You are on page 1of 9

Applied Catalysis A, General 588 (2019) 117281

Contents lists available at ScienceDirect

Applied Catalysis A, General


journal homepage: www.elsevier.com/locate/apcata

CdS nanorods anchored with CoS2 nanoparticles for enhanced T


photocatalytic hydrogen production
Jie Tanga,1, Bin Gaoa,1, Jinbo Pana,1, Lang Chena,⁎, Zihao Zhaoa, Sheng Shena, Jun-Kang Guoa,⁎,
Chak-Tong Aub, Shuang-Feng Yina,⁎
a
State Key Laboratory of Chemo/Biosensing and Chemometrics, College of Chemistry and Chemical Engineering, Provincial Hunan Key Laboratory for Cost-effective
Utilization of Fossil Fuel Aimed at Reducing Carbon-dioxide Emissions, Hunan University, Changsha, 410082, Hunan, People’s Republic of China
b
College of Chemistry and Chemical Engineering, Hunan Institute of Engineering, Xiangtan, 411104, Hunan, People’s Republic of China

A R T I C LE I N FO A B S T R A C T

Keywords: Herein, we report the use of cobalt sulfide (CoS2) as efficient and inexpensive co-catalyst of CdS nanorods for
CoS2/CdS composites photocatalytic water splitting. The aim is to explore the use of earth-abundant cobalt species to replace precious
Photocatalysis metals for the photocatalytic reactions. The results of first-principles DFT simulation and planar-averaged dif-
DFT calculation ferential charge density calculation reveal that at the CoS2/CdS interface, CoS2 has zero band gap which is a class
H2 production
nature of precious metals, and functions as electron trap to enhance the transfer of hot electrons from CdS to
CoS2. Owing to the merits of efficient charge separation, high exposure of active sites as well as large surface
area, the CoS2/CdS composites exhibit outstanding photocatalytic activity in H2 production under visible light
(˜58 mmol·g−1 h−1), which is about 19 times that of CdS nanorods alone and 3 times that of 1 wt%Pt/CdS under
the same conditions.

1. Introduction cheaper [4], such as transition-metal sulfides (e.g., WS2 [22], MoS2
[23–25], NiS [26,27], CoxSy [28,29], MnS [30]), and many have
With the ever increasing demand of energy, the burning of fossil showed excellent performance in a majority of photocatalytic systems.
fuels is inevitable and environmental pollution is a global concern. In our previous works, WS2 and MoS2 nanoplates were anchored on CdS
Extensive studies have been conducted in search of energy resources nanorods, and there was great improvement in hydrogen production
that are environment-friendly and renewable [1–4]. In this aspect, under visible light irradiation [31–33]. Recently, there were reports on
hydrogen is the most commendable. The photocatalytic generation of the coupling of cobalt sulfide (e.g., CoS [34], CoS2 [35], Co3S4 [36],
hydrogen from water splitting using inexhaustible solar energy is a Co9S8 [28,29]) with semiconductors to enhance the separation of
state-of-the-art strategy, and the use of an efficient photocatalyst is charge carriers for the promotion of photocatalytic activity. For ex-
critical [5–7]. Many attempts have been made to develop high-effi- ample, through rational design of Z-scheme system, Tan et al. devel-
ciency photocatalysts, such as phase and morphology adjustment [8], oped hierarchical CdS-supported Co9S8 nanoparticles that displayed
doping with metal and non-metal elements [9,10], fabrication of het- high activity and stability in photocatalytic hydrogrn evolution (PHE).
erojunctions [11,12] and the use of co-catalysts [13–16]. Among them, Furthermore, over CdS nanorods modified by cobalt-based co-catalysts
the introduction of a co-catalyst can not only suppress the recombina- (Co(OH)2, Co3O4, CoS, CoO), Lang et al. observed high photocatalytic
tion of charge carriers through electron trapping but also provide active activity and reported that the cobalt-containing species play a sig-
sites for proton reduction. Noble metals such as Pt, Pd, Au and Ag have nificant role in the trapping and transport of hot electrons. According to
been widely used as efficient co-catalysts for hydrogen production due these photocatalytic systems, cobalt sulfide as a co-catalyst can
to their low Fermi levels [17–21]. However, the high price and rela- strengthen the absorption of solar light, accelerate the separation of
tively rare resource of precious metals hampers their usage. Therefore, photogenerated charge carriers, and offer ample active sites for reac-
it is extremely urgent to explore alternatives that are economically tions [37]. Hou and coworkers reported a hydrogen production effi-
commendable. ciency of 1232 μmol·h−1 g−1 over CoS2/g-C3N4 hybrid composites [35].
Co-catalysts that are based on earth abundant materials are much Fu et al. improved the light harvesting ability of carbon quantum dots


Corresponding authors.
E-mail addresses: huagong042cl@163.com (L. Chen), jkguozd@126.com (J.-K. Guo), sf_yin@hnu.edu.cn (S.-F. Yin).
1
The authors contributed equally to this work.

https://doi.org/10.1016/j.apcata.2019.117281
Received 21 August 2019; Received in revised form 29 September 2019; Accepted 3 October 2019
Available online 08 October 2019
0926-860X/ © 2019 Elsevier B.V. All rights reserved.
J. Tang, et al. Applied Catalysis A, General 588 (2019) 117281

