Z-scheme Co3O4/BiOBr heterojunction for efficient photoreduction CO2 reduction

Xiaofan ZHANG Yu DUAN Meijie SHI Nan LU Renhong LI Xiaoqing YAN

Citation:  Xiaofan ZHANG, Yu DUAN, Meijie SHI, Nan LU, Renhong LI, Xiaoqing YAN. Z-scheme Co3O4/BiOBr heterojunction for efficient photoreduction CO2 reduction[J]. Chinese Journal of Inorganic Chemistry, 2025, 41(9): 1878-1888. doi: 10.11862/CJIC.20250079 shu

Z型异质结Co3O4/BiOBr的构筑及其光催化CO2还原性能

    通讯作者: 李仁宏,
    闫晓庆, yanxiaoqing927@126.com
  • 基金项目:

    国家自然科学基金 22172143

    浙江理工大学科研发展专项 23212093-Y

摘要: 通过简便的一步溶剂热法成功构建了Co3O4/BiOBr异质结光催化剂, 并用于高效选择性光催化CO2还原。优化后的Co3O4/BiOBr-0.8催化剂表现出优异的光催化性能, 其CO和CH4生成速率分别达到了112.2和5.5 μmol·g-1·h-1, 相较于原始BiOBr分别提高了6.3倍和3.9倍。表征结果表明, 该异质结具有显著增强的全光谱光吸收效率、优越的光电性能、较小的阻抗以及改善的光生载流子(e-/h+)分离效率。这些协同效应源于Z型异质结结构的形成, 该结构不仅促进了太阳光利用效率和电子还原能力, 同时有效抑制了载流子复合。

English

  • Photocatalysis for the conversion of CO2 to fuels and chemical materials holds considerable promise in reducing humanity's dependence on fossil fuels, which is fundamental to addressing energy shortages and mitigating CO2 production[1-3]. An efficient CO2 reduction photocatalyst requires broad-spectrum light absorption ability, high-efficiency charge separation, and robust redox capacity. Nevertheless, developing an efficient and highly selective CO2 catalyst remains a formidable challenge. This is mainly attributed to the high dissociation energy (750 kJ·mol-1) required to break the C=O double bond, as well as the complexity introduced by multiple reaction intermediates[2, 4]. As a result, the design of highly active photocatalytic systems for CO2 reduction continues to be a significant challenge.

    The solar photo-driven semiconductor photocatalysts have garnered considerable attention due to their optimal redox potential, environmental compatibility, sustainability, and exceptional efficiency[5]. However, the utilization of single-component catalysts is hindered by the rapid recombination of photogenerated carriers (h+/e-), which restricts their applications owing to the minimal quantum efficiency[6]. To overcome this challenge, researchers have devoted significant efforts to enhancing catalyst performance through techniques such as precious metal deposition[7], atom doping[8], or heterojunction construction[9-11]. Recently, the integration of heterojunctions through the coupling of semiconductors has emerged as a promising strategy to enhance the performance of photocatalysts. It is important to highlight that the construction of Z-scheme heterojunctions can improve the separation of h+/e- and electron transport efficiency while maintaining superior redox ability, such as Bi2O2.33-CdS[12], ZnIn2S4@Ni(OH)2/NiO[13], g‑C3N5/BiOBr[14]. However, it remains crucial to develop innovative Z‑scheme heterojunctions with enhanced activity for CO2 photocatalytic reduction via a facile synthesis process.

    The bismuth-based catalysts, particularly BiOX (X=Cl, Br, and I) with a layered ternary oxide structure, have garnered significant attention has been garnered in the fields of hydrogen evolution, pollutant photodegradation, and CO2 reduction owing to the unique structural characteristics and superior optical and electronic properties of these materials[15]. Among them, BiOBr has been extensively investigated owing to its relatively wide bandgap energy and the appropriate alignment of its conduction and valence band (VB). Although BiOBr demonstrates advantages in visible-light absorption and structural tunability, its practical application in CO2 photoreduction is limited by four intrinsic challenges: (ⅰ) inadequate light harvesting beyond 460 nm, (ⅱ) rapid charge recombination induced by interlayer polarization, (ⅲ) insufficient CO2 adsorption and activation, and (ⅳ) restricted product selectivity. These challenges motivated us to design and construct a Z-type heterojunction of Co3O4/BiOBr by fully exploiting the synergistic advantages of both components, where the narrow-bandgap Co3O4 facilitates visible-light response across the entire spectrum.

