Controlled Loading of MoS2 On Hierarchical Porous TiO2 For Enhanced Photocatalytic Hydrogen Evolution - Tiwari - 2021

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 13

pubs.acs.

org/JPCC Article

Controlled Loading of MoS2 on Hierarchical Porous TiO2 for


Enhanced Photocatalytic Hydrogen Evolution
Amritanjali Tiwari,* Amit Gautam, Saddam Sk, Deepak S. Gavali, Ranjit Thapa, and Ujjwal Pal*
Cite This: J. Phys. Chem. C 2021, 125, 11950−11962 Read Online

ACCESS Metrics & More Article Recommendations *


sı Supporting Information
See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

ABSTRACT: In this study, we garnered three important factors


simultaneously, namely, wormhole mesoporosity of TiO2 with well-
designed interfaces for effective charge transfers, precise loading of
MoS2 for plasmon induction, and increased surface area with
Downloaded via TSINGHUA UNIV on August 11, 2021 at 02:51:45 (UTC).

exposed surface atoms and active sites. The controlled loading of


MoS2 on porous TiO2 (MPT) forms a heterojunction that
effectively modulates the interface engineering and thereby greatly
enhances hydrogen evolution. The synthesis of a photocatalyst is
based on a simple hydrothermal process that is well characterized.
The resulting composite materials were tested for hydrogen
evolution reactions. At optimum loading, MPT10 induced a
maximum hydrogen evolution rate of 1376 μmol h−1 g−1 with
2.28% apparent quantum yield (AQY), which was 10-fold higher
compared to the MCT10 (MoS2-commercial TiO2) H2 evolution rate of 138 μmol h−1 g−1 with 0.23% AQY under similar reaction
conditions. The shorter decay component, lower emission intensity, and higher estimated lifetime of MPT10 suggest its superiority
over other materials. Density functional theory (DFT) calculations have further revealed the active sites of MPT and hierarchical
porous TiO2 (HPT) to support the experimental hydrogen evolution reaction (HER). This study suggests an avenue to design an
efficient noble-metal-free photocatalyst for solar fuel productions.

■ INTRODUCTION
Due to the decline in nonrenewable resources and increasing
with minimized use of precious metals has become an
increasingly important topic.
environmental pollution, significant attention has been given to The design of a binary support through controlled synthesis
renewable and clean energy domains. Hydrogen (H2) is enabled a greater number of active sites in HER, such as two-
considered as one of the most suitable energy carriers due to dimensional molybdenum sulfide (2D MoS2), which possesses
its higher energy density per unit mass in comparison to the distinctive optoelectronic and catalytic properties and shows a
other chemical fuels. In recent times, photocatalytic fission of wide spectral response ranging from visible to near-IR light. It
water has been considered an attractive solution for solar-to- improves the charge separation during photoirradiation of the
chemical H2 energy conversion.1−3 Therefore, the develop- materials.9,10 The controlled synthesis of 2D metal chalcoge-
ment of an excellent, stable, efficient, and economical nides such as MoS2, upon band-gap engineering in the
photocatalyst for ultrahigh H2 production efficiency is resultant heterostructure composites, gives rise to active sites
paramount to researchers. that enhance the photocatalytic reactions.11,12 The hetero-
Ever since the breakthrough research on H2 photogeneration structured hybrid materials can be made by judicious tuning of
from water using TiO2 under UV-light irradiation,4 an MoS2 and TiO2 morphology using specific synthesis processes.
enormous amount of research has been conducted on Further, it is noted that MoS2 possesses a direct band gap (1.90
photochemical H2 evolution using different semiconductor- eV),13−15 high carrier mobility,16 etc., which lead to versatile
based photocatalysts.5 However, the major drawbacks of TiO2 applications such as optoelectronic and nanoelectronic
are its wide band gap (3.2 eV) and its fast charge annihilation devices.17−19 The nanosized MoS2 edge defects favored H2
at photoinduced electron−hole pairs. Furthermore, Pt and
expensive noble metals are frequently used as cocatalysts for Received: March 3, 2021
water-splitting studies. A diverse range of supports and Revised: May 14, 2021
techniques have been employed, such as doping, band-gap Published: May 27, 2021
engineering, and use of different cocatalysts (Pt, Au, Ag, etc.),
for a higher reduction of water.6−8 Therefore, the quest for
more effective hydrogen evolution reaction (HER) catalysts

© 2021 American Chemical Society https://doi.org/10.1021/acs.jpcc.1c01922


11950 J. Phys. Chem. C 2021, 125, 11950−11962
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

adsorption.20−23 The sluggish electron-transfer process in TiO2 recombination, and sustainable H2 production over prolonged
is usually addressed using a proper cocatalyst. In this context, light irradiation due to the synergic effect of flowerlike MoS2
the main investigation here is the band-gap engineering of the and the fibrous wormhole mesoporous channel of titanium
TiO2 particulate for more light harvesting in the HER process. dioxide. The highest HER is obtained using MPT10 (MoS2 and
We then exploited the 2D structure of MoS2 nanoscale flakes HPT with a 1:10 weight ratio) at a rate of 1376 μmol h−1 g−1
with a few layers that allowed unprecedented control over bulk with 2.28% apparent quantum yield at 420 nm using the
materials, resulting in improved active sites and H2 evolution sacrificial electron-donor (SED) mixture of Na2S and Na2SO3.
performance.24−32 The layered structure of MoS2 formation Moreover, the photocurrent density of the MPT10 photo-
occurs through the Mo atom where electrons are sandwiched
catalyst is 0.55 μA cm−2, which is nearly 55 times higher than
between the two s-orbitals of the molecule, and its dimension
HPT (0.01 μA cm−2), 7.0 times greater compared to MPT5
reduced to the size of a monolayer on its transformation of the
band-gap pattern from indirect to direct energy gap. Therefore, (0.08 μA cm−2), about 1.7 times greater compared to MPT15
using MoS2 nanoscale flakes with a few layers is an important (0.33 μA cm−2), and 2.2 times higher than MPT20 (0.25 μA
strategy to enhance H2 production.24,25,33 In recent studies, cm−2). Hence, the present study enabled a non-noble approach
MoS2 was proved to have potential as a light-harvesting to develop intimate and large contact interfaces for a robust
semiconductor and exhibited some extent of plasmonic and efficient solar hydrogen generation system.


