Hybrid mesoporous Cu2ZnSnS4 (CZTS)–TiO2 photocatalyst for efficient photocatalytic conversion of CO2 into CH4 under solar irradiation

Kidon Kima, Abdul Razzaqa, Saurav Sorcara, Yiseul Parkb, Craig A. Grimesa and Su-Il In*a
aDepartment of Energy Systems Engineering, DGIST, 50-1, Sang-ri, Hyeonpung-myeon, Dalseoung-gun, Daegu-711-873, Republic of Korea. E-mail: insuil@dgist.ac.kr; Fax: +82-53-785-6409; Tel: +82-53-785-6417
bDivision of Nano and Energy Convergence Research, DGIST, Republic of Korea

Received 30th January 2016 , Accepted 4th April 2016

First published on 5th April 2016


Abstract

A major concern facing global society is the ongoing excessive release of CO2 into the atmosphere where it acts as a heat-trapping greenhouse gas. One approach to helping control atmospheric CO2 concentrations is to use solar energy to convert CO2 into useful products, namely hydrocarbons, by use of specifically designed photocatalytic materials. While numerous photocatalysts have been investigated for use in CO2 reduction, the field remains in it's infancy with, overall, relatively poor photoconversion efficiencies and product selectivity. This study reports the synthesis and design of a mesoporous noble metal free p-type Cu2ZnSnS4 (CZTS)/n-type TiO2 heterojunction photocatalyst for broad spectrum light absorption, enhanced charge separation and transfer that, in turn, enhances photocatalytic CO2 conversion. A maximum methane production rate of 118.75 ppm g−1 h−1 is observed, which represents a methane evolution rate approximately 12 times greater than that of pure TiO2. The key factors contributing to the enhanced photocatalytic performance seen in the mesoporous CZTS–TiO2 samples include improved light absorption, high surface area, and effective charge separation.


1. Introduction

The significant rise in atmospheric CO2 concentration since the onset of the industrial revolution, due primarily to the combustion of hydrocarbon fuels, has generated considerable, and in the opinion of the authors justified, concern.1 An effective approach to mitigate this problem would be to use solar energy to convert CO2 into hydrocarbon fuels compatible with the current energy infrastructure through photocatalysis.2 The concept is an appealing one, offering a ready means for storing and transporting solar energy; further, such fuels could be obtained on a sustainable, rather than one-shot, basis. To emphasize the underlying opportunity: the use of sunlight to efficiently convert CO2 and H2O into hydrocarbon fuels offers the dual advantages of minimizing the potential for severe global climatic disruptions while obtaining low-cost renewable fuels.3

Among photocatalysts, anatase TiO2-based materials offer advantages that include relatively low cost, excellent photocorrosion stability and photo-oxidation power, non-toxicity, and rapid transfer of photon-generated charges.4–7 A prodigious amount of work has been done to try and lower the bandgap of TiO2 in an attempt to better utilize solar energy. Such efforts include, for example, loading the TiO2 with metal8,9 or non-metal elements,10–15 and so on, all of which have been largely unsuccessful: that which lowers the bandgap lowers the mobility of the generated charges. However such efforts continue, and in the opinion of the authors serve as a worthy objective. For example, Zhu and co-workers have reported S-doped TiO2 possessing highly efficient photocatalytic performance under indoor sunlight irradiation for degradation of organic pollutants; the catalyst was reported to be highly stable even after several months of indoor sunlight illumination.16 The same authors have also reported a binary composite system harnessing the properties of two dimensional (2D) ultrathin carbon nitride (C3N4) nanosheets and NiS nanoparticles, with high hydrogen generation efficiency.17 Graphene based materials have been investigated, such as ultra-thin conductive reduced graphene oxide sheets coupled with Ag18 and rGO–TiO2 nanotubes19 aimed towards, respectively, photocatalytic oxidation of pollutants and CO2 photoreduction.

