Arun Current Applied Physics-Oct-2022
Arun Current Applied Physics-Oct-2022
net/publication/364257648
CITATIONS READS
6 285
6 authors, including:
All content following this page was uploaded by Arunangshu Biswas on 18 October 2022.
A R T I C L E I N F O A B S T R A C T
Keywords: The article presents a combined theoretical and experimental study attempting to show how Pd nanoparticles
Sensor (NPs) loading onto SnO2 substrate improves the acetone gas sensing performance. Pristine nanostructured SnO2
Acetone and Pd nanoparticles (Pd NPs) loaded SnO2 substrates have been prepared, characterized, and their acetone
Ethanol
sensing performances have been measured. Experimental measurements have shown that Pd NP loading onto
SnO2
Pd nanoparticles
SnO2 suppresses the interfering effects of ethanol, water vapors, etc., and enhances the acetone sensor response,
DFT reversibility, response/recovery speeds, and signal-to-noise ratio. Various parameters like the adsorption energy,
HOMO–LUMO energy gap, charge distribution, polarizability change, electrophilicity index, global hardness,
etc., of several model systems, have been computed by using DFT. The computed parameters have been corre
lated with the conductivity, local reactivity, sensor response and selectivity, response/recovery times, etc., of the
systems to understand the molecular-level effects of the Pd NP loading onto the SnO2 on the gas sensing process.
1. Introduction like good sensitivity, selectivity, and shorter response and recovery
times. The choice of sensing substrates plays a critical role in deciding
Detecting and monitoring various gases and volatile organic com the performance of the sensor, along with their cost, robustness, ease of
pounds (VOCs) find applications in the fields of safety and security, use, etc. It has led researchers to explore a variety of sensor materials for
environmental and industrial monitoring, agriculture and automotive gas-sensing devices [3–13]. Semiconducting metal oxides (SMOs) are
industry, and medical diagnosis and monitoring [1,2]. Although various among the favorite classes of materials that show the chemoresistance
sophisticated analytical tools and techniques are available for the ac property upon gas adsorptions [3–10]. For example, SnO2 is an n-type
curate determination of low concentrations of gases and VOCs, there are SMO widely used for detecting several reducing and oxidizing gases and
several disadvantages associated with such sophisticated instruments. VOCs [14–17]. SMO materials also have other favorable attributes like
These include their high cost, space and infrastructure requirement, low cost, ease of synthesis and fabrication, high thermal and chemical
complicated and time-consuming sample preparation steps, etc. Further, stability, reliability, and modest efficiency with excellent prospects of
these instruments are not favorable for routine and point-of-care use. miniaturization. However, the SMO-based sensors work by the basic
Thus, there is a great demand for various kinds of gas sensors for both principles of the conductance change of the substrate due to the
predictive and preventive measures. Therefore, designing and devel adsorption/surface reactions/desorption of oxygen and target gas mol
oping gas sensors with improved performance have emerged as a ecules. Therefore, SMO-based sensors also have several drawbacks, such
prominent area of research in recent times. as low target selectivity, long recovery/response times, weak response,
The requirement of the performance parameters of a gas sensor high operating temperature, etc., which need improvement.
varies depending on its use, the target and interfering gases, etc. How Various strategies have been explored to overcome such limitations
ever, a real-time high-performance gas sensor should have basic features and improve the performance of SMO-based sensing. For example,
* Corresponding author.
E-mail address: tapan.sau@iiit.ac.in (T.K. Sau).
https://doi.org/10.1016/j.cap.2022.10.003
Received 18 May 2022; Received in revised form 21 September 2022; Accepted 3 October 2022
Available online 8 October 2022
1567-1739/© 2022 Korean Physical Society. Published by Elsevier B.V. All rights reserved.
A. Biswas et al. Current Applied Physics 44 (2022) 131–143
researchers have explored complex metal oxides, composite systems, interactions [50–62]. Such information is essential for designing and
surface engineering, etc. [9,12,15,18–21] Since SMO-based sensing in developing efficient future gas sensors. We have employed DFT simu
volves surface adsorptions/reactions, substrate’s surface states, addi lations to systematically investigate the atomic and molecular level in
tives, etc. play critical roles in the sensor performance. It has led to two fluence of the Pd NP loading on the acetone sensing behavior of SnO2.
effective strategies for improving sensor performance: We have computed parameters like the adsorption and bandgap en
nanomaterials-based gas sensors and SMO-surface modifications by ergies, charge distribution, polarizability change, electrophilicity index,
noble metals. Nanomaterials have several features like a large global hardness, etc., for various ligand gas (viz., acetone, ethanol, ox
surface-to-volume ratio, higher surface-active sites, etc., offering ygen, moisture, and CO2) interactions on the pristine SnO2 as well as
favorable substrate surface-gas interactions and superior sensor perfor Pd-loaded SnO2 systems. The computational study reveals the molecular
mance [22–24]. On the other hand, SMO-surface modifications with aspects of the mechanisms of acetone sensor performance improvement
noble metal sensitizers like Pd, Pt, Au, Ag, etc., provide the catalytic and due to the Pd NP loading onto nanostructured SnO2 material.
spillover activities towards the substrate-gas interactions [14,24–28].
It is to be noted that despite the long history of using SMO-based 2. Experimental
sensors, mainly a trial-and-error approach has been used to select ma
terials for various gas sensing. The exact mechanisms of noble metal 2.1. Chemicals and reagents
sensitizers in improving gas sensing performance are yet to be under
stood. It has been proposed that noble metals activate the target gas and Stannous chloride (SnCl2.2H2O), glacial acetic acid (CH3COOH),
catalyze its oxidation on the SMO surface. For example, Pd can serve as concentrated aqueous ammonia, and polyethylene glycol-200 (PEG-
the initial adsorption site for oxygen adsorption via the formation of PdO 200) were purchased from LOBA Chemie. Potassium tetra
[29–31]. PdO being a p-type SMO favor an efficient electron withdrawal chloropalladate (II) (K2PdCl4) and sodium borohydride (NaBH4) were
from an n-type SMO like SnO2 particles boosting the sensor response purchased from Sigma-Aldrich. Millipore water was used for all
[30–32]. Pd loading has also been reported to suppress the humidity or experiments.
water vapor poisoning effect on the SnO2 surface [17,33]. On the other
hand, we hardly understand how one can enhance an SMO substrate’s 2.2. Preparation of nanostructured SnO2
sensor response/selectivity towards a given target gas over the other
structurally and/or chemically similar gas(es). Since the SMO-based Among various synthesis strategies such as solvothermal, hydro
sensing mechanism involves the oxidation of target gases by the thermal, spray pyrolysis, sol-gel, etc., employed for the preparation of
surface-adsorbed oxygen, the presence of more than one reducing gas SMOs, the sol-gel method is one of the most favored synthesis methods
poses a challenge to their SMO-based detection due to the interference. in terms of the ease and practicality of large-scale preparations. Here, for
For example, let us consider the acetone sensing by an SMO substrate in the synthesis of nanostructured SnO2 particles, we employed the facile
the presence of moisture (H2O), ethanol (CH3CH2OH), etc. Ethanol and sol-gel method. In a typical synthesis, a measured amount of solid
moisture are reducing in nature, like acetone. Furthermore, acetone and SnCl2.2H2O was added to 50 mL of hot Millipore water (at ~75 ◦ C) to
ethanol are structurally and chemically similar. Resolving the problem make a 10 mM solution. The mixture was stirred with the help of a
of target gas selectivity requires a deeper knowledge of the interactions magnetic stirrer for about 20 min. Then aq. ammonia was added drop
of the sensor substrates with oxygen and other ligand gas molecules. wise until white precipitation of Sn(OH)2 formed. Slowly the tempera
Therefore, the sensor performance improvement requires an under ture was brought down to ~60 ◦ C, and the reaction solution was left
standing of the microstructural sensor activation mechanisms, which, in undisturbed for 30 min to settle down the precipitate. The precipitate
turn, requires electronic and structural properties level information on was filtered and washed with Millipore water and diluted ammonium
the interactions of the sensing substrate and ligand gas(es). hydroxide. The residue was transferred into a beaker by adding 20 mL of
This article reports a combined theoretical and experimental study glacial acetic acid. Then 2 mL of polyethylene glycol 200 was added to
exploring the acetone sensing performance improvement of a nano the beaker and mixed with the help of a glass rod. The solution was then
structured SnO2 substrate by Pd nanoparticles (Pd NPs) loading. Acetone heated at 180 ◦ C with occasional stirring with the help of a glass rod to
sensors find uses in early diagnosis, health and environmental moni remove acetic acid completely. Next, the solution was left to cool down
toring, as well as industrial safety [34–41]. Acetone is commonly used in at room temperature. A light brown-colored sol appeared slowly. The sol
laboratories and chemical industries. However, it is a volatile organic transformed into a brown-yellow colored gel within 2–4 h. The gel was
liquid and toxic substance [42,43]. Exposure to acetone can cause transferred into a silica crucible and was calcined in a Muffle Furnace.