properties, which could be one of the key factors working
synergistically at the interface of junction architectures.24,34−36 METHODS
Accordingly, it is expected that MoS2-loaded semiconductors
Materials. Titanium(IV) isopropoxide (Ti{OCH(CH3)2}4,
would greatly enhance the photocatalytic activity and HER.
Several reports of MoS2−TiO2 heterostructured nanocompo- 99.99%), ammonium molybdate tetrahydrate
sites have been explored for improving the photocatalytic {(NH 4 ) 6 Mo 7 O 24 ·4H 2 O, 99.99%}, and thiourea (TU)
hydrogen evolution by modulating the titanium dioxide (NH2CSNH2, 99%) were procured from Sigma-Aldrich.
morphology and band gap through interband interac- Absolute ethanol and deionized (DI) water were supplied by
tion.35,37,38 For example, Xiang et al. reported that increasing Merck, Germany. All of the chemicals used in the present
the specific surface area of MoS2 nanosheets on TiO2-graphene study were used as received without further purification.
hybrid structures increased the photocatalytic activity with a Syntheses of MoS2@HPT Nanostructured Hetero-
H2 production rate of 165.3 μmol h−1 under ultraviolet light composites. The preparation process of hierarchical porous
illumination.38 Further, Liu et al. reported that MoS2 with a TiO2 (HPT) has been adopted from previous reports (details
growth-controlled process and loading of TiO2 had a in the Supporting Information).41,43 Further, the marigold-
substantial impact on H2 generation. They explained that at flower-like MoS2@HPT (MPT) nanostructured heterocompo-
optimized conditions, the TiO2/MoS2(E) composite exhibited sites were successfully prepared using a simple hydrothermal
impressive HER activity at a rate of 4300 μmol h−1 g−1, a much process. First, 0.1 g of ammonium molybdate and 0.4 g of
higher activity (36 times) compared to pristine TiO2.34 On the thiourea (TU) were charged in 50 mL of deionized water. The
other hand, the reduced TiO2 with the MoS2 composite was prepared solution was continuously stirred for 2 h. Then, 1.0 g
reported for the removal of methyl orange (MO) and H2
of the HPT catalyst was mixed into the reaction mixture and
evolution.24 In particular, Yang et al. described plasmonic
stirred at room temperature (RT) for the next 30 min to
properties in MoS2@TiO2 hybrid heterostructures, which
display impressive photocatalytic H2 generation.35 According obtain a uniform reaction mixture. It was then shifted into a
to Lin et al., MoS2/TiO2 at 10 wt % MoS2 and using methanol hydrothermal system (Teflon-lined stainless steel autoclave)
as the sacrificial electron donor showed a H2 evolution of 1.93 and charged at 210 °C for 24 h. Subsequently, the system was
mmol g−1 h−1.39 Yuan et al. reported that the 2D MoS2/g- kept at RT for a natural cooling process. The black precipitates
C3N4 materials exhibited a moderate activity of 1155 μmol h−1 of MPT composites were collected through filtration. The
g−1 with an apparent quantum yield of 6.8%.40 Visible-light- products retrieved after filtration were washed and centrifuged
induced H2 production systems over P-25 TiO2 loaded with at 3500 rpm with distilled water and ethanol and then kept
MoS2 heterostructured nanocomposites are well reported. under a vacuum for removal of the residual reactant at 60 °C
Thus, with reference to our earlier studies, we observed that for 24 h. The obtained MoS2@HPT composite powder was
controlled morphology and tunable optoelectronic properties then dried at 80 °C for 12 h. Other controlled loading
of the photocatalyst play a vital role in electron−hole pair catalysts, namely, MPT photocatalysts, were obtained over a
generation and its further movement. As is known, the precise weight ratio variation of MoS2 and HPT using a similar
application of hierarchical porous TiO2 with a layered MoS2 synthesis strategy of changing the mass ratios of the precursors.
nanocomposite photocatalyst for solar-light-harvesting photo- We further attempted to develop pure MoS2 through the same
catalytic H2 production has not been reported yet. process but without using the HPT (see details in Table S1).
Taking into account all of these critical parameters, we
The resulting MPT nanocomposites were labeled as MPT10,
focused on the synthesis of hierarchical porous TiO2 with
marigold-flower-like MoS2 nanocomposites via a simple and M5PT, MPT1, MPT5, MPT15, MPT20, and MPT25 as per the wt
well-characterized hydrothermal process. In previous reports, % ratio of MoS2 to HPT. The controlled porosity and
the detailed synthesis and photophysical properties of wormhole structural morphology, surface area, and hetero-
hierarchical porous TiO2 (HPT) and its HER capability have junction formation of MoS2/HPT were established using
been described.41,42 To achieve visible-light-harvesting H2 analytical tools like scanning electron microscopy (SEM),
production activity, the prepared MoS2−HPT photocatalyst transmission electron microscopy (TEM), X-ray diffraction
is noble metal free and possesses unique photophysical (XRD), X-ray photoelectron spectroscopy (XPS), and
properties: visible-light response, controlled electron−hole Brunauer−Emmett−Teller (BET).
11951 https://doi.org/10.1021/acs.jpcc.1c01922
J. Phys. Chem. C 2021, 125, 11950−11962
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 1. XRD of HPT, MoS2, and MoS2@HPT nanocomposites.

Figure 2. FESEM images of (a) HPT, (b) MoS2, (c) M5PT, (d) MPT1, (e) MPT5, (f) MPT10, (g) MPT15, (h) MPT20, and (i) MPT25.

■ RESULTS AND DISCUSSION


Phase Analysis. The structural phases of HPT, MoS2, and
Morphology Analysis. The structural morphologies of
HPT, MoS2, and MPT nanocomposites were examined by field
emission scanning electron microscopy (FESEM), and the
MPTx heterocomposites were illustrated from powder XRD,
images obtained are depicted in Figure 2. As shown in Figure
and the results are displayed in Figure 1. The diffraction peaks
2a, the HPT sample contained the wormhole fibrous
of HPT were obtained at 2θ values of 25.3, 38.1, 48.1, 54.1, morphology that led to multiple reflections and drove the
55.3, and 62.9° ascribed to the (101), (103), (200), (105), effective charge transfer toward the H2 production reaction. As
(211), and (204) planes of the anatase TiO2 (JCPDS 21- shown in Figure 2b, the as-synthesized MoS2 contained
1272), respectively.41,42 Further, the bare MoS2 sample reveals marigold-flower-like structures assembled with several thin
three diffraction peaks at 14.1, 33.5, and 59.3° corresponding nanopetals or nanosheets.46,47 The prepared M5PT, MPT1,
to the (002), (100), and (110) planes of the few-layered MPT5, and MPT10 heterocomposites show that some HPT
nanosheets of MoS2 (JCPDS card no. 65-1951).44,45 The nanomaterials were decorated at MoS2 nanosheets (Figure 2c−
diffraction peaks of TiO2 and MoS2 were clearly observed in i) after the HPT was incorporated on MoS2. Further, when the
MPT-hybrid composites, which suggests the successful weight ratio of HPT on MoS2 was increased to 15, the surface
formation of MoS2@HPT heterocomposites. It was noted morphology of the as-synthesized MPT nanocomposite
that with an increase of HPT, the intensity of the diffraction interestingly changed to a wormhole fibrous shape similar to
peak of TiO2 also increased proportionally, while the MoS2 HPT, which may be due to the higher volume ratio of the HPT
(002) plane decreased (Figure 1), suggesting that the on the MoS2 surface. As an FESEM result, it was understood
introduction of HPT substantially inhibits the (002) plane that the volume ratio plays a vital role in modulating the
growth of the MoS2 crystal in the heterocomposite.44 surface morphology.
11952 https://doi.org/10.1021/acs.jpcc.1c01922
J. Phys. Chem. C 2021, 125, 11950−11962
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 3. TEM images with low and high magnification of (a, b) HPT, (c, d) M5PT, (e, f) MPT1, (g, h) MPT5, (i, j) MPT10, and (k, l) MPT15
photocatalysts.