Loading of noble metals within or on TiO2 based materials as a co-catalyst is an effective but expensive way of improving electron–hole separation. We believe a sound approach to achieve a high performance photocatalyst in a cost effective manner is to combine n-type TiO2 with a non-toxic, earth-abundant and hence inexpensive p-type semiconductor that serves as a strong light-absorber. To that end, Cu2ZnSnS4, or CZTS, a p-type semiconductor with a direct bandgap of 1.5 eV, has recently emerged as an excellent photovoltaic material platform,20–22 however its use as a photocatalyst for light-to-chemical energy conversion remains unexplored. Motivated by the unique qualities of CZTS and the opportunity for achieving improved photocatalytic properties by realization of a p–n junction type heterostructured material we have investigated, and herein report on a hybrid mesoporous CZTS–TiO2 photocatalyst. The material manifests significant CO2 reduction properties that can be attributed to broad-spectrum light absorption, the high surface-area mesoporous structure, and the intrinsic p–n junction nature of the mesoporous material for improved charge separation. Notably, we synthesize the mesoporous CZTS–TiO2 without the use of any noble metal co-catalysts, yet still find significant CO2 reduction at ambient temperature and pressure.

Our CZTS–TiO2 synthesis strategy shown in Fig. 1 uses a simple, low-cost two-step process: CZTS nanoparticles are prepared by hot injection; the resulting CZTS nanoparticles are dispersed in toluene under an argon atmosphere, then vigorously mixed with TiCl4 in toluene resulting in a gelatinous solution, subsequently annealed within an argon atmosphere resulting in the desired photocatalyst. A series of hybrid CZTS–TiO2 samples were synthesized with varied amounts of CZTS in TiCl4 (TiO2 precursor). The obtained materials are labelled as CT1, CT2, CT3 and CT4 for the samples obtained with 1.9 mg, 3.8 mg, 5.7 mg and 7.6 mg of CZTS nanoparticles in 0.1 mL of TiCl4 respectively. We believe our work will help lead to more cost-effective, high-performance photocatalysts for the conversion of solar energy to chemical fuels.


image file: c6ra02763f-f1.tif
Fig. 1 Schematic showing the experimental strategy for synthesis of hybrid mesoporous CZTS–TiO2 photocatalyst.

2. Experimental section

2.1. Materials

Copper(II) acetylacetonate (99.99%), zinc acetate (99.99%), tin(IV) acetate (99%), oleylamine (OLA, 70%), trioctylphosphine oxide (TOPO, 99%) and titanium(IV) chloride (99.99%), and all chemicals, were purchased from Sigma Aldrich and used without further purification.

2.2. Synthesis of Cu2ZnSnS4 (CZTS) nanocrystals

CZTS nanocrystals were synthesized by the hot-injection method.23 All reaction conditions were kept under inert atmosphere to prevent the formation of oxides. In a typical synthesis, a mixture of 0.5 mmol Cu(acac)2, 0.25 mmol Zn(OAc)2, 0.25 mmol Sn(OAc)4, and 4.0 mL OLA was prepared in a 25.0 mL 3-neck round-bottom reaction flask and heated under vacuum to 150 °C for 30 minutes. A change in the color of the metals solution was observed from blue to brown-green. The temperature was lowered to 125 °C and the mixture remained at this temperature under vacuum until injection. In a 5.0 mL scintillating vial, 1.0 mM sulfur powder was sonicated for 30 minutes with 1.0 mL OLA to achieve its complete dissolution resulting in an orange-red solution. In a 50.0 mL 3-neck round–bottom reaction flask, 10 mM TOPO was heated to 300 °C on an argon Schlenk line. When the temperature of the TOPO reaction flask reached 100 °C, the flask was pumped and purged three times in order to remove any adsorbed water vapors and impurities, then flushed with Ar and kept under argon for the remainder of the experiment. When the TOPO reaction temperature reached 300 °C, the S and metal precursors were rapidly injected simultaneously via two gas-tight Luer-Lock syringes. The reaction solution immediately changed from clear and colorless to a clear dark brown solution. All the aliquots were taken out using a syringe. 18.0 mL methanol was added to the aliquots to precipitate the nanocrystals, followed by 10 minutes of centrifugation at 3500 rpm. The supernatant, containing unreacted material was discarded, while the remaining precipitate was then redispersed in 8.0 mL toluene to which 10.0 mL methanol was again added and the solution centrifuged for 10 minutes. This process was repeated for a total of three washes. The final precipitate was re-suspended in toluene and then centrifuged for 10 minutes to remove any agglomerated or bulk constituents.

