
photocatalysis: LEDs for environmental applications. Industrial & Engineering
Chemistry Research, 53(6), pp.2073-2084. https://doi.org/10.1021/ie404176g

[497] Li, Y., Zhang, D., Fan, J. and Xiang, Q., 2021. Highly crystalline carbon nitride hollow
spheres with enhanced photocatalytic performance. Chinese Journal of
Catalysis, 42(4), pp.627-636. https://doi.org/10.1016/S1872-2067(20)63684-1

[498] Navarro, R.M., Sanchez-Sanchez, M.C., Alvarez-Galvan, M.C., Del Valle, F. and
Fierro, J.L.G., 2009. Hydrogen production from renewable sources: biomass and
photocatalytic opportunities. Energy & Environmental Science, 2(1), pp.35-54.
https://doi.org/10.1039/B808138G

[499] Escobedo, S. and de Lasa, H., 2021. Synthesis and Performance of Photocatalysts for
Photocatalytic Hydrogen Production: Future Perspectives. Catalysts, 11(12), p.1505.
https://doi.org/10.3390/catal11121505

[500] Méndez, J.O., López, C.R., Melián, E.P., Díaz, O.G., Rodríguez, J.D., Hevia, D.F. and
Macías, M., 2014. Production of hydrogen by water photo-splitting over commercial
and synthesised Au/TiO2 catalysts. Applied Catalysis B: Environmental, 147, pp.439-
452. https://doi.org/10.1016/j.apcatb.2013.09.029

[501] Luo, H., Barrio, J., Sunny, N., Li, A., Steier, L., Shah, N., Stephens, I.E. and Titirici,
M.M., 2021. Progress and Perspectives in Photo‐and Electrochemical‐Oxidation of
Biomass for Sustainable Chemicals and Hydrogen Production. Advanced Energy
Materials, 11(43), p.2101180. https://doi.org/10.1002/aenm.202101180

[502] Wu, N.L. and Lee, M.S., 2004. Enhanced TiO2 photocatalysis by Cu in hydrogen
production from aqueous methanol solution. International Journal of Hydrogen
Energy, 29(15), pp.1601-1605. https://doi.org/10.1016/j.ijhydene.2004.02.013

[503] Pojanavaraphan, C., Luengnaruemitchai, A. and Gulari, E., 2012. Effect of support
composition and metal loading on Au catalyst activity in steam reforming of

194
methanol. International journal of hydrogen energy, 37(19), pp.14072-14084.
https://doi.org/10.1016/j.ijhydene.2012.06.107

[504] Navarrete, M., Cipagauta‐Díaz, S. and Gómez, R., 2019. Ga2O3/TiO2 semiconductors
free of noble metals for the photocatalytic hydrogen production in a water/methanol
mixture. Journal of Chemical Technology & Biotechnology, 94(11), pp.3457-3465.
https://doi.org/10.1002/jctb.5967

[505] Guayaquil-Sosa, J.F., Serrano-Rosales, B., Valadés-Pelayo, P.J. and de Lasa, H., 2017.
Photocatalytic hydrogen production using mesoporous TiO2 doped with Pt. Applied
Catalysis B: Environmental, 211, pp.337-348.
https://doi.org/10.1016/j.apcatb.2017.04.029

[506] Bashir, S., Wahab, A.K. and Idriss, H., 2015. Synergism and photocatalytic water
splitting to hydrogen over M/TiO2 catalysts: effect of initial particle size of
TiO2. Catalysis Today, 240, pp.242-247. https://doi.org/10.1016/j.cattod.2014.05.034

[507] Lin, Y., Jiang, Z., Zhu, C., Hu, X., Zhu, H., Zhang, X., Fan, J. and Lin, S.H., 2013. The
optical absorption and hydrogen production by water splitting of (Si, Fe)-codoped
anatase TiO2 photocatalyst. International journal of hydrogen energy, 38(13), pp.5209-
5214. https://doi.org/10.1016/j.ijhydene.2013.02.079

[508] Kuo, Y. and Klabunde, K.J., 2011. Hydrogen from ethanol solution under UV–visible
light. Photocatalysts produced by nitriding titanium nitride and indium oxide intimate
mixtures to form Ti–In nitride composites. Applied Catalysis B: Environmental, 104(3-
4), pp.245-251. https://doi.org/10.1016/j.apcatb.2011.03.025

[509] Strataki, N., Bekiari, V., Kondarides, D.I. and Lianos, P., 2007. Hydrogen production
by photocatalytic alcohol reforming employing highly efficient nanocrystalline titania
films. Applied Catalysis B: Environmental, 77(1-2), pp.184-189.
https://doi.org/10.1016/j.apcatb.2007.07.015

[510] Cui, W., Feng, L., Xu, C., Lü, S. and Qiu, F., 2004. Hydrogen production by
photocatalytic decomposition of methanol gas on Pt/TiO2 nano-film. Catalysis
communications, 5(9), pp.533-536. https://doi.org/10.1016/j.catcom.2004.06.011

[511] Rosseler, O., Shankar, M.V., Karkmaz-Le Du, M., Schmidlin, L., Keller, N. and Keller,
V., 2010. Solar light photocatalytic hydrogen production from water over Pt and
Au/TiO2 (anatase/rutile) photocatalysts: Influence of noble metal and porogen
promotion. Journal of Catalysis, 269(1), pp.179-190.
https://doi.org/10.1016/j.jcat.2009.11.006

[512] Meyer, K., Ranocchiari, M. and van Bokhoven, J.A., 2015. Metal organic frameworks
for photo-catalytic water splitting. Energy & Environmental Science, 8(7), pp.1923-
1937. https://doi.org/10.1002/solr.202100198

[513] Li, G., Liang, H., Xu, G., Li, C. and Bai, J., 2020. Controllable synthesized
heterojunction hollow nanotube of g-C3N4/CdS: enhance visible light catalytic
performance for hydrogen production. Journal of Physics and Chemistry of Solids, 145,
p.109549. https://doi.org/10.1016/j.jpcs.2020.109549

195
[514] Zha, Q., Li, M., Liu, Z. and Ni, Y., 2020. Hierarchical Co, Fe-MOF-74/Co/carbon cloth
hybrid electrode: simple construction and enhanced catalytic performance in full water
splitting. ACS sustainable chemistry & engineering, 8(32), pp.12025-12035.
https://doi.org/10.1021/acssuschemeng.0c02993

[515] Dubey, N., Rayalu, S.S., Labhsetwar, N.K. and Devotta, S., 2008. Visible light active
zeolite-based photocatalysts for hydrogen evolution from water. international journal
of hydrogen energy, 33(21), pp.5958-5966.
https://doi.org/10.1016/j.ijhydene.2008.05.095

[516] Enzweiler, H., Yassue-Cordeiro, P.H., Schwaab, M., Barbosa-Coutinho, E., Scaliante,
M.H.N.O. and Fernandes, N.R.C., 2020. Catalyst concentration, ethanol content and
initial pH effects on hydrogen production by photocatalytic water splitting. Journal of
photochemistry and photobiology A: Chemistry, 388, p.112051.
https://doi.org/10.1016/j.jphotochem.2019.112051

[517] Herrmann, J.M., 2010. Fundamentals and misconceptions in photocatalysis. Journal of


Photochemistry and Photobiology A: Chemistry, 216(2-3), pp.85-93.
https://doi.org/10.1016/j.jphotochem.2010.05.015

[518] Curcó, D., Giménez, J., Addardak, A., Cervera-March, S. and Esplugas, S., 2002.
Effects of radiation absorption and catalyst concentration on the photocatalytic
degradation of pollutants. Catalysis Today, 76(2-4), pp.177-188.
https://doi.org/10.1016/S0920-5861(02)00217-1

[519] Mu, F., Cai, Q., Hu, H., Wang, J., Wang, Y., Zhou, S. and Kong, Y., 2020. Construction
of 3D hierarchical microarchitectures of Z-scheme UiO-66-(COOH) 2/ZnIn2S4 hybrid
decorated with non-noble MoS2 cocatalyst: A highly efficient photocatalyst for
hydrogen evolution and Cr (VI) reduction. Chemical Engineering Journal, 384,
p.123352. https://doi.org/10.1016/j.cej.2019.123352

