You are on page 1of 6

Journal of Industrial and Engineering Chemistry 20 (2014) 3869–3874

Contents lists available at ScienceDirect

Journal of Industrial and Engineering Chemistry


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

Heterostructured metal sulfide (ZnS–CuS–CdS) photocatalyst for high


electron utilization in hydrogen production from solar water splitting
Eunpyo Hong, Duckchen Kim, Jung Hyeun Kim *
Department of Chemical Engineering, University of Seoul, 163 Siripdaero, Dongdaemun-gu, Seoul 130-743, South Korea

A R T I C L E I N F O A B S T R A C T

Article history: H2-production from solar water splitting attracts great attention due to its cleanness and renewable
Received 8 October 2013 characteristics in converting solar energy into chemical energy. A sequential fabrication approach for
Accepted 30 December 2013 metal sulfide photocatalysts consisting of three components (ZnS, CuS, and CdS) is designed in order to
Available online 8 January 2014
achieve efficient charge flows by introducing CuS and CdS on ZnS surfaces. ZnS particles are firstly
synthesized as a base material through colloidal precipitation, and then the ZnS surfaces are sequentially
Keywords: modified by adding CuS (cation exchange reaction) and CdS (ionic reaction) precursors. For the final
Solar water splitting
products (ZnS–CuS–CdS), atomic compositions are carefully analyzed. The ZnS–CuS–CdS composite
Hydrogen production
Metal sulfide photocatalyst
photocatalyst reaches a maximum H2-production rate (837.6 mmol/g h) at Cu 0.81 wt% and Cd 14.7 wt%
Charge separation under the standard solar irradiation (1 kW/m2, AM 1.5 G) compared to the ZnS only case (13.5 mmol/g h).
This highly enhanced photocatalytic activity is due to synergistic effects of heterostructured ZnS–CuS–
CdS photocatalyst which can improve charge flow and light absorption. This study shows a great
possibility of manufacturing multi-component metal sulfide systems for improving H2-production
activity from solar water splitting.
ß 2014 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights
reserved.

1. Introduction H2-production from aqueous solution was achieved by loading CuS


component which promote photoinduced interfacial charge
Investigations of H2-production from the solar water splitting transfer (IFCT) of electrons from the valence band of ZnS to the
have increased in recent researches due to its environmental CuS clusters, and the photocatalytic activity of the CuS/ZnS case
cleanness and unlimited utilization of solar energy. Various types was remarkably improved for the H2-production rate under visible
of photocatalysts such as TiO2 base [1–4], Z-scheme system base light condition (l  420 nm) with the 350 W xenon arc lamp [12].
[5,6], alkali metal base [7,8], and metal sulfide base [9–18] However, absorption range and intensity of the light irradiation is
materials for increasing the amount of H2-production from the still limited. Another strategy is a cationic doping with metal
solar water splitting have been reported since Fujishima and component such as copper (Cu) [10], cadmium (Cd) [13], and
Honda [19]. Among those photocatalyst materials, metal sulfides indium (In) [22] to improve activity and stability. These doped
are considered as a promising candidate because of their suitable compounds show great ability to produce hydrogen from water
band gap energy, band position, and catalytic activity. Particularly, splitting. Although these compounds have good activity, it has
zinc sulfide (ZnS) photocatalyst shows high activity on H2 some drawbacks in controlling the homogeneous dispersion of the
evolution because of the rapid generation of electron-hole pairs minor components with regular morphology and good crystallini-
by photon energy and the highly negative potential of excited ty. Additionally, absoption of these compounds are increased
electrons in the conduction band [20,21]. However, the ZnS’s sole compared to pure ZnS; however, its absorption is lower than
use is limited due to its relatively wide band gap energy (3.35 eV) narrow band gap materials such as CdS.
for the visible light condition [11]. In this study, we propose a new strategy for developing efficient
In order to solve this limited application, variety of researches photocatalyst for solar water splitting consisting of ZnS, CuS, and
has been conducted. Most recently, enhanced photocatalytic CdS materials. In order to make use of their own properties, ZnS–
activity of ZnS nanosheets for visible-light-driven photocatalytic CuS–CdS components were sequentially synthesized to form a
composite structure. In the composite, each component keeps their
own phases, so it can display their characteristics. The effects of
* Corresponding author. Tel.: +82 2 6490 2369; fax: +82 2 6490 2364.
thermal treatment, CuS and CdS contents, and its synergetic effect
E-mail address: jhkimad@uos.ac.kr (J.H. Kim). were also analyzed, and the composite materials were further used

