A new 0D–2D CsPbBr3–Co3O4 heterostructure photocatalyst with efficient charge separation for photocatalytic CO2 reduction

Xin Zhong a, Xinmeng Liang a, Xinyu Lin a, Jin Wang *ab, Malik Zeeshan Shahid *a and Zhengquan Li *ab
aKey Laboratory of the Ministry of Education for Advanced Catalysis Materials, Zhejiang Normal University, Jinhua, Zhejiang 321004, P. R. China. E-mail: zqli@zjnu.edu.cn; zeeshan-nano@zjnu.edu.cn; wangjin@zjnu.edu.cn
bZhejiang Institute of Optoelectronics, Zhejiang Normal University, Jinhua, Zhejiang 321004, P. R. China

Received 21st March 2023 , Accepted 17th April 2023

First published on 18th April 2023


Abstract

The effective spatial separation of photogenerated charge carriers is essential for realizing efficient CO2 conversion. Herein, a new CsPbBr3–Co3O4 heterostructure photocatalyst was rationally developed for photocatalytic CO2 reduction. A facile synthetic strategy based on electrostatic interactions was utilized. The results revealed that the CsPbBr3–Co3O4 hybrid exhibited a boosted evolution rate of 64.6 μmol g−1 h−1 (CO: 35.40 μmol g−1 h−1; CH4: 29.2 μmol g−1 h−1) with an electron consumption rate (Relectron) of 304.4 μmol g−1 h−1, surpassing pristine CsPbBr3 or Co3O4. The high activity mainly arises from efficient charge separation and the directional transfer of electrons from CsPbBr3 to Co3O4via an intimately coupled heterointerface. Notably, the surface features (derived from the unique morphology) expedited the CO2 adsorption and accumulation of electrons at the Co3O4 site which ultimately facilitated the conversion of CO2 over the CsPbBr3–Co3O4 composite. This approach provides a strategy to design and modulate highly active metal oxide and perovskite-based photocatalysts and presents great potential for constructing a heterointerface for CO2 reduction.


1. Introduction

The daily consumption of fossil fuels results in the emission of CO2, causing universal environmental and energy issues.1–4 Fixation of CO2 into value-added products such as CO, CH4, HCOOH, CH3OH, etc. via solar-driven catalysis, also known as artificial photosynthesis, is a clean and sustainable solution.5–7 However, it is challenging and suffers low conversion efficiency due to the high thermodynamic stability of CO2 molecules and the need for multi-electron transfer.8,9 Recently, instead of using single-component photocatalysts, research efforts have been diverted to designing their low-cost heterostructures. Many materials, such as C3N4,10,11 ZrO2,12 TiO2,13 Ta2O5,14 Nb2O5,15 metal sulfides,16 metal–organic frameworks,17 metal complexes,7 single-atom catalysts,18 MXenes,19 conducting polymers,20 metal halide perovskites,6,21etc., have been reported aiming at activity enhancement via optimizing the light-harvesting and charge carrier kinetics. However, the search for a more effective candidate photocatalyst has not stopped.

Among numerous materials, all inorganic metal halide perovskites, particularly cesium lead-bromide perovskite quantum dots (CsPbBr3 QDs), are extremely competitive photocatalysts for CO2 reduction.1,6,22,23 This is owing to their suitable energy band structure, small size, defect tolerance, and large carrier mobility as compared to other inorganic metal halide family members, e.g. CsPbCl3 and CsPbI3.1,6,24–43 However, pristine CsPbBr3 QDs suffer from instability and rapid recombination of electron–hole (e–h+) pairs, leading to low activity. To address this issue, Xu et al. first utilized the CsPbBr3 QDs/graphene composite for the photocatalytic reduction of CO2.43 After that, several efforts were devoted to enhance its performance. For instance, our group designed CsPbBr3 QDs/Bi2WO638 and CsPbBr3 QDs coupled with covalent triazine frameworks,36 and both exhibited enhanced charge separation and led to improved CO2 photoreduction. But the produced gas was mainly CO and only a minimum of CH4 was detected. This was possibly due to the lack of sufficient accumulation of reductive electrons at catalytic sites. So, we further our research to find some suitable materials which can form an intimately coupled interface with CsPbBr3 QDs and enable efficient photocatalysis to generate both CO and CH4, thereby overcoming the intrinsic issues of CsPbBr3 QDs.

We found that the transition metal oxide photocatalyst cobalt oxide (Co3O4) is an ideal non-precious catalyst for CO2 reduction as it exhibits suitable band alignment, efficient charge-carrier flux capability, and chemical/thermodynamic stability.44–52 However, the typical single component Co3O4 may suffer limited preservation of reductive electrons as well as hindered spatial separation of e–h+ pairs. Multiple morphologies have been constructed to meet such limitations, such as nanorods,53 ultrathin nanosheets,54 nanofibers,55 hollow dodecahedra,48 porous structures,56 and mesoporous two-dimensional (2D) hexagonal nanoplatelets (HPs) with various facets and catalytically active sites.45,57 For example, Gao et al. first developed [112] facet-rich Co3O4 HPs,52 and then later on Zhu et al. constructed the Co3O4/g-C3N4 (2D/2D) hybrid,57 both intended to facilitate the separation of charge carriers for photocatalytic CO2 conversion. Inspiringly, we intend to develop a high-performance Co3O4 heterostructure catalyst by simultaneously promoting charge separation and preserving the reductive electrons. It is widely accepted that constructing a heterostructure is efficient for channelizing and accelerating the separation and transfer of e–h+ pairs via a strongly coupled interface developed through proper band alignment and work functions (Φ).47,58–61 Therefore, constructing a hybrid between CsPbBr3 QDs and CO3O4 HPs probably can result in an improved photocatalytic performance toward CO2 reduction. To the best of our knowledge, there is no report on the fabrication of CsPbBr3 QDs on CO3O4 HPs, and it is desirable to design and construct a new heterostructure based on CsPbBr3 QDs and CO3O4 HPs for CO2 photoreduction.

Herein, a CsPbBr3–Co3O4 heterojunction photocatalyst is developed via electrostatic self-assembly between CO3O4 HPs and CsPbBr3 QDs for photocatalytic CO2 reduction. Investigations confirmed the successful formation of the CsPbBr3–Co3O4 hybrid. Diverse physicochemical and optoelectronic characterization studies revealed that: (i) CO3O4 HPs acted as a supporting matrix to collect electrons from CsPbBr3, rendering electron localization; (ii) the CsPbBr3–Co3O4 hybrid exhibited compact heterointerfaces, facilitating robust charge separation with hampered e–h+ pair recombination; (iii) the Co3O4 side offers numerous mesopores and catalytic sites, which facilitate the better capturing and activation of CO2 molecules. Consequently, CsPbBr3–Co3O4 showed improved activity for the generation of CO and CH4, with an evolving rate of 35.40 and 29.2 μmol g−1 h−1, respectively, surpassing their pristine counterparts i.e., CsPbBr3 QDs and CO3O4 HPs, as well as recent state-of-the-art photocatalysts (Table S1). This work highlights the rational design of new metal halide perovskite-based photocatalysts and addresses the critical issues regarding charge carrier kinetics to realize efficient solar-driven CO2 conversion.