[520] Ijaz, M. and Zafar, M., 2021. Titanium dioxide nanostructures as efficient
photocatalyst: Progress, challenges and perspective. International Journal of Energy
Research, 45(3), pp.3569-3589. https://doi.org/10.1002/er.6079

[521] Sharma, S., Agarwal, S. and Jain, A., 2021. Significance of hydrogen as economic and
environmentally friendly fuel. Energies, 14(21), p.7389.
https://doi.org/10.3390/en14217389

[522] Zhang, K., Hu, H., Shi, L., Jia, B., Huang, H., Han, X., Sun, X. and Ma, T., 2021.
Strategies for Optimizing the Photocatalytic Water‐Splitting Performance of Metal–
Organic Framework‐Based Materials. Small Science, 1(12), p.2100060.
https://doi.org/10.1002/smsc.202100060

[523] Wang, Y.C., Liu, X.Y., Wang, X.X. and Cao, M.S., 2021. Metal-organic frameworks
based photocatalysts: Architecture strategies for efficient solar energy
conversion. Chemical Engineering Journal, 419, p.129459.
https://doi.org/10.1016/j.cej.2021.129459

[524] Chen, X., Xiao, S., Wang, H., Wang, W., Cai, Y., Li, G., Qiao, M., Zhu, J., Li, H.,
Zhang, D. and Lu, Y., 2020. MOFs conferred with transient metal centers for enhanced

196
photocatalytic activity. Angewandte Chemie International Edition, 59(39), pp.17182-
17186. https://doi.org/10.1002/anie.202002375

[525] Horiuchi, Y., Tatewaki, K., Mine, S., Kim, T.H., Lee, S.W. and Matsuoka, M., 2020.
Linker defect engineering for effective reactive site formation in metal–organic
framework photocatalysts with a MIL-125 (Ti) architecture. Journal of Catalysis, 392,
pp.119-125. https://doi.org/10.1016/j.jcat.2020.09.017

[526] Hendon, C.H., Tiana, D., Fontecave, M., Sanchez, C., D’arras, L., Sassoye, C., Rozes,
L., Mellot-Draznieks, C. and Walsh, A., 2013. Engineering the optical response of the
titanium-MIL-125 metal–organic framework through ligand functionalization. Journal
of the American Chemical Society, 135(30), pp.10942-10945.
https://doi.org/10.1021/ja405350u

[527] Wang, H., Yuan, X., Wu, Y., Zeng, G., Chen, X., Leng, L., Wu, Z., Jiang, L. and Li,
H., 2015. Facile synthesis of amino-functionalized titanium metal-organic frameworks
and their superior visible-light photocatalytic activity for Cr (VI) reduction. Journal of
hazardous materials, 286, pp.187-194. https://doi.org/10.1016/j.jhazmat.2014.11.039

[528] Chen, T.F., Han, S.Y., Wang, Z.P., Gao, H., Wang, L.Y., Deng, Y.H., Wan, C.Q., Tian,
Y., Wang, Q., Wang, G. and Li, G.S., 2019. Modified UiO-66 frameworks with
methylthio, thiol and sulfonic acid function groups: The structure and visible-light-
driven photocatalytic property study. Applied Catalysis B: Environmental, 259,
p.118047. https://doi.org/10.1016/j.apcatb.2019.118047

[529] Mohammadnezhad, F., Kampouri, S., Wolff, S.K., Xu, Y., Feyzi, M., Lee, J.H., Ji, X.
and Stylianou, K.C., 2021. Tuning the optoelectronic properties of hybrid
functionalized MIL-125-NH2 for photocatalytic hydrogen evolution. ACS Applied
Materials & Interfaces, 13(4), pp.5044-5051. https://doi.org/10.1021/acsami.0c19345

[530] Nazir, M.A., Bashir, M.S., Jamshaid, M., Anum, A., Najam, T., Shahzad, K., Imran,
M., Shah, S.S.A. and ur Rehman, A., 2021. Synthesis of porous secondary metal-doped
MOFs for removal of Rhodamine B from water: Role of secondary metal on efficiency
and kinetics. Surfaces and Interfaces, 25, p.101261.
https://doi.org/10.1016/j.surfin.2021.101261

[531] Melillo, A., Cabrero-Antonino, M., Navalón, S., Álvaro, M., Ferrer, B. and Garcia, H.,
2020. Enhancing visible-light photocatalytic activity for overall water splitting in UiO-
66 by controlling metal node composition. Applied Catalysis B: Environmental, 278,
p.119345. https://doi.org/10.1016/j.apcatb.2020.119345

[532] Yasin, A.S., Li, J., Wu, N. and Musho, T., 2016. Study of the inorganic substitution in
a functionalized UiO-66 metal–organic framework. Physical Chemistry Chemical
Physics, 18(18), pp.12748-12754. https://doi.org/10.1039/C5CP08070C

[533] Wu, M., Gong, Y., Nie, T., Zhang, J., Wang, R., Wang, H. and He, B., 2019. Template-
free synthesis of nanocage-like g-C 3 N 4 with high surface area and nitrogen defects for
enhanced photocatalytic H 2 activity. Journal of Materials Chemistry A, 7(10),
pp.5324-5332. https://doi.org/10.1039/C8TA12076E

197
[534] Maarisetty, D. and Baral, S.S., 2020. Defect engineering in photocatalysis: formation,
chemistry, optoelectronics, and interface studies. Journal of Materials Chemistry
A, 8(36), pp.18560-18604. https://doi.org/10.1039/D0TA04297H

[535] Wang, Z.D., Zang, Y., Liu, Z.J., Peng, P., Wang, R. and Zang, S.Q., 2021. Opening
catalytic sites in the copper-triazoles framework via defect chemistry for switching on
the proton reduction. Applied Catalysis B: Environmental, 288, p.119941.
https://doi.org/10.1016/j.apcatb.2021.119941

[536] Zhou, Y., Zhang, Y., Lin, M., Long, J., Zhang, Z., Lin, H., Wu, J.C.S. and Wang, X.,
2015. Monolayered Bi2WO6 nanosheets mimicking heterojunction interface with open
surfaces for photocatalysis. Nature communications, 6(1), pp.1-8.
https://doi.org/10.1038/ncomms9340

[537] Zhang, C., Zhou, Y., Bao, J., Fang, J., Zhao, S., Zhang, Y., Sheng, X. and Chen, W.,
2018. Structure regulation of ZnS@ g-C3N4/TiO2 nanospheres for efficient
photocatalytic H2 production under visible-light irradiation. Chemical Engineering
Journal, 346, pp.226-237. https://doi.org/10.1016/j.cej.2018.04.038

[538] Yang, Y., Zeng, Z., Zhang, C., Huang, D., Zeng, G., Xiao, R., Lai, C., Zhou, C., Guo,
H., Xue, W. and Cheng, M., 2018. Construction of iodine vacancy-rich BiOI/Ag@ AgI
Z-scheme heterojunction photocatalysts for visible-light-driven tetracycline
degradation: transformation pathways and mechanism insight. Chemical Engineering
Journal, 349, pp.808-821. https://doi.org/10.1016/j.cej.2018.05.093

[539] Zhou, C., Lai, C., Huang, D., Zeng, G., Zhang, C., Cheng, M., Hu, L., Wan, J., Xiong,
W., Wen, M. and Wen, X., 2018. Highly porous carbon nitride by supramolecular
preassembly of monomers for photocatalytic removal of sulfamethazine under visible
light driven. Applied Catalysis B: Environmental, 220, pp.202-210.
https://doi.org/10.1016/j.apcatb.2017.08.055