2.3. Synthesis of hybrid mesoporous Cu2ZnSnS4 (CZTS)–TiO2 photocatalyst

As synthesized CZTS nanocrystals were oven dried at 80 °C overnight to remove toluene and obtain a powder of CZTS nanocrystals. CZTS–TiO2 samples were prepared by mixing various amounts of CZTS powder in fixed volume i.e. 0.1 mL of TiCl4 under argon atmosphere for 1 h. The samples were named as CT1, CT2, CT3 and CT4 standing for 1.9 mg, 3.8 mg, 5.7 mg and 7.6 mg of CZTS in 0.1 mL of TiCl4 respectively. The various obtained samples were dried at 50 °C in order to remove toluene and then were annealed at 350 °C (rate time: 5°C min−1, holding time 1 h under argon gas flow of 30 cm3 min−1).

2.4. Photocatalyst characterizations

The crystalline structure of the materials are analyzed by powder X-ray diffractometry (XRD, Panalytical, Empyrean; operating at 40 kV and 30 mA with Cu Kα radiation, λ = 1.54 Å, as an X-ray source, scanned with a rate of 1° min−1 in the range of 2θ = 10°–70°). High resolution images are obtained using field emission transmission electron microscopy (FE-TEM, Hitachi HF-3300) operating at 300 kV. For FE-TEM analysis, the samples are prepared by dispersing in ethanol followed by mild sonication. The dilute suspension is dropped on a TEM Cu mesh grid and allowed to dry overnight. The elemental composition of sample CT4 is measured using the energy dispersive spectroscopy (EDS) attachment of the Hitachi S-4800 FE-SEM; the electron energy was 15 keV for the EDS measurement. UV-visible diffuse reflectance spectroscopy (DRS) of the materials is measured using a Cary series UV-visible near IR spectrophotometer with a diffuse reflection accessory. The N2 sorption isotherms are measured at −196 °C on a Micromeritics ASAP 2000 apparatus, with the samples extensively degassed at 200 °C prior to the experiments. Surface areas are calculated by the Brunauer–Emmett–Teller (BET) equation. The pore volume is determined from the amount of N2 adsorbed at the highest relative pressure of (P/P0) = 0.99. The BJH pore diameter is obtained by applying the Barrett–Joyner–Halenda (BJH) equation to the desorption isotherm. The surface composition and oxidation states are determined by X-ray photoelectron spectroscopy (XPS, Thermo VG, K-alpha) using the Al Kα line (148[thin space (1/6-em)]606 eV) as the X-ray source.

2.5. Photocatalytic CO2 conversion test method

Photocatalytic CO2 conversion experiments are performed in a gas phase reactor as reported earlier.24 In short, the photoreactor (stainless steel with quartz window, volume = 15.4 cm3) is loaded with a photocatalyst material (50 mg) and then purged with CO2 gas (1000 ppm in He) and vacuum simultaneously to remove any air or other impurities. CO2 gas (1000 ppm in He) is passed through a water bubbler, forming a mixture of CO2 and H2O vapors which then enters the photoreactor. The photoreactor is exposed to simulated solar light irradiated by Xenon solar simulator (Oriel, LCS-100) with an AM 1.5 filter for variable durations and the reaction products (500 μL) analyzed by a Shimadzu GC-2014 gas chromatograph (Restek Rt-Q Bond column, ID = 0.53 mm, length = 30 m) equipped with flame ionization detector (FID) and thermal conductivity detector (TCD). A schematic showing the experimental setup for photocatalytic CO2 reduction is shown in Fig. S1. The hour-normalized photocatalytic CH4 evolution rate is calculated by:
Rate of CH4 evolution = amount of CH4 evolved (ppm)/amount of photocatalyst used (g)

To examine photocatalyst stability, under otherwise identical experimental conditions sample CT4 was illuminated for 5 h and 10 h durations.