To further understand the structural properties and confirm decreased, as given in Table 1. Further, the estimated BET
the composite constituents, HPT and MPT nanocomposites surface areas of M5PT, MPT1, MPT5, MPT10, MPT15, MPT20,
were analyzed by TEM and high-resolution transmission and MPT25 were found to be 22.08, 24.23, 35.47, 49.29, 46.80,
electron microscopy (HRTEM) (Figure 3). Figure 3a,b 37.84, and 33.65 m2 g−1, respectively. The BJH pore-size
demonstrates that the HPT displays the anatase phase of the distribution graphs of M5PT, MPT1, MPT5, MPT10, MPT15,
TiO2 (101) plane. The TEM images of M5PT, MPT1, MPT5, MPT20, and MPT25 nanocomposites are shown in Figures 4
and MPT10 heterocomposites illustrated in Figure 3c−j reveal and S5 (inset); the evaluated pore sizes were 2.86, 5.11, 16.35,
that the MoS2 layers decreased with an increase in the amount 5.44, 12.81, 13.38, and 14.11 nm, respectively (Table 1). It is
of HPT on the MoS2 surface. Further, the surface morphology interesting to mention that the surface area of the MoS2@HPT
of MoS2 changed from layer to particle at a higher volume ratio nanocomposites enhanced with an increase in the amount of
of porous TiO2 on the MoS2 surface, as shown in Figure 3k,l of both precursors. On the basis of the above results, it can be
the MPT15 photocatalyst. Moreover, each composite consists envisioned that porous networks exposed the active sites,
of two different lattice spacings, one with d = 0.354 nm, which which in turn play a significant role in improving the HER and
matches with the (101) plane of anatase TiO2, and the other reducing the charge recombination. The obtained results were
fringes with d = 0.65 nm, which matches with the (002) plane consistent with the SEM and XRD analyses.
of MoS2. Some cross-lattice fringes were observed in the XPS Analysis. Further, we performed XPS analysis to
HRTEM images, which clearly indicate the successful
examine the elemental composition and elemental states of the
formation of the heterojunction between MoS2 and hier-
best-performing material (MPT10), as shown in Figure 5. The
archical porous TiO2. The composite formation was also
full surface scan XPS spectrum also corroborated the fact that
confirmed by high-angle annular dark field (HAADF) and
energy dispersive X-ray (EDX) images (Figure S1). The the prepared photocatalyst was composed of Ti, O, Mo, and S
synergistic effects were beneficial for the fast photogenerated elements (Figure 5a). As illustrated in Figure 5b, the high-
electron transfer and further improvement in photocatalytic resolution XPS spectra of Ti 2p at 458.3 and 465.3 eV could be
performances.48 ascribed to Ti 2p3/2 and Ti 2p1/2, respectively. It was noted
Surface Analysis. The Brunauer−Emmett−Teller (BET) that the O 1s peak appeared at a binding energy of 531.5 eV
adsorption isotherms and Barrett−Joyner−Halenda (BJH) (Figure 5c). The obtained results were compared with that of
pore-size distribution patterns of different MPT nano- the previous literature reports.38 Further, the high-resolution
composites are shown in Figures 4 and S5. To achieve higher XPS spectra in Figure 5d demonstrate two characteristics peaks
photocatalytic H2 production efficiency, electron and hole ascending at 228.9 and 232.3 eV, which can be attributed to
transporting materials should have a higher surface area. The the energy values of Mo 3d5/2 and Mo 3d3/2, respectively,
N2 adsorption−desorption isotherms of all nanocomposites suggesting the presence of Mo4+ in the MPT sample. The
show type IV isotherm patterns with H3 hysteresis loops binding energy values of S 2p1/2 and S 2p3/2 were observed at
attributed to the mesoporous nature. We also found that the 163.5 and 162.3 eV, respectively (Figure 5e).35,37,38 Hence, the
BET specific surface areas increased with increasing optimal results obtained from XPS data reinforce the fact that MoS2 is
loading amount of HPT on the MoS2 surface and then successfully integrated onto the TiO2 surface and there is
11953 https://doi.org/10.1021/acs.jpcc.1c01922
J. Phys. Chem. C 2021, 125, 11950−11962
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 4. BET adsorption−desorption isotherm and Barrett−Joyner−Halenda (BJH) plots of MPTx nanocomposites.

Table 1. Photophysical Properties of MoS2-Hierarchical diffuse reflectance spectroscopy (DRS), as shown in Figure 6.
Porous TiO2 Nanocomposites The intrinsic absorption band located around 370 nm
specific surface area average pore size pore volume
corresponded to the porous TiO2 substrate, suggesting a
sample (m2 g−1)a (nm)b (cm3 g−1) band energy distance of ∼3.0 eV. Furthermore, a pronounced
M5PT 22.08 2.86 0.015 absorption edge ranging between 420 and 720 nm could be
MPT1 24.23 5.11 0.030 observed in the MoS2@HPT heterocomposites, which can be
MPT5 35.47 16.35 0.066 attributed to the plasmonic induction of MoS2 on the HPT
MPT10 49.29 5.44 0.150 surface. The band-gap energy of the prepared MoS2@HPT
MPT15 46.80 12.81 0.145
samples was determined by the Kubelka−Munk analysis and
MPT20 37.84 13.38 0.126
MPT25 33.65 14.11 0.118
extrapolations of the Tauc plot demonstrated in Figure
a S2.39,49,50 The estimated band-gap energy values of the
Specific surface area was calculated from the linear part of the BET
plot. bAverage pore diameter. prepared heterocomposite samples were found to be 1.95,
2.97, 2.95, 2.92, 3.0, and 3.01 eV corresponding to M5PT,
successful formation of MPT-hybrid composite elemental MPT1, MPT5, MPT10, MPT15, and MPT20 (Table 2 and Figure
characteristics. S2). The results suggest that the band-gap potential gradually
Optical Properties. The MoS2−HPT heterostructure’s narrows with an increase in the weight ratio of TiO2 to MoS2
solar-light-harvesting properties were examined by UV−vis in the composite. Moreover, the MPT10 photocatalyst has the
11954 https://doi.org/10.1021/acs.jpcc.1c01922
J. Phys. Chem. C 2021, 125, 11950−11962
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 5. XPS spectra of the MPT10 photocatalyst: (a) full scan, (b) Ti 2p, (c) O 1s, (d) Mo 3d, and (e) S 2p.

Figure 6. (a) UV−vis absorption spectra, (b) Mott−Schottky (MS) plot at 1000 Hz frequency of MPTx heterocomposites, and (c) Mott−Schottky
plot at 1000 Hz frequency of HPT.