1226-086X/$ – see front matter ß 2014 The Korean Society of Industrial and Engineering Chemistry. Published by Elsevier B.V. All rights reserved.
http://dx.doi.org/10.1016/j.jiec.2013.12.092
3870 E. Hong et al. / Journal of Industrial and Engineering Chemistry 20 (2014) 3869–3874

for H2-production by solar water splitting under the standard solar junction formation for effective electron transfer. Therefore, the
irradiation condition (1 kW/m2 with AM 1.5 G filter). ZnS–CuS particles were prepared first (by cation exchange
In addition, previous works about photocatalytic H2-production reaction), and then CdS components were loaded on ZnS–CuS
have been reported under different light illumination conditions particle surfaces as described above.
such as wavelength and intensity; thus, it is very difficult to
compare reported results with each other. Therefore, it is highly 2.3. Photocatalytic water splitting
recommended to use a general standard solar irradiation with
known power density. In a practical point of view, it is also very H2-production activity of all samples was examined in 150 mL
important to consider natural sunlight as a power source. The aqueous solution containing 0.1 g photocatalyst and 60 mM
condition for the AM 1.5 G spectra is generally used as the sodium sulfide nonahydrate under the standard solar power
representative average sun condition in solar industries for all density condition (1 kW/m2) with a solar simulator (Peccel
standardized testing. In this study, the current reference spectrum Technologies, PED-L11) equipped with 150 W xenon lamp and
adopted by the international terrestrial photovoltaics community AM 1.5 G filter. The standard light intensity was carefully
is used which is known in International Electrotechnical Commis- calibrated using the Si photodiode (PEC cell, BS-520). The Pyrex
sion (IEC) Standard 60904-3 and American Society for Testing and glass reactor which is a flat top cylindrical shape of 330 mL total
Materials (ASTM) Standard G159 [23]. volume was exposed to the simulated solar irradiation. Prior to
light illumination, the reactor was completely deaerated with a
2. Experimental nitrogen gas for 1 h. Photocatalytic reaction was performed for
150 min, and the evolved gas was collected every 30 min with a
2.1. Materials gas-tight syringe and analyzed by a gas chromatograph (Chrom-
pack, CP9001) equipped with a thermal conductivity detector and
Zinc nitrate hexahydrate (Zn(NO3)26H2O, 98%), sodium sulfide a 5 Å molecular sieve column.
nonahydrate (Na2S9H2O, 98%), cadmium nitrate tetrahydrate
(Cd(NO3)24H2O, 99%), and copper(II) nitrate trihydrate 2.4. Characterizations
(Cu(NO3)23H2O, 99–104%) were purchased from Aldrich and used
without further purification. Ethanol was obtained from Duksan The crystalline structures of the samples were analyzed by an
pure chemicals (C2H6OH, 99.9%), and deionized (DI) water was X-ray diffractometer (XRD, Rigaku, DMAX-2500) using Cu Ka
used in all experiments. irradiation with accelerating voltage and the applied current were
40 kV and 200 mA, respectively. High resolution transmission
2.2. Preparation of ZnS–CuS–CdS photocatalyst composites electron microscopy (HR-TEM) images were taken from FEI
COMPANY (Tecnai G2 F30 ST) operated at 300 kV. For determining
ZnS nanoparticle and thermal treatment: for the first stage, the the atomic composition of the samples, atomic absorption
ZnS colloidal particles were synthesized in ionic solutions spectrometer (AAS, Thermo Scientific, ICE3000 series) was used.