various health problems like eye/skin/respiratory tract irritation, First, it was ramped at 2 ◦ C min− 1 to 350 ◦ C and then heated for 2 h at
headache, fatigue, and nausea and can even damage our central nervous this temperature. Finally, the temperature was raised to 700 ◦ C and left
system at as low as 173 ppm acetone [43–45]. The presence of acetone in at this temperature for 30 min to obtain the desired product [63].
exhaled breath is an important chemical biomarker correlated with
several physiological and pathological processes in the human body. 2.3. Preparation of Pd NPs loaded SnO2
Breath-acetone is the end product of ketone body metabolism. It has
been linked to diabetes, physical activities and diet, acute kidney injury, For the formation and deposition of Pd NPs on the nanostructured
colorectal cancer, lung cancer, etc. [2,46,47] Recently, breath SnO2 in aqueous dispersion, we have employed room-temperature
VOC-based diagnosing and monitoring of human metabolism and health borohydride (NaBH4)-reduction of Pd(II) ions in K2PdCl4. This simple
conditions have gained momentum [2,46,48,49]. Breath-VOCs offer method allows rapid preparation of Pd NP-loaded SnO2 particles
noninvasive and inexpensive real-time sampling and analysis, avoiding avoiding high-temperature modifications of the NPs. Pd NPs formed and
inconveniences like interventional time-consuming sampling tech deposited directly on the nanostructured SnO2 particles. In solution,
niques, sample volume, dilution issues, etc. However, the detection of PdCl2−
4 ions can bind to the surface-hydroxyls of tin oxide particles and
breath-acetone among a plethora of volatile organic compounds (VOCs), are then reduced by the BH−4 ions forming Pd NPs on the surface of the
moisture, CO2, ethanol, etc., present in the exhaled breath poses several SnO2 particles. In a typical synthesis, 30 mg of SnO2 was first added to
challenges. 10 mL of Millipore water and was dispersed by magnetic stirring for 30
In recent times, several theoretical approaches, especially the density min. The required amount of 0.5 mM K2PdCl4 salt was added to the
functional theory (DFT), have emerged as powerful methods for dispersion. The solution color turned yellowish at this stage. Next, 10 mL
revealing the atomic and molecular level structural and electronic of freshly prepared ice-cold 3 mM NaBH4 solution was added to the
changes arising from the sensor substrates and the target/non-target gas reaction mixture. The reaction mixture was stirred for another 30 min.
132
A. Biswas et al. Current Applied Physics 44 (2022) 131–143
The solution color changed to grey, indicating the formation of Pd0 NPs. 3. Results and discussion
Then the reaction solution was left undisturbed for 2 h. The reaction
product was then filtered, and the residue was washed with Millipore 3.1. Phase structure and composition analysis
water. Finally, the residue was dried at room temperature. Ma et al.
reported that the Pd loading reached a saturation level at ~0.7 mol% 3.1.1. Powder XRD and XPS studies
precursor concentration for almost similar-sized SnO2 particles [64,65]. Fig. 1 shows typical powder XRD patterns of the synthesized samples
We present here data for two Pd concentrations such that one is slightly in the range of 2θ = 20◦ –80◦ . The experimental diffraction peaks can be
below, and the other is slightly above this saturation point, namely, 0.5 indexed to SnO2 (JCPDS 88–0287), Pd (JCPDS 46–1043), and PdO
mol% and 1.0 mol% Pd concentrations (henceforth l-Pd/SnO2 and (JCPDS 41–1107), suggesting the formation of tetragonal SnO2 phase
h-Pd/SnO2, respectively). and the presence of Pd0 and PdO [68–70]. The small line shifting in the
XRD pattern of the prepared samples with respect to the standard could
2.4. Characterizations be due to the micro-strains present in the nanostructured particles.
Particularly, the broader diffraction peaks with low intensity of Pd and
The synthesized particles were characterized by using powder X-ray PdO can be attributed to their small crystallite sizes. A few peak profiles
diffraction (XRD, Bruker, D8 Advance, Cu kα radiation, λ = 1.5406 Å), show asymmetric shapes. It could be due to a distribution of lattice
Raman scattering spectroscopy (EZRaman-N-785, Enwave Optronics, parameters arising from the defects and disorders in the lattice sites, a
Inc.), scanning electron microscopy (SEM, Carl Zeiss Ultra 55) equipped chemical variation, and/or a non-stoichiometric compound with a
with energy-dispersive X-ray spectroscopy (EDS), X-ray photoelectron particular phase width [71]. The mean crystallite sizes (D) were esti
spectroscopy (XPS) (PHI 5000 VersaProbe III), UV–Visible spectropho mated to be ~9.9 nm and ~9.8 nm, respectively, for SnO2 and Pd/PdO,
tometry (PerkinElmer Lambda-35), and Photoluminescence (PL) spec by using the Debye-Scherrer formula, D = 0.94 λ/β cos θ (where D is the
troscopy (Test Right Portable Spectroscopy Solutions). The BET surface diameter of the particle, λ is the source X-ray wavelength (Cu Kα) =
area measurements were carried out using Quantachrome Novawin 0.15406 nm, β is the full width (in radian) at half maximum of the
(Nova Station A) instrument. diffraction peak at diffraction angle 2θ).
We know that XPS provides near-surface chemical compositions of
2.5. Sensor fabrication and gas sensing measurements the samples. XPS confirmed the presence of Sn and O elements in the
SnO2 sample and Sn, O, and Pd elements in the Pd nanoparticle-loaded
The construction of the sensor element was similar to that reported in SnO2 sample. Fig. 2 shows survey as well as high-resolution XPS of Sn, O,
ref. [66]. A sketch of the electric circuit diagram used for the sensors’ and Pd elements. Sn 3d XPS (2(b)) shows Sn 3d5/2 and Sn 3d3/2 peaks at
signal measurement is shown in SI Scheme S1. The sensor was designed 486.88 eV and 495.25 eV, respectively, with a spitting energy separation
on a ceramic tube with preinstalled gold electrodes on it and platinum
wires attached to it. The ceramic tube was coated with 20 mg of the
substrate materials under test and placed in the testing circuit. A
nichrome wire was used as a heating element that generates ~200 ◦ C at
5V/0.32A. The heating offers enhanced sensitivity and performance to
the sensor elements via inter-particle sintering and enhanced trans
ducing and surface catalysis [67]. It is worth noting that 200 ◦ C falls at
the lower end of the operating temperature range used for most acetone
sensing experiments. The sensor element was placed in a 200 mL volume
box where the sensing gas was injected. To study the sensor recovery,
once the sensor response reached the stable state for a given period, the
test box was opened to re-expose the sensor element to fresh air when its
resistance returned to the base value. The sensor was tested with target
gas like acetone and interfering gas like ethanol in the 1–100 ppm
concentrations range. The relative humidity (RH) was 40–50% while
performing the experiments. Testing was also performed for up to 90%
moisture to check its interfering effects in the acetone sensing. All
measurements were repeated at least three times.