2. Experimental section

2.1. Synthesis of CsPbBr3 QDs

To obtain CsPbBr3 QDs, the reported method was followed (see also in Fig. S1).62 Briefly, for the Cs-OA stock solution, 0.2 g of cesium carbonate (Cs2CO3), 10 mL of octadecene (ODE, C18H36), and 0.6 mL of oleic acid (OA, C18H34O2) were loaded into a three-necked round bottom flask. Under Ar flow, the temperature was increased to 120 °C and a clear light-yellow solution was obtained. The temperature was increased to 150 °C just before the hot injection. For the Pb stock solution, 0.1380 g of lead bromide (PbBr2) and 10 mL of ODE were added into a 50 mL three-necked round bottom flask, under Ar flow, the temperature was increased to 120 °C to get a white turbid liquid and maintained it for 30 min. Next, the temperature was further increased to 165 °C, and 1.5 mL OA and 1.1 mL OAm were added to the Pb stocks. When a yellow homogeneous solution was obtained, 0.8 mL of the Cs-stock solution was swiftly injected into it. After 5 s, the mixture was immediately cooled down using an ice-water bath. The original solution was directly centrifuged at 8000 rpm for 5 min and further washed with ethyl acetate and isopropanol to remove the organic residue. Finally, the CsPbBr3 QDs collected and stored in n-hexane are added in an equal volume.

2.2. Preparation of Co3O4 HPs

To obtain Co3O4 HPs, the previously reported method was modified (see also in Fig. S2).52 Firstly, β-Co(OH)2 precipitates were prepared. 0.2379 g (5 mmol) of cobalt chloride hexahydrate (CoCl2·6H2O), 1.6823 g (60 mmol) of cyclohexamethylenetetramine (urotropine), 180 mL of deionized water and 20 mL of ethanol were added into a beaker under stirring. The mixture was heated in an oil bath to 90 °C for 1 h to obtain a pink solution. After cooling it to an ambient temperature, it was centrifuged at 9000 rpm for 15 min to obtain a pink precipitate of β-Co(OH)2. Subsequently, 50 mL of deionized water and 50 mL of ethanol were added, and the β-Co(OH)2 precipitate was redispersed by ultrasonication. The sample was collected by centrifuging at 9000 rpm for 15 min. This operation was repeated 3 times. The obtained precipitate was freeze-dried under vacuum for 6 hours. Finally, the precipitate was transferred into a cuboid crucible and put into a muffle furnace for calcination at 400 °C for 3 h to obtain a black powder of Co3O4 HPs.

2.3. Synthesis of the CsPbBr3–Co3O4 heterostructure

The CsPbBr3–Co3O4 heterostructure was prepared using a solution-processed approach at room temperature. First, 0.5 mg of the Co3O4 HP powder was dispersed into 250 μL of ethanol under uninterrupted stirring. Second, 400 μL (2 mg) of CsPbBr3 QDs were dispersed into 5 mL of ethyl acetate. Third, the CsPbBr3 QD solution was swiftly added into the solution of Co3O4 HPs and mechanically stirred at room temperature in the dark for 30 min. Finally, the mixture was ultrasonicated for another 30 min under ambient conditions. The obtained solution was centrifuged at 8000 rpm for 5 min, and the precipitate was vacuum dried overnight at 45 °C to obtain the CsPbBr3–Co3O4 heterojunction. This synthesis was performed with an optimum mass ratio (Co3O4 HPs[thin space (1/6-em)]:[thin space (1/6-em)]CsPbBr3 QDs) i.e., 1[thin space (1/6-em)]:[thin space (1/6-em)]4, and a similar method was followed to obtain various mass ratios such as 1[thin space (1/6-em)]:[thin space (1/6-em)]1, 1[thin space (1/6-em)]:[thin space (1/6-em)]2, 1[thin space (1/6-em)]:[thin space (1/6-em)]8, and 1[thin space (1/6-em)]:[thin space (1/6-em)]16. Step by step addition of Co3O4 HPs and CsPbBr3 QDs as well as the formation process of the CsPbBr3–Co3O4 heterojunction are also picturized and described in Fig. S3. The details of all materials and other characterization studies for physicochemical and optoelectronic properties are provided in the ESI.

3. Results and discussion

3.1. Illustration of the synthesis process and verification of the heterostructure formation

Firstly, CsPbBr3 QDs (0D) was precisely prepared via a typical hot-injection method (Fig. S1).62 Subsequently, Co3O4 HPs (2D) holding rich facets [112] were synthesized through the calcination of brucite-like cobalt hydroxide β-Co(OH)2 precipitates (Fig. S2).52 Finally, to obtain a 0D/2D CsPbBr3–Co3O4 heterojunction, their mixture at an optimum mass ratio 4[thin space (1/6-em)]:[thin space (1/6-em)]1 (CsPbBr3 QDs[thin space (1/6-em)]:[thin space (1/6-em)]Co3O4 HPs) was exclusively stirred in a solution of ethyl acetate and ethanol at room temperature (Scheme 1, see details in Fig. S3). The tactic used here is based on the Coulomb electrostatic assembly which enables the incorporation of CsPbBr3 QDs and Co3O4 HPs. The zeta potentials of pristine CsPbBr3 QDs and Co3O4 HPs are 4.80 mV and −3.10 mV, respectively (Fig. S4), which indicates oppositely charged surfaces encounter electrostatic attractions, resulting in the formation of CsPbBr3–Co3O4. Such a kind of electrostatic self-assembly would possess a strong heterointerface, beneficial for interfacial charge transfer.63
image file: d3qi00527e-s1.tif
Scheme 1 Illustration for the formation of the CsPbBr3–Co3O4 heterojunction via electrostatic attraction.

To confirm the successful formation of the CsPbBr3–Co3O4 hybrid, powder X-ray diffraction (XRD) was first conducted (Fig. 1a). Pristine CsPbBr3 QDs exhibit a typical cubic-phase (JCPDS card, No. 75-0412). Meanwhile, pristine Co3O4 HPs show diffraction patterns associated with the face-centred cubic phase of the spinel Co3O4 (JCPDS card No. 74-1657). Notably, strong signals of all peaks emerged and co-existed in the CsPbBr3–Co3O4 hybrid, illustrating the successful formation of the heterojunction and the crystal phases were well maintained. To confirm this claim, nitrogen adsorption–desorption analysis was performed (Table S2). It can be seen that after the precise formation of the heterojunction, the pore size and pore volume were reduced from 31.3 nm to 21.21 nm and 0.1325 cm3 g−1 to 0.0266 cm3 g−1 respectively. In addition, the BET surface area also decreased from 16.9 m2 g−1 to 5.016 m2 g−1. This result indicates that the mesopores of Co3O4 were occupied by CsPbBr3 QDs. Besides in the XRD, an exclusive color transformation was observed from black (Co3O4-HPs, Fig. 1b) and bright yellow (CsPbBr3 QDs, Fig. 1c) to dark grey (CsPbBr3–Co3O4, Fig. 1d). In addition, under ultraviolet irradiation, the CsPbBr3–Co3O4 heterojunction demonstrated a dark green fluorescence in comparison with the bright green fluorescence of CsPbBr3 QDs (Fig. S5). Such color transformations further endorse that the heterojunction was constructed successfully in a well-controlled manner via a current facile strategy.