11955 https://doi.org/10.1021/acs.jpcc.1c01922
J. Phys. Chem. C 2021, 125, 11950−11962
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Table 2. Photocatalytic Hydrogen Production Activity of assessed and compared in Figures 7a and S3. The bare porous
the Prepared Heterocomposites TiO2 revealed negligible H2 production activity under visible
light. However, MPTx heterostructures showed obvious
band gap H2 yield H2 rate AQY
catalyst (eV) (μmol gcat−1) (μmol h−1 gcat−1) (%) photocatalytic activities toward H2 production because of
M5PT 1.95 387 78 0.13
their impressive visible-light response and the surface
MPT1 2.97 598 120 0.20 plasmonic resonance of MoS2 nanosheets. A series of loading
MPT5 2.95 654 130 0.22 amounts of MoS2 nanosheets on the porous TiO2 surface were
MPT10 2.92 6879 1376 2.28 examined to calibrate the photocatalytic HER activity. The
MPT15 3.00 4725 945 1.57 superior photocatalytic response of MPT heterostructures
MPT20 3.01 2272 454 0.75 obtained at a loading ratio of 1 wt % MoS2 to 10 wt % porous
MCT 690 138 0.23 TiO2 at the rate of H2 evolution is 1376 μmol h−1 gcat−1, which
is higher than the other ratios of MPTx heterostructures.
Furthermore, as tabulated in Table 2, the MPT10 photocatalyst
maximum photon-harvesting ability and better charge-trans- exhibited a higher H2 production rate and apparent quantum
port properties toward solar energy conversion.34,35 yield (2.28%) compared to M5PT (0.13%), MPT1 (0.20%),
Mott−Schottky (MS) measurements were employed to MPT5 (0.22%), MPT15 (1.57%), and MPT20 (0.75%). It could
confirm the nanostructures in p−n junctions and determine be because the higher loading of MoS2 on the porous TiO2
the flat-band positions of MPTx samples. The Mott−Schottky surface hinders the photoexcited charge-transfer process
slopes positively shifted in the composite (Figure 6b) between HPT, assuming the porous TiO2 as the core and
compared to HPT (Figure 6c); simultaneously, the n-type the MoS2 nanosheet as the shell; however, lower loading leads
and p-type characteristics of the semiconductor, which were to low efficiency of the surface plasmon resonance or deficient
observed in the Mott−Schottky image of MPTx, confirmed the visible-light absorption. Moreover, on optimal loading of MoS2
formation of the p−n junction.51,52 The obtained flat-band on the porous TiO2 surface, it acted as an efficient light-
position (Vfb) results of the Mott−Schottky measurement for
harvester photocatalyst, which provided efficient charge
MPT10 was −0.63 eV vs Ag/AgCl (i.e., −0.71 eV vs NHE),
separation to prevent electron−hole recombination, which in
which was more negative compared to the electrode potential
turn improved the HER activity. The H2 production activity is
of H+/H2 (−0.41 V vs NHE, pH = 7)53,54 and fulfilled the
thermodynamic requirement in the HER process. In the same also examined in the presence and absence of SED. Figure S6
manner, the CB could be calculated as −0.51 eV vs NHE for shows the magnitude to be about 17 orders lower in true
MPT5, −0.44 eV vs NHE for MPT15, and −0.48 eV vs NHE water-splitting reactions, as expected. This is because the
for MPT20 nanomaterial samples. Further, the valenceband consumption rate of the positive h+ hole is the limiting kinetic
(VB) positions were at 2.44, 2.29, 2.56, and 2.53 eV vs NHE step.56,57
for MPT5, MPT10, MPT15, and MPT20, respectively. The The surface plasmonic effect of the MPT10 composite was
estimated energy band structures favored the suitability of further affirmed by a wavelength-dependent photocatalytic
efficient charge injection, separation, and transfer of photo- process under various monochromatic light sources (λ = 500,
induced electrons and holes.55 550, 600, 650, and 700 nm) using corresponding narrow band-
Photocatalytic Hydrogen Production Activity. Encour- pass filters. As displayed in Figure 7b, the spectral-dependent
aged by the above characterization results, the photocatalytic H2 evolution rates are about 17, 29, and 5.75 μmol h−1 gcat−1 at
experiments for H2 production were performed under 500, 550, and 600 nm, respectively, while no H2 evolution was
simulated solar light irradiation of a 300 W Xe lamp using found at 650 and 700 nm. The relative photoabsorption and
20 mg of the prepared MoS2@HPT in 30 mL of an aqueous hydrogen evolution consistency over simulated sunlight
suspension containing 0.35 M Na2S and 0.25 M Na2SO3. The signifies that the enhanced photoactivity was mainly due to
photocatalytic H2 evolution activities of MPTx samples were induction of the plasmonic effect between blue and green

Figure 7. (a) Histogram of comparative photocatalytic H2 generation activities of the MPTx photocatalyst in 30 mL of the mixed aqueous solution
containing 0.35 M Na2S and 0.25 M Na2SO3 as sacrificial agents under visible-light illumination as the lighting source. (b) Controlled H2 evolution
rate as a function of the wavelength by monochromatic light irradiation after 4 h.

11956 https://doi.org/10.1021/acs.jpcc.1c01922
J. Phys. Chem. C 2021, 125, 11950−11962
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 8. (a) Nyquist curves, (b) photocurrent density plot, (c) PL spectra, and (d) singlet excited-state lifetimes of the prepared photocatalysts in
methanol.

wavelengths. The higher H2 evolution rate obtained at 500 and emission analysis was employed to detect the charge carrier
550 nm suggests plasmon-induced enhancement in HER. trapping and electron−hole recombination of the prepared
Photoelectrochemical Studies. To gain further insights heterostructured materials during photocatalytic reactions
into the mechanism of photocatalytic H2 generation and the (Figure 8c). The photoluminescence spectra of bare TiO2,
separation of photogenerated electrons and holes in these MoS2, and MoS2@TiO2 nanocomposites excited at 320 nm are
MPTx heterostructures, we also performed electrochemical presented in Figure 8c. The results showed that the TiO2
impedance spectroscopy (EIS), photoluminescence (PL) emission peak intensities at around 575 and 590 nm were
emission spectroscopy, and transient photocurrent measure- remarkably quenched by varying the MoS2 content. The results
ments. In that, EIS analysis was employed to evaluate the can be ascribed to the beneficial interfacial photoexcited charge
internal resistance of the charge-transfer process of the transfers between the hierarchical porous TiO2 and MoS2 and
samples.58 The semicircular radius of Nyquist curves of the MoS2 that acted as an electron sink, which in turn inhibited
MPT10 exhibited a smaller diameter compared to those of the photogenerated charge recombination. The result corre-
MPT5, MPT15, and MPT20 (Figure 8a), resulting in a higher sponds to an efficient charge separation in MoS2−TiO2, which
photocurrent response. The charge carrier transport ability of enhanced the HER of the nanomaterial samples. To gain a
MPT10 is faster than those of MPT5, MPT15, and MPT20.50 clear understanding, we studied the electron−hole charge
Moreover, it was also confirmed by EIS that the MPT10 hybrid separation dynamics of bare MoS2, porous TiO2, and MPTx
showed the lowest charge-transfer resistance, which signifi- nanomaterials; the time-resolved fluorescence (TRF) decay
cantly accelerated the electron mobility compared to those of data were taken from the photoluminescence lifetime study
the other MPTx photocatalysts.36 The photocurrent density of using the time-correlated single photon counting (TCSPC)
the MPT10 photocatalyst was 0.55 μA cm−2, which was nearly analysis in methanol. The TRF decay curves were fitted to a
55 times that of HPT (0.01 μA cm−2), 7.0 times that of MPT5 triexponential decay function, which shows that three different
(0.08 μA cm−2), about 1.7 times that of MPT15 (0.33 μA emission paths impact the TRF decay in MPTx. The
cm−2), and 2.2 times that of MPT20 (0.25 μA cm−2) (Figure spectroscopic details are shown in Figures 8d and S7, and
8b). This suggests that the nanomaterial catalyst leads to their fluorescence transient (τ) fittings are tabulated in Table
separation of photogenerated charges. The continuous increase S2. The obtained spectroscopic value revealed that the optimal
of photocurrent to the initial time could be due to the MoS2 loading on TiO2 and the periodically patterned
photostability of the catalyst. Further, photoluminescence (PL) morphology provided the MPT10 photocatalyst with less
11957 https://doi.org/10.1021/acs.jpcc.1c01922
J. Phys. Chem. C 2021, 125, 11950−11962
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

Figure 9. (a) Side view of the MPT heterostructure. (b) Top view of the MPT heterostructure along with active sites shown by black circles. (c)
Free energy profile of the one-layer MPT heterostructure.

Figure 10. Side views of the charge density difference of hydrogen adsorbed at MPT, (a) O1, (b) O2, (c) O3, and (d) Ti sites.

trapping centers and prolonged lifetime of electrons at the free energy profile. There are total three steps, which are
activation level as well as a pronounced HER. Moreover, the involved in the estimation of the free energy profile. The first
average weighted lifetime of MPT10 (τavg = 1.19 ns) was (GH+) and final (G1/2H2) steps are expected to be in equilibrium
estimated to be significantly longer than those of MoS2 (τavg = at acidic conditions (pH = 0). The second step, hydrogen
0.07 ns), MPT5 (τavg = 0.15 ns), MPT15 (τavg = 0.20 ns), adsorption (H*), is a key step to define the free energy profile.
MPT20 (τavg = 0.21 ns), and porous TiO2 (τavg = 0.20 ns), The Gibbs free energy is calculated by the equation ΔGH = ΔE
which correlated with the higher photocatalytic performance − TΔS + ZPE, where ΔE denotes the change in the overall
observed for MPT10. It is interesting to note that a shorter energy (calculated from DFT) and ΔS and ZPE represent
decay component, a lower emission intensity, and a higher changes in entropy and zero-point energy difference between
estimated lifetime of MPT10 suggest that the separation of the the adsorbed and the gas phase of hydrogen, respectively.
photogenerated charge carrier efficiency becomes markedly However, the ΔS and ZPE of the adsorbed hydrogen are
better by creating the heterojunction and pronounced negligible compared to those of the gaseous phase at RT and
nonradiative rates in the MPT10 relative to the radiative rates ambient pressure.
than the other MPTx nanocomposites. These data are in Using the first principles approach, the Gibbs free energy
agreement with the trend of the series of materials tested for (ΔGH) has been calculated to describe the HER activity. As
the HER. per the experimental result, HPT matches with the anatase
Computational Study. The free energy of H2 adsorption phase of the TiO2 (1 0 1) plane. To mimic the TiO2/MoS2
(GH*) is a good indicator of the activity of H2 evolution in an heterostructure, we used the (3 × 4 × 1) anatase phase of the
acidic medium. The HER activity can be estimated using the TiO2 structure and (5 × 5 × 1) MoS2 with the lattice
11958 https://doi.org/10.1021/acs.jpcc.1c01922
J. Phys. Chem. C 2021, 125, 11950−11962
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