133
A. Biswas et al. Current Applied Physics 44 (2022) 131–143
Fig. 2. (a) XPS survey spectra of SnO2 and Pd nanoparticle-loaded SnO2 samples. High-resolution XPS of (b) Sn 3d, (c) Pd 3d, (d) Pd 3d deconvoluted XPS peaks, and
(e) O 1s (with deconvoluted peaks) of the Pd nanoparticle loaded SnO2 samples.
of 8.37 eV which is assigned to Sn4+ in SnO2 [69]. Fig. 2(c) shows the 3.1.2. UV–Vis and photoluminescence (PL) studies
high-resolution Pd 3d XPS with spin-orbit splitting peaks of Pd 3d5/2 and The UV–visible absorption and photoluminescence spectra collected
Pd 3d3/2 at 335.63 eV and 340.75 eV, respectively. A Pd 3d3/2 peak at from SnO2 and Pd nanoparticles loaded SnO2 samples are shown in
340.0–340.9 eV (3d5/2 peak at 334.8–335.6 eV) is attributed to Fig. 3. Asymmetric absorption occurs mainly in the UV region which can
(metallic) Pd0. The peaks shift to higher energy with an increase in the be deconvoluted into two bands (inset Fig. 3(a)) peaking at 295 nm
oxidation state of palladium [72]. For example, Pd2+ species is charac (4.20 eV) and 371 nm (3.34 eV). The latter band is close to the bulk SnO2
terized by the Pd 3d3/2 peak at 341.6–342.6 eV (3d5/2 peak at band gap energy of 3.6 eV (444.4 nm). The higher energy absorption
336.4–337.3 eV) whereas Pd4+ species is characterized by the Pd 3d3/2 could arise from the larger band gap of smaller SnO2 nanostructures,
peak at 343.2–344.2 eV (3d5/2 peak at 338.0–338.9 eV) [73]. From the since the quantum confinement effect in the semiconductor nanocrystals
multipeak fitting of the Pd 3d3/2 peak in Fig. 2(d), one can say that both shifts the band gap to higher energy when the particle size decreases.
(metallic) Pd0 and Pdδ+ (0 < δ ≤ 2) species are present in the Pd The presence of Pd dampens the absorbance and shifts the absorption
nanoparticle-loaded sample. It is in agreement with our other experi maximum to UV region further due to the Pd nanoparticles’ own ab
mental and computational studies. Similarly, the O 1s XPS (Fig. 2(e)) sorption in the UV region. The photoluminescence emission spectra
signal can be deconvoluted into two peaks at 530.88 eV and 532.13 eV. (Fig. 3(b)) of SnO2 nanoparticles were collected at 365 nm and 470 nm
The 532.13 eV peak can be assigned to the surface adsorbed oxygen excitations. The former excitation energy employed is slightly higher
species (O−x ) while the 530.88 eV peak is attributed to the oxygen ions in than the bulk band gap energy and the latter is lower than it. The PL
the crystal lattice (Olattice). It shows the presence of surface-adsorbed exhibits maximal emissions at 395 nm (3.14 eV) and 494 nm (2.51 eV),
oxygen species (O−x ), in addition to lattice oxygen (Olattice) ions in the respectively. The former emission is very close to the bulk band gap of
samples [72,74]. SnO2 which can be associated with a band-to-band recombination pro
cess. The lower energy emission can be attributed to the excitation of the
134
A. Biswas et al. Current Applied Physics 44 (2022) 131–143
Fig. 3. (a) UV–Vis absorption and (b) photoluminescence spectra (with 365 and 470 nm excitations) collected from SnO2 and Pd nanoparticle loaded SnO2 samples.
electrons from valence band to –OH group related defect energy states,
oxygen vacancies, etc. in the SnO2 nanoparticles. The lower PL intensity
in the presence of Pd nanoparticles could be due to the heavy-atom ef
fect PL quenching. The optical absorption and photoluminescence
spectra were similar to that previously reported in the literature
[75–78].
135
A. Biswas et al. Current Applied Physics 44 (2022) 131–143
also show vibrations below the 800 cm− 1 range. As shown in Fig. 4, 3.2. Acetone sensing performance: Experimental study
many peaks in the 240 cm− 1 – 495 cm− 1 region could be assigned to
different Pd–O vibrations [88]. Surface superoxide species (O−2 ) adsor The sensor responses of the prepared SnO2 and Pd/SnO2 sensor
bed on the various sites are observed to have O–O stretching vibration substrates were experimentally measured by comparing the sensor sig
frequencies in the 1010–1180 cm− 1 range [81,89,90]. The O–O vibra nals before (in the air) and after the exposure to the target gas (including
tion of peroxide adspecies (O2−2 ) appears in the range 640–970 cm
− 1
air). Exposure of an n-type SMO (e.g., SnO2) surface to the air results in
[89,90]. The bands above ~1525 cm− 1 have been assigned to molecular oxygen adsorption. Raman study has confirmed the presence of chem
O2 species. On the other hand, the bands above ~1300 cm− 1 have been isorbed oxygen species on the substrate surface (Fig. 4). Oxygen
2 (0 < δ < 1)
attributed to dioxygen adspecies of intermediate forms, Oδ− adsorption causes the withdrawal of carrier electrons from the con
[89,91]. Bands peaking at 1012, 1036, 1070, 1096, and 1164 cm− 1 can duction band of the SMO. It results in a decrease in the carrier concen
be ascribed to the stretching frequencies of superoxo (O−2 )-like surface tration and thus an increase in electrical resistance of the substrate.
species [90,92,93]. Raman peaks at 652, 680, 726, 822, and 880 cm− 1 When the SMO sensor is exposed to a reducing target gas, the surface
correspond to the O–O stretching vibrational frequencies of different adsorbed oxygen species reacts with it and gets oxidized (e.g., acetone is
peroxo (O2−2 ) adspecies on Pd NP surfaces [92–94]. Higher Raman in oxidized to CO2 and H2O). The process releases the electrons to the
tensity of several surface oxygen species on the Pd/SnO2 substrate than conduction band of SMO, resulting in an increase in the conductance (or
the pristine SnO2 shows enhanced oxygen adsorptions after loading of a decrease in the resistance) of the SMO substrate. Since the resistance
Pd NPs, which can produce improved sensor response. value decreased upon exposure to the reducing gas, the sensor response,
S, was calculated using the relation: S (%) = (Ra − Rg) × 100/Rg, where
3.1.4. SEM-EDS and BET surface area measurement studies Ra is the sensor resistance in the air (original signal or base resistance)
Fig. 5 shows typical SEM images of the prepared SnO2 and Pd/SnO2 and Rg is the resistance in the presence of the target gas along with air.
samples. The roughness of the surface due to the nanoparticulate forms Fig. 7 shows the typical sensor response (S) of the pristine SnO2 and
of the materials can be seen in all samples. The nanoscale roughness of Pd/SnO2 sensors in the range of 1–100 ppm acetone vapor. It can be seen
the sensor materials results in a more accessible surface area and active that the sensor response increased with increasing acetone vapor con
sites for the adsorption and diffusion of the target gas molecules, which centration. The increase in the sensor response with the increasing
augment the sensitivity and response rate of the sensors. The size of concentration of the target gas can be attributed to the enhanced surface
particles on the surface ranged from ~15 nm to ~25 nm. Pd NPs are reactions due to the increased surface coverage of the gas molecules,
hardly distinguishable from the SnO2 background. The lack of contrast gradually reaching a saturation. First, let us check the developed sen
between Pd and SnO2 phases could be due to minor differences in the sor’s sensitivity to a gas (as defined by the ‘Rg/Ra’ ratio) and the signal-
atomic numbers of 46Pd and 50Sn [17]. However, to demonstrate the to-noise ratio (S/N) ratio. It is known that when the sensitivity to a gas
existence of Pd, elemental composition analysis and mapping using (Rg/Ra) is less than 0.8, the detection of the gas is possible. The ‘Rg/Ra’
electron energy-dispersive x-ray spectroscopy (EDS) were carried out. ratio value was slightly below 0.8 even at the concentration of 1 ppm
EDS mapping shed light on the dispersion of Pd NPs on the surface of acetone for the prepared Pd/SnO2 sensor substrate. Further, the devel
SnO2. Fig. 6 shows a typical EDS elemental composition analysis and oped sensors showed a high S/N ratio. The S/N was calculated as fol
mapping of Sn, O, and Pd on the sample surfaces. The elemental map lows: S/N = [(Average sensor response of the sample) – (Average sensor
ping showed a nearly homogeneous distribution of Pd NPs on the sample response of the blank)]/(Standard Deviation of the blank). Taking
surface. A scrutiny of the distribution of the elements in the sample pristine SnO2 as the blank, we calculated S/N ≈ 17 and ≈ 20 for 1 ppm
clearly shows that Sn and O are relatively evenly distributed across the acetone with l-Pd/SnO2 and h-Pd/SnO2 sensors substrates, respectively.
sample, while Pd shows discrete distribution. It suggests that Pd forms The effects of Pd NP loading become clear from comparing the sensor
particles on the SnO2 background. It is interesting to compare the atom% responses of the pristine SnO2 and Pd/SnO2 sensor substrates. Pd NP
compositions of the elements obtained from the EDS study (SI Table T1). loading onto SnO2 improved both the sensor response and the acetone
For the pristine SnO2 particles, the experimentally determined O:Sn selectivity. The pristine SnO2 sensor showed a marginally higher sensor
atom% ratio was 2.2, close to that of the stoichiometric SnO2 compound. response for ethanol than acetone. It is consistent with the slightly
However, the O:Sn atom% ratio of l-Pd/SnO2 ≫ h-Pd/SnO2 ≈ (pristine) higher calculated Ead value for ethanol binding to the pristine SnO2.