image file: d3qi00527e-f1.tif
Fig. 1 (a) Comparison of the XRD pattern for the as-prepared products within the 2 theta range 10°–80°, where pink circles and blue diamonds denote the corresponding peaks in the Co3O4–CsPbBr3 heterojunction originating from pristine products. Exclusive color transformation of the as-synthesized products where (b) Co3O4 HPs black, (c) CsPbBr3 QDs bright yellow, and (d) CsPbBr3–Co3O4 dark grey. (e) The TEM image of pristine CsPbBr3 QDs and the corresponding inset with high magnification. (f) The SEM image with low magnification and the inset is the TEM image for a single particle of Co3O4 HPs showing a hexagonal platelet-like morphology. (g) The TEM image of a Co3O4–CsPbBr3 heterojunction and the corresponding inset which was captured from the central side of a random particle and (h) the image taken from the boundary side. Finely dispersed CsPbBr3 QDs are denoted with yellow circles and the pores are denoted with green marked areas. (i) and (j) HRTEM images, and (k–p) HAADF and EDX mapping results for the Co3O4–CsPbBr3 heterojunction.

Transmission electron microscopy (TEM) and scanning electron microscopy (SEM) analysis further confirmed the formation of the heterojunction. CsPbBr3 QDs exhibited an average size of about 12 nm (Fig. 1e and inset), whereas Co3O4 exhibits a hexagonal platelet-like structure with lateral sizes of ∼4 ± 1 μm, and a thickness of about ∼50 ± 10 nm, a dominant facet (112), and mesopores (Fig. 1f and inset). Importantly, after incorporating CsPbBr3 QDs into Co3O4 HPs, several facts were noticed. (i) Both CsPbBr3 QDs and Co3O4 HPs co-existed with well-defined morphologies, and their sizes were preserved without any further ripening, verifying that the synthesis was well-controlled (Fig. 1g). (ii) Besides electrostatic interactions, the huge surface energy of CsPbBr3 QDs could drive its face-to-face attaching to Co3O4 HPs. (iii) CsPbBr3 QDs were well dispersed (denoted by yellow circles) all over Co3O4 HPs, both on the inner side (Fig. 1g and inset) and on the boundary sides (Fig. 1h) and this would facilitate the formation of a rich heterointerface. (iv) Even though most of the mesopores were occupied by CsPbBr3 QDs, still many mesopores could be observed as denoted by the marked area in green (Fig. 1g, h, and the inset).

Such accessible mesopores are not only beneficial for supporting CO2 adsorption but also facilitate the exposure of the innermost lattice surfaces for the rapid transfer of photoinduced e to the outermost active surface sites.52,57 More clear evidence regarding the formation of the heterointerface was collected via high-resolution TEM analysis (HRTEM, Fig. 1i and j). The spacing in the lattice fringes was found to match well with both constituents, that is, 0.58 nm and 0.285 nm being the corresponding planes of CsPbBr3 and Co3O4 respectively (Fig. 1j). Accordingly, the HAADF and energy dispersive X-ray (EDX) mapping analysis depicts the precise incorporation and distribution of all elements including Cs, Pb, Br, Co, and O in the CsPbBr3–Co3O4 heterojunction (Fig. 1k–p). The above results are in-line with the XRD and TEM analysis, validating the formation of the heterojunction with finely dispersed CsPbBr3 QDs on the support matrix of Co3O4 HPs.

Next, surface chemical states and interfacial interaction in the CsPbBr3–Co3O4 heterojunction were evaluated via high-resolution X-ray photoelectron spectroscopy (XPS) of Cs 3d, Pb 4f, Br 3d, Co 2p, and O 1s. It is noteworthy that, in comparison with CsPbBr3 QDs, the binding energy of Cs 3d (Fig. 2a), Pb 4f (Fig. 2b), and Br 3d (Fig. 2c) faced a positive shift in the heterojunction. This result strongly endorsed the close interfacial contact between QDs and HPs. Meanwhile, the XPS of Co 2p spectra signals were comparatively examined for CsPbBr3–Co3O4 and Co3O4 HPs (Fig. 2d), where two binding energy values at 796 ± 0.2 eV and 780 ± 0.2 eV fit to Co 2p1/2 and Co 2p3/2, respectively, which normally attributed to two main regions i.e., Co2+ and Co3+.44 As is known the theoretical value for the atomic ratio Co2+/Co3+ is 0.5 (i.e., perfect Co3O4), but this value may increase in the presence of surface defects or oxygen vacancies (OVs).44 Herein, no signal corresponding to defects in O 1s spectra (Fig. 2e) was found, which means that Co3O4 synthesized here is perfect and in an equally balanced state. Notably, the absence of OVs also indicates that the high crystallinity of Co3O4 was maintained after forming the CsPbBr3–Co3O4 heterojunction. Meanwhile, a slight negative shift towards the lower binding energy was observed, demonstrating the interaction between Co3O4 and CsPbBr3 (Fig. 2d). Moreover, relative O 1s spectra disclosed two main peaks that fit with the hydroxyl species and Co–O bond respectively. In addition, the obvious co-existence of CsPbBr3 QDs and Co3O4 HPs is supported through the XPS survey spectrum (Fig. 2f). All the characteristic peaks are assigned to Cs, Pb, Br, Co, and O (highlighted with bars), affirming the formation of the CsPbBr3–Co3O4 heterojunction.


image file: d3qi00527e-f2.tif
Fig. 2 Comparative representation of high-resolution XPS spectra for (a) Cs 3d, (b) Pb 4f, (c) Br 3d, (d) Co 2p, and (e) O 1s in all the synthesized products. (f) The XPS survey spectrum highlights the co-existence of CsPbBr3 QDs and Co3O4 HPs in the CsPbBr3–Co3O4 heterojunction.

3.2. Investigation of the photocatalytic activity for CO2 reduction

The activity of the as-synthesized CsPbBr3–Co3O4 heterojunction for CO2 photocatalytic reduction is evaluated under the solid–gas environment using 50 μL of water as the proton source (see details in the ESI section 1.3).36,38,63 In particular, no sacrificial agent was used, and visible light was irradiated during photocatalysis. As shown in Fig. 3a, the pristine components only can produce limited CO and negligible CH4, calculated to be 14.23 and 0.39 (for CsPbBr3 QDs) and 9.52 and 0.46 (for Co3O4 HPs) μmol g−1 h−1, respectively. Notably, the CsPbBr3–Co3O4 heterojunction delivered an enhanced performance for the evolution of CO and CH4, which is 35.40 and 29.2 μmol g−1 h−1, respectively. The corresponding electron consumption rate (Relectron, see the formula in Table S2) was found to be 304.4 μmol g−1 h−1, which is 9.51 and 13.4 fold larger than that of pristine CsPbBr3 QDs and Co3O4 HPs, respectively. The maximum cumulative production of CO and CH4 after a 6-hour reaction was 211.58 and 175.31 μmol g−1, respectively (Fig. S6). The current performance of the CsPbBr3–Co3O4 heterojunction was found to be exceeding that of the pristine counter components as well as recent state-of-the-art photocatalysts (Table S1). Moreover, altering the dosage ratio of Co3O4[thin space (1/6-em)]:[thin space (1/6-em)]CsPbBr3 in the heterojunction could greatly influence the activity, for instance, the upper limit of gases was only reached with an optimum mass ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]4 (Fig. S7).
image file: d3qi00527e-f3.tif
Fig. 3 (a) A comparative depiction of the catalytic activity in terms of the produced gases CO and CH4 using the as-synthesized CsPbBr3 QDs, Co3O4 HPs, and CsPbBr3–Co3O4 heterojunction. (b) Controlled experiments were performed under different reaction conditions to produce CO and CH4 to examine the origin of the evolved gases. (c) Stability test of the CsPbBr3–Co3O4 heterojunction for 5 consecutive cycles.