[540] Yang, Y., Zhang, C., Huang, D., Zeng, G., Huang, J., Lai, C., Zhou, C., Wang, W.,
Guo, H., Xue, W. and Deng, R., 2019. Boron nitride quantum dots decorated ultrathin
porous g-C3N4: intensified exciton dissociation and charge transfer for promoting
visible-light-driven molecular oxygen activation. Applied Catalysis B:
Environmental, 245, pp.87-99. https://doi.org/10.1016/j.apcatb.2018.12.049

[541] Wang, R., Gu, L., Zhou, J., Liu, X., Teng, F., Li, C., Shen, Y. and Yuan, Y., 2015.
Quasi‐polymeric metal–organic framework UiO‐66/g‐C3N4 heterojunctions for
enhanced photocatalytic hydrogen evolution under visible light irradiation. Advanced
Materials Interfaces, 2(10), p.1500037. https://doi.org/10.1002/admi.201500037

[542] Karthik, P., Balaraman, E. and Neppolian, B., 2018. Efficient solar light-driven H2
production: post-synthetic encapsulation of a Cu2O co-catalyst in a metal–organic
framework (MOF) for boosting the effective charge carrier separation. Catalysis
Science & Technology, 8(13), pp.3286-3294 https://doi.org/10.1039/C8CY00604K

[543] Zhang, M., Shang, Q., Wan, Y., Cheng, Q., Liao, G. and Pan, Z., 2019. Self-template
synthesis of double-shell TiO2@ ZIF-8 hollow nanospheres via sonocrystallization with
enhanced photocatalytic activities in hydrogen generation. Applied Catalysis B:
Environmental, 241, pp.149-158. https://doi.org/10.1016/j.apcatb.2018.09.036

198
[544] Zhang, L., Zhang, J., Zeng, G., Dong, H., Chen, Y., Huang, C., Zhu, Y., Xu, R., Cheng,
Y., Hou, K. and Cao, W., 2018. Multivariate relationships between microbial
communities and environmental variables during co-composting of sewage sludge and
agricultural waste in the presence of PVP-AgNPs. Bioresource technology, 261, pp.10-
18. https://doi.org/10.1016/j.biortech.2018.03.089

[545] Liu, Y., Liu, Z., Huang, D., Cheng, M., Zeng, G., Lai, C., Zhang, C., Zhou, C., Wang,
W., Jiang, D. and Wang, H., 2019. Metal or metal-containing nanoparticle@ MOF
nanocomposites as a promising type of photocatalyst. Coordination Chemistry
Reviews, 388, pp.63-78. https://doi.org/10.1016/j.ccr.2019.02.031

[546] Guo, W., Lv, H., Chen, Z., Sullivan, K.P., Lauinger, S.M., Chi, Y., Sumliner, J.M.,
Lian, T. and Hill, C.L., 2016. Self-assembly of polyoxometalates, Pt nanoparticles and
metal–organic frameworks into a hybrid material for synergistic hydrogen
evolution. Journal of Materials Chemistry A, 4(16), pp.5952-5957.
https://doi.org/10.1039/C6TA00011H

[547] Tian, J., Xu, Z.Y., Zhang, D.W., Wang, H., Xie, S.H., Xu, D.W., Ren, Y.H., Wang, H.,
Liu, Y. and Li, Z.T., 2016. Supramolecular metal-organic frameworks that display high
homogeneous and heterogeneous photocatalytic activity for H2 production. Nature
communications, 7(1), pp.1-9. https://doi.org/10.1038/ncomms11580

[548] Han, J., Wang, D., Du, Y., Xi, S., Chen, Z., Yin, S., Zhou, T. and Xu, R., 2016.
Polyoxometalate immobilized in MIL-101 (Cr) as an efficient catalyst for water
oxidation. Applied Catalysis A: General, 521, pp.83-89.
https://doi.org/10.1016/j.apcata.2015.10.015

[549] Shah, W.A., Waseem, A., Nadeem, M.A. and Kögerler, P., 2018. Leaching-free
encapsulation of cobalt-polyoxotungstates in MIL-100 (Fe) for highly reproducible
photocatalytic water oxidation. Applied Catalysis A: General, 567, pp.132-138.
https://doi.org/10.1016/j.apcata.2018.08.002

[550] Paille, G., Gomez-Mingot, M., Roch-Marchal, C., Lassalle-Kaiser, B., Mialane, P.,
Fontecave, M., Mellot-Draznieks, C. and Dolbecq, A., 2018. A fully noble metal-free
photosystem based on cobalt-polyoxometalates immobilized in a porphyrinic metal–
organic framework for water oxidation. Journal of the American Chemical
Society, 140(10), pp.3613-3618. https://doi.org/10.1021/jacs.7b11788

[551] El-Bery, H.M. and Abdelhamid, H.N., 2021. Photocatalytic hydrogen generation via
water splitting using ZIF-67 derived Co3O4@ C/TiO2. Journal of Environmental
Chemical Engineering, 9(4), p.105702. https://doi.org/10.1016/j.jece.2021.105702

[552] Su, Y., Li, S., He, D., Yu, D., Liu, F., Shao, N. and Zhang, Z., 2018. MOF-derived
porous ZnO nanocages/rGO/carbon sponge-based photocatalytic microreactor for
efficient degradation of water pollutants and hydrogen evolution. ACS Sustainable
Chemistry & Engineering, 6(9), pp.11989-11998.
https://doi.org/10.1021/acssuschemeng.8b02287

199
[553] Xu, J., Qi, Y., Wang, C. and Wang, L., 2019. NH2-MIL-101 (Fe)/Ni (OH)2-derived C,
N-codoped Fe2P/Ni2P cocatalyst modified g-C3N4 for enhanced photocatalytic
hydrogen evolution from water splitting. Applied Catalysis B: Environmental, 241,
pp.178-186. https://doi.org/10.1016/j.apcatb.2018.09.035

[554] Li, R., Wu, S., Wan, X., Xu, H. and Xiong, Y., 2016. Cu/TiO2 octahedral-shell
photocatalysts derived from metal–organic framework@ semiconductor hybrid
structures. Inorganic Chemistry Frontiers, 3(1), pp.104-110.
https://doi.org/10.1039/C5QI00205B

[555] Nasalevich, M.A., Becker, R., Ramos-Fernandez, E.V., Castellanos, S., Veber, S.L.,
Fedin, M.V., Kapteijn, F., Reek, J.N.H., Van Der Vlugt, J.I. and Gascon, J., 2015. Co@
NH 2-MIL-125 (Ti): cobaloxime-derived metal–organic framework-based composite
for light-driven H2 production. Energy & Environmental Science, 8(1), pp.364-375.
https://doi.org/10.1039/C4EE02853H
[556] Toyao, T., Saito, M., Horiuchi, Y., Mochizuki, K., Iwata, M., Higashimura, H. and
Matsuoka, M., 2013. Efficient hydrogen production and photocatalytic reduction of
nitrobenzene over a visible-light-responsive metal–organic framework
photocatalyst. Catalysis Science & Technology, 3(8), pp.2092-2097.
https://doi.org/10.1039/C3CY00211J

[557] Toyao, T., Saito, M., Dohshi, S., Mochizuki, K., Iwata, M., Higashimura, H., Horiuchi,
Y. and Matsuoka, M., 2014. Development of a Ru complex-incorporated MOF
photocatalyst for hydrogen production under visible-light irradiation. Chemical
Communications, 50(51), pp.6779-6781. https://doi.org/10.1039/C4CC02397H

[558] Xu, J., Gao, J., Wang, C., Yang, Y. and Wang, L., 2017. NH2-MIL-125 (Ti)/graphitic
carbon nitride heterostructure decorated with NiPd co-catalysts for efficient
photocatalytic hydrogen production. Applied Catalysis B: Environmental, 219, pp.101-
108. https://doi.org/10.1016/j.apcatb.2017.07.046