(Zn2+ + S2) from zinc nitrate hexahydrate (Zn(NO3)26H2O) and The light absorption spectra of the photocatalysts were measured
sodium sulfide nonahydrate (Na2S9H2O). In a typical preparation by UV–vis spectrometer (Scinco, S-4100) with a diffuse reflectance
procedure of the chemical bath, 300 mL aqueous solution of 0.05 M module. The chemical composition and binding states of the
zinc source was added dropwise into the same amount of 0.05 M samples were analyzed from X-ray photoelectron spectra (XPS, PHI
sodium sulfide aqueous solution. Then, white colored ZnS particles 5000 VersaProve, Ulvac-PHI) measured by monochromatic Mg Ka
were formed in the solution and it was further stirred for 10 min. radiation as the X-ray source. Prior to data analysis, the binding
After stirring, it was carefully rinsed with DI-water and ethanol, energies were calibrated using C–C component of the C 1s peak at
and dried in a room air condition. The dried ZnS particles were 284.6 eV.
thermally treated at 500 8C for 30 min in air condition, and the ZnS
samples as-prepared and after thermal treatment are named as Z0 3. Results and discussion
and Z500, respectively.
ZnS–CuS composites: for the second stage, CuS clusters were For the first stage, ZnS particles synthesized by facile ionic
introduced on the preformed ZnS particle surfaces by cation reaction (Zn2+ + S2), XRD patterns are shown in Fig. 1. The XRD
exchange reaction with a series of feedstock (0.0207, 0.0414, and peaks from the as-prepared ZnS photocatalyst (Z0) show cubic ZnS
0.0828 mmol) of copper(II) nitrate trihydrate (Cu(NO3)23H2O), crystalline structure with 2u values at 28.5608 (1 1 1), 47.5068
which was mixed drop by drop to 20 mL ethanol suspension (2 2 0), and 56.3708 (3 1 1), which are in good agreement with the
containing 0.3 g ZnS particles. Rapid color change was observed literature values (JCPDS no. 71-5971). After thermal treatment
during mixing process from white to deep green. After stirring for (Z500) at 500 8C for 30 min in air, the diffraction peaks of the cubic
10 min, the ZnS–CuS particles were cleaned and dried for further ZnS phase become strong. It means that crystallinity and grain size
experiments. are improved but cubic crystalline structure is not changed. In
ZnS–CdS composites: to add the CdS sensitizers on ZnS surfaces, addition, several new peaks appeared at 2u values at 31.7378
a simple ionic reaction (Cd2+ + S2) was used with ZnS particles as (1 0 0), 34.3788 (0 0 2), and 36.2158 (1 0 1) which are well matched
nucleation sites. The CdS particles were introduced on the as- with hexagonal phase ZnO crystals (JCPDS no. 76-0704), and it can
formed ZnS particle surfaces. The ZnS particles (0.25 g) were be attributed to the surface oxidation during thermal treatment
dispersed in various concentrations (pre-determined amount: process. Surface oxidation of ZnS particles is an exchange reaction
0.25, 0.50, 0.75, and 1.00 mmol) of the sodium sulfide nonahydrate between oxygen and sulfur as explained in Eq. (1) [24]:
aqueous solution (100 mL), and then cadmium nitrate tetrahydrate
2ZnS þ 3O2 ! 2ZnO þ 2SO2 " (1)
aqueous solution was slowly added (molar ratio of the Cd and S
was 1:1). The composite particles were also washed and dried for The ZnS crystalline domain size (D) can be calculated via
further experiments. Scherrer formula [25]:
ZnS–CuS–CdS composites (final product): to fabricate the final
ZnS–CuS–CdS photocatalyst composites, above mentioned proce- kl
D¼ ; (2)
dures were sequentially applied for desired serial chemical bcosu
E. Hong et al. / Journal of Industrial and Engineering Chemistry 20 (2014) 3869–3874 3871