SnO2 substrates. The higher oxygen content in l-Pd/SnO2 material However, Pd NP loading (i.e., Pd/SnO2 substrate) significantly increased
samples could arise from the enhanced oxygen adsorption from the the acetone sensor response with respect to ethanol, unlike the pristine
ambiance. It is corroborated by the Raman spectroscopic studies. The SnO2 case (Fig. 7(a)). Pd/SnO2 sensor showed a more than two times
specific surface areas of the samples were determined by using the higher response than the pristine SnO2 sensor. However, a doubling of
Brunauer-Emmett-Teller (BET) method. Both SnO2 and Pd nanoparticles the Pd content produced only a minor increase in the sensor response,
loaded SnO2 samples showed similar BET surface area of 73.34 m2/g and which could be due to a diminished spillover effect due to a greater
69.06 m2/g, respectively. coverage of the SnO2 surface by Pd NPs. The slope of the plot of the
sensor response (S%) vs. gas concentration also gives a measure of the
sensor’s sensitivity. Pd/SnO2 sensors exhibited an almost linear and
Fig. 5. SEM images of the (a) SnO2 and (b) Pd/SnO2 samples.
136
A. Biswas et al. Current Applied Physics 44 (2022) 131–143
Fig. 6. A typical SEM-EDS elemental distribution on the surface of a synthesized Pd/SnO2 sample: (a) Sn (La1) in green, (b) O (Ka1) in red, and (c) Pd (La1) in yellow.
(For interpretation of the references to color in this figure legend, the reader is referred to the Web version of this article.)
137
A. Biswas et al. Current Applied Physics 44 (2022) 131–143
138
A. Biswas et al. Current Applied Physics 44 (2022) 131–143
Table 1
Calculated adsorption energy (Ead), HOMO-LUMO energies, HOMO-LUMO energy gap (Eg), recovery time (tr), electrophilicity index (ω), and polarizability change
(Δα) of various systems.
System Ead (eV) EHOMO (eV) ELUMO (eV) Eg (eV) ω (eV) Δα tra (s)
Table 2 where ν0 stands for the attempt frequency of the gas molecule, k is for
Pd–O and O–O bonds distances (in Å) for the adsorbed O2 species. the Boltzmann’s constant, and T is the Kelvin temperature of the system.
If the attempt frequencies of the ligand molecules are assumed to be in
System D (O–O) (Å) D (Pd—OL) (Å)
(in O2) (OL in O2) the same order of magnitude, the equation shows that a longer recovery
time arises from a higher (negative) Ead value. One can find that the
O2 1.205 –
SnO2–O2 1.219 –
presence of Pd NC considerably reduces the Ead values of acetone- and
Pd6–O2 1.319 2.038 ethanol-bound systems (Table 1) compared to the pristine SnO2 case.
SnO2–Pd6–O2 1.326 2.044 Therefore, the Pd NC loading onto SnO2 improves the recovery speed
compared to the pristine SnO2. As mentioned above, the response/re
covery rates are also related to the reactivity properties of the target
namely, moisture (H2O), acetone, and ethanol. The adsorption energies
gases on the substrates. A higher reactivity of the ligand gas favors a
decrease in the following order: SnO2-Eth > SnO2-Act > SnO2–H2O >
faster sensor response/recovery, which was mentioned above in the
SnO2–Pd6–H2O ≈ SnO2–Pd6-Eth ≈ SnO2–Pd6-Act. It shows that the
context of the adsorbed O2 species and will be further discussed later.
adsorption is significantly more exothermic on the pristine SnO2 than
that in the presence of Pd NC. In general, higher exothermic adsorption
3.3.2. HOMO-LUMO gap (Eg) and polarizability change
means better selectivity and rapid response [50]. At the same time, the
The frontier molecular orbitals (FMOs), namely, the highest occu
stronger adsorption also suggests a longer recovery time on the pristine
pied molecular orbitals (HOMOs) and the lowest unoccupied molecular
sensor substrate. The low Ead values or a weaker binding of the ligands in
orbitals (LUMOs) and their energy gaps (Eg = ELUMO − EHOMO) are very
the presence of Pd NC favor the faster recovery of the sensor [100]. The
useful for predicting and interpreting the molecular level interactions,
lowering of H2O Ead values in the presence of the Pd NC implies its roles
qualitative comparison of the chemical stabilities and reactivities of
in decreasing the H2O sensitivity of the pristine SnO2. It is consistent
clusters and molecules, and the electrical conductivity of the clusters
with the earlier report claiming that Pd loading suppressed the water
[55,56]. Table 1 lists the HOMO-LUMO energies of relevant systems. In
vapor poisoning of the SnO2 surface [33]. Similarly, from a comparison
SMO-based gas sensing, the sensor response is associated with the
of the corresponding Ead values, one can find that the presence of the Pd
electrical conductivity change of the sensor substrate resulting from the
NC also reduces the interference effects of CO2. However, SnO2–Pd6-Act,
ligand-binding events. A smaller energy gap (Eg) of the clusters offers an
SnO2–Pd6-Eth, and SnO2–Pd6–H2O have similar Ead values. Therefore,
easier transfer of electrons from the valence band to the conduction
from the adsorption energy point of view, the SnO2–Pd6 substrate shows
band. We know that the intrinsic conductivity (σ) of a semiconductor
almost similar selectivity towards these reducing gases. Later, we shall
system is mainly controlled by Eg/kT ratio and related to Eg as follows
see that the Pd loading influences various electronic properties of the
[101]:
systems, which favor acetone selectivity.
( )
We have seen above that the adsorption/desorption process and the − Eg
σ ∝exp
ease of target gas reactions play important roles in deciding the times or 2kT
speeds of the sensor response and recovery. Faster response and recov
ery of the sensor (i.e., shorter response/recovery times) are desired for where Eg, k, and T are the semiconductor HOMO-LUMO band gap,
practical gas sensors. The recovery time, tr, for gas desorption from the Boltzmann’s constant, and thermodynamic temperature, respectively. It
sensor materials has been related to the adsorption energy, Ead, as fol shows that a smaller Eg value gives rise to higher electrical conductivity,
lows [56]: σ . In other words, the σ and hence the sensor’s sensitivity will increase if
( ) the Eg value decreases upon the Pd NC and the ligand gas binding onto
1 − Ead SnO2 [102].
tr = .exp
ν0 kT First, let us consider the effects of ligand binding onto the pristine
139
A. Biswas et al. Current Applied Physics 44 (2022) 131–143
SnO2. We can find that only O2 adsorption produces a significant energy of the acetone and ethanol ligands bound systems and lowers the
decrease in the Eg values, whereas acetone adsorption onto pristine SnO2 LUMO energy of the corresponding O2-bound systems (Table 1 and
results in a slight decrease in Eg value. However, the Eg values either Fig. S5). The SnO2–O2 system can behave as an oxidant to SnO2–Pd6-Act
slightly increased or barely changed in the cases of other ligand bind since the SnO2–O2 system has a lower LUMO energy than SnO2–Pd6-Act
ings, although these ligands showed favorable exothermic adsorptions. HOMO energy. A clear measure of the electron donation-acceptance
It implies that the pristine SnO2 is not favorably sensitive to the detec tendency of the systems can be obtained from the comparison of the
tion of these gas molecules (except the acetone adsorption), and the global electrophilicity index, ω, (defined as ω = μ2/2η, where μ is the
conductivity change is the result of the O2 adsorption and desorption electronic chemical potential and η is the chemical hardness as described
processes on the pristine SnO2. above) values [103,104]. A stronger electron-acceptor tendency is
On the other hand, Pd NC loading caused a substantial drop in the Eg characterized by a larger electrophilicity index (ω) value.