[559] Cao, A., Zhang, L., Wang, Y., Zhao, H., Deng, H., Liu, X., Lin, Z., Su, X. and Yue, F.,
2018. 2D–2D heterostructured UNiMOF/g-C3N4 for enhanced photocatalytic H2
production under visible-light irradiation. ACS Sustainable Chemistry &
Engineering, 7(2), pp.2492-2499. https://doi.org/10.1021/acssuschemeng.8b05396
[560] Wang, Y., Zhang, Y., Jiang, Z., Jiang, G., Zhao, Z., Wu, Q., Liu, Y., Xu, Q., Duan, A.
and Xu, C., 2016. Controlled fabrication and enhanced visible-light photocatalytic
hydrogen production of Au@ CdS/MIL-101 heterostructure. Applied Catalysis B:
Environmental, 185, pp.307-314. https://doi.org/10.1016/j.apcatb.2015.12.020

[561] Zhou, J.J., Wang, R., Liu, X.L., Peng, F.M., Li, C.H., Teng, F. and Yuan, Y.P., 2015.
In situ growth of CdS nanoparticles on UiO-66 metal-organic framework octahedrons
for enhanced photocatalytic hydrogen production under visible light
irradiation. Applied Surface Science, 346, pp.278-283.
https://doi.org/10.1016/j.apsusc.2015.03.210
[562] Zhang, B., Zhang, J., Tan, X., Shao, D., Shi, J., Zheng, L., Zhang, J., Yang, G. and Han,
B., 2018. MIL-125-NH2@ TiO2 core–shell particles produced by a post-solvothermal
route for high-performance photocatalytic H2 production. ACS applied materials &
interfaces, 10(19), pp.16418-16423. https://doi.org/10.1021/acsami.8b01462

200
[563] Tilgner, D. and Kempe, R., 2017. A Plasmonic Colloidal Photocatalyst Composed of a
Metal–Organic Framework Core and a Gold/Anatase Shell for Visible‐Light‐Driven
Wastewater Purification from Antibiotics and Hydrogen Evolution. Chemistry–A
European Journal, 23(13), pp.3184-3190. https://doi.org/10.1002/chem.201605473
[564] Su, Y., Zhang, Z., Liu, H. and Wang, Y., 2017. Cd0. 2Zn0. 8S@ UiO-66-NH2
nanocomposites as efficient and stable visible-light-driven photocatalyst for H2
evolution and CO2 reduction. Applied Catalysis B: Environmental, 200, pp.448-457.
https://doi.org/10.1016/j.apcatb.2016.07.032
[565] Liu, H., Zhang, J. and Ao, D., 2018. Construction of heterostructured ZnIn 2S4@ NH2-
MIL-125 (Ti) nanocomposites for visible-light-driven H2 production. Applied Catalysis
B: Environmental, 221, pp.433-442. https://doi.org/10.1016/j.apcatb.2017.09.043

[566] Bag, P.P., Wang, X.S., Sahoo, P., Xiong, J. and Cao, R., 2017. Efficient photocatalytic
hydrogen evolution under visible light by ternary composite CdS@ NU-
1000/RGO. Catalysis Science & Technology, 7(21), pp.5113-5119.
https://doi.org/10.1039/C7CY01254C

[567] Bai, C., Bi, J., Wu, J., Meng, H., Xu, Y., Han, Y. and Zhang, X., 2018. Fabrication of
noble‐metal‐free g‐C3N4‐MIL‐53 (Fe) composite for enhanced photocatalytic H2‐
generation performance. Applied Organometallic Chemistry, 32(12), p.e4597.
https://doi.org/10.1002/aoc.4597

[568] Wang, Z., Jin, Z., Yuan, H., Wang, G. and Ma, B., 2018. Orderly-designed Ni2P
nanoparticles on g-C3N4 and UiO-66 for efficient solar water splitting. Journal of
colloid and interface science, 532, pp.287-299.
https://doi.org/10.1016/j.jcis.2018.07.138
[569] Liang, Q., Jin, J., Liu, C., Xu, S., Yao, C. and Li, Z., 2018. Fabrication of the ternary
heterojunction Cd 0.5 Zn 0.5 S@ UIO-66@ g-C3N4 for enhanced visible-light
photocatalytic hydrogen evolution and degradation of organic pollutants. Inorganic
chemistry frontiers, 5(2), pp.335-343. https://doi.org/10.1039/C7QI00638A
[570] Pan, Y., Li, D. and Jiang, H.L., 2018. Sodium‐Doped C3N4/MOF Heterojunction
Composites with Tunable Band Structures for Photocatalysis: Interplay between Light
Harvesting and Electron Transfer. Chemistry–A European Journal, 24(69), pp.18403-
18407. https://doi.org/10.1002/chem.201803555
[571] Zhao, X., Feng, J., Liu, J., Shi, W., Yang, G., Wang, G.C. and Cheng, P., 2018. An
Efficient, Visible‐Light‐Driven, Hydrogen Evolution Catalyst NiS/ZnxCd1− xS
Nanocrystal Derived from a Metal–Organic Framework. Angewandte
Chemie, 130(31), pp.9938-9942. https://doi.org/10.1002/anie.201805425

[572] Reddy, D.A., Kim, E.H., Gopannagari, M., Ma, R., Bhavani, P., Kumar, D.P. and Kim,
T.K., 2018. Enhanced photocatalytic hydrogen evolution by integrating dual co-
catalysts on heterophase CdS nano-junctions. ACS Sustainable Chemistry &
Engineering, 6(10), pp.12835-12844. https://doi.org/10.1021/acssuschemeng.8b02098
[573] Reddy, D.A., Park, H., Gopannagari, M., Kim, E.H., Lee, S., Kumar, D.P. and Kim,
T.K., 2018. Designing CdS mesoporous networks on Co‐C@ Co9S8 double‐shelled
nanocages as redox‐mediator‐free Z‐scheme photocatalyst. ChemSusChem, 11(1),
pp.245-253. https://doi.org/10.1002/cssc.201701643
[574] Reddy, D.A., Kim, H.K., Kim, Y., Lee, S., Choi, J., Islam, M.J., Kumar, D.P. and Kim,
T.K., 2016. Multicomponent transition metal phosphides derived from layered double
hydroxide double-shelled nanocages as an efficient non-precious co-catalyst for

201
hydrogen production. Journal of Materials Chemistry A, 4(36), pp.13890-13898.
https://doi.org/10.1039/C6TA05741A
[575] Majeed, I., Nadeem, M.A., Badshah, A., Kanodarwala, F.K., Ali, H., Khan, M.A.,
Stride, J.A. and Nadeem, M.A., 2017. Titania supported MOF-199 derived Cu–Cu2O
nanoparticles: highly efficient non-noble metal photocatalysts for hydrogen production
from alcohol–water mixtures. Catalysis Science & Technology, 7(3), pp.677-686.
https://doi.org/10.1039/C6CY02328B
[576] Mondal, I. and Pal, U., 2016. Synthesis of MOF templated Cu/CuO@ TiO2
nanocomposites for synergistic hydrogen production. Physical Chemistry Chemical
Physics, 18(6), pp.4780-4788. https://doi.org/10.1039/C5CP06292F
[577] Yan, B., Zhang, L., Tang, Z., Al-Mamun, M., Zhao, H. and Su, X., 2017. Palladium-
decorated hierarchical titania constructed from the metal-organic frameworks NH2-
MIL-125 (Ti) as a robust photocatalyst for hydrogen evolution. Applied Catalysis B:
Environmental, 218, pp.743-750. https://doi.org/10.1016/j.apcatb.2017.07.020
[578] Chen, H., Gu, Z.G., Mirza, S., Zhang, S.H. and Zhang, J., 2018. Hollow Cu–TiO2/C
nanospheres derived from a Ti precursor encapsulated MOF coating for efficient
photocatalytic hydrogen evolution. Journal of Materials Chemistry A, 6(16), pp.7175-
7181. https://doi.org/10.1039/C8TA01034J
[579] Xu, J., Gao, J., Qi, Y., Wang, C. and Wang, L., 2018. Anchoring Ni2P on the UiO‐66‐
NH2/g‐C3N4‐derived C‐doped ZrO2/g‐C3N4 Heterostructure: Highly Efficient
Photocatalysts for H2 Production from Water Splitting. ChemCatChem, 10(15),
pp.3327-3335. https://doi.org/10.1002/cctc.201800353
[580] Xu, J.Y., Zhai, X.P., Gao, L.F., Chen, P., Zhao, M., Yang, H.B., Cao, D.F., Wang, Q.
and Zhang, H.L., 2016. In situ preparation of a MOF-derived magnetic carbonaceous
catalyst for visible-light-driven hydrogen evolution. RSC advances, 6(3), pp.2011-
2018. https://doi.org/10.1039/C5RA23838B
[581] Sun, X., Yu, Q., Zhang, F., Wei, J. and Yang, P., 2016. A dye-like ligand-based metal–
organic framework for efficient photocatalytic hydrogen production from aqueous
solution. Catalysis Science & Technology, 6(11), pp.3840-3844.
https://doi.org/10.1039/C5CY01716E
[582] An, Y., Liu, Y., An, P., Dong, J., Xu, B., Dai, Y., Qin, X., Zhang, X., Whangbo, M.H.
and Huang, B., 2017. NiII Coordination to an Al‐Based Metal–Organic Framework
Made from 2‐Aminoterephthalate for Photocatalytic Overall Water
Splitting. Angewandte Chemie, 129(11), pp.3082-3086.
https://doi.org/10.1002/anie.201612423
[583] Xiao, Z., Sun, Y., Bao, Y., Sun, Y., Zhou, R. and Wang, L., 2019. Two new inorganic–
organic hybrid zinc phosphate frameworks and their application in fluorescence sensor
and photocatalytic hydrogen evolution. Journal of Solid State Chemistry, 269, pp.575-
579. https://doi.org/10.1016/j.jssc.2018.10.038