    The Co3O4/BiOBr heterojunctions with various molar ratios of Co3O4 and BiOBr were synthesized using a facile chemical method at ambient temperature. The CO2 photoreduction efficiency was systematically investigated by evaluating the influence of Co3O4/BiOBr molar ratios, H2O dosage, irradiation time, and sacrificial reagent in the absence of a photosensitizer. Remarkably, the Co3O4/BiOBr-0.8 heterojunction exhibited an impressive CO evolution rate of 112.2 μmol·g-1·h-1, which was 6.3 times higher than pristine BiOBr. Experimental findings unequivocally demonstrate that the as-synthesized Co3O4/BiOBr heterojunction possesses a Z-scheme band structure, thereby facilitating efficient charge separation dynamics, superior redox capability, and enhanced adsorption and activation capacity towards CO2, leading to a substantial enhancement in its photocatalytic performance for CO2 reduction.

    The reagents utilized in this investigation were of analytical grade and employed without any additional purification steps (Supporting information).

    The Bi(NO3)3·5H2O solution was prepared by dissolving 1 mmol Bi(NO3)3·5H2O in 45 mL ethylene glycol and subjected to ultrasonication for 45 min. Subsequently, solution A was formed by dissolving 1 mmol cetyltrimethylammonium bromide (CTAB) and 0.2 g polyvinylpyrrolidone (PVP) in the aforementioned Bi(NO3)3·5H2O solution. Separate batches of ethylene glycol (24 mL each) were used to disperse solutions containing varying amounts (0, 0.2, 0.4, 0.6, 0.8, and 1 mmol) of Co3O4, respectively. The thoroughly mixed Co3O4 solution was then combined with the solution A under continuous stirring for 30 min to form solution B. The resulting mixture was kept in an oven at a temperature of 180 ℃ for 10 h. The resulting precipitates were cooled to room temperature before being collected through filtration, washed multiple times using deionized water, and dried at 60 ℃ for 10 h to obtain Co3O4/BiOBr powders exhibiting varying molar ratios of Co3O4 to BiOBr (0.2∶1, 0.4∶1, 0.6∶1, 0.8∶1, and 1∶1), which were referred to as Co3O4/BiOBr-0.2, Co3O4/BiOBr-0.4, Co3O4/BiOBr-0.6, Co3O4/BiOBr-0.8, and Co3O4/BiOBr-1.0, respectively. The pure BiOBr was prepared using the same method without the addition of Co3O4.

    The photocatalytic reduction of CO2 measurement process and photoelectrochemical measurement are shown in Supporting information.

    The SEM images in Fig.1a and 1b exhibit the morphology of Co3O4/BiOBr‑0.8, revealing a distinctive nanoflower structure with incorporated Co3O4 nanospheres. Furthermore, TEM image confirms the presence of a special nanoflower with nanosphere morphology in Co3O4/BiOBr-0.8 (Fig.1c), demonstrating evident integration between the nanospheres and nanosheets compared to pristine BiOBr (Fig.S1). The high temperature and pressure environment hinders the growth of the synthesis process and benefits the formation of heterogeneous interfaces. High-resolution transmission electron microscopy (HRTEM) analysis of Co3O4/ BiOBr‑0.8 revealed an intimate contact interface between the heterojunctions, as evidenced by lattice spacings of 0.28 and 0.27 nm corresponding to the Co3O4 (220) plane and BiOBr (003) plane[16-17], respectively (Fig.1d). The distribution of Bi (green), O (yellow), Br (orange), and Co (purple) elements in the Co3O4/BiOBr-0.8 heterojunction was found to be uniform (Fig.1e and 1f). Additionally, Fig.S1 shows the SEM image of pure BiOBr, which shows the nanoflower structure combined with nanosheets.

    Figure 1

    Figure 1.  (a, b) SEM images, (c) TEM image, (d) HRTEM image, and (e, f) elemental mapping images of Co3O4/BiOBr-0.8

    The XRD patterns presented in Fig.2a display characteristic peaks corresponding to Co3O4/BiOBr composites with varying molar ratios. These peaks confirm the presence of tetragonal structures, specifically BiOBr (PDF No.09‑0393) and Co3O4 (PDF No.42‑1467)[17]. The sharpness of these peaks indicated a high degree of crystallinity across all samples examined. Moreover, the intensity of diffraction peaks associated with Co3O4 showed a gradual increase, suggesting its coexistence within the composite materials alongside BiOBr. Additionally, higher concentrations of Co3O4 result in enhanced intensities at the (311) plane peak position, while lower concentrations of BiOBr lead to reduced intensities at the (101) and (113) plane peaks due to interface interactions between Co3O4 and BiOBr.