values compared to the pristine SnO2 (SI Fig. S5). Comparing the Eg In these sensor systems, the chemisorbed O2 species (S–O2) are the
values of the pristine SnO2 and SnO2–Pd6 systems shows that the Pd electron-acceptors, and the other S-L species are electron donors. A
loading resulted in approx. 70% decrease in the Eg values. Therefore, Pd comparison shows that the electron-acceptor systems, S–O2, have high
NC loading can produce a significant increase in the electrical conduc electrophilicity index values than the electron donors, S-L. The higher
tivity and hence the sensor response of the sensor substrate. It is electron affinity of oxygen plays a crucial role compared to acetone,
important to note that the Eg values increased upon Act and Eth binding ethanol, CO2, and water molecules. When the SMO is exposed to a
to SnO2–Pd6, and the increase was more in the case of ethanol. It sug reducing target gas (viz., ethanol, acetone), the surface-adsorbed oxygen
gests that Pd loading will produce a higher sensor response for acetone species will oxidize the reducing gas or be replaced by competitive
than ethanol, consistent with the experimental observations. adsorption, which will release the captured electrons to the conduction
The DFT calculated dispersion of isodensities of frontier orbitals of band of the SMO. It leads to an increase in the conductance or decrease
all species (SI Fig. S6) shows two different patterns of electronic distri in the resistance of the sensor substrate. It can be noted that Eth system
butions. The HOMO and LUMO were mainly distributed over the sub has a lower ω and the pristine SnO2-Eth system has a higher Ead energy
strate, SnO2–Pd6, when there was a weak interaction and small change and a lower ω than the pristine SnO2-Act system (Table 1). It is consistent
in the HOMO-LUMO energy gap. On the other hand, a uniform distri with our experimental observations that ethanol at lower concentrations
bution over the entire complex and notable change in the orbital density showed a marginally higher sensor response than acetone on the pristine
were observed in the cases where there were stronger interactions like SnO2 (see Fig. 7). The effects of the Pd NC loading can be assessed by
O2. Therefore, the Pd NC loading onto SnO2 substrate provides sites for comparing the electrophilicity index values of SnO2 and SnO2–Pd6 sys
co-adsorption of target gas and O2, which is favorable for their reactions. tems (Table 1). One can see that the electron-acceptor tendency in
The HOMO− LUMO energies have also been correlated with the creases upon Pd NC loading, which is mainly determined by the
chemical stability and reactivity of various chemical systems [103,104]. electrophilicity of the Pd6 NC. O2 binding further increases the elec
The knowledge of the chemical reactivities of the ligand-bound clusters trophilicity while the other ligands decrease it. The decrease in the
is valuable in the context of the response/recovery times, operating electrophilicity index due to the binding of the ligands onto SnO2–Pd6
temperature, etc. Parameters like the chemical hardness (η) and Mul followed the order: Act > Eth > H2O > CO2. It is to be noted that the
liken electronegativity (χ), which have been defined as η = (ELUMO – presence of the Pd NC caused a reversal in the electrophilicity of the
EHOMO)/2 = Eg/2 and χ = − (ELUMO + EHOMO)/2 (= − μ, electronic ethanol and acetone systems with respect to their pristine systems. It
chemical potential), are useful reactivity descriptor parameters [103, suggests that the electron donation tendency of acetone increases due to
104]. One can see that while the Mulliken electronegativity represents the Pd loading onto SnO2. Thus, the Pd loading onto the SnO2 substrate
the average valence electron energy, the chemical hardness (η) is a results in acetone selectivity over ethanol and water vapor.
measure of the breadth of the HOMO− LUMO energy difference. In other The changes in the total polarizability of the ligand complexes were
words, the chemical hardness (η) represents the resistance to the elec computed. The polarizability, α, of a species was calculated as follows: α
tron distribution change and hence the chemical stability and reactivity = (αXX + αYY + αZZ)/3, where αXX, αYY, and αZZ are polarizability com
of a system. Thus, the clusters with smaller Eg values correspond to ponents along the X, Y, and Z axes, respectively. The change in total
lower chemical hardness (i.e., conversely, higher global softness). In polarizability, Δα, was defined as: Δα = [α(S-L) – α(S) – α(L)], where α(i)
other words, smaller Eg value clusters are softer where electron density stands for the polarizability (α) of species, i, and S = SnO2, Pd6 or
changes easily, showing higher chemical reactivity and vice versa. SnO2–Pd6, and L = Act, Eth, O2, and H2O). Δα values are given in
It can be seen that SnO2 with a high Eg value is chemically relatively Table 1. It can be seen that the Pd NC loading affected the polarizability
stable. In fact, the higher chemical stability of the sensor building block of the substrate clusters to different extents depending on the system.
is essential for designing robust SMO-based sensors. On the other hand, The polarizability change was significant in the cases of O2 adsorption
Pd6 NC, as expected, is highly reactive or chemically less stable. How onto the substrate and acetone or ethanol binding. It should be noted
ever, as discussed above, in the context of conductivity change, the that the Pd NC loading caused a considerable difference in the polariz
comparison of the SnO2 and SnO2-L systems showed hardly any change ability of the acetone and ethanol systems. Since softer molecular sys
in the Eg values upon ligand binding to SnO2 (except the SnO2–O2 and tems where electron density changes more easily could show higher
SnO2-Act cases). A comparison of the Eg values of SnO2–Pd6-L and SnO2- chemical reactivity, the higher polarizability could be one of the factors
L systems showed that the SnO2–Pd6-L systems have relatively low Eg of the higher acetone selectivity of Pd NC-loaded SnO2 substrates over
values. Based on the chemical hardness (η) values, it can also be ethanol compared to the pristine SnO2 substrates. Selectivity is a crucial
concluded that the presence of the Pd NC loading is favorable for the factor in sensing, especially in the presence of gases with similar phys
higher chemical reactivities of the corresponding systems, which can icochemical properties to the target gas.
significantly improve the sensor performance like the recovery time,
operation temperature, etc. It also implies that Pd NC loading onto SnO2 3.3.3. Mulliken charges
substrate improves the acetone sensing performance compared to the Let us recall that the conductivity change arising from the bandgap
pristine SnO2. (Eg) change due to the ligand-binding was not favorable for sensing in
In these sensing reactions, reducing gases are the electron-donors, the case of most ligands. The electronic charge transfer between the
while the chemisorbed O2 species are the electron-acceptors. There substrate and the adsorbed molecules plays a critical role in determining
fore, the energy gap between the HOMO of the ligand adsorbed system the conductivity and the performance of the chemoresistive sensors
and the LUMO of the O2-adsorbed system can shed light on this charge [105]. We have employed the Mulliken population analysis to evaluate
donation process. It can be seen that the Pd NC loading raises the HOMO the direction and magnitude of the electronic charge transfer due to the
140
A. Biswas et al. Current Applied Physics 44 (2022) 131–143
various ligand binding. It can be seen that ligand O2 acts as the charge Table 3
acceptor, and the other ligands act as the charge donors (Table 3). The Calculated Mulliken charge transfer from the substrate (S) to ligand (L).
charge withdrawal due to the oxygen adsorption is vital in the System (S-L) Charge (e) transferred to Ligand (L)
SMO-based chemoresistive gas sensors. The Mulliken charge analysis
SnO2–O2 − 0.021
showed that O2 adsorption onto the pristine SnO2 caused only a small SnO2-Act 0.198
charge separation. The presence of Pd NC caused large charge separa SnO2-Eth 0.141
tions between the substrate and the adsorbed O2. Since the reaction of SnO2–H2O 0.097
the adsorbed oxygen species with the target gas releases back the SnO2–CO2 0.049
Pd6–O2 − 0.283
negative charge to the substrate, the higher charge separation gives rise Pd6–H2O 0.073
to a greater change in the electrical conductivity and hence a higher Pd6–CO2 0.035
sensor response. Pd NC loading enhanced the charge separation between Pd6-Act 0.125
the donor ligands and the loaded substrates. A scrutiny of the charge Pd6-Eth 0.108
SnO2–Pd6 0.443
distributions of the SnO2–Pd6-L systems showed that Pd loading caused
SnO2–Pd6–O2 − 0.284
higher charge separations in acetone and ethanol systems compared to SnO2–Pd6-Act 0.234
the other donor ligands, namely, H2O and CO2 systems. Thus, one can SnO2–Pd6-Eth 0.231
conclude that the sensor performance improvement involves several SnO2–Pd6–H2O 0.135
factors like the Ead, Eg, reactivity, charge separation, etc. SnO2–Pd6–CO2 0.061
141
A. Biswas et al. Current Applied Physics 44 (2022) 131–143
[10] X. Gao, T. Zhang, An overview: facet-dependent metal oxide semiconductor gas [39] T.D.C. Minh, D.R. Blake, P.R. Galassetti, The clinical potential of exhaled breath
sensors, Sens. Actuators, B 277 (2018) 604–633. analysis for diabetes mellitus, Diabetes Res. Clin. Pract. 97 (2) (2012) 195–205.