More and/or less dosage of CsPbBr3 than the optimum value in the heterojunction is not conducive to producing CO and CH4 in high yields, maybe due to the insufficient amount and/or blockage of active sites respectively, and this phenomenon was found to be consistent with the literature.38 Additionally, control experiments were performed to confirm the origin of CO, and CH4 during photocatalysis (Fig. 3b). First, under an argon atmosphere, traces of CO or CH4 was found, which shows that feed-stock CO2 is necessary to run the photocatalytic reduction reaction.36,38,63 Second, nothing was detected under the dark conditions, which again confirms light irradiation is required to obtain CO and CH4 from the reduction of CO2. Finally, the optimum result could only be achieved using the CsPbBr3–Co3O4 heterojunction with a mass ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]4 as the photocatalyst, CO2 gas as the feedstock, and solar light illumination (λ > 400 nm). To further verify the origin of the carbon source of the photo-generated CO and CH4, 13CO2 isotope labelling was performed and the GC-MS result is displayed in Fig. S8. Evidently, the signals belonging to 13CO (m/z 29) and 13CH4 (m/z 17) are derived from 13CO2, validating that CO2 feedstock is the actual carbon source. Likewise, the oxidation of water to oxygen further validates the continuous supply of protons during CO2 conversion (Fig. S9). Furthermore, the CsPbBr3–Co3O4 heterojunction can remain stable during the consecutive 5 cycles (Fig. 3c). The morphological and structural characteristics of the CsPbBr3–Co3O4 heterojunction are well-maintained after photocatalysis (Fig. S10a and S10b).

3.3. Unveiling of the origins for improved activity

The origins of the possible enhanced activity were investigated and have been described systematically. The CO2 adsorption on the surface of the photocatalyst is a primary step that would further govern its activation and conversion. And it is well-known that CsPbBr3 QDs have a large surface area and are capable of adsorbing more CO2 on their surface.32,43 As verified by the TEM analysis, the size of CsPbBr3 QDs remained unchanged after forming a hybrid with Co3O4 HPs, indicating that it would exhibit similar CO2 adsorption activity on its surface. Likewise, Co3O4 HPs also possess numerous adsorptions sites derived from their structural mesopores. Although most of the mesopores are occupied by CsPbBr3 QDs (Fig. S11 and Table S2), additional porosities are still accessible in the heterojunction (as shown in the green marked area in Fig. 1g and h). These porosities would actively endure the CO2 adsorption.52,57

As the light absorption range is a key factor in solar-driven catalysis, to investigate the light-harvesting ability of the as-prepared components, diffuse reflectance spectra (DRS) were recorded. Both CsPbBr3 QDs and Co3O4 HPs show a strong visible light response within the wavelength range of 400–600 nm, and their corresponding energy band gaps (Eg), determined to be 2.36 eV and 2.31 eV, respectively, are well-consistent with the literature (Fig. 4a and b and corresponding insets).38,52 Where; Eg is estimated through Tauc plots using the equation αhν = A(Eg)n (see details in section 1.5 of the ESI). It is to be noted that E1g = 1.83 eV is attributed to the O2− → Co3+ excitation and is not a real energy bandgap of Co3O4 HPs.52,64–66 The enhanced light absorption ability of both components in the CsPbBr3–Co3O4 heterojunction supports its higher catalytic activity. Moreover, the valence and conduction band positions (EVB and ECB) strongly define the thermodynamic feasibility of CO2 reduction reactions. According to the reported results, the ECB of Co3O4 and CsPbBr3 are −0.64 eV52 and −0.99 eV,36 respectively, and correspondingly, the calculated EVB for Co3O4 HPs and CsPbBr3 QDs are 1.67 eV and 1.37 eV, respectively. Hence, the consequent CsPbBr3–Co3O4 heterojunction possess efficient light harvesting ability with staggered band alignment as depicted in Fig. 5a, further endorsing an uninterrupted CO2 conversion.


image file: d3qi00527e-f4.tif
Fig. 4 Tauc plots for energy bands and their corresponding insets depicting the diffuse reflectance spectra curves for (a) CsPbBr3 QDs and (b) Co3O4 HPs. (c) Steady-state photoluminescence spectra, and (d) time-resolved photoluminescence spectroscopy of CsPbBr3 QDs and the CsPbBr3–Co3O4 heterojunction. (e) Electrochemical impedance spectra and (f) transient photocurrent spectra of all the as-prepared products.

image file: d3qi00527e-f5.tif
Fig. 5 (a) Redox potentials and band alignment of the CsPbBr3–Co3O4 heterojunction. The in situ XPS analysis of (b) Cs 3d, (c) Pb 4f, (d) Br 3d, (e) Co 2p, and (f) O 1s, showing the electron-rich and electron-deficient sides in the CsPbBr3–Co3O4 heterojunction.

The rapid radiative recombination of photoinduced e–h+ pairs in pristine CsPbBr3 QDs can lead to low CO2 conversion efficiency.24 This was exclusively observed, where the intense photoluminescence (PL) spectra signal for pristine CsPbBr3 (Fig. 4c) confirmed its low catalytic activity. However, after immobilizing CsPbBr3 QDs on Co3O4 HPs, the results dramatically changed towards low PL intensity, identifying the significantly hindered recombination of e–h+ pairs over the CsPbBr3–Co3O4 heterojunction (Fig. 4c and Fig. S12). This finding matched well with the enhanced photocatalytic performance. This also means, combining such photocatalysts can substantially quench the radiative recombination of the metal halide perovskite, thereby exhibiting high activity. Likewise, to investigate the charge transfer kinetics, PL decay was also inquired via time-resolved photoluminescence spectroscopy (TRPL, curves were fitted with a bi-exponential function, Table S3 and Fig. 4d). The CsPbBr3–Co3O4 heterojunction demonstrated a shorter lifetime (τavg = 45.04 ns) in comparison with pristine CsPbBr3 QDs (τavg = 47.43 ns). This outcome further verified the restrained recombination of e–h+ pairs and their efficient separation in the CsPbBr3–Co3O4 hybrid. Besides charge recombination, charge transfer kinetics and their directional transfer are decisive to realize an improved CO2 conversion.61,67,68 Therefore, these were investigated for the CsPbBr3–Co3O4 heterojunction through electrochemical impedance spectroscopy (EIS) and transient photocurrent responses (TPR). The smallest arc radius was seen for the CsPbBr3–Co3O4 heterojunction followed by CsPbBr3 QDs and Co3O4 HPs (Fig. 4e). The markedly inhibited resistance in CsPbBr3–Co3O4 would promote charge transfer. Likewise, the CsPbBr3–Co3O4 shows the highest photocurrent with high repeatability (Fig. 4f). This result further supports the accelerated separation of e–h+ pairs as well as highlights that an accessible route builds up (i.e., heterointerface) by decorating CsPbBr3 QDs on Co3O4 HPs.