(CDs) by adding CoMo-S in the photocatalytic system [38]. Reddy and and ethanol for three times, and dried at 80 °C overnight. Altogether
coworkers also used C@CoS2 as co-catalyst to enhance the PHE activity seven CoS2/CdS composites with Co:Cd molar ratio of 1:0.5, 1:1, 1:2,
of CdS nanorods [37]. It is hence envisaged that similar to noble metals, 1:3, 1:4, 1:5, and 1:6 were synthesized, and are herein denoted as
cobalt sulfides may facilitate the transfer of photogenerated charge CC0.5, CC1, CC2, CC3, CC4, CC5, and CC6, respectively. The pure CoS2
carriers at the heterojunctions. catalyst was synthesized following the above steps but without the
It was reported that cobalt disulfide (CoS2) with intrinsic metallic addition of CdS.
conductivity (6.7 × 103 S cm−1 at 300 K) displayed remarkable elec-
trocatalytic and photoelectrocatalytic activities [39]. Using metallic 2.1.3. Synthesis of physically mixed CoS2-CdS composite
cobalt pyrite CoS2 (micro- and nanostructure) as catalysts, Matthew For comparison purpose, a sample of CoS2-CdS composite was
et al. observed good electrocatalytic performance for hydrogen evolu- prepared by physical mixing (Co:Cd = 1:4, denoted as CC4(P)). Briefly,
tion reaction (HER) [40]. Using Si microwires surrounded by cobalt 1 mmol of as-prepared CoS2 powder and 4 mmol of CdS nanorods were
disulfide as photocathode for solar hydrogen evolution, Chen and thoroughly ground in an agate mortar for 30 min, and the collected
coworkers found that CoS2 not only served as protective layer, but also powder was dried at 80 °C overnight.
as a co-catalyst to reduce the recombination of charge carriers [41]. In
other words, CoS2 can promote the migration and separation efficiency
2.1.4. Synthesis of Pt/CdS composite
of photoelectrons. Despite the presence of cobalt sulfide is known to be
The Pt/CdS composite was prepared by loading 1 wt% of Pt on CdS
beneficial for electron transfer from semiconductors, the exact nature of
according to the photoreduction method reported by Zhang et al. [43].
charge separation at the composite interfaces is always neglected.
First, 50 mL of H2PtCl6·6H2O solution containing 2 mg of Pt was pre-
In this work, we facilely anchored CoS2 nanoparticles on CdS na-
pared. Then 200 mg of CdS nanorods was added to the solution, and the
norods by an in-situ solvothermal method, and analyzed the band gap
mixture was allowed to stand at 25 °C for 2 h. The solution was then
as well as the shift direction of interfacial electrons at the CoS2/CdS
illuminated using a 300 W Xe lamp (with a cut-off filter to remove light
heterojunction by DFT calculations. Owing to large surface area, high
with λ < 400 nm) for 2 h with stirring. Finally the yellow powder was
exposure of active sites, and effective transfer of photogenerated elec-
collected, washed with deionized water and absolute ethyl alcohol for
trons from CdS to CoS2, the CoS2/CdS composite gave a H2 production
three times, and dried at 80 °C overnight.
rate of ˜58 mmol·g−1 h−1 (λ ≥ 400 nm), much higher than that of CdS
or Pt/CdS.
2.2. Characterization

2. Experimental section Powder X-ray diffraction (XRD, D/MAX-2000/PC, λ = 0.15406 nm,


Rigaku Corporation) was employed for the characterization of crystal
2.1. Preparation structure. The morphology and elemental mapping analyses were
conducted over a field emission scanning electron microscope (FE-SEM)
2.1.1. Synthesis of CdS nanorods (Hitachi S-4800) and a transmission electron microscope (TEM, accel-
CdS nanorods were prepared according to the solvothermal method erating voltage: 100 kV, Tecnai G2 F20). The element contents of
described by Jiang et al. [42]. Typically, 20.23 mmol of CdCl2·2.5H2O samples were calculated according to the results of ICP-OES analysis
and 60.69 mmol of CH4N2S were dissolved in ethanediamine (60 mL). (Varian 720). The surface composition and chemical states of elements
Then the solution was transferred to a polyphenylene-lined autoclave of was analyzed by means of X-ray photoelectron spectroscopy (XPS,
100 mL and heated at 160 °C for 50 h. The yellow product was collected Thermo SCIENTIFIC K-Alpha). The specific surface area (BET) of sam-
by filtration, washed with de-ionized water and ethanol for three times, ples was obtained by using a Gemini VII 2390 instrument
and dried at 80 °C for 12 h. (Micromeritics Instrument Corp.). UV–vis diffuse reflectance spectra of
samples were recorded using a spectrophotometer (Cary 100, Agilent).
2.1.2. Synthesis of CoS2/CdS composites Photoluminescence (PL) measurement was performed on a fluorescence
The synthetic process of CoS2/CdS composites is illustrated in spectrophotometer (excitation light: 350 nm). An electrochemical ana-
Scheme 1. Briefly, 1 mmol of Co(NO3)2·6H2O and 2 mmol of thiourea lyzer (CHI660D, Shanghai Chenhua) was employed to record photo-
were dissolved in 80 mL of ethylene glycol, and the mixture was stirred electrochemical activity, using Ag/AgCl (3 M KCl) as reference elec-
for 10 min to form a pink solution. Then a certain amount of the as- trode, Pt wire as counter electrode, prepared sample (20 mg powder,
prepared CdS nanorods was added with stirring in a period of 10 min. 0.5 mL Nafion, 1.5 mL alcohol, and dried at 80 °C for 2 h) as working
The mixture was then subject to ultrasonic treatment with stirring for electrode, and sodium sulfate (Na2SO4, 0.2 M) as electrolyte. The data
60 min. The as-generated yellow solution was transferred into a poly- for the plotting of M-S curves were measured using sodium sulfate
phenylene-lined autoclave of 100 mL and heated at 180 °C for 24 h. The electrolyte solution (0.2 M, pH = 7), and the amplitude voltage was
final product was collected by filtration, washed with deionized water 5 mV at a selected frequency. The final potentials vs. Ag/AgCl were

Scheme 1. Synthetic procedure of CoS2/CdS composites.