Fig. 1. This thermal treatment greatly increased the H2-production


rate from 13.5 mmol/g h (Z0) to 75.0 mmol/g h (Z500). It can be due
to the improved crystallinity which is a crucial factor in utilization
of photoexited electrons for H2-production. Although photocata-
lytic reactions are occurred at the surface of photocatalyst, the
crystallinity is directly involved in movements of the excited
electron-hole pairs. In this respect, Jang et al. [26–28] reported that
CdS NW/TiO2 NP showed considerably lower activity in H2-
production than CdS (bulk)/TiO2 NP composite photocatalysts
because of the low crystallinity of CdS NW. Another possible
reason for improved H2-production through thermal treatment is a
formation of oxide layer on ZnS surfaces. Many studies in H2-
production from ZnS/ZnO composites revealed that relative ratio of
ZnS/ZnO and surface properties of the particles are also important
factors, and this heterostructure improves photocatalytic activity
[24,29,30]. The improvement of photocatalytic activity from the
ZnS/ZnO composites could be attributed to the increased charge
separation process and the expanded absorption range of solar
irradiation. As pointed in the previous work, Zhao et al. [24]
Fig. 1. XRD patterns of ZnS particles (a) before (Z0) and (b) after (Z500) thermal reported that ZnS base material was thermally treated at 500 8C to
treatment. form ZnS–ZnO composite structure because ZnO crystals were not
formed below 500 8C. Thus, we used 500 8C thermal treatment
where k is the shape factor of the particle (k = 0.98), l is the X- condition for ZnS–ZnO crystal formation.
ray wavelength (1.5406 Å), b is the half-maximum peak width, and First step to improve photocatalytic activity of the Z500 sample
u is the Bragg diffraction angle. Average crystalline sizes of ZnS is by introducing CuS clusters to ZnS surfaces by cationic exchange
particles before and after thermal treatment are 5.5 nm and reaction. CuS cluster formation mechanism is shown below
12.5 nm, respectively. Eq. (3)–(5) [31]. Initially, due to the dynamic equilibrium of
Fig. 2 shows the high resolution transmission electron Eq. (3) in ethanol, there is a small amount of S2 ions around the
microscope (HR-TEM) images of ZnS particles before (Fig. 2a) solid ZnS particles. As the Cu2+ is added to solution, the CuS clusters
and after (Fig. 2b) thermal treatment. Selected crystalline are formed on the ZnS surfaces. Fig. 2c and d shows CuS loaded ZnS
boundaries of the ZnS particles are marked in Fig. 2a and b, nanoparticles. The CuS particles are attached on ZnS surfaces and
approximately. ZnS nanoparticles with crystal size about 5 nm are well dispersed (only one or two CuS particles are observed on ZnS
agglomerated as shown in Fig. 2a. After thermal treatment, the particles).
crystalline size is increased bigger than 10 nm shown in Fig. 2b.
These results are well corresponding to the XRD data shown in ZnS $ Zn2þ þ S2 (3)

Fig. 2. HR-TEM images of (a) before (Z0) and (b) after (Z500) thermal treatment, and magnified ZnS–CuS samples (c) and (d).
3872 E. Hong et al. / Journal of Industrial and Engineering Chemistry 20 (2014) 3869–3874

Fig. 3. Hydrogen production from ZnS (Z500) and ZnS–CuS composites as a function
of Cu content. Inset is a schematic energy diagram for the ZnS–CuS composite. Fig. 4. XPS spectra of Cu 2p of ZnS–CuS composites before and after light irradiation
(for 150 min) without sacrificial agent (S2 ions).

Cu2þ þ S2 $ CuS (4)


photo-generated electron transfers, and it improves the charge
ZnS þ Cu2þ $ CuS þ Zn2þ (5) separation efficiency through ZnS–CuS + hy ! ZnS(h+)–CuS(e).