mismatch of 4.66%, called MPT, as shown in Figure 9a. To


understand the HER activity, we studied two different
structures, namely, pristine MoS2 and MPT. For pristine
MoS2, the S site is considered as an active site, while in the case
of the MPT heterostructure, O1, O2, O3, and Ti sites were
considered as active sites, as shown in Figure 9b. O1, O2, O3,
and Ti sites can be defined as follows. O1 is the layer-one O
atom, which is attached to two Ti atoms. The bond lengths of
O with these Ti atoms are 1.79 and 1.91 Å. O2 is the layer-two
atom, which is attached to three Ti atoms having bond lengths
of 1.97, 2.02, and 1.97 Å. O3 is the layer-three atom, which is
also attached to three Ti atoms having the same bond length as
the O2 atom. The Ti site atom is attached to five O atoms and
one Ti atom having bond lengths of 1.976, 2.025, 1.979, 1.91,
1.79, and 2.73 Å. The free energy of hydrogen adsorption,
ΔGH, for pristine MoS2 is 1.93 eV, while in the case of the
MPT heterostructure, O1, O2, and O3 sites are more favorable
for the HER activity with ΔGH values of 0.42, 0.52, and 0.88 Figure 11. Schematic illustration of photocatalytic processes for
eV, as shown in Figure 9c. The ΔGH value of MoS2 agrees well efficient solar-to-H2 conversion.
with reported works.59−61 For comparison, we calculated the
free energy of H adsorption on O1, O2, and O3 sites of the
HPT structure (anatase TiO2 surface of the 101 plane), and it
is estimated as 0.57, 0.65, and 1.01 eV, respectively (Figure transformed from the CB of hierarchical porous TiO2 to
S4). The free energy of hydrogen adsorption, ΔGH, for the flowerlike MoS2 without any cocatalyst. This was due to the
pristine TiO2 layer and the three-layer MPT was further comparative energy difference between TiO2CB (ECBTiO2 =
examined to consider the modeling of the TiO2/MoS2 −0.51 V vs NHE, pH 7)62 and MoS2 (ECBMoS2 = −0.16 V vs
heterostructure with similar parameters and induced strain as NHE, pH 7).63 It is interesting to note that the interfacial
taken for the one-layer MPT (Figure S8). The obtained free contact area of MoS2@HPT nanocomposites plays a key role
energy change (ΔGH) values of the three-layer MPT for O1, in determining the superior H2 production performances of the
O2, and O3 sites are 0.49, 0.61, and 0.98 eV, respectively, prepared materials, which is favorable for charge transfer from
which is higher than the free energy change (ΔGH) value of the TiO2 to MoS2 for H2 generation, resulting in improvement of
one-layer MPT and lower than the pristine TiO2 values for O1, photocatalytic H2 production activities (Figure 11). Eventually,
O2, and O3 sites of 0.57, 0.65, and 1.01 eV, respectively, as the electron on the surface of MoS2 can react with the
tabulated in Table S3. Overall, we observed that the MoS2 absorbed protons for efficient H2 evolution, while the holes in
layer helps in adsorption of the H atom by optimizing the free the VBs were regenerated to consume electrons from SED
energy. (Na2S + Na2SO3) and drove the photocatalytic process in a
To understand the electron-transfer pathway, we calculated repeated manner.
the charge density difference of the MPT heterostructure.
From the charge density difference of O1, O2, O3, and Ti sites
of the MPT heterostructure, it is clear that there is a small
■ CONCLUSIONS
In this work, it was shown that hierarchical porous TiO2 can be
charge transfer from hydrogen to oxygen sites (Figure 10), interfaced successfully with marigold-flower-like MoS2 flakes
whereas for Ti atoms, there is a negligible charge transfer with intriguing photophysical properties, viz., visible-light
because Ti shares charge with five O and one Ti atom (Figure response, controlled electron−hole recombination, and
10d), which is the main reason for not being able to adsorb the sustainable H2 production over prolonged light irradiation
hydrogen atom. To better understand the charge transfer and due to the synergic effect of flowerlike MoS2 and the fibrous
adsorption of the hydrogen atom on the MPT surface, we wormhole mesoporous channel of TiO2. This investigation
studied the occupancy of electrons on each oxygen atom. The provides useful information on the precisely controllable
electron occupancies of O1, O2, and O3 sites of the MPT are synthesis of MoS2 and HPT-based hybrid heterojunction
6.84e, 7.05e, and 7.06e, respectively. It confirms why O1 is materials for extensive photocatalytic and other energy-
more active toward HER compared to the other sites. Less harvesting applications. TEM results showed that MoS2
occupancy on the O1 atom increases the affinity toward the H nanoflower layers decreased on decreasing the weight ratio
atom, and it is found that it gains more charge during of MoS2 on the porous TiO2 surface and EDS mapping
adsorption of the H atom. Overall, from the DFT study, it is confirmed the elemental composition as Mo, S, Ti, and O and
clear that MPT heterostructure O sites are more favorable for spatial distribution in the heterostructured sample. The
HER reactivity, while pristine MoS2 is not feasible for HER superior photocatalyst MPT10 at the optimal weight ratio
activity. enhanced the hydrogen production results at a rate of 1376
Mechanistic Pathway. Based on the results and μmol g−1 h−1 and a photocurrent density of 0.55 μA cm−2. The
discussion, a plausible mechanism for H2 production over results show the pronounced plasmonic effects on the
the MPT photocatalyst is illustrated in Figure 11. It was photoactivities and highlight the role of the heterojunction
observed that the photoexcited electron was initiated upon structure in photocatalytic H2 production. We believe that the
light illumination and the excited electrons from the VB of present work would prove useful to extend the light-harvesting
TiO2 were injected into the CB of TiO2 for the creation of capability, suppress the recombination process, and enhance
charge pairs (electron−hole). The photogenerated electrons the photocatalytic performance.
11959 https://doi.org/10.1021/acs.jpcc.1c01922
J. Phys. Chem. C 2021, 125, 11950−11962
The Journal of Physical Chemistry C


pubs.acs.org/JPCC Article

ASSOCIATED CONTENT (3) Kim, J. H.; Hansora, D.; Sharma, P.; Jang, J.-W.; Lee, J. S.
*
sı Supporting Information
Toward Practical Solar Hydrogen Production−an Artificial Photo-
synthetic Leaf-to-Farm Challenge. Chem. Soc. Rev. 2019, 48, 1908−
The Supporting Information is available free of charge at 1971.
https://pubs.acs.org/doi/10.1021/acs.jpcc.1c01922. (4) Fujishima, A.; Honda, K. Electrochemical Photolysis of Water at
Structural characterization, apparent quantum yield, and a Semiconductor Electrode. Nature 1972, 238, 37.
turnover number calculation; HAADF and EDX images (5) Sathish, M.; Viswanathan, B.; Viswanath, R.; Gopinath, C. S.
of MPT10; Kubelka−Munk plots of MPTx; time-course Synthesis, Characterization, Electronic Structure, and Photocatalytic
photocatalytic H2 production curves; true water-splitting Activity of Nitrogen-Doped Tio2 Nanocatalyst. Chem. Mater. 2005,
graph; free energy profile of anatase TiO2; and lifetime 17, 6349−6353.
graphs of the photocatalysts (PDF) (6) Sivaranjani, K.; Gopinath, C. S. Porosity Driven Photocatalytic