2
J. Tang, et al. Applied Catalysis A, General 588 (2019) 117281

translated into the standard hydrogen electrode using the below CoS2/CdS samples are displayed in Fig. 1b, which indicates the pre-
equation [44]: sence of Cd and S on CdS, Co and S on CoS2, and Co, Cd, S on the CoS2/
CdS composites. The Co 2p spectrum of CoS2 shows deconvoluted peaks
ENHE = EAg/AgCl + 0.197
at 797.6, 793.9, 781.5, and 778.7 eV as well as two shake-up satellite
We carried out the density functional theory (DFT) calculations with peaks (Fig. 1c). The peaks at 797.6 eV and 781.5 eV are Co 2p1/2 and
the Perdew-Burke-Ernzerhof (PBE) exchange-correlation fuctional in 2p3/2 signals of CoS2, while those at 793.9 eV and 778.7 eV are Co 2p1/2
the Vienna ab initio simulation package (VASP) and employed the and 2p3/2 signals of cobalt oxide [46,47]. As for CdS, it shows Cd 3d
empirical correction method proposed by Grimme (DFT-D2) to describe peaks at 410.8 eV and 404.1 eV (Fig. 1d), and S 2p peaks at 161.5 eV
the van deer Waals (vdW) interaction. The Heyd-Scuseria-Ernzerhof and 160.4 eV (Fig. 1e), in agreement with literature data [31,42].
(HSE06) hybrid functional was adopted to investigate the band gap due Compared with the S 2p spectrum of CdS, that of CoS2 is more com-
to the underestimation of PBE. The energy cutoff for plane-wave basis plicated (Fig. 1e). The CoS2 peaks at 163.2 eV and 161.9 eV are as-
set and the Monkhorst-Pack k-point grid were set as 350 eV and 3 × 3 cribable to the S 2p1/2 and 2p3/2 signals of CoS2, which overlap strongly
× 1, respectively. A vacuum space of over 15 Å along the c axis was with those of SeC species. As for the peak at 168.9 eV, it is attributable
constructed to minimize the interaction between adjacent layers. The to the S 2p1/2 and 2p3/2 signals of SeO entities [38,47,48]. The XPS
geometric structure was optimized until the convergence standard of spectra of the hybrid CoS2/CdS samples (i.e. CC2, CC4 and CC6) are
total energy was 1.0 × 10–5 eV and the force on per atom was smaller rather similar (Fig. 1f,g,h). The Cd 3d3/2 and Cd 3d5/2 peaks are at
than 0.05 eV∙Å–1. 411.6 eV and 404.8 eV, respectively (Fig. 1d), which are consistent with
the Cd2+ signals of CdS in literature [13,31]. The Co 2p signals at
2.3. Photocatalytic hydrogen generation 796.6 eV and 780.6 eV are assignable to Co-S of CoS2, while those at
794.0 eV and 778.9 eV) to Co-O of cobalt oxide, and the two peaks in
Photocatalytic hydrogen production experiments were carried out purple color are attributed to shake-up satellites [45,47]. In Fig. 1h, the
in a sealed quartz reactor (250 mL) equipped with pumping facility. The peaks at 161.1 eV and 163.0 eV are ascribed to the S 2p3/2 and 2p1/2 of
reactor top was illuminated with a 300 W Xe lamp (HSX-F300, Co-S bond, and the signals at 168.8 eV was caused by SeO bond, which
Perfectlight, Beijing) with a cut-off filter (λ ≥ 400 nm). First, 20 mg of is in agreement with previous reports [29,43]. Compared to CdS, CC4
photocatalyst was put in the reactor lumen together with an aqueous exhibits positive shift (˜0.8 eV) in Cd 3d and S 2p binding energy, while
solution containing lactic acid or Na2S/Na2SO3 (50 mL, 20 vol%). Then compared to CoS2, there is negative shift (˜0.8 eV) of Co 2p and S 2p
the gases were removed using a vacuum pump, and the reaction system binding energy, which indicate electron transfer from CdS to CoS2
was kept in the dark with constant stirring for 30 min to establish ad- [28,49,50]. Based on the XPS results, it is deduced that there are close
sorption-and-desorption equilibrium. The production of hydrogen was interaction between CdS and CoS2, forming interfaces in CC2, CC4 and
hourly monitored using a gas chromatograph (Agilent, 7820A, 5A CC6 that act as ‘electron bridges’ to accelerate charge transfer.
molecular sieve column) equipped with a thermal conductivity de- According to the XRD results, the most exposed surfaces of the
tector. The apparent quantum yield (A.Q.Y.) was measured upon CoS2/CdS composites are those of the CoS2 (200) and CdS (001) planes
monochromatic light irradiation according to the method reported in (Fig. 1a). We hence conducted the first-principles DFT simulation
our previous work [31] using the following equations: (Fig. 2a and b) based on the structures of CoS2 (200) and CdS (001)
surfaces: cubic phase with Pa-3 space group for CoS2 and hexagonal
number of reacted electrons
A.Q. Y. = × 100% phase with P63mc space group for CdS. The geometric structure of
number of incident photons
CoS2/CdS is illustrated in Fig. 2c, and the lattice mismatch is no more
2 × (number of evolved hydrogen molecules) than 2.53%. The calculated binding energy of CoS2/CdS is less than
= × 100% zero, which indicates that CoS2 and CdS can form a heterostructure that
number of incident photons
is thermodynamically stable. We performed first-principle calculation
to explore the innate characteristics of charge transfer and separation.
3. Results and discussion The Fermi levels (EF) of the exposed CdS (001) and CoS2 (200) surfaces
are estimated as follows:
The XRD patterns of CdS, CoS2 and CoS2/CdS samples are compiled
Φ = E vac − EF
in Fig. 1a. The peaks of CdS nanorods match well with those of hex-
agonal phase CdS (JCPDS 41-1049) [42], while the main peaks of CoS2 Where Φ is the work function, which is calculated to be 5.17 eV and
nanoparticles are assignable to those of cubic CoS2 (JCPDS 41-1471) 5.23 eV, respectively, for CdS (001) and CoS2 (200) surfaces
[45], suggesting that pure CdS and CoS2 were successfully prepared. As (Fig. 2d − e), and Evac is the energy of a stationary electron at the va-
for the CoS2/CdS composites, the characteristic peaks of CdS can be cuum level (assumed to be 0 eV). Accordingly, the EF of CdS (001) and
detected but those of CoS2 cannot. The absence of CoS2 signal could be CoS2 (200) surfaces can be calculated to be -5.17 and -5.23 eV, re-
due to its high dispersion, low content, and/or poor crystallinity. Si- spectively. The results are consistent with the work of Xu et al., which
milar phenomena were reported in the works of Chen et al. [13] and revealed that the work function (vs. vacuum level) of CdS and CoS2 was
Jiang et al. [42]. It is worth pointing out that the relative intensity of 5.18 eV and 5.49 eV, respectively [51]. It is obvious that the EF of CoS2
the (100) peak (2θ = 24.807°) of CdS as well as that of the composites (200) surface is more negative than that of CdS (001) surface. Thus,
are much higher than that of standard CdS (in the PDF card), indicating when the CdS (001) and CoS2 (200) surfaces are in contact, electrons
the high exposure of the CdS (100) plane. Similarly, the XRD pattern of tend to migrate from CdS to CoS2 for potential equilibrium, restraining
CoS2 indicates the high exposure of the (200) plane. Furthermore, the the recombination of charge carriers as a result. To further demonstrate
actual composition of the CoS2/CdS composites was verified by ICP- the charge transfer at the CoS2/CdS interface, the planar-averaged
OES analysis (Table S1, supporting information, SI†), and the genuine differential charge density was calculated and the result is shown in Fig.
Co:Cd molar ratios are quite in agreement with the nominal ones. S1 (SI†). The positive value represents charge accumulation, while the
Overall, the results of XRD and ICP analyses evidence the successful negative value indicates charge depletion. It is apparent that at the
fabrication of the CoS2/CdS composites. Furthermore, because there is CoS2/CdS interface, electrons primarily transfer from the CdS layer to
no detection of impurity phases, it is deduced that the coupling of CoS2 the CoS2 layer.
with CdS has little effect on the crystallinity and phase structure of CdS. As shown in Fig. 3a and b, the SEM images of pure CdS clearly show
We performed XPS analysis to investigate the chemical states and nanorods with an average length of 800 nm. The surface of the na-
composition of surface elements. The survey spectra of CdS, CoS2 and norods is smooth, showing no impurity or pores [44]. On the other