[584] Zhen, W., Gao, H., Tian, B., Ma, J. and Lu, G., 2016. Fabrication of low adsorption
energy Ni–Mo cluster cocatalyst in metal–organic frameworks for visible
photocatalytic hydrogen evolution. ACS Applied Materials & Interfaces, 8(17),
pp.10808-10819 https://doi.org/10.1021/acsami.5b12524
[585] Iizuka, T., Honjo, K. and Uemura, T., 2019. Enhanced mechanical properties of a
metal–organic framework by polymer insertion. Chemical Communications, 55(5),
pp.691-694. https://doi.org/10.1039/C8CC08922A

[586] Sun, Y., Hu, Z., Zhao, D. and Zeng, K., 2017. Mechanical properties of
microcrystalline metal–organic frameworks (MOFs) measured by bimodal amplitude

202
modulated-frequency modulated atomic force microscopy. ACS applied materials &
interfaces, 9(37), pp.32202-32210.https://doi.org/10.1021/acsami.7b06809
[587] Yuan, S., Feng, L., Wang, K., Pang, J., Bosch, M., Lollar, C., Sun, Y., Qin, J., Yang,
X., Zhang, P. and Wang, Q., 2018. Stable metal–organic frameworks: design, synthesis,
and applications. Advanced Materials, 30(37),
p.1704303.https://doi.org/10.1002/adma.201704303
[588] Moghadam, P.Z., Rogge, S.M., Li, A., Chow, C.M., Wieme, J., Moharrami, N.,
Aragones-Anglada, M., Conduit, G., Gomez-Gualdron, D.A., Van Speybroeck, V. and
Fairen-Jimenez, D., 2019. Structure-mechanical stability relations of metal-organic
frameworks via machine learning. Matter, 1(1), pp.219-
234.https://doi.org/10.1016/j.matt.2019.03.002
[589] Rieth, A.J., Wright, A.M. and Dincă, M., 2019. Kinetic stability of metal–organic
frameworks for corrosive and coordinating gas capture. Nature Reviews
Materials, 4(11), pp.708-725.https://doi.org/10.1038/s41578-019-0140-1
[590] Burtch, N.C., Jasuja, H. and Walton, K.S., 2014. Water stability and adsorption in
metal–organic frameworks. Chemical reviews, 114(20), pp.10575-
10612.https://doi.org/10.1021/cr5002589
[591] Volkringer, C., Falaise, C., Devaux, P., Giovine, R., Stevenson, V., Pourpoint, F.,
Lafon, O., Osmond, M., Jeanjacques, C., Marcillaud, B. and Sabroux, J.C., 2016.
Stability of metal–organic frameworks under gamma irradiation. Chemical
Communications, 52(84), pp.12502-12505.https://doi.org/10.1039/C6CC06878B
[592] Healy, C., Patil, K.M., Wilson, B.H., Hermanspahn, L., Harvey-Reid, N.C., Howard,
B.I., Kleinjan, C., Kolien, J., Payet, F., Telfer, S.G. and Kruger, P.E., 2020. The thermal
stability of metal-organic frameworks. Coordination Chemistry Reviews, 419,
p.213388.https://doi.org/10.1016/j.ccr.2020.213388
[593] Bosch, M., Zhang, M. and Zhou, H.C., 2014. Increasing the stability of metal-organic
frameworks. Adv. Chem, 2014(182327.10),
p.1155.http://dx.doi.org/10.1155/2014/182327
[594] DeStefano, M.R., Islamoglu, T., Garibay, S.J., Hupp, J.T. and Farha, O.K., 2017.
Room-temperature synthesis of UiO-66 and thermal modulation of densities of defect
sites. Chemistry of Materials, 29(3), pp.1357-
1361.https://doi.org/10.1021/acs.chemmater.6b05115
[595] Li, H., Shi, W., Zhao, K., Li, H., Bing, Y. and Cheng, P., 2012. Enhanced hydrostability
in Ni-doped MOF-5. Inorganic Chemistry, 51(17), pp.9200-
9207.https://doi.org/10.1021/ic3002898
[596] Zhai, Q.G., Bu, X., Zhao, X., Li, D.S. and Feng, P., 2017. Pore space partition in metal–
organic frameworks. Accounts of Chemical Research, 50(2), pp.407-
417.https://doi.org/10.1021/acs.accounts.6b00526
[597] Jiang, H.L., Makal, T.A. and Zhou, H.C., 2013. Interpenetration control in metal–
organic frameworks for functional applications. Coordination Chemistry
Reviews, 257(15-16), pp.2232-2249.https://doi.org/10.1016/j.ccr.2013.03.017
[598] Feng, R., Jia, Y.Y., Li, Z.Y., Chang, Z. and Bu, X.H., 2018. Enhancing the stability and
porosity of penetrated metal–organic frameworks through the insertion of coordination
sites. Chemical Science, 9(4), pp.950-955.https://doi.org/10.1039/C7SC04192F
[599] Cohen, S.M., 2012. Postsynthetic methods for the functionalization of metal–organic
frameworks. Chemical reviews, 112(2), pp.970-
1000.https://doi.org/10.1021/cr200179u
[600] Castells-Gil, J., Novio, F., Padial, N.M., Tatay, S., Ruiz-Molina, D. and Marti-
Gastaldo, C., 2017. Surface functionalization of metal–organic framework crystals with