    Figure 2

    Figure 2.  (a) XRD patterns of Co3O4/BiOBr; (b) Br3d, (c) Bi4f, and (d) O1s XPS spectra of Co3O4, BiOBr, and Co3O4/BiOBr-0.8

    The chemical valence states and elemental composition of the synthesized samples were characterized using X-ray photoelectron spectroscopy (XPS). Fig.S2a presents the full XPS spectrum of Co3O4/BiOB-0.8, confirming the presence of Bi, O, Br, and Co elements. As illustrated in Fig.2b, the peaks corresponding to Br3d5/2 and Br3d3/2 exhibited binding energies of 68.2 and 69.2 eV, respectively, in both pure BiOBr and Co3O4/BiOBr-0.8, with no significant shifts observed. In Fig.2c, the characteristic peaks of pure BiOBr at 159.3 and 164.6 eV in the Bi4f spectrum correspond to Bi3+4f7/2 and Bi3+4f5/2, respectively. The peaks at 157.6 eV (Bi4f7/2) and 162.9 eV (Bi4f5/2) associate with Bi(3-x)+ species. For Co3O4/BiOB-0.8, the characteristic peaks at 159.1 and 164.4 eV also correspond to Bi3+4f7/2 and Bi3+4f5/2, while the peaks at 157.8 eV (Bi4f7/2) and 163.1 eV (Bi4f5/2) reflect the presence of Bi(3-x)+ [18]. In Fig.2d, the O1s spectrum of Co3O4 at 531.1 eV can be attributed to the Co—O bond and the adsorption of the H2O group on the surface of the catalyst at 532.9 eV. The O1s spectrum of BiOBr can be fitted with three distinct peaks at 529.6, 530.9, and 532.4 eV, which are associated with the Bi—O bond, oxygen vacancies (OV), and the adsorption of H2O groups, respectively. Notably, for the Co3O4/BiOBr‑0.8 sample, the three peaks at 529.8, 531.6, and 532.5 eV correspond to the M—O (M: metal) bond, OV, and adsorption of H2O groups, respectively. It was observed that the intensity of the M—O peak increased, while the concentration of OV slightly decreased. The high-resolution XPS spectrum of the Co2p (Fig.S2b) exhibited two characteristic peaks, which can be associated with Co2+2p1/2 (797.3 eV), Co2+2p3/2 (781.8 eV), Co3+2p1/2 (795.2 eV), and Co3+2p3/2 (780.0 eV). This indicates the coexistence of Co2+ and Co3+ species in Co3O4. Notably, the Co2p peaks in the Co3O4/BiOBr-0.8 composite exhibited a slight shift towards lower binding energies (Co2+2p1/2: 797.1 eV, Co2+2p3/2: 781.6 eV; Co3+2p1/2: 795.0 eV, Co3+2p3/2: 779.8 eV) compared to those of pure Co3O4, suggesting the enhanced electron density around the cobalt sites[19]. This shift implies the formation of a heterojunction interface that facilitates electron transfer from BiOBr to Co3O4. In summary, BiOBr and Co3O4 are primarily interconnected through Bi—O and Co—O bonds, indicating the formation of robust chemical bonds at the interface of the Co3O4/BiOBr-0.8 heterojunction.

    Electron paramagnetic resonance (EPR) spectroscopy was applied to BiOBr and Co3O4/BiOBr-0.8, as illustrated in Fig.S3. Both samples exhibited a pronounced signal at g=2.002, with the peak intensity of OV decreased with the Co3O4 content combined. This suggests that the Co3O4 content can effectively modulate the concentration of OV in the composites. Furthermore, the variations in OV and Bi(3-x)+ species were successfully characterized. These findings indicate that during heterojunction formation, Bi(3-x)+ ions bind to the oxygen atoms of Co3O4, thereby modulating the electron cloud density and facilitating electron transfer to the Co3O4 phase[5]. This interaction leads to an increased Bi3+/Bi(3-x)+ ratio and a decreased EPR intensity, providing robust evidence for the successful synthesis of the Co3O4/BiOBr-0.8 composites.

    The UV-Vis diffuse reflectance spectra (DRS) of pure Co3O4, pure BiOBr, and Co3O4/BiOBr heterojunctions with different molar ratios are illustrated in Fig.3a. Forming a heterojunction with Co3O4 enhances the visible-light absorbance of BiOBr. Notably, Co3O4/BiOBr-0.8 exhibited a markedly improved light absorption response in the visible spectrum, which can be attributed to the synergistic effect resulting from the intimate coupling between Co3O4 and BiOBr. The band gap energies of pure BiOBr and Co3O4 were estimated by the following Eq.1.