3
J. Tang, et al. Applied Catalysis A, General 588 (2019) 117281

Fig. 1. (a) XRD patterns, (b) XPS survey spectra, and high-resolution XPS spectra of samples: (c) (f) Cd 3d, (d) (g) S 2p, (e) (h) Co 2p.

hands, pure CoS2 exists as microparticles approximately 1.5 μm in size 0.24 nm are attributed to the (210) plane of CoS2 and (102) plane of
(Fig. 3c and d), which are assemblies of nanoparticles. The morphology CdS, respectively. The results are in line with those of previous reports
of CoS2/CdS heterostructure is displayed in Fig. 3e and f. It is clear that [13,53,54]. The elemental mapping (Fig. 4e − h) and EDX analysis
the CoS2 nanoparticles (about 15 nm in size) are evenly distributed on results (Fig. S3, SI†) of CC4 further disclose homogeneous distribution
the CdS nanorods with no sight of aggregation, and hence could not be of Cd, S and Co elements which confirms the high dispersion of CoS2
detected in XRD analysis. Furthermore, the BET surface area of CdS and nanoparticles on CdS.
CC4 is 32.7 m2 g−1 and 34.7 m2 g−1, respectively (Fig. S2, SI†). Despite The absorption coefficient of the composites is determined ac-
the difference in specific surface area is not apparent, there are much cording to the following equation:
more active sites on the latter due to the presence of CoS2. 1
The TEM and HRTEM images of pristine CdS nanorods and CC4 are α (ω) = ( 2 ) ω [ ε1 (ω)2 + ε2 (ω)2 − ε1 (ω)] 2
displayed in Fig. 4. In consistent with the SEM results, the TEM image of
Fig. 4a clearly shows that the surface of CdS nanorods is smooth. The Where ε1 (ω) and ε2 (ω) are the real and imaginary parts of dielectric
HRTEM image (Fig. 4b) reveals crystal lattice fringes of 0.36 nm which function, respectively, and ω is defined as the function of optical fre-
is assignable to the (100) lattice plane of CdS [52]. The images of CC4 quency. As shown in Fig. 5a, CdS has poor optical absorption and needs
in Figs. 4c and d exhibit the presence of particles approximately 15 nm improvement in the 1.6 − 3.1 eV region for visible light absorption,
in size on the CdS nanorod, and their presence does not alter the while the absorption coefficient of CoS2 is higher than that of CdS.
structure of the CdS nanorod. The lattice spacings of 0.25 nm and Upon coupling of CdS and CoS2, there is significant enhancement of
light absorption. In order to disclose the energy band structure of CdS