Fig. 3 shows H2-production rate from Z500 and ZnS–CuS CuS=ZnS þ hv ! CuSðe Þ=ZnSðhþ Þ (6)
composite particles. Hydrogen evolution from the ZnS–CuS
composites shows much higher than from Z500 and it is also 2CuS þ 2e ! Cu2 S þ S2 (7)
varied as a function of Cu contents. By loading the CuS particles on
the ZnS surfaces, electrons can be exited much easily from the Cu2 S þ 2Hþ þ S2 ! 2CuS þ H2 (8)
valence band of ZnS to CuS (2.94 eV, [12]) compared to the band to
band transition (3.35 eV, [12]), as shown in Fig. 3 inset diagram. Second step is to introduce relatively narrow band gap material
This interfacial charge transfer (IFCT) phenomenon is also reported for improving visible light absorption. CdS semiconductor material
by other studies [12,31–35] between ZnS and CuS components. is used as a sensitizer considering the band gap energy (2.45 eV,
Additionally, this interfacial charge separation suppresses the [37]) and band position (valence and conduction bands). The CdS
possible electron-hole recombination, and the loaded CuS clusters particles formed on ZnS surfaces can expand light harvesting
play a significant role as reaction sites for H2-production. H2- region and form a potential gradient at the interface of CdS and
production rate shows the highest value (210.6 mmol/g h) at Cu ZnS. Therefore, recombination process of the photo excited
0.33 wt%. However, the higher amount of CuS loading than the electron–hole pairs can be reduced. Fig. 5a shows the amount of
optimum value rather decreases the photocatalytic activity of the evolved hydrogen from ZnS–CdS composites as a function of Cd
ZnS–CuS because excessively formed CuS (1) blocks the incoming content, and the highest H2-production rate (208 mmol/g h) is
light to the ZnS base material and thus reduces electron observed at 24.2 wt% Cd. It is also much enhanced H2-production
generation; (2) diminishes photocatalytic activity of CuS clusters rate compared to the ZnS only case (Z500). This is because CdS has
by increasing their cluster size, by reducing the number of active relatively smaller band gap (2.45 eV) than that of ZnS (3.35 eV) by
ZnS surface sites, and by increasing charge recombination centers itself; it enhances electron excitation from visible light under solar
between ZnS–CuS and CuS–CuS junctions. It is very similar irradiation and leads high charge separation. Fig. 5b shows UV–vis
phenomenon with the metal loading on a semiconductor particle spectra of Z500 and ZnS–CdS composites as a function of Cd
such as Pt–TiO2 and Au–TiO2 [36]. content. Diffuse reflective spectra of ZnS–CdS photocatalyst are
Photoinduced IFCT process includes chemical state change of highly increased compared to Z500 sample, and its absorption
the CuS. The major electron transfer mechanisms on photocata- intensity is gradually increased with increasing Cd content for
lysts under light illumination are shown below Eq. (6)– (8) [12]. visible light region. Although, light absorption is increased with
Evidence of this change is demonstrated by X-ray photoelectron increasing Cd content, an excess use of CdS rather decreases the
spectroscopy (XPS) in Fig. 4. The XPS spectrum of Cu in the 2p H2-production rate possibly due to the increased electron loss by
region before the solar irradiation shows the binding energies of long pathways through CdS clusters [3,38].
932.8 eV and 952.8 eV for Cu 2p3/2 and Cu 2p1/2, and these are From above results of ZnS–CuS and ZnS–CdS, it clearly reveals
typical values for Cu2+ in CuS materials. However, after the solar that CuS and CdS components can improve photocatalytic activity
irradiation for 150 min without sacrificial agent (S2 ions); the XPS by conjugation with ZnS base material as a composite form.
spectrum shows slight shifts of the binding energies to the lower Optimum amounts of the components are also important to use for
region (932.6 eV and 952.5 eV) for Cu 2p3/2 and Cu 2p1/2, and it high H2-production, and the H2-production activities from the
indicates the presence of Cu1+ as reported by other researcher also optimum compositions including experimental conditions are
[12,34]. Therefore, some amounts of Cu2+ in CuS are possibly summarized in Table 1 (S1S3). However, combination of these
reduced to Cu2S, and it further demonstrates the electron transfer individual optimum amounts of the components leads only 3.6
from the valence band of ZnS to CuS for the partial reduction. times increment (S4) of the H2-production. It could be possibly
Therefore, the heterojunction between the ZnS body and CuS attributed to the disbalance of the electron–hole flows that the
clusters is considered as a favorable connection site for the number of CuS clusters are not enough for draining electrons
E. Hong et al. / Journal of Industrial and Engineering Chemistry 20 (2014) 3869–3874 3873