3. Results and discussion

3.1. Characterization of hybrid mesoporous CZTS–TiO2 photocatalyst

Fig. 2 shows a transmission electron microscopy (TEM) image of as-synthesized CZTS nanoparticles; one sees anisotropic shapes with an average size of 13 nm. High resolution transmission electron microscopy (HRTEM) indicates an interplanar spacing of 3.1 Å corresponding to d112 of CZTS (Fig. 2b).20 The selected area electron diffraction (SAED) pattern supports the polycrystalline nature of CZTS by displaying rings indexed to d112, d220 and d312 of CZTS, respectively (Fig. 2c).25 The TEM image of representative hybrid CZTS–TiO2, sample CT4, synthesized from 7.6 mg CZTS in 0.1 mL TiCl4, clearly demonstrate the post-annealing mesoporous nature (Fig. 2d). The obtained mesoporosity of the hybrid system can be attributed to the removal of the capping ligands during annealing, imparting both the mesoporous nature and hybridization of TiO2 with CZTS. Fig. 2e is a HRTEM image of a hybrid CZTS–TiO2 sample, showing a lattice spacing of 3.1 Å and 3.5 Å, corresponding to d112 plane of CZTS and d101 plane of TiO2 respectively.20,26 The selected area electron diffraction (SAED) pattern show rings which can be indexed to d112 plane of CZTS,20 and d101 plane of anatase TiO2 (ref. 25 and 27) further assuring the formation of hybrid CZTS–TiO2 structure (Fig. 2f). The mesoporous nature of all annealed CZTS–TiO2 samples is confirmed by TEM (Fig. S2).
image file: c6ra02763f-f2.tif
Fig. 2 TEM, HRTEM and SAED patterns of: (a–c) as synthesized CZTS nanoparticles; and (d–f) CZTS–TiO2 sample CT4.

Fig. 3 shows the elemental analysis of sample CT4 using energy dispersive X-ray scanning transmission electron microscopy (EDX-STEM); the representative sample consists of Ti, O, Cu, Zn, Sn and S, confirming the formation of a hybrid CZTS–TiO2 photocatalyst. Further, the uniform color suggests the uniform distribution of elements.


image file: c6ra02763f-f3.tif
Fig. 3 EDX-STEM elemental mapping image of CT4 sample showing composition of: (a) titanium, (b) oxygen, (c) copper, (d) zinc, (e) tin, and (f) sulphur.

The atomic composition of sample CT4, obtained using FESEM-EDS, is shown in Fig. S3; the rectangular region scan shows strong peaks associated with Cu, Zn, Sn, S, Ti and O. However, the atomic ratios of Sn and S in CZTS are less than anticipated, which we suggest can be attributed to diffusion of SnS/S2 gases due to in situ annealing while being FESEM-EDS imaged,28 or while sample-annealing under argon atmosphere for CT4 synthesis. We are not the only investigators to encounter this; a loss of Sn and S is reported to be one of the critical problems encountered during CZTS synthesis.29,30

X-ray diffraction (XRD) of as-synthesized TiO2, CZTS, and hybrid CZTS–TiO2 photocatalysts are shown in Fig. 4. CZTS nanoparticles indicate peaks at 2θ values of 28.53°, 47.33°, and 56.17° corresponding to d112, d220 and d312 of Kesterite CZTS.31 The XRD pattern of the TiO2 sample prepared from TiCl4 clearly shows peaks at 2θ values of 25.28°, 37.80°, 48.05°, 53.89° and 55.06° corresponding to d101, d004, d200, d105 and d211 of anatase TiO2.26 All CZTS–TiO2 samples show peaks corresponding to anatase TiO2, however the CZTS peaks are diffuse and appear as a relatively broad hump at 2θ = 28.53°, a behavior that might be attributed to the small size of CZTS nanoparticles. Upon increasing the CZTS amount within the hybrid CZTS–TiO2 samples, CT4 synthesized from 7.6 mg CZTS in 0.1 mL TiCl4, the XRD peak for Kesterite CZTS at 2θ = 28.53° becomes more prominent, confirming the CZTS presence.


image file: c6ra02763f-f4.tif
Fig. 4 XRD pattern for as-synthesized CZTS, TiO2 and hybrid CZTS–TiO2 samples (CT1, CT2, CT3 and CT4 stands for 1.9, 3.8, 5.7 and 7.6 mg of CZTS in 0.1 mL of TiCl4 respectively).