4
J. Tang, et al. Applied Catalysis A, General 588 (2019) 117281

Fig. 2. Side and top views of (a) CoS2, (b) CdS and (c) CoS2/CdS; calculated electrostatic potentials along z direction of (d) CdS and (e) CoS2.

and CoS2, we carried out calculation using density functional theory Mott–Schottky curves were plotted to qualitatively analyze the p–n
(DFT). The electronic band structure of CoS2 and CdS are shown in junctions in the absence of illumination as illustrated in Fig. S4 (SI†).
Fig. 5b and c. It is clear that the band gap of CoS2 is 0 eV, which is in The C−2/F−2-E plot of CdS displays a positive slope, which is char-
consistent with the result of Chen et al. [55]. Meanwhile, the band gap acteristic of n-type semiconductor. By obtaining the X intercept by
of CdS is 1.52 eV, in agreement with the calculation result of Garg et al. means of extrapolation of Mott-Schottky plots (in supporting informa-
that planar CdS is a semiconductor with a band gap of 1.60 eV [56]. The tion), the flat-band potential (Vfb) vs. NHE of CdS at pH = 7.0 is
agreement indicates that our calculation adopting the PBE approach is roughly -0.719 V. Using the band gap of CdS (2.38 eV) estimated from
appropriate. The outcomes (Fig. 5) predict that CoS2 has good light- the UV–vis DRS graph, the valence band of CdS is found to be +1.661 V
harvesting ability and good photoelectron conductivity (Eg = 0 eV at vs. NHE at pH = 7.0. With built-in electric field at the CoS2/CdS het-
the CoS2/CdS interface) at the CoS2/CdS heterojunction. It is hence erojunction, the photoelectrons quickly transfer from CdS to CoS2,
theoretically confirmed that the metallic property of CoS2 is crucial in leading to enhanced H2 production under visible light irradiation.
the transfer and separation of photoelectrons for photocatalytic evolu- The photoluminescence (PL) spectra of CoS2/CdS composites are
tion of hydrogen in water splitting. shown in Fig. 6c. That of CdS shows an apparent peak at about 537 nm
The UV–vis diffuse reflectance spectra of the as-prepared samples due to the direct recombination of photogenerated charge carriers. A
are depicted in Fig. 6a. Pure CdS nanorods show obvious absorption in lower PL signal means higher separation efficiency of photo-excited
the region of 400–530 nm as previously reported [15,42], and CoS2 charge carriers. Apparently, the loading of CoS2 nanoparticles on the
exhibits absorption that covers almost the entire visible light region CdS nanorods results in lowering of PL intensity. Among the CoS2/CdS
[45]. As for the composites, they show light absorption pattern similar composites, CC4 shows the lowest PL intensity and hence is the most
to that of CdS but with enhanced absorption in the 530–700 nm region, efficient in preventing the recombination of charge carriers. To further
in agreement with the DFT calculation results (Fig. 5a). The phenom- confirm the mobility of photogenerated charge carriers in the CoS2/CdS
enon could be ascribed to the strong absorption capacity of CoS2 in samples, transient photocurrent was recorded and analyzed using
visible light region and the interfacial interaction between CdS and electrodes coated with the samples under visible light irradiation
CoS2. In addition, the band gaps of the CoS2/CdS samples can be ob- (Fig. 6d). It is clear that photocurrent could only be obtained upon light
tained using the following equation [15]: irradiation. Specifically, pure CdS nanorods is the lowest in photo-
current intensity while CC4 the highest, giving an intensity order of
αhv = A(hv-Eg)n/2
CdS < CC2 < CC6 < CC4. Again, the results demonstrate that the
where α, h, v, A and Eg stand for the absorption coefficient, Planck’s loading of CoS2 nanoparticles on CdS nanorods can efficiently enhance
constant, light frequency, proportionality constant, band gap, respec- the transfer and separation of photoexcited electrons and holes.
tively; and n is dependent on the type of semiconductor. According to Photocatalytic hydrogen evolution under visible light illumination
the Kubelka–Munk plots in Fig. 6b, the band gap values of CdS and CC4 was used to assess the photocatalytic performance of the CoS2/CdS
are 2.38 eV and 2.37 eV, respectively, which are larger than the theo- composites. As illustrated in Fig. 7a, there is no detection of H2 over
retical value (1.52 eV). It is plausibly because the former are experi- CoS2 nanoparticles owing to its class nature of precious metals, while
mental values based on bulk samples while the latter is calculated value the activity of CdS is limited (3.1 mmol·g−1 h−1 after an irradiation of
based on a small model of layered CdS at the CoS2/CdS interface. The 3 h). Gratifyingly, all the CoS2/CdS composites show activity higher

5
J. Tang, et al. Applied Catalysis A, General 588 (2019) 117281

Fig. 3. SEM images of (a) (b) CdS nanorods, (c) (d) CoS2 and (e) (f) CC4.

Fig. 4. TEM images of (a) (b) CdS nanorods, (c) (d) CC4 and (e)–(h) the relevant EDX mappings of Cd, S, Co.