3.4. Uncovering charge redistribution on the intimately coupled heterointerface and the corresponding catalytic mechanism

To further shed light on the interfacial charge-carrier flux, and their transfer route, particularly the accumulation of reductive electrons, the in situ XPS spectra were conducted under dark and visible light irradiation, respectively. As can be seen, the binding energies corresponding to Cs 3d, Pb 4f, and Br 3d exclusively underwent a positive shift (Fig. 5a–c) under light, which demonstrates a decrease of the electron density in the CsPbBr3 side. In the meantime, the peak attributed to Co 2p and O 1s shifted to lower binding energy, suggesting an increased electron density at the Co3O4 side (Fig. 5e and f). Such patterns verify the continuous transfer of photoinduced electrons from CsPbBr3 to Co3O4. In other words, Co3O4 plays a role as a supporting matrix to evoke and gather electrons from CsPbBr3 QDs, allowing electron localization on the active surface of Co3O4. At the same time, an uninterrupted separation of e–h+ pairs occurs in CsPbBr3 owing to the significantly improved light harvesting ability and abundant exposed surface. Therefore, the electron-rich and electron-deficient phenomena occur simultaneously to balance the electron redistribution in the CsPbBr3–Co3O4 heterojunction. Notably, except for the negative shift in O 1s spectra (Fig. 5f), no new peak belonging to OVs appeared, validating the structural permanence of Co3O4 in the heterojunction.

Furthermore, the work function (Φ) is another pivotal factor for charge carrier kinetics, particularly at the heterointerfaces. The Φ for CsPbBr3 QDs and Co3O4 HPs are 4.39 eV and 5.74 eV respectively.33,52 So, a larger Φ value of Co3O4 refers to its lower Fermi level (Ef) than that of CsPbBr3, implying that the flow of electrons is from CsPbBr3 to Co3O4 which is consistent with the outcomes of in situ XPS. This would hasten the extraction of reductive electrons from CsPbBr3 to Co3O4, thus suppressing the recombination of e–h+ pairs. In a word, CsPbBr3 facilitated the rapid and continuous electron transfer via shorter diffusion pathways due to the size effect, and Co3O4 spontaneously accepted the electrons and accumulated on its active surface to exclusively participate in CO2 reduction under the irradiation of solar light.

Based on the above results, the proposed CO2 photoreduction mechanism over the CsPbBr3–Co3O4 hybrid has been described. As illustrated in Fig. 6a, CsPbBr3 and Co3O4 in the heterojunction have a staggered band alignment (type-II). In addition, the CsPbBr3–Co3O4 heterojunction holds enough negative CB and positive VB potentials to reduce CO2 and oxidize water, respectively. Under light irradiation, electrons and holes can be generated in both CsPbBr3 and Co3O4. Subsequently, the photoinduced electrons are transferred from CB of CsPbBr3 to CB of CO3O4 (Fig. 6b), causing electron localization at the CO3O4 side, making it suitable to conduct the CO2 reduction reaction (as mentioned earlier in the in situ XPS analysis section). In the meantime, holes are transported in opposite directions i.e., from VB of CO3O4 to VB of CsPbBr3, building a hole-rich CsPbBr3 side, favourable for oxidizing H2O to O2 and H+. Accordingly, the breaking of C[double bond, length as m-dash]O bonds in CO2 occurred at the electron-rich Co3O4 side, where the coupling of electron and protons took place, for instance, 2e and 8e get coupled with 2H+ and 8H+ to produce CO and CH4, respectively, as presented in eqn (1) and (2). Thus, the effective charge separation enables the uninterrupted electron–proton integrative reaction to realize highly efficient CO2 photocatalytic reduction over the CsPbBr3–Co3O4 heterojunction.

 
CO2 + 2H+ + 2e → CO + H2O(1)
 
CO2 + 8H+ + 8e → CH4 + 2H2O(2)


image file: d3qi00527e-f6.tif
Fig. 6 (a) Photoexcited charge separation and their transfer route over the CsPbBr3–Co3O4 heterojunction under visible light irradiation. (b) Schematic diagram depicting the continuous flow of electrons from CsPbBr3 towards Co3O4 enabling the photocatalytic conversion of CO2.

4. Conclusions

In conclusion, a facile solution-processed room-temperature method was developed to fabricate a new CsPbBr3–Co3O4 heterojunction photocatalyst for visible-light-driven CO2 conversion. The successful formation of the CsPbBr3–Co3O4 hybrid was confirmed by XRD, TEM, HRTEM and XPS analysis. The enhanced generation of CO (35.40 μmol g−1 h−1) and CH4 (29.2 μmol g−1 h−1) with a high Relectron rate (304.4 μmol g−1 h−1) was realized on the CsPbBr3–Co3O4 hybrid, which outperformed the counter products and some state-of-the-art photocatalysts. The formation of staggered band alignment and an intimately contacted heterointerface led to robust separation and directional transfer of charge carriers. Because of the uninterrupted flow of reductive e, an electron enrichment zone was obtained at the active site of Co3O4, which ensured the efficient conversion of CO2. This work describes the potential of designing a new perovskite-based heterojunction that could enable efficient charge separation to achieve better solar-driven CO2 conversion.

Author contributions

Xin Zhong: methodology, investigation, data curation, and writing the original draft. Xinmeng Liang: methodology, investigation, data curation, and writing the original draft. Xinyu Lin: methodology, investigation, and validation. Malik Zeeshan Shahid: writing the original draft, visualization, investigation, and data curation. Jin Wang: supervision, conceptualization, descriptions, and funding acquisition. Zhengquan Li: supervision, conceptualization, descriptions, and funding acquisition.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was financially supported by the National Natural Science Foundation of China (21701143, 21975223), the Natural Science Foundation of Zhejiang Province (LGG19B010002, LZ22B030002), and the Industrial Key Projects of Jinhua City (2021A22383).