203
catechol coatings for enhanced moisture tolerance. ACS applied materials &
interfaces, 9(51), pp.44641-44648.https://doi.org/10.1021/acsami.7b15564
[601] Hou, L., Wang, L., Zhang, N., Xie, Z. and Dong, D., 2016. Polymer brushes on metal–
organic frameworks by UV-induced photopolymerization. Polymer Chemistry, 7(37),
pp.5828-5834.https://doi.org/10.1039/C6PY01008C
[602] Dodson, R.A., Wong-Foy, A.G. and Matzger, A.J., 2018. The metal–organic
framework collapse continuum: insights from two-dimensional powder X-ray
diffraction. Chemistry of Materials, 30(18), pp.6559-
6565.https://doi.org/10.1021/acs.chemmater.8b03378
[603] He, X., Lei, L. and Dai, Z., 2021. Green hydrogen enrichment with carbon membrane
processes: Techno-economic feasibility and sensitivity analysis. Separation and
Purification Technology, 276, p.119346. https://doi.org/10.1016/j.seppur.2021.119346
[604] Tesfaye, T., Ayele, M., Ferede, E., Gibril, M., Kong, F. and Sithole, B., 2021. A techno-
economic feasibility of a process for extraction of starch from waste avocado
seeds. Clean Technologies and Environmental Policy, 23(2), pp.581-595.
https://doi.org/10.1007/s10098-020-01981-1
[605] De Jong, S., Hoefnagels, R., Faaij, A., Slade, R., Mawhood, R. and Junginger, M., 2015.
The feasibility of short‐term production strategies for renewable jet fuels–a
comprehensive techno‐economic comparison. Biofuels, Bioproducts and
Biorefining, 9(6), pp.778-800. https://doi.org/10.1002/bbb.1613
[606] Qureshi, F., Yusuf, M., Kamyab, H., Vo, D.V.N., Chelliapan, S., Joo, S.W. and
Vasseghian, Y., 2022. Latest eco-friendly avenues on hydrogen production towards a
circular bioeconomy: Currents challenges, innovative insights, and future
perspectives. Renewable and Sustainable Energy Reviews, 168, p.112916.
https://doi.org/10.1016/j.rser.2022.112916
[607] Levene, J.I., Mann, M.K., Margolis, R.M. and Milbrandt, A., 2007. An analysis of
hydrogen production from renewable electricity sources. Solar energy, 81(6), pp.773-
780. https://doi.org/10.1016/j.solener.2006.10.005
[608] Rad, M.A.V., Ghasempour, R., Rahdan, P., Mousavi, S. and Arastounia, M., 2020.
Techno-economic analysis of a hybrid power system based on the cost-effective
hydrogen production method for rural electrification, a case study in Iran. Energy, 190,
p.116421. https://doi.org/10.1016/j.energy.2019.116421
[609] Abbott, D., 2009. Keeping the energy debate clean: How do we supply the world's
energy needs? Proceedings of the IEEE, 98(1), pp.42-66.
https://doi.org/10.1109/JPROC.2009.2035162
[610] Ng, K.H., Lai, S.Y., Cheng, C.K., Cheng, Y.W. and Chong, C.C., 2021. Photocatalytic
water splitting for solving energy crisis: Myth, Fact or Busted?. Chemical Engineering
Journal, 417, p.128847. https://doi.org/10.1016/j.cej.2021.128847
[611] Ringsmuth, A.K., Landsberg, M.J. and Hankamer, B., 2016. Can photosynthesis enable
a global transition from fossil fuels to solar fuels, to mitigate climate change and fuel-
supply limitations?. Renewable and Sustainable Energy Reviews, 62, pp.134-163.
https://doi.org/10.1016/j.rser.2016.04.016
[612] Doğan, B. and Erol, D., 2019, October. The future of fossil and alternative fuels used
in automotive industry. In 2019 3rd International Symposium on Multidisciplinary
Studies and Innovative Technologies (ISMSIT) (pp. 1-8). IEEE.
https://doi.org/10.1109/ISMSIT.2019.8932925
[613] Pinaud, B.A., Benck, J.D., Seitz, L.C., Forman, A.J., Chen, Z., Deutsch, T.G., James,
B.D., Baum, K.N., Baum, G.N., Ardo, S. and Wang, H., 2013. Technical and economic
feasibility of centralized facilities for solar hydrogen production via photocatalysis and

204
photoelectrochemistry. Energy & Environmental Science, 6(7), pp.1983-2002.
https://doi.org/10.1039/C3EE40831K
[614] Parra, D., Valverde, L., Pino, F.J. and Patel, M.K., 2019. A review on the role, cost and
value of hydrogen energy systems for deep decarbonisation. Renewable and
Sustainable Energy Reviews, 101, pp.279-294.
https://doi.org/10.1016/j.rser.2018.11.010
[615] Abe, J.O., Popoola, A.P.I., Ajenifuja, E. and Popoola, O.M., 2019. Hydrogen energy,
economy and storage: review and recommendation. International journal of hydrogen
energy, 44(29), pp.15072-15086. https://doi.org/10.1016/j.ijhydene.2019.04.068
[616] Staffell, I., Scamman, D., Abad, A.V., Balcombe, P., Dodds, P.E., Ekins, P., Shah, N.
and Ward, K.R., 2019. The role of hydrogen and fuel cells in the global energy
system. Energy & Environmental Science, 12(2), pp.463-491.
https://doi.org/10.1039/c8ee01157e
[617] Dincer, I. and Acar, C., 2016. A review on potential use of hydrogen in aviation
applications. International Journal of Sustainable Aviation, 2(1), pp.74-100.
https://doi.org/10.1504/ijsa.2016.076077
[618] Uyar, T.S. and Beşikci, D., 2017. Integration of hydrogen energy systems into
renewable energy systems for better design of 100% renewable energy
communities. International Journal of Hydrogen Energy, 42(4), pp.2453-2456.
https://doi.org/10.1016/j.ijhydene.2016.09.086
[619] Council, H., 2020. Path to hydrogen competitiveness: a cost perspective.
http://refhub.elsevier.com/S0360-3199(21)04729-7/sref38
[620] Council, H., 2017. Hydrogen scaling up: A sustainable pathway for the global energy
transition. http://refhub.elsevier.com/S0360-3199(21)04729-7/sref39
[621] Lund, P.D., Lindgren, J., Mikkola, J. and Salpakari, J., 2015. Review of energy system
flexibility measures to enable high levels of variable renewable electricity. Renewable
and sustainable energy reviews, 45, pp.785-807.
https://doi.org/10.1016/j.rser.2015.01.057
[622] Loisel, R., Baranger, L., Chemouri, N., Spinu, S. and Pardo, S., 2015. Economic
evaluation of hybrid off-shore wind power and hydrogen storage system. International
Journal of hydrogen energy, 40(21), pp.6727-6739.
https://doi.org/10.1016/j.ijhydene.2015.03.117
[623] Dawood, F., Anda, M. and Shafiullah, G.M., 2020. Hydrogen production for energy:
An overview. International Journal of Hydrogen Energy, 45(7), pp.3847-3869.
https://doi.org/10.1016/j.ijhydene.2019.12.059
[624] Chi, J. and Yu, H., 2018. Water electrolysis based on renewable energy for hydrogen
production. Chinese Journal of Catalysis, 39(3), pp.390-394.
https://doi.org/10.1016/S1872-2067(17)62949-8
[625] James, B.D., Baum, G.N., Perez, J. and Baum, K.N., 2009. Technoeconomic analysis
of photoelectrochemical (PEC) hydrogen production. DOE report.
[626] Kanto, T., 2022. Renewable hydrogen storage and supply options for large-scale
industrial users in Finland (Doctoral dissertation, University of Oulu).
http://jultika.oulu.fi/files/nbnfioulu-202207123249.pdf
[627] Bhandari, R. and Shah, R.R., 2021. Hydrogen as energy carrier: Techno-economic
assessment of decentralized hydrogen production in Germany. Renewable Energy, 177,
pp.915-931. https://doi.org/10.1016/j.renene.2021.05.149
[628] Abdin, Z. and Mérida, W., 2019. Hybrid energy systems for off-grid power supply and
hydrogen production based on renewable energy: A techno-economic analysis. Energy
Conversion and management, 196, pp.1068-1079.
https://doi.org/10.1016/j.enconman.2019.06.068

205
[629] Frowijn, L.S. and van Sark, W.G., 2021. Analysis of photon-driven solar-to-hydrogen
production methods in the Netherlands. Sustainable Energy Technologies and
Assessments, 48, p.101631. https://doi.org/10.1016/j.seta.2021.101631
[630] DeSantis, D., Mason, J.A., James, B.D., Houchins, C., Long, J.R. and Veenstra, M.,
2017. Techno-economic analysis of metal–organic frameworks for hydrogen and
natural gas storage. Energy & Fuels, 31(2), pp.2024-2032.DOI:
10.1021/acs.energyfuels.6b02510
[631] Luo, H., Cheng, F., Huelsenbeck, L. and Smith, N., 2021. Comparison between
conventional solvothermal and aqueous solution-based production of UiO-66-NH2:
Life cycle assessment, techno-economic assessment, and implications for CO2 capture
and storage. Journal of Environmental Chemical Engineering, 9(2), p.105159.DOI:
10.1016/j.jece.2021.105159
[632] Corredor, J., Rivero, M.J., Rangel, C.M., Gloaguen, F. and Ortiz, I., 2019.
Comprehensive review and future perspectives on the photocatalytic hydrogen
production. Journal of Chemical Technology & Biotechnology, 94(10), pp.3049-
3063.https://doi.org/10.1002/jctb.6123