6
J. Tang, et al. Applied Catalysis A, General 588 (2019) 117281

Fig. 5. (a) The optical absorption spectra of


CoS2, CdS and CoS2/CdS; electronic band
structures of (b) CoS2 and (c) CdS calculated by
PBE functional. The Fermi level is placed at zero
and marked by horizontal green dotted lines.
The vertical red dash lines G, M and K represent
the high-symmetry k points: (0, 0, 0), (0, 1/2, 0)
and (-1/3, 2/3, 0), respectively. (For inter-
pretation of the references to colour in this
figure legend, the reader is referred to the web
version of this article).

than that of CdS (Fig. 7b). Among the prepared catalysts, CC4 is the there is apparent decrease in H2 production across the four cycles.
most efficient, giving a H2 production rate of 58.1 mmol·g−1 h−1, Nonetheless, when Na2S/Na2SO3 (0.35 M Na2S/0.25 M Na2SO3) is used,
which is ca. 19 times that of CdS nanorods, four times that of CC4(P) there is no apparent loss, confirming the high stability of CC4 in alka-
prepared by physical mixing (14.7 mmol·g−1 h−1), and three times that line solution. We employed the XRD, TEM and XPS techniques to
of 1 wt%Pt/CdS (19.8 mmol·g−1 h−1). The results reveal that CoS2 can analyze CC4 before and after reaction using either lactic acid or Na2S/
replace noble metals as an effective co-catalyst for photocatalytic gen- Na2SO3 as sacrificial agent. The XRD patterns suggest no change of
eration of hydrogen in water splitting. Among the CdS-based photo- phase structures after reaction (Fig. S6, SI†). There is no change in the
catalysts using Co-containing compounds as co-catalysts for H2 pro- Cd 3d spectra neither after reaction (Fig S7, SI†). Hence the XRD and
duction under visible light irradiation, CC4 is moderate in activity. XPS results confirm the stability of CdS nanorods. However, there is
However, among the reported CdS photocatalysts modified by cobalt obvious loss of Co 2p signal after recycling in lactic acid solution (Fig.
sulfides, CC4 is the highest in activity as well as apparent quantum yield S7, SI†), and the depletion of Co species may result in the loss of
(A.Q.Y., 39.6% at 400 nm) (Table S2, SI†). photocatalytic activity. The loss of CoS2 nanopartices in lactic acid
To evaluate the reusability of CC4, recycling experiments were solution was further confirmed by the TEM images (Fig. S8a and b, SI†).
performed (Fig. S5a, SI†). When lactic acid is used as sacrificial agent, In the case of using Na2S/Na2SO3 as sacrificial agent, Co 2p signal

Fig. 6. (a) UV–vis diffuse reflectance spectra (DRS), (b) the corresponding (αhv)2 versus hv curves, (c) photoluminescence spectra (excitation light: 350 nm), and (d)
transient photocurrent response in 0.5 M Na2SO4 aqueous solution under visible light (λ > 400 nm) of pure CdS and CoS2/CdS composites.

7
J. Tang, et al. Applied Catalysis A, General 588 (2019) 117281

Fig. 7. (a) Extent of photocatalytic H2 evolution, and (b) photocatalytic H2 evolution rate over the as-prepared samples under visible light irradiation. Reaction
conditions: 100 mL (10 vol% LA) aqueous solution, 20 mg catalyst, 300 W Xe-lamp (λ ≥ 400 nm) for 3 h.

for visible-light-driven water splitting.