The UV-vis diffuse reflectance spectra (DRS) of bare TiO2 annealed under air and argon atmosphere is shown in Fig. S4 and it exhibits a steep absorption edge in the range of 380–420 nm, corresponding to the fundamental bandgap of 3.2 eV.27 That the as-prepared TiO2 nanoparticles absorb in the visible region can be attributed to the Urbach band tail, arising from the momentary localization of excitons due to phonon interaction.34,35

Fig. 5a shows the absorption spectrum of pure CZTS nanoparticles, showing broad absorption in the visible light range. The bandgap of the CZTS nanoparticles is estimated using the Tauc relationship, resulting in a band gap of 1.5 eV, a value consistent with reported literature values (1.45–1.6 eV).23 In case of hybrid CZTS–TiO2, a prominent enhancement in light absorption is observed in the visible light range (400–700 nm) with the increase in CZTS (Fig. 5b); the degree of CZTS content strongly affects the visible light absorption of TiO2 leading to a red shift in the absorption spectrum.


image file: c6ra02763f-f5.tif
Fig. 5 UV-vis DRS spectra for: (a) as-synthesized CZTS nanoparticles, with the inset showing the determined band gap value of 1.5 eV; (b) CZTS–TiO2 samples CT1, CT2, CT3 and CT4, which stand for, respectively, 1.9, 3.8, 5.7 and 7.6 mg of CZTS in 0.1 mL of TiCl4.

Further evidence indicating a hybrid CZTS–TiO2 material comes from X-ray photoelectron spectroscopy (XPS) data of sample CT4 (Fig. 6). The Ti 2p region shows two main peaks located at 463.9 eV and 458.3 eV corresponding to the Ti 2p1/2 and Ti 2p3/2 states respectively (Fig. 6a), while the O 1s region exhibits peak around 530.9 eV attributed to Ti(IV) ion bound oxygen, in TiO2 (Fig. 6b).32 Cu 2p region with the main characteristic peaks located at 931.2 eV and 951 eV, attributed to Cu 2p1/2 and Cu 2p3/2 electronic states respectively with the peak separation of 19.8 eV, assuring the presence of Cu(I) (Fig. 6c).33 The XPS spectrum of Zn 2p splitting into two peaks appearing at 1020.9 eV and 1043.9 eV, can be assigned to Zn 2p1/2 and Zn 2p3/2 respectively (Fig. 6d). These two peaks display a peak separation of 23 eV, indicating presence of Zn(II).34 The XPS spectrum of Sn 3d with the peaks at 486.1 eV and 494.5 eV are attributed to Sn 3d3/2 and Sn 2d5/2 of Sn(IV), Fig. 6e.36 The S 2p region exhibits two main peaks appearing at 162.38 eV and 163.53 eV (Fig. 6f); these peaks can be assigned to S 2p1/2 and S 2p3/2 with a peak separation of 1.15 eV confirming the presence of S.37 XPS spectra for the as-prepared TiO2 nanoparticles and pure CZTS nanoparticles were investigated to confirm the presence of all the respective elements in their expected oxidation states (Fig S5 and S6); it is observed that all the corresponding peaks appearing for sample CT4, our representative sample, are in good agreement with the peaks of as-synthesized CZTS and TiO2, thus assuring no significant change in the electronic states of the respective materials.


image file: c6ra02763f-f6.tif
Fig. 6 X-ray photoelectron spectroscopy (XPS) of CT4 sample with regions of: (a) titanium, (b) oxygen, (c) copper, (d) zinc, (e) tin and (f) sulphur.

The Brunauer–Emmett–Teller (BET) surface areas are also measured for all hybrid CZTS–TiO2 samples (Fig. S7 & Table S1). The N2-sorption isotherms of all samples, shown in Fig. S5, are found to be of type IV according to IUPAC classification.38 The samples initially exhibit a monolayer adsorption, which is followed by a steep rise indicating the multilayer adsorption phenomenon. The multilayer adsorption is then followed by condensation within the mesoporous material at high relative pressures. Due to difference in pressure of capillary evaporation and condensation, the isotherm exhibits H3 hysteresis loops revealing the presence of plate like particles with slit-like pores.38 Among all samples, sample CT4 shows the highest surface area of 92.71 m2 g−1. Further, the surface area of the hybrid CZTS–TiO2 samples increase with the addition of CZTS nanoparticles, indicating the CZTS nanoparticles are largely responsible for inducing the mesoporous nature of the hybrid samples.