References

  1. J. Wang, Y. Shi, Y. Wang and Z. Li, Rational design of metal halide perovskite nanocrystals for photocatalytic CO2 reduction: Recent advances, challenges, and prospects, ACS Energy Lett., 2022, 7, 2043–2059 CrossRef CAS.
  2. S. Yin, X. Zhao, E. Jiang, Y. Yan, P. Zhou and P. Huo, Boosting water decomposition by sulfur vacancies for efficient CO2 photoreduction, Energy Environ. Sci., 2022, 15, 1556–1562 RSC.
  3. Y. Feng, C. Wang, P. Cui, C. Li, B. Zhang, L. Gan, S. Zhang, X. Zhang, X. Zhou, Z. Sun, K. Wang, Y. Duan, H. Li, K. Zhou, H. Huang, A. Li, C. Zhuang, L. Wang, Z. Zhang and X. Han, Ultrahigh photocatalytic CO2 reduction efficiency and selectivity manipulation by single-tungsten-atom oxide at the atomic step of TiO2, Adv. Mater., 2022, 34, 2109074 CrossRef CAS PubMed.
  4. J. Wang, L. Xiong, Y. Bai, Z. J. Chen, Q. Zheng, Y. Y. Shi, C. Zhang, G. C. Jiang and Z. Q. Li, Mn-doped perovskite nanocrystals for photocatalytic CO2 reduction: Insight into the role of the charge carriers with prolonged lifetime, Sol. RRL, 2022, 6, 2200294 CrossRef CAS.
  5. S. Xu, Q. Shen, J. Zheng, Z. Wang, X. Pan, N. Yang and G. Zhao, Advances in biomimetic photoelectrocatalytic reduction of carbon dioxide, Adv. Sci., 2022, 2203941 CrossRef CAS PubMed.
  6. S. H. Teo, C. H. Ng, Y. H. Ng, A. Islam, S. Hayase and Y. H. Taufiq-Yap, Resolve deep-rooted challenges of halide perovskite for sustainable energy development and environmental remediation, Nano Energy, 2022, 99, 107401 CrossRef CAS.
  7. E. Gong, S. Ali, C. B. Hiragond, H. S. Kim, N. S. Powar, D. Kim, H. Kim and S.-I. In, Solar fuels: research and development strategies to accelerate photocatalytic CO2 conversion into hydrocarbon fuels, Energy Environ. Sci., 2022, 15, 880–937 RSC.
  8. Y. Wang, E. Chen and J. Tang, Insight on reaction pathways of photocatalytic CO2 conversion, ACS Catal., 2022, 12, 7300–7316 CrossRef CAS PubMed.
  9. Ž. Kovačič, B. Likozar and M. Huš, Photocatalytic CO2 reduction: A review of Ab initio mechanism, kinetics, and multiscale modeling simulations, ACS Catal., 2020, 10, 14984–15007 CrossRef.
  10. H. Bian, D. Li, S. Wang, J. Yan and S. F. Liu, 2D-C3N4 encapsulated perovskite nanocrystals for efficient photo-assisted thermocatalytic CO2 reduction, Chem. Sci., 2022, 13, 1335–1341 RSC.
  11. S. D. Yang, H. Y. Li, H. M. Li, H. M. Li, W. S. Qi, Q. Zhang, J. Zhu, P. Zhao and L. Chen, Rational design of 3D carbon nitrides assemblies with tunable nano-building blocks for efficient visible-light photocatalytic CO2 conversion, Appl. Catal., B, 2022, 316, 121612 CrossRef CAS.
  12. M. Gao, J. Zhang, P. Zhu, X. Liu and Z. Zheng, Unveiling the origin of alkali metal promotion in CO2 methanation over Ru/ZrO2, Appl. Catal., B, 2022, 314, 121476 CrossRef CAS.
  13. L. Collado, P. Reñones, J. Fermoso, F. Fresno, L. Garrido, V. Pérez-Dieste, C. Escudero, M. D. Hernández-Alonso, J. M. Coronado, D. P. Serrano and V. A. D. O’Shea, The role of the surface acidic/basic centers and redox sites on TiO2 in the photocatalytic CO2 reduction, Appl. Catal., B, 2022, 303, 120931 CrossRef CAS.
  14. S. Wei, Q. Heng, Y. Wu, W. Chen, X. Li and W. Shangguan, Improved photocatalytic CO2 conversion efficiency on Ag loaded porous Ta2O5, Appl. Surf. Sci., 2021, 563, 150273 CrossRef CAS.
  15. X. Lin, S. Xia, L. Zhang, Y. Zhang, S. Sun, Y. Chen, S. Chen, B. Ding, J. Yu and J. Yan, Fabrication of flexible mesoporous black Nb2O5 nanofiber films for visible-light-driven photocatalytic CO2 Reduction into CH4, Adv. Mater., 2022, 34, 2200756 CrossRef CAS PubMed.
  16. J. Wang, S. Lin, N. Tian, T. Ma, Y. Zhang and H. Huang, Nanostructured metal sulfides: classification, modification strategy, and solar-driven CO2 reduction application, Adv. Funct. Mater., 2020, 31, 2008008 CrossRef.
  17. C. I. Ezugwu, S. Liu, C. Li, S. Zhuiykov, S. Roy and F. Verpoort, Engineering metal-organic frameworks for efficient photocatalytic conversion of CO2 into solar fuels, Coord. Chem. Rev., 2022, 450, 214245 CrossRef CAS.
  18. C. B. Hiragond, N. S. Powar, J. Lee and S. I. In, Single-atom catalysts (SACs) for photocatalytic CO2 reduction with H2O: activity, product selectivity, stability, and surface chemistry, Small, 2022, 18, 2201428 CrossRef CAS PubMed.
  19. X. J. Liu, T. Q. Chen, Y. H. Xue, J. C. Fan, S. L. Shen, M. S. A. A. Hossain, M. A. Amin, L. K. Pan, X. T. Xu and Y. Yamauchi, Nanoarchitectonics of MXene/semiconductor heterojunctions toward artificial photosynthesis via photocatalytic CO2 reduction, Coord. Chem. Rev., 2022, 459, 214440 CrossRef CAS.
  20. S.-H. Wang, F. Khurshid, P.-Z. Chen, Y.-R. Lai, C.-W. Cai, P.-W. Chung, M. Hayashi, R.-J. Jeng, S.-P. Rwei and L. Wang, Solution-processable naphthalene diimide-based conjugated polymers as organocatalysts for photocatalytic CO2 reaction with extremely stable catalytic activity for over 330 hours, Chem. Mater., 2022, 34, 4955–4963 CrossRef CAS.
  21. S. Chen, H. Yin, P. Liu, Y. Wang and H. Zhao, Stabilisation and performance enhancement strategies for halide perovskite photocatalysts, Adv. Mater., 2022, 35, 2203836 CrossRef PubMed.
  22. H. Huang, D. Verhaeghe, B. Weng, B. Ghosh, H. Zhang, J. Hofkens, J. A. Steele and M. B. J. Roeffaers, Metal halide perovskite based heterojunction photocatalysts, Angew. Chem., Int. Ed., 2022, 61, e202203261 CAS.
  23. S. Purohit, K. L. Yadav and S. Satapathi, Metal halide perovskite heterojunction for photocatalytic hydrogen generation: progress and future opportunities, Adv. Mater. Interfaces, 2022, 9, 2200058 CrossRef CAS.