[11] X.-Z. Song, F.-F. Sun, S.-T. Dai, X. Lin, K.-M. Sun, X.-F. Wang, Hollow NiFe2O4 [40] S. Ghosal, P. Bhattacharyya, Irreversible n to p transition and corresponding
microspindles derived from Ni/Fe bimetallic MOFs for highly sensitive acetone performance improvement of RGO/TiO2 nanotubes hybrid vapor sensor devices
sensing at low operating temperatures, Inorg. Chem. Front. 5 (5) (2018) by varying electrophoretic deposition time, IEEE Trans. Nanotechnol. 17 (6)
1107–1114. (2018) 1098–1105.
[12] J. Lu, Y. Xie, F. Luo, H. Fu, X. Huang, Y. Liu, H. Liu, Heterostructures of [41] N. Teshima, J. Li, K. Toda, P.K. Dasgupta, Determination of acetone in breath,
mesoporous hollow Zn2SnO4/SnO2 microboxes for high-performance acetone Anal. Chim. Acta 535 (1–2) (2005) 189–199.
sensors, J. Alloys Compd. 844 (2020), 155788. [42] A. Koo, R. Yoo, S.P. Woo, H.-S. Lee, W. Lee, Enhanced acetone-sensing properties
[13] M. Yin, Y. Wang, S. Liu, Synthesis of Fe2O3–ZnWO4 nanocomposites and their of Pt-decorated al-doped ZnO nanoparticles, Sens. Actuators, B 280 (2019)
enhanced acetone sensing performance, J. Alloys Compd. 831 (2020), 154713. 109–119.
[14] X. Kou, N. Xie, F. Chen, T. Wang, L. Guo, C. Wang, Q. Wang, J. Ma, Y. Sun, [43] Q. Jia, H. Ji, Y. Zhang, Y. Chen, X. Sun, Z. Jin, Rapid and selective detection of
H. Zhang, G. Lu, Superior acetone gas sensor based on electrospun SnO2 acetone using hierarchical ZnO gas sensor for hazardous odor markers
nanofibers by Rh doping, Sens. Actuators, B 256 (2018) 861–869. application, J. Hazard Mater. 276 (2014) 262–270.
[15] S. Shao, H. Wu, S. Wang, Q. Hong, R. Koehn, T. Wu, W.-F. Rao, Highly crystalline [44] S.T. Navale, Z.B. Yang, C. Liu, P.J. Cao, V.B. Patil, N.S. Ramgir, R.S. Mane, F.
and ordered nanoporous SnO2 thin films with enhanced acetone sensing property J. Stadler, Enhanced acetone sensing properties of titanium dioxide nanoparticles
at room temperature, J. Mater. Chem. C Mater. Opt. Electron. Devices 3 (41) with a sub-ppm detection limit, Sens. Actuators, B 255 (2018) 1701–1710.
(2015) 10819–10829. [45] H. Zhang, H. Qin, P. Zhang, Y. Chen, J. Hu, Low concentration acetone gas
[16] M. Parthibavarman, B. Renganathan, D. Sastikumar, Development of high sensing properties of 3 Wt% Pd-doped SmCoxFe1-XO3 nanocrystalline powders
sensitivity ethanol gas sensor based on Co-doped SnO2 nanoparticles by under UV light illumination, Sens. Actuators, B 260 (2018) 33–41.
microwave irradiation technique, Curr. Appl. Phys. 13 (7) (2013) 1537–1544. [46] H. Haick, Y.Y. Broza, P. Mochalski, V. Ruzsanyi, A. Amann, Assessment, origin,
[17] S.-W. Tsai, J.-C. Chiou, Improved crystalline structure and H2S sensing and implementation of breath volatile cancer markers, Chem. Soc. Rev. 43 (5)
performance of CuO–Au–SnO2 thin film using SiO2 additive concentration, Sens. (2014) 1423–1449.
Actuators, B 152 (2) (2011) 176–182. [47] M. Righettoni, A. Tricoli, S.E. Pratsinis, Si WO3 sensors for highly selective
[18] J. Yang, J. Liu, Y. Xu, X. Li, J. Wu, Y. Han, Z. Wang, X. Zhang, Enhanced selective detection of acetone for easy diagnosis of diabetes by breath analysis, Anal.
acetone-sensing performance of hierarchical hollow SnO2/α-Fe2O3 microcubes, Chem. 82 (2010) 3581–3587.
J. Mater. Chem. C Mater. Opt. Electron. Devices 7 (38) (2019) 11984–11990. [48] P. Zhu, S. Li, X. Jiang, Q. Wang, F. Fan, M. Yan, Y. Zhang, P. Zhao, J. Yu,
[19] M. Epifani, S. Kaciulis, A. Mezzi, T. Zhang, J. Arbiol, P. Siciliano, A. Landström, Noninvasive and wearable respiration sensor based on organic semiconductor
I. Concina, A. Moumen, E. Comini, C. Xiangfeng, Rhodium as efficient additive for film with strong electron affinity, Anal. Chem. 91 (15) (2019) 10320–10327.
boosting acetone sensing by TiO2 nanocrystals. Beyond the classical view of noble [49] P. Zhu, Y. Wang, P. Ma, S. Li, F. Fan, K. Cui, S. Ge, Y. Zhang, J. Yu, Low-power
metal additives, Sens. Actuators, B 319 (2020), 128338. and high-performance trimethylamine gas sensor based on n-n heterojunction
[20] S. Wei, S. Li, R. Wei, S. Liu, W. Du, Different morphologies of WO3 and their microbelts of perylene diimide/CdS, Anal. Chem. 91 (9) (2019) 5591–5598.
exposed facets-dependent acetone sensing properties, Sens. Actuators, B 329 [50] Z. Zhao, Y. Yong, R. Gao, S. Hu, Q. Zhou, X. Su, Y. Kuang, X. Li, Adsorption,
(2021), 129188. sensing and optical properties of molecules on BC3 monolayer: first-principles
[21] Y.J. Hong, J.-W. Yoon, J.-H. Lee, Y.C. Kang, One-pot synthesis of Pd-loaded SnO2 calculations, Mater. Sci. Eng. B 271 (2021), 115266.
yolk-shell nanostructures for ultraselective methyl benzene sensors, Chemistry 20 [51] X. Wang, Y. Yong, W. Yang, A. Zhang, X. Xie, P. Zhu, Y. Kuang, Adsorption, gas-
(10) (2014) 2737–2741. sensing, and optical properties of molecules on a diazine monolayer: a first-
[22] S. Yang, G. Lei, H. Xu, Z. Lan, Z. Wang, H. Gu, Metal oxide based heterojunctions principles study, ACS Omega 6 (17) (2021) 11418–11426.
for gas sensors: a Review, Nanomaterials 11 (4) (2021) 1026. [52] W. Chen, Y. Gui, T. Li, H. Zeng, L. Xu, Z. Ding, Gas-sensing properties and
[23] B. Saruhan, R. Lontio Fomekong, S. Nahirniak, Review: influences of mechanism of Pd-GaNNTs for air decomposition products in ring main unit, Appl.
semiconductor metal oxide properties on gas sensing characteristics, Front. Sens. Surf. Sci. 531 (2020), 147293.
2 (2021), 657931. [53] Y. Yong, X. Su, Y. Kuang, X. Li, Z. Lu, B40 and M@B40 (M=Li and Ba) fullerenes as
[24] C. Liu, Q. Kuang, Z. Xie, L. Zheng, The effect of noble metal (Au, Pd and Pt) potential molecular sensors for acetone detection: a first-principles study, J. Mol.
nanoparticles on the gas sensing performance of SnO2-based sensors: a case study Liquids 264 (2018) 1–8.
on the 221 high-index faceted SnO2octahedra, CrystEngComm 17 (33) (2015) [54] Y. Yong, X. Su, H. Cui, Q. Zhou, Y. Kuang, X. Li, Two-dimensional tetragonal GaN
6308–6313. as potential molecule sensors for no and NO2 detection: a first-principle study,
[25] S.M. Majhi, P. Rai, Y.T. Yu, Facile approach to synthesize Au@ZnO core-Shell ACS Omega 2 (2017) 8888–8895.
nanoparticles and their application for highly sensitive and selective gas sensors, [55] L.E. Gálvez-González, J.A. Alonso, L.O. Paz-Borbón, A. Posada-Amarillas, H2
ACS Appl. Mater. Interfaces 7 (2015) 9462–9468. adsorption on Cu4-XMX (M = Au, Pt; X = 0–4) clusters: similarities and differences
[26] L. Xiao, S. Shu, S. Liu, A facile synthesis of Pd-doped SnO2 hollow microcubes as predicted by density functional theory, J. Phys. Chem. C 123 (51) (2019)
with enhanced sensing performance, Sens. Actuators, B 221 (2015) 120–126. 30768–30780.