3.2. Photocatalytic CO2 reduction

Photocatalytic activities of all hybrid CZTS–TiO2 samples, TiO2, and CZTS samples are investigated by examining their ability to drive the photocatalytic conversion of CO2 and water vapor into hydrocarbon fuels, under simulated AM1.5G solar light, with TiO2 (synthesized from TiCl4 precursor) and pure CZTS nanoparticles used as baseline reference samples (Fig. 7). Gas chromatography (GC) analysis of the products showed predominantly methane as a main hydrocarbon product. The rate of methane production from each sample is obtained after 1 h of solar illumination.
image file: c6ra02763f-f7.tif
Fig. 7 Methane production rate for 1 h illumination from pure CZTS, TiO2 and hybrid CZTS–TiO2 samples under AM1.5G illumination.

It is obvious that all hybrid mesoporous CZTS–TiO2 samples show an increase in CH4 yield. On increasing the amount of CZTS in the hybrid samples CZTS–TiO2 the CH4 evolution increases, with highest value of methane evolution found for sample CT4 (118.75 ppm g−1 h−1), respectively 11.97 times and 9.7 times higher than TiO2 (9.92 ppm g−1 h−1) and CZTS (12.29 ppm g−1 h−1).

Illumination of sample CT4 in an Ar/H2O (g) atmosphere under similar conditions served as a control test, in which one can see a negligible amount of CH4, suggesting any organic impurities adsorbed on the surface of hybrid CZTS–TiO2 composites having negligible involvement in the photocatalytic CH4 evolution. Turnover numbers (TON) and turnover frequencies (TOF) for the hybrid CZTS–TiO2 photocatalysts are obtained and presented in Table S2. Amongst hybrid photocatalyts, for 1 h illumination sample CT4 yielded a TON of 1.48 and TOF of 1.48 h−1. Under extended illumination it is observed that the TON slightly increases after 5 h, reaching a value of 1.67, then decreases to 0.90 for 10 h of continuous illumination.

It can be inferred from the control experiments that the CH4 evolved during our normal experiments is mainly due to photoreduction of CO2 and not a product of organic oxidation or surface bound organic species. The improved photocatalytic performance of our hybrid CZTS–TiO2 samples can be inferred to: (1) improve light absorption; (2) enhance the surface area; and (3) improved charge separation via electric-field induced drift per formation of a p–n junction at the interface of p-type CZTS and n-type TiO2.

Based upon these factors, an illustration of energy levels and suggested process for conversion of CO2 and water vapor into methane is shown in Fig. 8. As revealed by XPS, hybrid samples contain both CZTS and TiO2; the conduction band edge of CZTS is higher than the TiO2 conduction band edge, while the TiO2 valence band edge is more positive than the valence band edge of CZTS and the O2/H2O redox potential.39 Upon illumination, the photo-generated electrons in CZTS can easily flow to the conduction band of TiO2, and can reduce the adsorbed CO2 whereas the holes migrate in the opposite direction to oxidize adsorbed H2O. Photoreduction of CO2 is a multi-electron process, requiring reaction of 8 holes (h+) with adsorbed H2O, producing O2 and H+ ions and reducing the CO2 adsorbed on photocatalyst surface via reaction of 8 electrons (e) and 8 protons (H+ ions), evolving CH4 as the dominant product.40


image file: c6ra02763f-f8.tif
Fig. 8 Schematic illustration for photocatalytic CO2 conversion into methane by hybrid CZTS–TiO2 samples with proposed energy level diagram.

Sample stability is investigated by testing sample CT4 under up to 10 h of continuous illumination for CO2 photoreduction (Fig. S8). It is observed that the methane production rate after 5 h illumination slightly increases from the 1 h value, whereas a 40% decrease in the methane production rate is observed after 10 h of continuous illumination. We are unsure of the origin of such behavior, but such behavior is not unusual in the realm of photocatalysts.

4. Conclusions

We have successfully synthesized a hybrid mesoporous CZTS–TiO2 photocatalyst using a simple two-step process. The hybrid CZTS–TiO2 samples exhibit improved light absorption and mesoporosity mainly induced by CZTS content. The as-prepared samples are characterized by various analytical techniques such as XRD, TEM, UV-vis DRS spectra, BET and XPS. A considerable enhancement in CO2 conversion into methane under AM1.5G illumination is observed for the mesoporous CZTS–TiO2 samples, which we attribute to improved light absorption, large surface areas which in turn promote higher interfacial reactions, and intrinsic formation of a p–n junction between the CZTS and TiO2 regions with formation with suitable bandgap edge positions for improved charge separation. Our material synthesis strategy, and coupling of favorable semiconductors suggests an effective approach for design of high performance noble metal free photocatalyst materials.