4. Conclusion

In summary, CoS2 nanoparticles were successfully loaded on the


surface of CdS nanorods using a simple in-situ solvothermal method.
With the deposition of an optimal amount CoS2 nanoparticles, CdS
nanorods exhibit high apparent quantum yield (39.6%, λ = 400 nm)
and outstanding efficiency for H2 production (˜58.1 mmol·g−1 h−1)
under visible light, much higher than that of CdS nanorods deposited
with 1 wt%Pt (19.8 mmol·g−1 h−1). Based on the results of DFT cal-
culation and characterization, it is deduced that with class nature of
precious metals, cobalt disulfide acts as an effective co-catalyst that can
not only efficiently enhance the separation and migration of photo-
generated charge carriers but also offer ample surface sites for proton
reduction. This article demonstrates that CoS2 is a commendable al-
Fig. 8. Possible mechanism of visible-light-driven H2 production over CdS na-
ternative to noble metals as co-catalyst for the enhancement of photo-
norods modified with CoS2 nanoparticles using hole scavenger.
catalytic H2 production under visible light irradiation. Moreover, the
findings provide a new synthetic route for the coupling of zero di-
similar to that of fresh CC4 can still be observed after use (Fig. S7, SI†). mensional nanoparticles with two dimensional semiconductors.
Also, uniform dispersion of CoS2 nanoparticles on CdS can be found in
the TEM image of CC4 after 4 cycles of use (Fig. S8c and d, SI†). The Acknowledgements
results indicate the stability of CC4 when Na2S/Na2SO3 is adopted as
sacrificial agent. This project was financially supported by the NSFC
Overall, CdS nanorods exhibit low activity in photocatalytic hy- (Grants21725602, 21776064, and21671062), and Innovative Research
drogen production due to poor separation efficiency of photogenerated Groups of Hunan Province (Grant 2019JJ10001). C. T. Au thanks HNU
charge carriers. When CoS2 nanoparticles are used as co-catalyst, the for an adjunct professorship.
modified CdS nanorods exhibit higher hydrogen production efficiency
under visible light irradiation. It is because CoS2 nanoparticles can Appendix A. Supplementary data
strengthen the absorption of visible light and enhance the separation
efficiency of photogenerated charge carriers. Furthermore, the loading Supplementary material related to this article can be found, in the
of CoS2 nanoparticles with class nature of precious metals can enrich online version, at doi:https://doi.org/10.1016/j.apcata.2019.117281.
the CdS surface with sites for proton reduction. Based on the above
results, a possible mechanism is proposed for HER under visible light as References
illustrated in Fig. 8. Upon visible light illumination, electrons and holes
are generated on the surface of CdS nanorods. Then the photoexcited [1] M.F. Kuehnel, D.W. Wakerley, K.L. Orchard, E. Reisner, Angew. Chem. Int. Ed. 54
charge carriers are subject to immediate separation and/or transition (2015) 9627–9631.
under the influence of internal electric field. Part of the photogenerated [2] K. Maeda, K. Teramura, D. Lu, T. Takata, N. Saito, Y. Inoue, K. Domen, Nature 440
(2006) 295–295.
electrons in the CdS valence band (VB) transfer to the CdS conduction [3] X.F. Qian, D. Yue, Z. Tian, M. Reng, Y. Zhu, M. Kan, T. Zhang, Y. Zhao, Appl. Catal.
band (CB) and ulteriorly migrate to the catalyst surface, leaving holes in B Environ. 193 (2016) 16–21.
the VB of CdS. Finally, lactic acid molecules (or Na2S/Na2SO3) are [4] R.E. Blankenship, D.M. Tiede, J. Barber, G.W. Brudvig, G. Fleming, M. Ghirardi,
Science 332 (2011) 805–809.
oxidized by the VB holes and water molecules are reduced to hydrogen [5] S.S. Yi, X.B. Zhang, B.R. Wulan, J.M. Yan, Q. Jiang, Energy Environ. Sci. 11 (2018)
by the CB electrons. The presence of CoS2 nanoparticles on CdS can 3128–3156.
lower the migration resistance of photogenerated electrons for facile [6] X. Zou, Y. Zhang, Chem. Soc. Rev. 44 (2015) 5148–5180.
[7] J. Tian, Q. Liu, A.M. Asiri, X. Sun, J. Am. Chem. Soc. 136 (2014) 7587–7590.
transit to CoS2 nanoparticles, consequently enhancing the separation of [8] X.B. Chen, L. Liu, P.Y. Yu, S.S. Mao, Science 331 (2011) 746–750.
charge carriers and promoting the photocatalytic performance of CdS [9] Y. Zhu, T. Cao, C. Cao, J. Luo, W. Chen, L. Zheng, J. Dong, J. Zhang, Y. Han, Z. Li,