Fig. 6. Schematic illustration of band energy levels and charge transfers for the ZnS–
CuS–CdS photocatalyst.

transfer efficiency to CuS reaction sites. Further decreased CdS


content (S7) rather reduces H2-production down to 6 times. Thus,
as expected, the composition of the multi-component photo-
catalyst system should be precisely optimized to achieve high
efficiency by evaluating H2-production rate because electron-hole
generation, charge separation, and electron movement can be
totally different from the two components cases.
Based on the H2-production results, a schematic of the three
component system is shown in Fig. 6 illustrating charge separation
and electron flow (CdS ! ZnS ! CuS). So, CuS clusters act as
electron wells where CuS reduction takes place (2CuS + 2e !
Cu2S + S2), and the reduced Cu2S state returns to CuS while
producing H2 (Cu2S + 2H+ + S2 ! 2CuS + H2). Therefore, it is very
important to have an optimum amount of CuS clusters for this
Fig. 5. Hydrogen production (a) and UV–vis absorbance spectra (b) as a function of multi-component photocatalytic material (CuS deficiency: elec-
Cd content from the ZnS–CdS photocatalyst system. tron-hole recombination increased, excess CuS: incident light
blocked and photocatalyst activity reduced).
Fig. 7a demonstrates the effect of CuS and CdS additions to ZnS
excited not only from the valence band of ZnS but also from the on the UV–vis light absorbance. The absorbance increment (S2 and
valence band of CdS. In other words, the electrons generated from S6) at wavelengths higher than 400 nm can be attributed to the
ZnS and CdS under solar irradiation can be too crowded to transfer interfacial electron transfer from the valence band of ZnS to CuS,
efficiently to CuS cluster sites, and thus some of them can be lost by and also a significant increase (S3) in the absorption step at
combining with holes presented in ZnS. Our primary goal is to wavelengths from 400 nm to 600 nm can be due to the CdS
achieve synergistic H2-production from ZnS–CuS–CdS composite particles having smaller band gap than that of ZnS. Therefore, it
photocatalyst under the standard solar irradiation. Therefore, it is clearly indicates the sequential addition of CuS and CdS
necessary to test this assumption by increasing CuS amount or components onto the ZnS particle surfaces instead of incorporating
decreasing CdS amount for the three components system. H2- into the ZnS lattice sites schematically illustrated in Fig. 6. We
production rates from the sample S5 (increased CuS) and the S6 performed additional XRD analyses for the ZnS–CuS(S2), ZnS–
(decreased CdS) are 6.1 times and 11.2 times higher than the S1 CdS(S3), and ZnS–CuS–CdS(S6) samples (not shown in this paper),
case, and so it convincingly demonstrates the enhancement of but there is no additional noticeable new peaks observed from CuS
electron utilizations for H2-production by improving the electron and CdS crystals. It is mainly due to the formation of amorphous

Table 1
H2-production rate for various Cu and Cd compositions in ZnS–CuS–CdS photocatalysts.

Sample Cu Cd H2-production rate (mmol/g h)

Feedstock (mmol) Measured content (wt%)b Feedstock (mmol) Measured content (wt%)b

S1a – – – – 75.0
S2 0.021 0.33 – – 210.6
S3 – – 0.75 24.2 208.3
S4 0.021 0.27 0.75 21.2 265.7
S5 0.041 0.85 0.75 21.1 459.6
S6 0.041 0.81 0.50 14.7 837.6
S7 0.041 0.80 0.25 6.5 452.1
a
Z500 (thermal treatment at 500 8C for 30 min).
b
wt% measured by atomic absorption spectrometer (AAS, Thermo Scientific, ICE3000 series).
3874 E. Hong et al. / Journal of Industrial and Engineering Chemistry 20 (2014) 3869–3874