Activity of Wormhole Mesoporous Tio 2-X N X in Direct Sunlight. J.
Mater. Chem. 2011, 21, 2639−2647.
AUTHOR INFORMATION (7) Wang, C.; Thompson, R. L.; Baltrus, J.; Matranga, C. Visible
Corresponding Authors Light Photoreduction of Co2 Using Cdse/Pt/Tio2 Heterostructured
Amritanjali Tiwari − Department of Energy and Catalysts. J. Phys. Chem. Lett. 2010, 1, 48−53.
Environmental Engineering, CSIR-Indian Institute of (8) Wang, W.-N.; An, W.-J.; Ramalingam, B.; Mukherjee, S.;
Chemical Technology, Hyderabad 500007, India; Niedzwiedzki, D. M.; Gangopadhyay, S.; Biswas, P. Size and Structure
orcid.org/0000-0001-5547-3903; Email: amritanjali4@ Matter: Enhanced Co2 Photoreduction Efficiency by Size-Resolved
gmail.com Ultrafine Pt Nanoparticles on Tio2 Single Crystals. J. Am. Chem. Soc.
2012, 134, 11276−11281.
Ujjwal Pal − Department of Energy and Environmental
(9) Zhou, W.; Yin, Z.; Du, Y.; Huang, X.; Zeng, Z.; Fan, Z.; Liu, H.;
Engineering, CSIR-Indian Institute of Chemical Technology, Wang, J.; Zhang, H. Synthesis of Few-Layer Mos2 Nanosheet-Coated
Hyderabad 500007, India; Academy of Scientific and Tio2 Nanobelt Heterostructures for Enhanced Photocatalytic
Innovative Research (AcSIR), CSIR-Human Resource Activities. Small 2013, 9, 140−147.
Development Centre (CSIR-HRDC) Campus, Ghaziabad (10) Jackman, M. J.; Thomas, A. G.; Muryn, C. Photoelectron
201002, Uttar Pradesh, India; orcid.org/0000-0002- Spectroscopy Study of Stoichiometric and Reduced Anatase Tio2
2110-4242; Email: upal03@gmail.com (101) Surfaces: The Effect of Subsurface Defects on Water
Adsorption at near-Ambient Pressures. J. Phys. Chem. C 2015, 119,
Authors 13682−13690.
Amit Gautam − Department of Energy and Environmental (11) Wang, S.; Guan, B. Y.; Yu, L.; Lou, X. W. Rational Design of
Engineering, CSIR-Indian Institute of Chemical Technology, Three-Layered Tio2@ Carbon@ Mos2 Hierarchical Nanotubes for
Hyderabad 500007, India; Academy of Scientific and Enhanced Lithium Storage. Adv. Mater. 2017, 29, No. 1702724.
Innovative Research (AcSIR), CSIR-Human Resource (12) Xiao, S.; Lu, Y.; Li, X.; Xiao, B. Y.; Wu, L.; Song, J. P.; Xiao, Y.
Development Centre (CSIR-HRDC) Campus, Ghaziabad X.; Wu, S. M.; Hu, J.; Wang, Y.; et al. Hierarchically Dual-Mesoporous
201002, Uttar Pradesh, India Tio2 Microspheres for Enhanced Photocatalytic Properties and
Saddam Sk − Department of Energy and Environmental Lithium Storage. Chem. - Eur. J. 2018, 24, 13246−13252.
Engineering, CSIR-Indian Institute of Chemical Technology, (13) Mak, K. F.; Lee, C.; Hone, J.; Shan, J.; Heinz, T. F. Atomically
Hyderabad 500007, India Thin Mos 2: A New Direct-Gap Semiconductor. Phys. Rev. Lett. 2010,
Deepak S. Gavali − Department of Physics, SRM University 105, No. 136805.
AP, Amaravati 522240, Andhra Pradesh, India (14) Splendiani, A.; Sun, L.; Zhang, Y.; Li, T.; Kim, J.; Chim, C.-Y.;
Ranjit Thapa − Department of Physics, SRM UniversityAP, Galli, G.; Wang, F. Emerging Photoluminescence in Monolayer Mos2.
Amaravati 522240, Andhra Pradesh, India; orcid.org/ Nano Lett. 2010, 10, 1271−1275.
(15) Tongay, S.; Zhou, J.; Ataca, C.; Lo, K.; Matthews, T. S.; Li, J.;
0000-0002-9285-0525
Grossman, J. C.; Wu, J. Thermally Driven Crossover from Indirect
Complete contact information is available at: toward Direct Bandgap in 2d Semiconductors: Mose2 Versus Mos2.
https://pubs.acs.org/10.1021/acs.jpcc.1c01922 Nano Lett. 2012, 12, 5576−5580.
(16) Kang, K.; Xie, S.; Huang, L.; Han, Y.; Huang, P. Y.; Mak, K. F.;
Notes Kim, C.-J.; Muller, D.; Park, J. High-Mobility Three-Atom-Thick
The authors declare no competing financial interest. Semiconducting Films with Wafer-Scale Homogeneity. Nature 2015,

■ ACKNOWLEDGMENTS
The authors acknowledge the institutional in-house project P-
520, 656−660.
(17) Radisavljevic, B.; Radenovic, A.; Brivio, J.; Giacometti, V.; Kis,
A. Single-Layer Mos 2 Transistors. Nat. Nanotechnol. 2011, 6, 147.
(18) Wang, Q. H.; Kalantar-Zadeh, K.; Kis, A.; Coleman, J. N.;
07 and the externally funded DAE/BRNS Nos. 34/14/03/ Strano, M. S. Electronics and Optoelectronics of Two-Dimensional
2018-BRNS/340 and DST/TMD/HFC/2k18/60. The au- Transition Metal Dichalcogenides. Nat. Nanotechnol. 2012, 7, 699.
thors also acknowledge the High Performance Computing (19) Chhowalla, M.; Shin, H. S.; Eda, G.; Li, L.-J.; Loh, K. P.; Zhang,
Center, SRM IST, for providing the computational facility. H. The Chemistry of Two-Dimensional Layered Transition Metal
CSIR-IICT Communication No: IICT/Pubs./2020/273.


Dichalcogenide Nanosheets. Nat. Chem. 2013, 5, 263.
(20) Li, Y.; Wang, H.; Xie, L.; Liang, Y.; Hong, G.; Dai, H. Mos2
REFERENCES Nanoparticles Grown on Graphene: An Advanced Catalyst for the
(1) Li, A.; Zhu, W.; Li, C.; Wang, T.; Gong, J. Rational Design of Hydrogen Evolution Reaction. J. Am. Chem. Soc. 2011, 133, 7296−
Yolk−Shell Nanostructures for Photocatalysis. Chem. Soc. Rev. 2019, 7299.
48, 1874−1907. (21) Min, S.; Lu, G. Sites for High Efficient Photocatalytic
(2) Wang, Z.; Li, C.; Domen, K. Recent Developments in Hydrogen Evolution on a Limited-Layered Mos2 Cocatalyst Confined
Heterogeneous Photocatalysts for Solar-Driven Overall Water on Graphene Sheets―the Role of Graphene. J. Phys. Chem. C
Splitting. Chem. Soc. Rev. 2019, 48, 2109−2125. 2012, 116, 25415−25424.