  24. L. Y. Wu, Y. F. Mu, X. X. Guo, W. Zhang, Z. M. Zhang, M. Zhang and T. B. Lu, Encapsulating perovskite quantum dots in iron-based metal-organic frameworks (MOFs) for efficient photocatalytic CO2 reduction, Angew. Chem., Int. Ed., 2019, 58, 9491–9495 CrossRef CAS PubMed.
  25. X. Zhu, L. Ge, Y. Wang, M. Li, R. Zhang, M. Xu, Z. Zhao, W. Lv and R. Chen, Recent advances in enhancing and enriching the optical properties of Cl-based CsPbX3 nanocrystals, Adv. Opt. Mater., 2021, 9, 2100058 CrossRef CAS.
  26. M. Ou, W. Tu, S. Yin, W. Xing, S. Wu, H. Wang, S. Wan, Q. Zhong and R. Xu, Amino–assisted anchoring of CsPbBr3 perovskite quantum dots on porous g-C3N4 for enhanced photocatalytic CO2 reduction, Angew. Chem., Int. Ed., 2018, 57, 13570–13574 CrossRef CAS PubMed.
  27. M. Madi, M. Tahir and S. Tasleem, Advances in structural modification of perovskite semiconductors for visible light assisted photocatalytic CO2 reduction to renewable solar fuels: A review, J. Environ. Chem. Eng., 2021, 9, 106264 CrossRef CAS.
  28. Y. F. Mu, W. Zhang, X. X. Guo, G. X. Dong, M. Zhang and T. B. Lu, Water-tolerant lead halide perovskite nanocrystals as efficient photocatalysts for visible-light-driven CO2 reduction in pure water, ChemSusChem, 2019, 12, 4769–4774 CrossRef CAS PubMed.
  29. L. Clinckemalie, D. Valli, M. B. J. Roeffaers, J. Hofkens, B. Pradhan and E. Debroye, Challenges and opportunities for CsPbBr3 perovskites in low- and high-energy radiation detection, ACS Energy Lett., 2021, 6, 1290–1314 CrossRef CAS.
  30. J. Wang, J. Liu, Z. Du and Z. Li, Recent advances in metal halide perovskite photocatalysts: Properties, synthesis and applications, J. Energy Chem., 2021, 54, 770–785 CrossRef CAS.
  31. G. Gao, Q. Xi, H. Zhou, Y. Zhao, C. Wu, L. Wang, P. Guo and J. Xu, Novel inorganic perovskite quantum dots for photocatalysis, Nanoscale, 2017, 9, 12032–12038 RSC.
  32. Q. A. Akkerman, T. P. T. Nguyen, S. C. Boehme, F. Montanarella, D. N. Dirin, P. Wechsler, F. Beiglbock, G. Raino, R. Erni, C. Katan, J. Even and M. V. Kovalenko, Controlling the nucleation and growth kinetics of lead halide perovskite quantum dots, Science, 2022, 377, 1406–1412 CrossRef CAS PubMed.
  33. Z. Zhang, L. Li, Y. Jiang and J. Xu, Step-scheme photocatalyst of CsPbBr3 quantum dots/BiOBr nanosheets for efficient CO2 Photoreduction, Inorg. Chem., 2022, 61, 3351–3360 CrossRef CAS PubMed.
  34. X. Y. Yue, L. Cheng, J. J. Fan and Q. J. Xiang, 2D/2D BiVO4/CsPbBr3 S-scheme heterojunction for photocatalytic CO2 reduction: Insights into structure regulation and Fermi level modulation, Appl. Catal., B, 2022, 304, 120979 CrossRef CAS.
  35. C. C. Lin, T. R. Liu, S. R. Lin, K. M. Boopathi, C. H. Chiang, W. Y. Tzeng, W. C. Chien, H. S. Hsu, C. W. Luo, H. Y. Tsai, H. A. Chen, P. C. Kuo, J. Shiue, J. W. Chiou, W. F. Pong, C. C. Chen and C. W. Chen, Spin-polarized photocatalytic CO2 reduction of Mn-Doped perovskite nanoplates, J. Am. Chem. Soc., 2022, 144, 15718–15726 CrossRef CAS PubMed.
  36. Q. Wang, J. Wang, J. C. Wang, X. Hu, Y. Bai, X. Zhong and Z. Li, Coupling CsPbBr3 quantum dots with covalent triazine frameworks for visible-light-driven CO2 reduction, ChemSusChem, 2021, 14, 1131–1139 CrossRef CAS PubMed.
  37. L. Ding, B. Borjigin, Y. Li, X. Yang, X. Wang and H. Li, Assembling an affinal 0D CsPbBr3/2D CsPb2Br5 architecture by synchronously in situ growing CsPbBr3 QDs and CsPb2Br5 nanosheets: enhanced activity and reusability for photocatalytic CO2 reduction, ACS Appl. Mater. Interfaces, 2021, 13, 51161–51173 CrossRef CAS PubMed.
  38. J. Wang, J. Wang, N. Li, X. Du, J. Ma, C. He and Z. Li, Direct Z-scheme 0D/2D heterojunction of CsPbBr3 quantum Dots/Bi2WO6 nanosheets for efficient photocatalytic CO2 reduction, ACS Appl. Mater. Interfaces, 2020, 12, 31477–31485 CrossRef CAS PubMed.
  39. Y. Jiang, H. Y. Chen, J. Y. Li, J. F. Liao, H. H. Zhang, X. D. Wang and D. B. Kuang, Z-Scheme 2D/2D heterojunction of CsPbBr3/Bi2WO6 for improved photocatalytic CO2 reduction, Adv. Funct. Mater., 2020, 30, 2004293 CrossRef CAS.
  40. Z. J. Chen, Y. G. Hu, J. Wang, Q. Shen, Y. H. Zhang, C. Ding, Y. Bai, G. C. Jiang, Z. Q. Li and N. Gaponik, Boosting photocatalytic CO2 reduction on CsPbBr3 perovskite nanocrystals by immobilizing metal complexes, Chem. Mater., 2020, 32, 1517–1525 CrossRef CAS.
  41. S. Wan, M. Ou, Q. Zhong and X. Wang, Perovskite-type CsPbBr3 quantum dots/UiO-66(NH2) nanojunction as efficient visible-light-driven photocatalyst for CO2 reduction, Chem. Eng. J., 2019, 358, 1287–1295 CrossRef CAS.
  42. Z.-C. Kong, J.-F. Liao, Y.-J. Dong, Y.-F. Xu, H.-Y. Chen, D.-B. Kuang and C.-Y. Su, Core@Shell CsPbBr3@Zeolitic imidazolate framework nanocomposite for efficient photocatalytic CO2 reduction, ACS Energy Lett., 2018, 3, 2656–2662 CrossRef CAS.
  43. Y. F. Xu, M. Z. Yang, B. X. Chen, X. D. Wang, H. Y. Chen, D. B. Kuang and C. Y. Su, A CsPbBr3 perovskite quantum dot/Graphene oxide composite for photocatalytic CO2 reduction, J. Am. Chem. Soc., 2017, 139, 5660–5663 CrossRef CAS PubMed.
  44. Q. Zhang, P. Yang, H. Zhang, J. Zhao, H. Shi, Y. Huang and H. Yang, Oxygen vacancies in Co3O4 promote CO2 photoreduction, Appl. Catal., B, 2022, 300, 120729 CrossRef CAS.
  45. Z. Feng, X. Zhu, J. Yang, K. Zhong, Z. Jiang, Q. Yu, Y. Song, Y. Hua, H. Li and H. Xu, Inherent facet-dominant effect for cobalt oxide nanosheets to enhance photocatalytic CO2 reduction, Appl. Surf. Sci., 2022, 578, 151848 CrossRef CAS.
  46. P. Yang, Q. Zhang, Z. Yi, J. Wang and H. Yang, Rational electronic control of carbon dioxide reduction over cobalt oxide, J. Catal., 2020, 387, 119–128 CrossRef CAS.