Acknowledgements

The authors gratefully acknowledges the support by the DGIST R&D Program of the Ministry of Education, Science and Technology of Korea (16-BD-0404) and by Basic Science research program through the National Research Foundation of Korea funded by the Ministry of Science, ICT & Future Planning (2013R1A1A008678 & 2014K1A3A1A47067086). This research was also supported by the Technology Development Program to Solve Climate Changes of the National Research Foundation (NRF) funded by the Ministry of Science, ICT & Future Planning(2015M1A2A2074670). The authors are thankful to DGIST R&D Program and MIREBraiN Program of the Ministry of Science, ICT & Future Planning (15-01-HRSS01 & 201510023).

References

  1. P. Tans, Trends in atmospheric carbon dioxide, http://www.esrl.noaa.gov/gmd/ccgg/trends/, accessed, 07, 2015 Search PubMed.
  2. S. N. Habisreutinger, L. S. Mende and J. Stolarczyk, Angew. Chem., Int. Ed., 2013, 52, 7372–7408 CrossRef CAS PubMed.
  3. Q. Liu, Y. Zhou, J. Kou, X. Chen, Z. Tian, J. Gao, S. Yan and Z. Zou, J. Am. Chem. Soc., 2010, 132, 14385–14387 CrossRef CAS PubMed.
  4. M. R. Hoffmann, S. T. Martin, W. Choi and D. W. Bahnemann, Chem. Rev., 1995, 95, 69–96 CrossRef CAS.
  5. S. U. Khan, M. Al-Shahry and W. B. Ingler Jr, Science, 2002, 297, 2243–2244 CrossRef CAS PubMed.
  6. X. Wu, Z. Chen, G. Q. M. Lu and L. Wang, Adv. Funct. Mater., 2011, 21, 4167–4172 CrossRef CAS.
  7. P. Yaron, Appl. Catal., B, 2010, 99, 448–460 CrossRef.
  8. S. K. Dutta, S. K. Mehetor and N. Pradhan, J. Phys. Chem. Lett., 2015, 6, 936–944 CrossRef CAS PubMed.
  9. N. Roy, Y. Sohn, K. M. Leung and D. Pradhan, J. Phys. Chem. C, 2014, 118, 29499–29506 CAS.
  10. X. Liu, J. Chen, J. Li and C. Burda, Chemosphere, 2005, 61, 11–18 CrossRef PubMed.
  11. J. C. Yu, L. Zhang, Z. Zheng and J. Zhao, Chem. Mater., 2003, 15, 2280–2286 CrossRef CAS.
  12. T. Hirai, K. Suzuki and I. Komasawa, J. Colloid Interface Sci., 2001, 244, 262–265 CrossRef CAS.
  13. D. Chatterjee and A. Mahata, Appl. Catal., B, 2001, 33, 119–125 CrossRef CAS.
  14. J. Wang, T. Ma, G. Zhang, Z. Zhang, X. Zhang, Y. Jiang, G. Zhao and P. Zhang, Catal. Commun., 2007, 8, 607–611 CrossRef CAS.
  15. S. K. Parayil, A. Razzaq, S. M. Park, C. A. Grimes and S. l. In, Appl. Catal., A, 2015, 498, 205–213 CrossRef CAS.
  16. M. Zhu, C. Zhai, L. Qiu, A. S. Paton, Y. Du and M. C. Goh, ACS Sustainable Chem. Eng., 2015, 3, 3123–3129 CrossRef CAS.
  17. Y. Lu, D. Chu, M. Zhu, Y. Du and P. Yang, Phys. Chem. Chem. Phys., 2015, 17, 17355–17361 RSC.
  18. S. Song, B. Cheng, N. Wu, A. Meng, S. Cao and J. Yu, Appl. Catal., B, 2016, 181, 71–78 CrossRef CAS.
  19. A. Razzaq, C. A. Grimes and S. I. In, Carbon, 2016, 98, 537–544 CrossRef CAS.
  20. X. Xin, M. He, W. Han, J. Jung and Z. Lin, Angew. Chem., Int. Ed., 2011, 50, 11739–11742 CrossRef CAS PubMed.