11960 https://doi.org/10.1021/acs.jpcc.1c01922
J. Phys. Chem. C 2021, 125, 11950−11962
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

(22) Kibsgaard, J.; Chen, Z.; Reinecke, B. N.; Jaramillo, T. F. (38) Xiang, Q.; Yu, J.; Jaroniec, M. Synergetic Effect of Mos2 and
Engineering the Surface Structure of Mos 2 to Preferentially Expose Graphene as Cocatalysts for Enhanced Photocatalytic H2 Production
Active Edge Sites for Electrocatalysis. Nat. Mater. 2012, 11, 963. Activity of Tio2 Nanoparticles. J. Am. Chem. Soc. 2012, 134, 6575−
(23) Yang, X.; Huang, H.; Kubota, M.; He, Z.; Kobayashi, N.; Zhou, 6578.
X.; Jin, B.; Luo, J. Synergetic Effect of Mos2 and G-C3n4 as (39) Lin, Y.; Ren, P.; Wei, C. Fabrication of Mos 2/Tio 2
Cocatalysts for Enhanced Photocatalytic H2 Production Activity of Heterostructures with Enhanced Photocatalytic Activity. CrystEng-
Tio2. Mater. Res. Bull. 2016, 76, 79−84. Comm 2019, 21, 3439−3450.
(24) Liu, X.; Xing, Z.; Zhang, H.; Wang, W.; Zhang, Y.; Li, Z.; Wu, (40) Yuan, Y.-J.; Shen, Z.; Wu, S.; Su, Y.; Pei, L.; Ji, Z.; Ding, M.;
X.; Yu, X.; Zhou, W. Fabrication of 3 D Mesoporous Black Tio2/ Bai, W.; Chen, Y.; Yu, Z.-T.; et al. Liquid Exfoliation of G-C3n4
Mos2/Tio2 Nanosheets for Visible-Light-Driven Photocatalysis. Nanosheets to Construct 2d-2d Mos2/G-C3n4 Photocatalyst for
ChemSusChem 2016, 9, 1118−1124. Enhanced Photocatalytic H2 Production Activity. Appl. Catal., B
(25) Islam, M. J.; Reddy, D. A.; Han, N. S.; Choi, J.; Song, J. K.; 2019, 246, 120−128.
Kim, T. K. An Oxygen-Vacancy Rich 3d Novel Hierarchical Mos 2/ (41) Tiwari, A.; Krishna, N. V.; Giribabu, L.; Pal, U. Hierarchical
Bioi/Agi Ternary Nanocomposite: Enhanced Photocatalytic Activity Porous Tio2 Embedded Unsymmetrical Zinc−Phthalocyanine
through Photogenerated Electron Shuttling in a Z-Scheme Manner. Sensitizer for Visible-Light-Induced Photocatalytic H2 Production. J.
Phys. Chem. Chem. Phys. 2016, 18, 24984−24993. Phys. Chem. C 2018, 122, 495−502.
(26) Binwal, D. C.; Pramoda, K.; Zak, A.; Kaur, M.; Chithaiah, P.; (42) Tiwari, A.; Duvva, N.; Rao, V. N.; Venkatakrishnan, S. M.;
Rao, C. Nanocomposites of 1d Mos2 with Polymer-Functionalized Giribabu, L.; Pal, U. Tetrathiafulvalene Scaffold-Based Sensitizer on
Nanotubes of Carbon and Borocarbonitride, and Their Her Activity. Hierarchical Porous Tio2: Efficient Light-Harvesting Material for
ACS Appl. Energy Mater. 2021, 2339−2347. Hydrogen Production. J. Phys. Chem. C 2019, 123, 70−81.
(27) Zhang, Y.; Shi, J.; Ding, X.; Wu, J.; Zheng, Y.-Z.; Tao, X. Stable (43) Naik, B.; Kim, S. M.; Jung, C. H.; Moon, S. Y.; Kim, S. H.; Park,
Mixed-Organic-Cation Perovskite Ma1−X Fa X Pbi3 Integrated with J. Y. Enhanced H2 Generation of Au-Loaded, Nitrogen-Doped Tio2
Mos2 for Enhanced Visible-Light Photocatalytic H2 Evolution. Ind. Hierarchical Nanostructures under Visible Light. Adv. Mater. Interfaces
Eng. Chem. Res. 2020, 59, 20667−20675. 2014, 1, No. 1300018.
(28) Zhou, C.; Wang, R.; Jiang, C.; Chen, J.; Wang, G. Dynamically (44) Luo, Z.; Wu, D.; Xu, B.; Xu, H.; Cai, Z.; Peng, J.; Weng, J.; Xu,
Optimized Multi-Interface Novel Bisi-Promoted Redox Sites Spatially S.; Zhu, C.; Wang, F.; et al. Two-Dimensional Material-Based
Separated N−P−N Double Heterojunctions Bisi/Mos2/Cds for Saturable Absorbers: Towards Compact Visible-Wavelength All-Fiber
Hydrogen Evolution. Ind. Eng. Chem. Res. 2019, 58, 7844−7856. Pulsed Lasers. Nanoscale 2016, 8, 1066−1072.
(29) Xun, W.; Wang, Y.; Fan, R.; Mu, Q.; Ju, S.; Peng, Y.; Shen, M. (45) Zhang, S.; Yang, H.; Gao, H.; Cao, R.; Huang, J.; Xu, X. One-
Activating the Mos2 Basal Plane toward Enhanced Solar Hydrogen Pot Synthesis of Cds Irregular Nanospheres Hybridized with Oxygen-
Generation Via in Situ Photoelectrochemical Control. ACS Energy Incorporated Defect-Rich Mos2 Ultrathin Nanosheets for Efficient
Lett. 2021, 6, 267−276. Photocatalytic Hydrogen Evolution. ACS Appl. Mater. Interfaces 2017,
(30) Guan, W.; Li, Y.; Zhong, Q.; Liu, H.; Chen, J.; Hu, H.; Lv, K.; 9, 23635−23646.
Gong, J.; Xu, Y.; Kang, Z.; et al. Fabricating Mapbi3/Mos2 (46) Zhu, X.; Liang, X.; Fan, X.; Su, X. Fabrication of Flower-Like
Composites for Improved Photocatalytic Performance. Nano Lett. Mos 2/Tio 2 Hybrid as an Anode Material for Lithium Ion Batteries.
2021, 21, 597−604. RSC Adv. 2017, 7, 38119−38124.
(31) Hendi, A. H.; Osman, A. M.; Khan, I.; Saleh, T. A.; Kandiel, T. (47) Pandey, K.; Yadav, P.; Singh, D.; Gupta, S. K.; Sonvane, Y.;
A.; Qahtan, T. F.; Hossain, M. K. Visible Light-Driven Photo- Lukačević, I.; Kim, J.; Kumar, M. First Step to Investigate Nature of
electrocatalytic Water Splitting Using Z-Scheme Ag-Decorated Mos2/ Electronic States and Transport in Flower-Like Mos 2: Combining
Rgo/Niwo4 Heterostructure. ACS Omega 2020, 31644−31656. Experimental Studies with Computational Calculations. Sci. Rep.
(32) Lim, K. R. G.; Handoko, A. D.; Johnson, L. R.; Meng, X.; Lin, 2016, 6, No. 32690.
M.; Subramanian, G. S.; Anasori, B.; Gogotsi, Y.; Vojvodic, A.; Seh, Z. (48) Quan, L. N.; Jang, Y. H.; Jang, Y. J.; Kim, J.; Lee, W.; Moon, J.
W. 2h-Mos2 on Mo2ct X Mxene Nanohybrid for Efficient and H.; Kim, D. H. Mesoporous Carbon-Tio2 Beads with Nanotextured
Durable Electrocatalytic Hydrogen Evolution. ACS Nano 2020, 14, Surfaces as Photoanodes in Dye-Sensitized Solar Cells. ChemSusChem
16140−16155. 2014, 7, 2590−2596.
(33) Dong, Y.; Chen, S. Y.; Lu, Y.; Xiao, Y. X.; Hu, J.; Wu, S. M.; (49) Paul, K. K.; Sreekanth, N.; Biroju, R. K.; Pattison, A. J.;
Deng, Z.; Tian, G.; Chang, G. G.; Li, J.; et al. Hierarchical Mos2@ Escalera-López, D.; Guha, A.; Narayanan, T. N.; Rees, N. V.; Theis,
Tio2 Heterojunctions for Enhanced Photocatalytic Performance and W.; Giri, P. Strongly Enhanced Visible Light Photoelectrocatalytic
Electrocatalytic Hydrogen Evolution. Chem. - Asian J. 2018, 13, Hydrogen Evolution Reaction in an N-Doped Mos 2/Tio 2 (B)
1609−1615. Heterojunction by Selective Decoration of Platinum Nanoparticles at
(34) He, H.; Lin, J.; Fu, W.; Wang, X.; Wang, H.; Zeng, Q.; Gu, Q.; the Mos 2 Edge Sites. J. Mater. Chem. A 2018, 6, 22681−22696.
Li, Y.; Yan, C.; Tay, B. K.; et al. Mos2/Tio2 Edge-on Heterostructure (50) Paul, K. K.; Sreekanth, N.; Biroju, R. K.; Narayanan, T. N.; Giri,
for Efficient Photocatalytic Hydrogen Evolution. Adv. Energy Mater. P. Solar Light Driven Photoelectrocatalytic Hydrogen Evolution and
2016, 6, No. 1600464. Dye Degradation by Metal-Free Few-Layer Mos2 Nanoflower/Tio2
(35) Guo, L.; Yang, Z.; Marcus, K.; Li, Z.; Luo, B.; Zhou, L.; Wang, (B) Nanobelts Heterostructure. Sol. Energy Mater. Sol. Cells 2018,
X.; Du, Y.; Yang, Y. Mos 2/Tio 2 Heterostructures as Nonmetal 185, 364−374.
Plasmonic Photocatalysts for Highly Efficient Hydrogen Evolution. (51) Wang, Z.; Nayak, P. K.; Caraveo-Frescas, J. A.; Alshareef, H. N.
Energy Environ. Sci. 2018, 11, 106−114. Recent Developments in P-Type Oxide Semiconductor Materials and
(36) Liang, J.; Wang, C.; Zhao, P.; Wang, Y.; Ma, L.; Zhu, G.; Hu, Devices. Adv. Mater. 2016, 28, 3831−3892.
Y.; Lu, Z.; Xu, Z.; Ma, Y.; et al. Interface Engineering of Anchored (52) Swain, G.; Sultana, S.; Moma, J.; Parida, K. Fabrication of
Ultrathin Tio2/Mos2 Heterolayers for Highly-Efficient Electro- Hierarchical Two-Dimensional Mos2 Nanoflowers Decorated Upon
chemical Hydrogen Production. ACS Appl. Mater. Interfaces 2018, Cubic Cain2s4 Microflowers: Facile Approach to Construct Novel
10, 6084−6089. Metal-Free P−N Heterojunction Semiconductors with Superior
(37) Yuan, Y.-J.; Fang, G.; Chen, D.; Huang, Y.; Yang, L.-X.; Cao, Charge Separation Efficiency. Inorg. Chem. 2018, 57, 10059−10071.
D.-P.; Wang, J.; Yu, Z.-T.; Zou, Z.-G. High Light Harvesting (53) Zhang, J.; Chen, X.; Takanabe, K.; Maeda, K.; Domen, K.;
Efficiency Cuins 2 Quantum Dots/Tio 2/Mos 2 Photocatalysts for Epping, J. D.; Fu, X.; Antonietti, M.; Wang, X. Synthesis of a Carbon
Enhanced Visible Light Photocatalytic H 2 Production. Dalton Trans. Nitride Structure for Visible-Light Catalysis by Copolymerization.
2018, 47, 5652−5659. Angew. Chem., Int. Ed. 2010, 49, 441−444.