[27] D. Singh, V.S. Kundu, A.S. Maan, Structural, morphological and gas sensing study [56] G. Li, X. Chen, Z. Zhou, F. Wang, H. Yang, J. Yang, B. Xu, B. Yang, D. Liu,
of palladium doped tin oxide nanoparticles synthesized via hydrothermal Theoretical insights into the structural, relative stable, electronic, and gas sensing
technique, J. Mol. Struct. 1100 (2015) 562–569. properties of PbnAun (n = 2–12) clusters: a DFT study, RSC Adv. 7 (72) (2017)
[28] B.-Y. Kim, J.S. Cho, J.-W. Yoon, C.W. Na, C.-S. Lee, J.H. Ahn, Y.C. Kang, J.-H. Lee, 45432–45441.
Extremely sensitive ethanol sensor using Pt-doped SnO2 hollow nanospheres [57] A. Granja-DelRío, J.A. Alonso, M.J. López, Competition between palladium
prepared by Kirkendall diffusion, Sens. Actuators, B 234 (2016) 353–360. clusters and hydrogen to saturate graphene vacancies, J. Phys. Chem. C 121 (20)
[29] K.D. Schierbaum, U.K. Kirner, J.F. Geiger, W. Göpel, Schottky-barrier and (2017) 10843–10850.
conductivity gas sensors based upon Pd/SnO2 and Pt/TiO2, Sens. Actuators, B 4 [58] M. Blanco-Rey, J.I. Juaristi, M. Alducin, M.J. López, J.A. Alonso, Is spillover
(1–2) (1991) 87–94. relevant for hydrogen adsorption and storage in porous carbons doped with
[30] M. Batzill, U. Diebold, The surface and materials science of tin oxide, Prog. Surf. palladium nanoparticles? J. Phys. Chem. C 120 (31) (2016) 17357–17364.
Sci. 79 (2–4) (2005) 47–154. [59] Y. Yong, C. Li, X. Li, T. Li, H. Cui, S. Lv, Ag7Au6 cluster as a potential gas sensor
[31] S. Matsushima, Y. Teraoka, N. Miura, N. Yamazoe, Electronic interaction between for CO, HCN, and NO detection, J. Phys. Chem. C 119 (13) (2015) 7534–7540.
metal additives and tin dioxide in tin dioxide-based gas sensors, Jpn. J. Appl. [60] M. Hu, D.P. Linder, M. Buongiorno Nardelli, A. Striolo, Hydrogen adsorption on
Phys. 27 (10R) (1988) 1798–1802. platinum–gold bimetallic nanoparticles: a density functional theory study,
[32] X. Yang, H. Gao, L. Zhao, T. Wang, P. Sun, F. Liu, G. Lu, Enhanced gas sensing J. Phys. Chem. C 117 (29) (2013) 15050–15060.
properties of monodisperse Zn2SnO4 octahedron functionalized by PdO [61] T.M. Inerbaev, Y. Kawazoe, S. Seal, Theoretical calculations of hydrogen
nanoparticals, Sens. Actuators, B 266 (2018) 302–310. adsorption by SnO2 (110) surface: effect of doping and calcination, J. Appl. Phys.
[33] F. Gyger, A. Sackmann, M. Hübner, P. Bockstaller, D. Gerthsen, H. Lichtenberg, 107 (10) (2010), 104504.
J.-D. Grunwaldt, N. Barsan, U. Weimar, C. Feldmann, Pd@SnO2 and SnO2@Pd [62] I. Swart, F.M.F. de Groot, B.M. Weckhuysen, P. Gruene, G. Meijer, A. Fielicke, H2
Core@shell nanocomposite sensors, Part. Part. Syst. Char. 31 (5) (2014) 591–596. adsorption on 3d transition metal clusters: a combined infrared spectroscopy and
[34] K. Musa-Veloso, S.S. Likhodii, E. Rarama, S. Benoit, Y.-M.C. Liu, D. Chartrand, density functional study, J. Phys. Chem. A 112 (6) (2008) 1139–1149.
R. Curtis, L. Carmant, A. Lortie, F.J.E. Comeau, S.C. Cunnane, Breath acetone [63] H. Köse, A.O. Aydin, H. Akbulut, The effect of temperature on grain size of SnO2
predicts plasma ketone bodies in children with epilepsy on a ketogenic diet, nanoparticles synthesized by sol-gel method, Acta Phys. Pol., A 125 (2) (2014)
Nutrition 22 (1) (2006) 1–8. 345–347.
[35] S.S. Likhodii, K. Musa, S.C. Cunnane, Breath acetone as a measure of systemic [64] N. Ma, K. Suematsu, M. Yuasa, T. Kida, K. Shimanoe, Effect of water vapor on Pd-
ketosis assessed in a rat model of the ketogenic diet, Clin. Chem. 48 (1) (2002) loaded SnO2 nanoparticles gas sensor, ACS Appl. Mater. Interfaces 7 (10) (2015)
115–120. 5863–5869.
[36] T. Toyooka, S. Hiyama, Y. Yamada, A prototype portable breath acetone analyzer [65] N. Ma, K. Suematsu, M. Yuasa, K. Shimanoe, Pd size effect on the gas sensing
for monitoring fat loss, J. Breath Res. 7 (3) (2013), 036005. properties of Pd-loaded SnO2 in humid atmosphere, ACS Appl. Mater. Interfaces 7
[37] Z. Wang, C. Wang, Is breath acetone a biomarker of diabetes? A historical Review (28) (2015) 15618–15625.
on breath acetone measurements, J. Breath Res. 7 (3) (2013), 037109. [66] C. Wang, X. Cui, J. Liu, X. Zhou, X. Cheng, P. Sun, X. Hu, X. Li, J. Zheng, G. Lu,
[38] A. Thati, A. Biswas, S. Roy Chowdhury, T.K. Sau, Breath acetone-based Design of superior ethanol gas sensor based on Al-doped NiO nanorod-flowers,
noninvasive detection of blood glucose levels, Int. J. Smart Sens. Intell. Syst. 8 (2) ACS Sens. 1 (2) (2016) 131–136.
(2015) 1244–1260. [67] N. Yamazoe, New approaches for improving semiconductor gas sensors, Sens.
Actuators, B 5 (1–4) (1991) 7–19.
142
A. Biswas et al. Current Applied Physics 44 (2022) 131–143
[68] M.D. Obradović, Z.M. Stanćić, U.Č Lačnjevac, V.V. Radmilović, A. Gavrilović- [86] R.G. Drabeski, J.V. Gunha(Visualization), A. Novatski, G.B. de Souza, S.
Wohlmuther, V.R. Radmilović, S. Lj Gojković, Electrochemical oxidation of M. Tebcherani, E.T. Kubaski, D.T. Dias, Raman and photoacoustic spectroscopies
ethanol on palladium-nickel nanocatalyst in alkaline media, Appl. Catal. B of SnO2 thin films deposited by spin coating technique, Vib. Spectrosc. 109
Environ. 189 (2016) 110–118. (2020), 103094.
[69] P. Duan, Q. Duan, Q. Peng, K. Jin, J. Sun, Design of ultrasensitive gas sensor based [87] A. Dieguez, A. Romano-Rodriguez, J.R. Morante, U. Weimar, M. Schweizer-
on self-assembled Pd-SnO2/rGO porous ternary nanocomposites for ppb-level Berberich, W. Göpel, Morphological analysis of nanocrystalline SnO2 for gas
hydrogen, Sens. Actuators, B 369 (2022), 132280. sensor applications, Actuators B Chem 31 (1–2) (1996) 1–8.