[633] Lee, J., Farha, O.K., Roberts, J., Scheidt, K.A., Nguyen, S.T. and Hupp, J.T., 2009.
Metal–organic framework materials as catalysts. Chemical Society Reviews, 38(5),
pp.1450-1459. https://doi.org/10.1039/B807080F
[634] Ding, M., Cai, X. and Jiang, H.L., 2019. Improving MOF stability: approaches and
applications. Chemical Science, 10(44), pp.10209-10230.
https://doi.org/10.1039/C9SC03916C
[635] Feng, L., Wang, K.Y., Day, G.S., Ryder, M.R. and Zhou, H.C., 2020. Destruction of
metal–organic frameworks: Positive and negative aspects of stability and
lability. Chemical Reviews, 120(23), pp.13087-13133.
https://doi.org/10.1021/acs.chemrev.0c00722
[636] Kudo, A. and Miseki, Y., 2009. Heterogeneous photocatalyst materials for water
splitting. Chemical Society Reviews, 38(1), pp.253-
278.https://doi.org/10.1039/B800489G
[637] Cao, S., Piao, L. and Chen, X., 2020. Emerging photocatalysts for hydrogen
evolution. Trends in Chemistry, 2(1), pp.57-70.
https://doi.org/10.1016/j.trechm.2019.06.009
[638] Zhang, F., Yao, H., Zhao, Y., Li, X., Zhang, G. and Yang, Y., 2017. Mixed matrix
membranes incorporated with Ln-MOF for selective and sensitive detection of
nitrofuran antibiotics based on inner filter effect. Talanta, 174, pp.660-666.
https://doi.org/10.1016/j.talanta.2017.07.007
[639] Howarth, A.J., Liu, Y., Li, P., Li, Z., Wang, T.C., Hupp, J.T. and Farha, O.K., 2016.
Chemical, thermal and mechanical stabilities of metal–organic frameworks. Nature
Reviews Materials, 1(3), pp.1-15. https://doi.org/10.1038/natrevmats.2015.18
[640] Yan, Y., He, T., Zhao, B., Qi, K., Liu, H. and Xia, B.Y., 2018. Metal/covalent–organic
frameworks-based electrocatalysts for water splitting. Journal of Materials Chemistry
A, 6(33), pp.15905-15926. https://doi.org/10.1039/C8TA05985C
[641] Hirscher, M., Yartys, V.A., Baricco, M., von Colbe, J.B., Blanchard, D., Bowman Jr,
R.C., Broom, D.P., Buckley, C.E., Chang, F., Chen, P. and Cho, Y.W., 2020. Materials
for hydrogen-based energy storage–past, recent progress and future outlook. Journal of
Alloys and Compounds, 827, p.153548. https://doi.org/10.1016/j.jallcom.2019.153548
[642] Zhang, X., Dong, H., Sun, X.J., Yang, D.D., Sheng, J.L., Tang, H.L., Meng, X.B. and
Zhang, F.M., 2018. Step-by-step improving photocatalytic hydrogen evolution activity
of NH2–UiO-66 by constructing heterojunction and encapsulating carbon

206
nanodots. ACS Sustainable Chemistry & Engineering, 6(9), pp.11563-11569.
https://doi.org/10.1021/acssuschemeng.8b01740

[643] Tian, L., Yang, X., Liu, Q., Qu, F. and Tang, H., 2018. Anchoring metal-organic
framework nanoparticles on graphitic carbon nitrides for solar-driven photocatalytic
hydrogen evolution. Applied Surface Science, 455, pp.403-409.
https://doi.org/10.1016/j.apsusc.2018.06.014

[644] Wang, Z., Jin, Z., Wang, G. and Ma, B., 2018. Efficient hydrogen production over
MOFs (ZIF-67) and g-C3N4 boosted with MoS2 nanoparticles. International Journal of
Hydrogen Energy, 43(29), pp.13039-13050.
https://doi.org/10.1016/j.ijhydene.2018.05.099
[645] Jiang, Z., Liu, J., Gao, M., Fan, X., Zhang, L. and Zhang, J., 2017. Assembling Polyoxo‐
Titanium Clusters and CdS Nanoparticles to a Porous Matrix for Efficient and Tunable
H2‐Evolution Activities with Visible Light. Advanced Materials, 29(5), p.1603369.
https://doi.org/10.1002/adma.201603369
[646] Wang, Y., Yu, Y., Li, R., Liu, H., Zhang, W., Ling, L., Duan, W. and Liu, B., 2017.
Hydrogen production with ultrahigh efficiency under visible light by graphene well-
wrapped UiO-66-NH2 octahedrons. Journal of Materials Chemistry A, 5(38),
pp.20136-20140. https://doi.org/10.1039/C7TA06341E
[647] Liu, D., Jin, Z., Zhang, Y., Wang, G. and Ma, B., 2018. Light harvesting and charge
management by Ni4S3 modified metal− organic frameworks and rGO in the process of
photocatalysis. Journal of colloid and interface science, 529, pp.44-52.
https://doi.org/10.1016/j.jcis.2018.06.001

[648] Shi, X., Zhang, J., Cui, G., Deng, N., Wang, W., Wang, Q. and Tang, B., 2018.
Photocatalytic H2 evolution improvement for H free-radical stabilization by
electrostatic interaction of a Cu-BTC MOF with ZnO/GO. Nano Research, 11(2),
pp.979-987. https://doi.org/10.1007/s12274-017-1710-4
[649] Shen, L., Luo, M., Huang, L., Feng, P. and Wu, L., 2015. A clean and general strategy
to decorate a titanium metal–organic framework with noble-metal nanoparticles for
versatile photocatalytic applications. Inorganic Chemistry, 54(4), pp.1191-1193.
https://doi.org/10.1021/ic502609a
[650] Kampouri, S., Nguyen, T.N., Ireland, C.P., Valizadeh, B., Ebrahim, F.M., Capano, G.,
Ongari, D., Mace, A., Guijarro, N., Sivula, K. and Sienkiewicz, A., 2018. Photocatalytic
hydrogen generation from a visible-light responsive metal–organic framework system:
the impact of nickel phosphide nanoparticles. Journal of Materials Chemistry A, 6(6),
pp.2476-2481. https://doi.org/10.1039/C7TA10225A
[651] Shen, C.C., Liu, Y.N., Wang, X., Fang, X.X., Zhao, Z.W., Jiang, N., Ma, L.B., Zhou,
X., Cheang, T.Y. and Xu, A.W., 2018. Boosting visible-light photocatalytic H2
evolution via UiO-66-NH2 octahedrons decorated with ultrasmall NiO
nanoparticles. Dalton Transactions, 47(33), pp.11705-11712.
https://doi.org/10.1039/C8DT02681E
[652] Li, D., Yu, S.H. and Jiang, H.L., 2018. From UV to near‐infrared light‐responsive
metal–organic framework composites: plasmon and upconversion enhanced
photocatalysis. Advanced Materials, 30(27), p.1707377.
https://doi.org/10.1002/adma.201707377
[653] Liu, X.L., Wang, R., Zhang, M.Y., Yuan, Y.P. and Xue, C., 2015. Dye-sensitized MIL-
101 metal organic frameworks loaded with Ni/NiOx nanoparticles for efficient visible-

207
light-driven hydrogen generation. APL materials, 3(10), p.104403.
https://doi.org/10.1063/1.4922151