11961 https://doi.org/10.1021/acs.jpcc.1c01922
J. Phys. Chem. C 2021, 125, 11950−11962
The Journal of Physical Chemistry C pubs.acs.org/JPCC Article

(54) Wang, J.; Yu, Y.; Zhang, L. Highly Efficient Photocatalytic


Removal of Sodium Pentachlorophenate with Bi3o4br under Visible
Light. Appl. Catal., B 2013, 136−137, 112−121.
(55) Faraji, M.; Yousefi, M.; Yousefzadeh, S.; Zirak, M.; Naseri, N.;
Jeon, T. H.; Choi, W.; Moshfegh, A. Z. Two-Dimensional Materials in
Semiconductor Photoelectrocatalytic Systems for Water Splitting.
Energy Environ. Sci. 2019, 12, 59−95.
(56) Mateo, D.; Esteve-Adell, I.; Albero, J.; Royo, J. F. S.; Primo, A.;
Garcia, H. 111 Oriented Gold Nanoplatelets on Multilayer Graphene
as Visible Light Photocatalyst for Overall Water Splitting. Nat.
Commun. 2016, 7, No. 11819.
(57) Hainer, A. S.; Hodgins, J. S.; Sandre, V.; Vallieres, M.; Lanterna,
A. E.; Scaiano, J. C. Photocatalytic Hydrogen Generation Using
Metal-Decorated Tio2: Sacrificial Donors Vs True Water Splitting.
ACS Energy Lett. 2018, 3, 542−545.
(58) Banerjee, S.; Mohapatra, S. K.; Das, P. P.; Misra, M. Synthesis
of Coupled Semiconductor by Filling 1d Tio2 Nanotubes with Cds.
Chem. Mater. 2008, 20, 6784−6791.
(59) Zhang, K.; Jin, B.; Gao, Y.; Zhang, S.; Shin, H.; Zeng, H.; Park,
J. H. Aligned Heterointerface-Induced 1t-Mos2 Monolayer with near-
Ideal Gibbs Free for Stable Hydrogen Evolution Reaction. Small
2019, 15, No. 1804903.
(60) Ali, T.; Wang, X.; Tang, K.; Li, Q.; Sajjad, S.; Khan, S.; Farooqi,
S. A.; Yan, C. Sns2 Quantum Dots Growth on Mos2: Atomic-Level
Heterostructure for Electrocatalytic Hydrogen Evolution. Electrochim.
Acta 2019, 300, 45−52.
(61) Zhuang, P.; Yue, H.; Dong, H.; Zhou, X. Effects of a Ni
Cocatalyst on the Photocatalytic Hydrogen Evolution Reaction of
Anatase Tio 2 by First-Principles Calculations. New J. Chem. 2020, 44,
5428−5437.
(62) Balasubramanian, S.; Wang, P.; Schaller, R. D.; Rajh, T.;
Rozhkova, E. A. High-Performance Bioassisted Nanophotocatalyst for
Hydrogen Production. Nano Lett. 2013, 13, 3365−3371.
(63) Han, S.; Kwon, H.; Kim, S. K.; Ryu, S.; Yun, W. S.; Kim, D.;
Hwang, J.; Kang, J.-S.; Baik, J.; Shin, H.; et al. Band-Gap Transition
Induced by Interlayer Van Der Waals Interaction in Mos 2. Phys. Rev.
B 2011, 84, No. 045409.

11962 https://doi.org/10.1021/acs.jpcc.1c01922
J. Phys. Chem. C 2021, 125, 11950−11962

You might also like