[70] N. Mohri, B. Oschmann, N. Laszczynski, F. Mueller, J. von Zamory, M.N. Tahir, [88] B.A. Banse, B.E. Koel, Interaction of oxygen with Pd(111): high effective O2
S. Passerini, R. Zentel, W. Tremel, Synthesis and characterization of carbon pressure conditions by using nitrogen dioxide, Surf. Sci. 232 (3) (1990) 275–285.
coated sponge-like tin oxide (SnOx) films and their application as electrode [89] J.G.S. Carrettin, A. Corma, Spectroscopic evidence for the supply of reactive
materials in lithium-ion batteries, J. Mater. Chem. 4 (2016) 612–619. oxygen during CO oxidation catalyzed by gold supported on nanocrystalline
[71] K. Vijayarangamuthu, S. Rath, Nanoparticle size, oxidation state, and sensing CeO2, J. Am. Chem. Soc. 127 (10) (2005) 3286–3287.
response of tin oxide nanopowders using Raman spectroscopy, J. Alloys Compd. [90] R.Q. Long, H.L. Wan, In situ confocal microprobe Raman spectroscopy study of
610 (2014) 706–712. CeO2/BaF2 catalyst for the oxidative coupling of methane, J. Chem. Soc., Faraday
[72] N. Chen, D. Deng, Y. Li, X. Liu, X. Xing, X. Xiao, Y. Wang, TiO2 nanoparticles Trans. 93 (2) (1997) 355–358.
functionalized by Pd nanoparticles for gas-sensing application with enhanced [91] R.Q. Long, Y.P. Huang, H.L. Wan, Surface oxygen species over cerium oxide and
butane response performances, Sci. Rep. 7 (2017) 7692. their reactivities with methane and ethane by means of in situ confocal
[73] C. Wang, N. Xu, T.-T. Liu, W. Xu, H. Guo, Y. Li, P. Bai, X.-P. Wu, X.-Q. Gong, microprobe Raman spectroscopy, J. Raman Spectrosc. 28 (1) (1997) 29–32.
X. Liu, S. Mintova, Mechanical pressure-mediated Pd active sites formation in [92] H. Nakatsuji, H. Nakai, Y. Fukunishi, Dipped adcluster model for chemisorptions
NaY zeolite catalysts for indirect oxidative carbonylation of methanol to dimethyl and catalytic reactions on a metal surface: image force correction and
carbonate, J. Catal. 396 (2021) 269–280. applications to Pd–O2 adclusters, J. Chem. Phys. 95 (1) (1991) 640–647.
[74] J.-C. Dupin, D. Gonbeau, P. Vinatier, A. Levasseur, Systematic XPS studies of [93] E. Hasselbrink, H. Hirayama, A. de Meijere, F. Weik, M. Wolf, G. Ertl,
metal oxides, hydroxides and peroxides, Phys. Chem. Chem. Phys. 2 (2000) Photodissociation and photodesorption of O2 adsorbed on Pd(111), Surf. Sci.
1319–1324. 269–270 (1992) 235–246.
[75] L. Mazeina, Y.N. Picard, J.D. Caldwell, E.R. Glaser, S.M. Prokes, Growth and [94] R. Imbihl, J.E. Demuth, Adsorption of oxygen on a Pd(111) surface studied by
photoluminescence properties of vertically aligned SnO2 nanowires, J. Cryst. high resolution electron energy loss spectroscopy (EELS), Surf. Sci. 173 (2–3)
Growth 311 (2009) 3158–3162. (1986) 395–410.
[76] N.C. Horti, M.D. Kamatagi, N.R. Patil, M.N. Wari, S.R. Inamdar, [95] A.A. Rydosz, Negative correlation between blood glucose and acetone measured
Photoluminescence properties of SnO2 nanoparticles: effect of solvents, Optik – in healthy and type 1 diabetes mellitus patient breath, J. Diabetes Sci. Technol. 9
Int. J. Light Electron Optics 169 (2018) 314–320. (4) (2015) 881–884.
[77] B.V. Ramana, A. Das, S. Amirthapandian, S.K. Dhara, A.K. Tyagi, [96] M.C. Valero, P. Raybaud, P. Sautet, Interplay between molecular adsorption and
Photoluminescence of oxygen vacancies and hydroxyl group surface metal–support interaction for small supported metal clusters: CO and C2H4
functionalized SnO2 nanoparticles, Phys. Chem. Chem. Phys. 17 (2015) adsorption on Pd4/γ-Al2O3, J. Catal. 247 (2) (2007) 339–355.
9794–9801. [97] O.C. Gagne, F.C. Hawthorne, Bond-length distributions for ions bonded to
[78] J. Jeong, S.-P. Choi, C.I. Chang, D.C. Shin, J.S. Park, B.-T. Lee, Y.-J. Park, H.- oxygen: results for the transition metals and quantification of the factors
J. Song, Photoluminescence properties of SnO2 thin films grown by thermal CVD, underlying bond-length variation in inorganic solids, IUCrJ 7 (2020) 581–629.
Solid State Commun. 127 (2003) 595–597. [98] R. Huacuja, D.J. Graham, C.M. Fafard, C.-H. Chen, B.M. Foxman, D.E. Herbert,
[79] H. Luo, L.Y. Liang, H.T. Cao, Z.M. Liu, F. Zhuge, Structural, chemical, optical, and G. Alliger, C.M. Thomas, O.V. Ozerov, Reactivity of a Pd(I)-Pd(I) dimer with O2:
electrical evolution of SnO(x) films deposited by reactive Rf magnetron monohapto Pd superoxide and dipalladium peroxide in equilibrium, J. Am.
sputtering, ACS Appl. Mater. Interfaces 4 (10) (2012) 5673–5677. Chem. Soc. 133 (11) (2011) 3820–3823.
[80] E. Fazio, F. Neri, S. Savasta, S. Spadaro, S. Trusso, Surface-enhanced Raman [99] R. Long, K. Mao, X. Ye, W. Yan, Y. Huang, J. Wang, Y. Fu, X. Wang, X. Wu, Y. Xie,
scattering of SnO2 bulk material and colloidal solutions, Phys. Rev. B Condens. Y. Xiong, Surface facet of palladium nanocrystals: a key parameter to the
Matter 85 (19) (2012). activation of molecular oxygen for organic catalysis and cancer treatment, J. Am.
[81] V.V. Pushkarev, V.I. Kovalchuk, J.L. d’Itri, Probing defect sites on the CeO2 Chem. Soc. 135 (8) (2013) 3200–3207.
surface with dioxygen, J. Phys. Chem. B 108 (17) (2004) 5341–5348. [100] Q. Ren, Y.-Q. Cao, D. Arulraj, C. Liu, D. Wu, W.-M. Li, A.-D. Li, Review—resistive-
[82] X. Mathew, J.P. Enriquez, C. Mejía-García, G. Contreras-Puente, M.A. Cortes- type hydrogen sensors based on zinc oxide nanostructures, J. Electrochem. Soc.
Jacome, J.A. Toledo Antonio, J. Hays, A. Punnoose, Structural modifications of 167 (6) (2020), 067528.
SnO2 due to the incorporation of Fe into the lattice, J. Appl. Phys. 100 (2006), [101] S.S. Li, Semiconductor Physical Electronics, Springer, New York, NY, 2006.
073907. [102] Z. Zhao, Y. Yong, Q. Zhou, Y. Kuang, X. Li, Gas-sensing properties of the SiC
[83] S.H. Sun, G.W. Meng, G.X. Zhang, T. Gao, B.Y. Geng, L.D. Zhang, J. Zuo, Raman monolayer and bilayer: a density functional theory study, ACS Omega 5 (21)
scattering study of rutile SnO2 nanobelts synthesized by thermal evaporation of (2020) 12364–12373.
Sn powders, Chem. Phys. Lett. 376 (1–2) (2003) 103–107. [103] R.G. Parr, R.G. Pearson, Absolute hardness: companion parameter to absolute
[84] M. Aliahmad, M. Dehbashi, Ni-doped SnO nanoparticles synthesized by chemical electronegativity, J. Am. Chem. Soc. 105 (26) (1983) 7512–7516.
Co-precipitation method, Iran. J. Energy Environ. (2013) 49–52. [104] B. Konate, S.T. Affi, N. Ziao, DFT study of dimerization sites in imidazo[1,2-a]
[85] M.X. Gu, L.K. Pan, B.K. Tay, C.Q. Sun, Atomistic origin and temperature Pyridinyl-chalcone series, Comput. Chem. (2021) 1–17, 09 (01).
dependence of Raman optical redshift in nanostructures: a broken bond rule, [105] T. Jiang, Q. He, M. Bi, X. Chen, H. Sun, L. Tao, First-principles calculations of
J. Raman Spectrosc. 38 (6) (2007) 780–788. adsorption sensitivity of Au-doped MoS2 gas sensor to main characteristic gases in
oil, J. Mater. Sci. 56 (2021) 13673–13683.
143