  47. L. Huang, B. Li, B. Su, Z. Xiong, C. Zhang, Y. Hou, Z. Ding and S. Wang, Fabrication of hierarchical Co3O4@CdIn2S4 p–n heterojunction photocatalysts for improved CO2 reduction with visible light, J. Mater. Chem. A, 2020, 8, 7177–7183 RSC.
  48. L. Wang, J. Wan, Y. Zhao, N. Yang and D. Wang, Hollow multi-shelled structures of Co3O4 dodecahedron with unique crystal orientation for enhanced photocatalytic CO2 reduction, J. Am. Chem. Soc., 2019, 141, 2238–2241 CrossRef CAS PubMed.
  49. J. Y. Choi, C. K. Lim, B. Park, M. Kim, A. Jamal and H. Song, Surface activation of cobalt oxide nanoparticles for photocatalytic carbon dioxide reduction to methane, J. Mater. Chem. A, 2019, 7, 15068–15072 RSC.
  50. W. Chen, B. Han, C. Tian, X. Liu, S. Liang, H. Deng and Z. Lin, MOFs-derived ultrathin holey Co3O4 nanosheets for enhanced visible light CO2 reduction, Appl. Catal., B, 2019, 244, 996–1003 CrossRef CAS.
  51. X. Huang, Q. Shen, J. Liu, N. Yang and G. Zhao, A CO2 adsorption-enhanced semiconductor/metal-complex hybrid photoelectrocatalytic interface for efficient formate production, Energy Environ. Sci., 2016, 9, 3161–3171 RSC.
  52. C. Gao, Q. Meng, K. Zhao, H. Yin, D. Wang, J. Guo, S. Zhao, L. Chang, M. He, Q. Li, H. Zhao, X. Huang, Y. Gao and Z. Tang, Co3O4 Hexagonal platelets with controllable facets enabling highly efficient visible-light photocatalytic reduction of CO2, Adv. Mater., 2016, 28, 6485–6490 CrossRef CAS PubMed.
  53. L. Wang, J. Deng, Z. Lou and T. Zhang, Nanoparticles-assembled Co3O4 nanorods p-type nanomaterials: One-pot synthesis and toluene-sensing properties, Sens. Actuators, B, 2014, 201, 1–6 CrossRef CAS.
  54. L. Li, C. Zhang, R. Zhang, X. Gao, S. He, M. Liu, X. Li and W. Chen, 2D ultrathin Co3O4 nanosheet array deposited on 3D carbon foam for enhanced ethanol gas sensing application, Sens. Actuators, B, 2017, 244, 664–672 CrossRef CAS.
  55. A. Aljabour, H. Coskun, D. H. Apaydin, F. Ozel, A. W. Hassel, P. Stadler, N. S. Sariciftci and M. Kus, Nanofibrous cobalt oxide for electrocatalysis of CO2 reduction to carbon monoxide and formate in an acetonitrile-water electrolyte solution, Appl. Catal., B, 2018, 229, 163–170 CrossRef CAS.
  56. Y. Gu, J. Ding, X. Tong, H. Yao, R. Yang and Q. Zhong, Photothermal catalyzed hydrogenation of carbon dioxide over porous nanosheet Co3O4, J. CO2 Util., 2022, 61, 102003 CrossRef CAS.
  57. X. Zhu, H. Ji, J. Yi, J. Yang, X. She, P. Ding, L. Li, J. Deng, J. Qian, H. Xu and H. Li, A specifically exposed cobalt oxide/carbon nitride 2D heterostructure for carbon dioxide photoreduction, Ind. Eng. Chem. Res., 2018, 57, 17394–17400 CrossRef CAS.
  58. W. Yang, G. Ma, Y. Fu, K. Peng, H. Yang, X. Zhan, W. Yang, L. Wang and H. Hou, Rationally designed Ti3C2 MXene@TiO2/CuInS2 Schottky/S-scheme integrated heterojunction for enhanced photocatalytic hydrogen evolution, Chem. Eng. J., 2022, 429, 132381 CrossRef CAS.
  59. X.-T. Liu, B.-H. Li, X.-J. Wang, Y.-L. Li, J. Zhao, Y.-P. Li and F.-T. Li, Enhanced schottky effect in the Ni2P cocatalyst via work function up-shift induced by MoO2 for boosting photocatalytic hydrogen evolution, ACS Sustainable Chem. Eng., 2022, 10, 10627–10640 CrossRef CAS.
  60. K. Liu, H. Zhang, T. Fu, L. Wang, R. Tang, Z. Tong and X. Huang, Construction of BiOBr/Ti3C2/exfoliated montmorillonite Schottky junction: New insights into exfoliated montmorillonite for inducing MXene oxygen functionalization and enhancing photocatalytic activity, Chem. Eng. J., 2022, 438, 135609 CrossRef CAS.
  61. Y. Li, Q. Zhao, Y. Zhang, Y. Li, L. Fan, F.-T. Li and X. Li, In-situ construction of sequential heterostructured CoS/CdS/CuS for building “electron-welcome zone” to enhance solar-to-hydrogen conversion, Appl. Catal., B, 2022, 300, 120763 CrossRef CAS.
  62. J. Li, L. Xu, T. Wang, J. Song, J. Chen, J. Xue, Y. Dong, B. Cai, Q. Shan, B. Han and H. Zeng, 50-Fold EQE Improvement up to 6.27% of solution-processed all-inorganic perovskite CsPbBr3 QLEDs via surface ligand density control, Adv. Mater., 2017, 29, 1603885 CrossRef PubMed.
  63. N. Li, X. Chen, J. Wang, X. Liang, L. Ma, X. Jing, D. L. Chen and Z. Li, ZnSe Nanorods-CsSnCl3 perovskite heterojunction composite for photocatalytic CO2 reduction, ACS Nano, 2022, 16, 3332–3340 CrossRef CAS PubMed.
  64. D. Barreca, C. Massignan, S. Daolio, M. Fabrizio, C. Piccirillo, L. Armelao and E. Tondello, Composition and microstructure of cobalt oxide thin films obtained from a novel cobalt(II) precursor by chemical vapor deposition, Chem. Mater., 2001, 13, 588–593 CrossRef CAS.
  65. Z. Hu, L. Hao, F. Quan and R. Guo, Recent developments of Co3O4-based materials as catalysts for the oxygen evolution reaction, Catal. Sci. Technol., 2022, 12, 436–461 RSC.
  66. Y. Xu, F. Zhang, T. Sheng, T. Ye, D. Yi, Y. Yang, S. Liu, X. Wang and J. Yao, Clarifying the controversial catalytic active sites of Co3O4 for the oxygen evolution reaction, J. Mater. Chem. A, 2019, 7, 23191–23198 RSC.
  67. W. He, L. Liu, T. Ma, H. Han, J. Zhu, Y. Liu, Z. Fang, Z. Yang and K. Guo, Controllable morphology CoFe2O4/g-C3N4 p-n heterojunction photocatalysts with built-in electric field enhance photocatalytic performance, Appl. Catal., B, 2022, 306, 121107 CrossRef CAS.
  68. S. Bai, J. Jiang, Q. Zhang and Y. Xiong, Steering charge kinetics in photocatalysis: intersection of materials syntheses, characterization techniques and theoretical simulations, Chem. Soc. Rev., 2015, 44, 2893–2939 RSC.

Footnotes

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d3qi00527e
These authors contributed equally.

This journal is © the Partner Organisations 2023
Click here to see how this site uses Cookies. View our privacy policy here.