  21. L. Shi, C. Pei, Y. Xu and Q. Li, J. Am. Chem. Soc., 2011, 133, 10328–10331 CrossRef CAS PubMed.
  22. H. Yang, L. A. Jauregui, G. Zhang, Y. P. Chen and Y. Wu, Nano Lett., 2012, 12, 540–545 CrossRef CAS PubMed.
  23. S. C. Riha, B. A. Parkinson and A. L. Prieto, J. Am. Chem. Soc., 2009, 131, 12054–12055 CrossRef CAS PubMed.
  24. H. Kim, A. Razzaq, H. J. Heo and S. I. In, Rapid Communication in Photoscience, 2013, 2, 64–66 CrossRef.
  25. N. Muhunthan, O. P. Singh, S. Singh and V. N. Singh, Int. J. Photoenergy, 2013, 752012, 1–7 CrossRef.
  26. P. A. Sedach, T. J. Gordon, S. Y. Sayed, T. Fürstenhaupt, R. Sui, T. Baumgartner and P. C. P. Berlinguette, J. Mater. Chem., 2010, 20, 5063–5069 RSC.
  27. D. R. Coronado, G. R. Gattorno, M. E. E. Pesqueira, C. Cab, R. Coss and G. Oskam, Nanotechnology, 2008, 19, 1–10 Search PubMed.
  28. P. Kush, S. K. Ujjain, N. C. Mehra, P. Jha, R. K. Sharma and S. Deka, ChemPhysChem, 2013, 14, 2793–2799 CrossRef CAS PubMed.
  29. J. J. Scragg, T. Ericson, T. Kubart, M. Edoff and C. P. Björkman, Chem. Mater., 2011, 23, 4625–4633 CrossRef CAS.
  30. A. Weber, R. Mainz and H. W. Schock, J. Appl. Phys., 2010, 107, 013516-1–013516-6 CrossRef.
  31. S. P. Lim, A. Pandikumar, N. M. Huang, H. N. Lim, G. Gud and T. L. Ma, RSC Adv., 2014, 4, 48236–48244 RSC.
  32. T. Kameyama, T. Osaki, K. Okazaki, T. Shibayama, A. Kudo, S. Kuwabata and T. Torimoto, J. Mater. Chem., 2010, 20, 5319–5324 RSC.
  33. X. Yin, C. Tang, M. Chen, S. Adams, H. Wang and H. Gong, J. Mater. Chem. A, 2013, 1, 7927–7932 CAS.
  34. B. Lu, X. Wang, G. Cai, L. Wen, Y. Song and X. Zhao, J. Hazard. Mater., 2009, 169, 1112–1118 CrossRef PubMed.
  35. H. Tang, F. Lévy, H. Berger and P. E. Schmid, Phys. Rev. B: Condens. Matter Mater. Phys., 1995, 52, 7771–7774 CrossRef CAS.
  36. S. Sarkar, K. Bhattacharjee, G. C. Das and K. K. Chattopadhyay, CrystEngComm, 2014, 16, 2634–2644 RSC.
  37. P. Dai, X. Shen, Z. Lin, Z. Z. Feng, H. Xu and J. Zhan, Chem. Commun., 2010, 46, 5749–5751 RSC.
  38. M. Kruk and M. Jaroniec, Chem. Mater., 2001, 13, 3169–3183 CrossRef CAS.
  39. S. P. Dilsaver, M. D. Reichert, B. L. Hallmark, M. J. Thompson and J. Vela, J. Phys. Chem. C, 2014, 118, 21226–21234 Search PubMed.
  40. S. K. Parayil, A. Razzaq, S. M. Park, H. R. Kim, C. A. Grimes and S. I. In, Appl. Catal., A, 2015, 498, 205–213 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available: Schematic of experimental setup for photocatalytic CO2 conversion test; TEM images of all hybrid CZTS–TiO2 samples; FE-SEM EDS of CT4 sample; UV-vis DRS of as synthesized TiO2 (air and Ar annealed); high resolution XPS of as-synthesized TiO2 nanoparticles, and as-synthesized CZTS nanoparticles; BET surface areas for all hybrid CZTS–TiO2 samples; stability tests of CT4 sample; turnover numbers and turnover frequencies for all hybrid CZTS–TiO2 samples. See DOI: 10.1039/c6ra02763f

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.