that the solar light causes (1) interfacial charge transfer from the
valence band of ZnS to CuS, which results in the partial reduction of
CuS to Cu2S for enhancing H2-production, (2) efficient electron
excitation and separation in CdS due to the relatively smaller band
gap than that of ZnS, and thus the separated electrons flow to CuS
cluster sites by reducing electron-hole recombination possibility.
Improved photocatalytic activity at wavelengths higher than
400 nm is also observed in UV-vis absorbance spectra due to (1)
absolute absorbance increased by CuS addition, (2) significant
absorbance step (400 nm < l < 600 nm) appeared by CdS addi-
tion. This study demonstrates a synergistic effect of the three metal
sulfide system fabricated in a sequential manner on the photo-
catalytic H2-production under the standard solar irradiation
condition.

Acknowledgment

This work was supported by the Human Resources Develop-


ment program (No. 20124010203260) of the Korea Institute of
Energy Technology Evaluation and Planning (KETEP) grant funded
by the Korea government Ministry of Trade, Industry and Energy.

References

[1] E. Hong, J. Choi, J.H. Kim, Thin Solid Films 527 (2013) 363–368.
[2] H. Park, W. Choi, M. Hoffmann, J. Mater. Chem. 18 (2008) 2379–2385.
[3] E. Hong, J.H. Kim, S. Yu, J.H. Kim, Korean J. Chem. Eng. 28 (2011) 1684–1687.
[4] A. Pérez-Larios, R. Lopez, A. Hernández-Gordillo, F. Tzompantzi, R. Gómez, L.
Torres-Guerra, Fuel 100 (2012) 139–143.
[5] M. Higashi, R. Abe, A. Ishikawa, T. Takata, B. Ohtani, K. Domen, Chem. Lett. 37
(2008) 138–139.
[6] H. Kato, M. Hori, R. Konta, Y. Shimodaira, A. Kudo, Chem. Lett. 33 (2004) 1348–
1349.
[7] H. Kato, A. Kudo, Catal. Lett. 58 (1999) 153–155.
[8] A. Kudo, H. Kato, Chem. Lett. 26 (1997) 867–868.
[9] T. Arai, S.-i. Senda, Y. Sato, H. Takahashi, K. Shinoda, B. Jeyadevan, K. Tohji, Chem.
Fig. 7. UV–vis absorbance spectra (a) and time course of the H2-production activity Mater. 20 (2008) 1997–2000.
(b) from the four different types of photocatalysts. [10] A. Kudo, M. Sekizawa, Catal. Lett. 58 (1999) 241–243.
[11] J.F. Reber, K. Meier, J. Phys. Chem. 88 (1984) 5903–5913.
[12] J. Zhang, J. Yu, Y. Zhang, Q. Li, J.R. Gong, Nano Lett. 11 (2011) 4774–4779.
[13] K. Zhang, D. Jing, C. Xing, L. Guo, Int. J. Hydrogen Energy 32 (2007) 4685–4691.
structure or due to the existence of small amount of crystals. In the [14] W. Zhang, R. Xu, Int. J. Hydrogen Energy 34 (2009) 8495–8503.
[15] C. Xing, Y. Zhang, W. Yan, L. Guo, Int. J. Hydrogen Energy 31 (2006) 2018–2024.
case of CuS, Cu contents in S2 and S6 are only 0.33 and 0.81 wt%, [16] W. Zhang, Z. Zhong, Y. Wang, R. Xu, J. Phys. Chem. C 112 (2008) 17635–17642.
and thus it can be hardly detected by XRD analysis even if CuS [17] G. Liu, L. Zhao, L. Ma, L. Guo, Catal. Commun. 9 (2008) 126–130.
crystals are formed. In case of CdS, Cd contents are relatively large [18] H. Lee, Y. Park, M. Kang, J. Ind. Eng. Chem. 19 (2012) 1162–1168.
[19] A. Fujishima, K. Honda, Nature 238 (1972) 37–38.