8
J. Tang, et al. Applied Catalysis A, General 588 (2019) 117281

C. Chen, Q. Peng, D. Wang, Y. Li, ACS Catal. 8 (2018) 10004–10011. [31] J. He, L. Chen, F. Wang, Y. Liu, P. Chen, C.T. Au, S.F. Yin, ChemSusChem 9 (2016)
[10] Y. Shi, Y. Zhou, D.R. Yang, W.X. Xu, C. Wang, F.B. Wang, J.J. Xu, X.H. Xia, 624–630.
H.Y. Chen, J. Am. Chem. Soc. 139 (2017) 15479–15485. [32] J. He, L. Chen, Z.Q. Yi, C.T. Au, S.F. Yin, Ind. Eng. Chem. Res. 55 (2016)
[11] S.P. Adhikari, Z.D. Hood, H. Wang, R. Peng, A. Krall, H. Li, V.W. Chen, K.L. More, 8327–8333.
Z. Wu, S. Geyer, A. Lachgar, Appl. Catal. B Environ. 217 (2017) 448–458. [33] J. He, L. Chen, Z.Q. Yi, D. Ding, C.T. Au, S.-F. Yin, Catal. Commun. 99 (2017)
[12] J. Yu, J. Low, W. Xiao, P. Zhou, M. Jaroniec, J. Am. Chem. Soc. 136 (2014) 79–82.
8839–8842. [34] Z. Wang, J. Peng, X. Feng, Z. Ding, Z. Li, Catal. Sci. Technol. 7 (2017) 2524–2530.
[13] H. Chen, D. Jiang, Z. Sun, R.M. Irfan, L. Zhang, P. Du, Catal. Sci. Technol. 7 (2017) [35] H. Li, P. Deng, Y. Hou, Mater. Lett. 229 (2018) 217–220.
1515–1522. [36] H. Wang, Z. jin, Sustain. Energy Fuels 3 (2019) 173–183.
[14] K. He, J. Xie, Z.Q. Liu, N. Li, X. Chen, J. Hu, X. Li, J. Mater. Chem. A 6 (2018) [37] D.A. Reddy, E.H. Kim, M. Gopannagari, R. Ma, P. Bhavani, D.P. Kumar, T.K. Kim,
13110–13122. ACS Sustain. Chem. Eng. 6 (2018) 12835–12844.
[15] S. Ma, Y. Deng, J. Xie, K. He, W. Liu, X. Chen, X. Li, Appl. Catal. B Environ. 227 [38] Y.Y. Fu, C. Zhu, C. Liu, M. Zhang, H. Wang, W. Shi, H. Huang, Y. Liu, Z. Kang, Appl.
(2018) 218–228. Catal. B Environ. 226 (2018) 295–302.
[16] J.H. Yang, D. Wang, H.X. Han, C. Li, Acc. Chem. Res. 46 (2013) 1900–1909. [39] Y. Zhu, L.F. Song, N. Song, M.X. Li, C. Wang, X.F. Lu, ACS Sustain. Chem. Eng. 7
[17] K. Wu, Z. Chen, H. Lv, H. Zhu, C.L. Hill, T. Lian, J. Am. Chem. Soc. 136 (2014) (2019) 2899–2905.
7708–7716. [40] M.S. Faber, R. Dziedzic, M.A. Lukowski, N.S. Kaiser, Q. Ding, S. Jin, J. Am. Chem.
[18] L. Zhang, X. Fu, S. Meng, X. Jiang, J. Wang, S. Chen, J. Mater. Chem. A 3 (2015) Soc. 136 (2014) 10053–10061.
23732–23742. [41] C.J. Chen, P.T. Chen, M. Basu, J. Mater. Chem. A 3 (2015) 23466–23476.
[19] I. Majeed, U. Manzoor, F.K. Kanodarwala, M.A. Nadeem, E. Hussain, H. Ali, [42] D. Jiang, Z. Sun, H. Jia, D. Lu, P. Du, J. Mater. Chem. A 4 (2016) 675–683.
A. Badshah, J.A. Stride, M.A. Nadeem, Catal. Sci. Technol. 8 (2018) 1183–1193. [43] F. Zhang, H.Q. Zhuang, J. Song, Y.L. Men, Y.X. Pan, S.H. Yu, Appl. Catal. B Environ.
[20] J.M. Lee, H.B. Jin, I.Y. Kim, Y.K. Jo, J.W. Hwang, K.K. Wang, M.G. Kim, Y.R. Kim, 226 (2018) 103–110.
S.J. Hwang, Small 11 (2015) 5771–5780. [44] D. Lang, F. Cheng, Q. Xiang, Catal. Sci. Technol. 6 (2016) 6207–6216.
[21] Z. Song, B. Hong, X. Zhu, F. Zhang, S. Li, J. Ding, X. Jiang, J. Bao, C. Gao, S. Sun, [45] S. Peng, L. Li, S.G. Mhaisalkar, M. Srinivasan, S. Ramakrishna, Q. Yan,
Appl. Catal. B Environ. 238 (2018) 248–254. ChemSusChem 7 (2014) 2212–2220.
[22] M. Gopannagari, D.P. Kumar, D.A. Reddy, S. Hong, M.I. Song, T.K. Kim, J. Catal. [46] S. Ji, T. Li, Z.D. Gao, Y.Y. Song, J.J. Xu, Chem. Commun. 54 (2018) 8765–8768.
351 (2017) 153–160. [47] L. Shi, D. Li, P. Yao, J. Yu, C. Li, B. Yang, C. Zhu, J. Xu, Small 14 (2018) 1802716.
[23] J. Chen, X.J. Wu, L. Yin, B. Li, X. Hong, Z. Fan, B. Chen, C. Xue, H. Zhang, Angew. [48] Q. Pan, Y. Liu, L. Zhao, Chem. Eng. J. 351 (2018) 603–612.
Chem. Int. Ed. 54 (2015) 1210–1214. [49] C. Yang, X. Liang, X. Ou, Q. Zhang, H.S. Zheng, F. Zheng, J.H. Wang, K. Huang,
[24] X. Zhang, Z. Lai, C. Tan, H. Zhang, Angew. Chem. Int. Ed. 55 (2016) 8816–8838. M. Liu, Adv. Funct. Mater. 29 (2019) 1807971.
[25] J.H. Lin, Y.H. Tsao, M.H. Wu, T.M. Chou, Z.H. Lin, J.M. Wu, Nano Energy 31 (2017) [50] M. Cai, J. Han, Y. Lin, W. Liu, X. Luo, H. Zhang, M. Zhong, Electrochim. Acta 287
575–581. (2018) 1–9.
[26] C. Li, H. Wang, S.B. Naghadeh, J.Z. Zhang, P. Fang, Appl. Catal. B Environ. 227 [51] Y. Xu, M.A.A. Schoonen, Am. Mineral. 85 (2000) 543–556.
(2018) 229–239. [52] X.L. Yin, G.Y. He, B. Sun, W.J. Jiang, D.J. Xue, A.D. Xia, L.J. Wan, J.S. Hu, Nano
[27] Y. Zhang, Z. Peng, S. Guan, X. Fu, Appl. Catal. B Environ. 224 (2018) 1000–1008. Energy 28 (2016) 319–329.
[28] B.C. Qiu, Q.H. Zhu, M.M. Du, L.G. Fan, M.Y. Xing, J.L. Zhang, Angew. Chem. Int. [53] Y. Pan, X. Cheng, L. Gong, L. Shi, T. Zhou, Y. Deng, H. Zhang, ACS Appl. Mater.
Ed. 56 (2017) 1–6. Interfaces 10 (2018) 31441–31451.
[29] P. Tan, Y. Liu, A. Zhu, W. Zeng, H. Cui, J. Pan, ACS Sustain. Chem. Eng. 6 (2018) [54] Y. Pan, X. Cheng, L. Gong, L. Shi, H. Zhang, Nanoscale 10 (2018) 20813–20820.
10385–10394. [55] C.J. Chen, P.T. Chen, Mrinmoyee Basu, J. Mater. Chem. A 3 (2015) 23466–23476.
[30] Q. Yuan, D. Liu, N. Zhang, W. Ye, H. Ju, L. Shi, R. Long, J. Zhu, Y. Xiong, Angew. [56] P. Garg, S. Kumar, I. Choudhuri, A. Mahata, B. Pathak, J. Phys. Chem. C 120 (2016)
Chem. Int. Ed. 56 (2017) 4206–4210. 7052–7060.

You might also like