[654] Yuan, Y.P., Yin, L.S., Cao, S.W., Xu, G.S., Li, C.H. and Xue, C., 2015. Improving
photocatalytic hydrogen production of metal–organic framework UiO-66 octahedrons
by dye-sensitization. Applied Catalysis B: Environmental, 168, pp.572-576.
https://doi.org/10.1016/j.apcatb.2014.11.007

[655] Chen, Y.F., Tan, L.L., Liu, J.M., Qin, S., Xie, Z.Q., Huang, J.F., Xu, Y.W., Xiao, L.M.
and Su, C.Y., 2017. Calix [4] arene based dye-sensitized Pt@ UiO-66-NH2 metal-
organic framework for efficient visible-light photocatalytic hydrogen
production. Applied Catalysis B: Environmental, 206, pp.426-433.
https://doi.org/10.1016/j.apcatb.2017.01.040
[656] Kong, X.J., Lin, Z., Zhang, Z.M., Zhang, T. and Lin, W., 2016. Hierarchical integration
of photosensitizing metal–organic frameworks and nickel‐containing polyoxometalates
for efficient visible‐light‐driven hydrogen evolution. Angewandte Chemie
International Edition, 55(22), pp.6411-6416. https://doi.org/10.1002/anie.201600431
[657] He, F., Chen, G., Zhou, Y., Yu, Y., Li, L., Hao, S. and Liu, B., 2016. ZIF-8 derived
carbon (C-ZIF) as a bifunctional electron acceptor and HER cocatalyst for g- C3N4:
construction of a metal-free, all carbon-based photocatalytic system for efficient
hydrogen evolution. Journal of Materials Chemistry A, 4(10), pp.3822-3827.
https://doi.org/10.1039/C6TA00497K
[658] Kumar, D.P., Choi, J., Hong, S., Reddy, D.A., Lee, S. and Kim, T.K., 2016. Rational
synthesis of metal–organic framework-derived noble metal-free nickel phosphide
nanoparticles as a highly efficient cocatalyst for photocatalytic hydrogen
evolution. ACS Sustainable Chemistry & Engineering, 4(12), pp.7158-7166.
https://doi.org/10.1021/acssuschemeng.6b02032
[659] Lan, G., Zhu, Y.Y., Veroneau, S.S., Xu, Z., Micheroni, D. and Lin, W., 2018. Electron
injection from photoexcited metal–organic framework ligands to Ru2 secondary
building units for visible-light-driven hydrogen evolution. Journal of the American
Chemical Society, 140(16), pp.5326-5329. https://doi.org/10.1021/jacs.8b01601

[660] Wu, Z.L., Wang, C.H., Zhao, B., Dong, J., Lu, F., Wang, W.H., Wang, W.C., Wu, G.J.,
Cui, J.Z. and Cheng, P., 2016. A semi‐conductive copper–organic framework with two
types of photocatalytic activity. Angewandte Chemie, 128(16), pp.5022-5026.
http://dx.doi.org/10.1002/anie.201508325

[661] Chen, D.M., Sun, C.X., Liu, C.S. and Du, M., 2018. Stable layered semiconductive Cu
(I)–organic framework for efficient visible-light-driven Cr (VI) reduction and H2
evolution. Inorganic Chemistry, 57(13), pp.7975-7981.
https://doi.org/10.1021/acs.inorgchem.8b01137

[662] Pattengale, B., Yang, S., Lee, S. and Huang, J., 2017. Mechanistic probes of zeolitic
imidazolate framework for photocatalytic application. ACS Catalysis, 7(12), pp.8446-
8453. https://doi.org/10.1021/acscatal.7b02467
[663] Li, H., Yao, S., Wu, H.L., Qu, J.Y., Zhang, Z.M., Lu, T.B., Lin, W. and Wang, E.B.,
2018. Charge-regulated sequential adsorption of anionic catalysts and cationic
photosensitizers into metal-organic frameworks enhances photocatalytic proton
reduction. Applied Catalysis B: Environmental, 224, pp.46-
52.https://doi.org/10.1016/j.apcatb.2017.10.031

208
[664] Kampouri, S., Nguyen, T.N., Ireland, C.P., Valizadeh, B., Ebrahim, F.M., Capano, G.,
Ongari, D., Mace, A., Guijarro, N., Sivula, K. and Sienkiewicz, A., 2018. Photocatalytic
hydrogen generation from a visible-light responsive metal–organic framework system:
the impact of nickel phosphide nanoparticles. Journal of Materials Chemistry A, 6(6),
pp.2476-2481. https://doi.org/10.1021/acsami.8b10010
[665] Fan, K., Jin, Z., Wang, G., Yang, H., Liu, D., Hu, H., Lu, G. and Bi, Y., 2018.
Distinctive organized molecular assemble of MoS2, MOF and Co3O4, for efficient dye-
sensitized photocatalytic H2 evolution. Catalysis Science & Technology, 8(9), pp.2352-
2363. https://doi.org/10.1039/C8CY00380G
[666] Zhang, R., Liu, Y., Wang, J., Wang, Z., Wang, P., Zheng, Z., Qin, X., Zhang, X., Dai,
Y. and Huang, B., 2019. Post-synthetic platinum complex modification of a triazine
based metal organic frameworks for enhanced photocatalytic H2 evolution. Journal of
Solid State Chemistry, 271, pp.260-265. https://doi.org/10.1016/j.jssc.2019.01.006
[667] Feng, Y., Chen, C., Liu, Z., Fei, B., Lin, P., Li, Q., Sun, S. and Du, S., 2015. Application
of a Ni mercaptopyrimidine MOF as highly efficient catalyst for sunlight-driven
hydrogen generation. Journal of Materials Chemistry A, 3(13), pp.7163-7169.
https://doi.org/10.1039/C5TA00136F
[668] Fang, X., Shang, Q., Wang, Y., Jiao, L., Yao, T., Li, Y., Zhang, Q., Luo, Y. and Jiang,
H.L., 2018. Single Pt atoms confined into a metal–organic framework for efficient
photocatalysis. Advanced Materials, 30(7), p.1705112.
https://doi.org/10.1002/adma.201705112
[669] Zheng, S.T., Zuo, F., Wu, T., Irfanoglu, B., Chou, C., Nieto, R.A., Feng, P. and Bu, X.,
2011. Cooperative Assembly of Three‐Ring‐Based Zeolite‐Type Metal–Organic
Frameworks and Johnson‐Type Dodecahedra. Angewandte Chemie International
Edition, 50(8), pp.1849-1852. https://doi.org/10.1002/anie.201006882
[670] Guo, W., Lv, H., Chen, Z., Sullivan, K.P., Lauinger, S.M., Chi, Y., Sumliner, J.M.,
Lian, T. and Hill, C.L., 2016. Self-assembly of polyoxometalates, Pt nanoparticles and
metal–organic frameworks into a hybrid material for synergistic hydrogen
evolution. Journal of Materials Chemistry A, 4(16), pp.5952-5957.
https://doi.org/10.1039/C6TA00011H
[671] Guo, Y., Zhang, J., Dong, L.Z., Xu, Y., Han, W., Fang, M., Liu, H.K., Wu, Y. and Lan,
Y.Q., 2017. Syntheses of Exceptionally Stable Aluminum (III) Metal–Organic
Frameworks: How to Grow High‐Quality, Large, Single Crystals. Chemistry–A
European Journal, 23(61), pp.15518-15528. https://doi.org/10.1002/chem.201703682
[672] Liu, H., Xu, C., Li, D. and Jiang, H.L., 2018. Photocatalytic hydrogen production
coupled with selective benzylamine oxidation over MOF composites. Angewandte
Chemie, 130(19), pp.5477-5481. https://doi.org/10.1002/ange.201800320
[673] Wang, Y., Ling, L., Zhang, W., Ding, K., Yu, Y., Duan, W. and Liu, B., 2018. A
Strategy to Boost H2 Generation Ability of Metal–Organic Frameworks: Inside‐Outside
Decoration for the Separation of Electrons and Holes. ChemSusChem, 11(4), pp.666-
671. https://doi.org/10.1002/cssc.201702316

209

View publication stats

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