in S3 and S6 (24.2 and 14.7 wt%). Nevertheless, there were no [20] A. Hernández-Gordillo, F. Tzompantzi, R. Gómez, Int. J. Hydrogen Energy 37
additional crystal peaks found, and thus the CdS particles were (2012) 17002–17008.
believed to be amorphous structure. The time course of H2- [21] J.S. Hu, L.L. Ren, Y.G. Guo, H.P. Liang, A.M. Cao, L.J. Wan, C.L. Bai, Angew. Chem. 117
(2005) 1295–1299.
production over the four typical photocatalysts is shown in Fig. 7b
[22] S. Shen, L. Zhao, Z. Zhou, L. Guo, J. Phys. Chem. C 112 (2008) 16148–16155.
for 150 min under the standard solar irradiation. As explained in [23] V. Shrotriya, G. Li, Y. Yao, T. Moriarty, K. Emery, Y. Yang, Adv. Funct. Mater. 16
Table 1, three components’ system including 0.81 wt% Cu and (2006) 2016–2023.
[24] X. Zhao, M. Li, X. Lou, Mater. Sci. Semicond. Process. 16 (2012) 489–494.
14.7 wt% Cd have the highest production rate (837.6 mmol/g h),
[25] C. Chen, P. Liu, C. Lu, Chem. Eng. J. 144 (2008) 509–513.
and its photocatalytic activity shows a linearly increasing tendency [26] J.S. Jang, W. Li, S.H. Oh, J.S. Lee, Chem. Phys. Lett. 425 (2006) 278–282.
for the experimental time period. Therefore, the ZnS–CuS–CdS [27] J.S. Jang, H.G. Kim, U.A. Joshi, J.W. Jang, J.S. Lee, Int. J. Hydrogen Energy 33 (2008)
system has stable photocatalytic activity in H2-production from 5975–5980.
[28] J.S. Jang, S.M. Ji, S.W. Bae, H.C. Son, J.S. Lee, J. Photochem. Photobiol. A: Chem. 188
solar water splitting within our experimental condition. (2007) 112–119.
[29] D. Wu, Y. Jiang, Y. Yuan, J. Wu, K. Jiang, J. Nanopart. Res. 13 (2011) 2875–2886.
4. Conclusions [30] H. Ma, J. Han, Y. Fu, Y. Song, C. Yu, X. Dong, Appl. Catal. B: Environ. 102 (2011) 417–
423.
[31] J. Yu, J. Zhang, S. Liu, J. Phys. Chem. C 114 (2010) 13642–13649.
In summary, heterostructured ZnS–CuS–CdS photocatalyst was [32] H. Irie, K. Kamiya, T. Shibanuma, S. Miura, D.A. Tryk, T. Yokoyama, K. Hashimoto, J.
prepared as a composite form in a sequential manner (ZnS ! ZnS– Phys. Chem. C 113 (2009) 10761–10766.
[33] X. Qiu, M. Miyauchi, H. Yu, H. Irie, K. Hashimoto, J. Am. Chem. Soc. 132 (2010)
CuS ! ZnS–CuS–CdS), and they were used for water splitting H2- 15259–15267.
production under the standard solar irradiation (1 kW/m2, AM [34] M. Lee, K. Yong, Nanotechnology 23 (2012) 194014–194019.
1.5 G). The surface deposited CuS and CdS components enhance the [35] H. Yu, H. Irie, K. Hashimoto, J. Am. Chem. Soc. 132 (2010) 6898–6899.
[36] G. Bamwenda, S. Tsubota, T. Nakamura, M. Haruta, J. Photochem. Photobiol. A:
photocatalytic H2-production activity significantly as compared
Chem. 89 (1995) 177–189.
with the ZnS only case. The optimum Cu and Cd loading contents [37] J. Hou, Z. Wang, W. Kan, S. Jiao, H. Zhu, R. Kumar, J. Mater. Chem. 22 (2012) 7291–
are determined to be 0.81 wt% and 14.7 wt%, respectively and the 7299.
[38] Y.-F. Lin, Y.-J. Hsu, Appl. Catal. B: Environ. 130–131 (2012) 93–98.
corresponding H2-production rate is 838 mmol/g h. It is believed

You might also like