    $ αhν = A(αhν - E_{g})^{n/2} $

    (1)

    Figure 3

    Figure 3.  (a) UV-Vis DRS of different catalysts; (b) Band gaps, (c) Mott-Schottky curves, and (d) VB-XPS spectra of BiOBr and Co3O4

    where α, , A, and Eg represent the absorbance, photon energy, constant, and bandgap energy, respectively[20]. The band gaps of pristine BiOBr and Co3O4 were 3.1 and 1.8 eV, respectively, which are consistent with the values reported in the literature (Fig.3b)[21]. Furthermore, Mott-Schottky curves were used to evaluate the flat band potentials (Efb) of pristine BiOBr and Co3O4. The positive slopes observed in the Mott-Schottky curves indicate that both materials exhibited n-type semiconductor characteristics (Fig.3c). The flat band potentials of pristine BiOBr and Co3O4 were -1.47 and -1.08 V (vs Ag/AgCl) or -0.86 and -0.47 V (vs NHE), respectively. Subsequently, the conduction band (CB) of n-type semiconductors is close to the Efb, so that the CB potentials of BiOBr and Co3O4 were estimated to be -0.86 and -0.47 V vs (NHE), respectively. The VB potential of BiOBr and Co3O4 can be calculated using the Eg and VB energy, which were 2.24 and 1.33 V (vs NHE), respectively. The VB offset relative to the Fermi level for pure BiOBr and Co3O4 can be quantified through VB-XPS analysis (Fig.3d), revealing that the Fermi levels of pure BiOBr and Co3O4 were 0.61 and 0.17 eV, respectively.

    The photocatalytic performance of CO2 reduction was assessed on pure BiOBr, pure Co3O4, and Co3O4/BiOBr composites with different molar ratios under 5 h irradiation (Fig.4a). The results indicated that CO was the primary product of CO2 photocatalytic reduction[22]. Among all catalysts tested, Co3O4/BiOBr-0.8 exhibited the best photocatalytic performance, achieving CO evolution rates of 112.2 μmol·g-1·h-1 and CH4 evolution rates of 5.5 μmol·g-1·h-1 (Fig.S4), which were 6.3-fold and 3.9-fold higher than BiOBr alone. Furthermore, among the various molar ratios of Co3O4/BiOBr heterojunctions, the evolution rates of CO and CH4 showed a trend of rising first and then declining. Co3O4/BiOBr-0.8 demonstrated the highest photocatalytic activity, suggesting an optimal ratio of 0.8∶1 for maximizing the efficiency of the Co3O4/BiOBr heterojunction. Fig.4b illustrates that under experimental conditions such as an Ar atmosphere devoid of CO2, or the absence of H2O, light, or catalyst individually, trace reduction products were detected. This indicates that CO and CH4 production produced by CO2 molecules and confirms that all four factors—CO2, H2O, light, and catalyst—are indispensable for a successful photoreduction reaction. Fig.4c presents a comparative test of the photocatalytic activity for CO2 photoreduction between the Co3O4/BiOBr heterojunction and a mechanical mixture of pure Co3O4 and BiOBr (Co3O4+ 0.8BiOBr). The data demonstrated that the CO productions from the Co3O4/BiOBr heterojunction were markedly higher compared to those from the mechanical mixing, underscoring the significant role of the heterojunction structure in enhancing CO2 photoreduction efficiency[23]. Furthermore, Fig.4d illustrates that the CO evolution of Co3O4/BiOBr‑0.8 was significantly higher than that of BiOBr, with a consistent increase observed over time. Additionally, it is confirmed that the reduction of CO follows first-order reaction kinetics, suggesting that this step is rate-limiting in the overall reaction (CO2+2e-+2H+→CO+H2O). In contrast, the production of CH4 was significantly more challenging due to the requirement for a greater number of electrons and protons (CO2+8e-+8H+→CH4+2H2O). As a result, the production of CO was markedly higher compared to CH4, as the reduction processes leading to these products likely proceed via parallel pathways.

    Figure 4

    Figure 4.  (a) Productions of CO on various catalysts; (b) Productions of CO on Co3O4/BiOBr-0.8 with different reaction conditions; (c) Productions of CO on mechanical mixture and Co3O4/BiOBr-0.8; (d) Productions of CO on Co3O4/BiOBr-0.8 and BiOBr with different irradiation times

    Fig. 5a demonstrates that Co3O4/BiOBr-0.8 exhibited a minor decrease in CO production during the seven cycle reaction processes, yet it maintained overall stability over the entire duration of 28 h. The XRD pattern of the fresh Co3O4/BiOBr-0.8 showed significant similarity to that of the used Co3O4/BiOBr-0.8 after seven cycles of photocatalytic reactions (Fig. 5b). However, a slight attenuation in the peak intensity corresponding to Co3O4 was observed. This indicates that the catalyst possesses remarkable recyclability and stability. To elucidate the direction of electron transfer during the reaction process, XPS analyses were conducted on the catalyst both before and after the cyclic reaction. Notably, as shown in Fig. 5c, the Bi4f spectra exhibited a systematic shift toward lower binding energy, which is indicative of an electron-accumulating state. As illustrated in Fig. 5d, the Co element demonstrated a pronounced shift toward higher binding energy, reflecting electron depletion. Furthermore, the peak intensity of Co3+ exhibited significant attenuation, leading to a marked decline in the Co3+/Co2+ ratio. Collectively, these observations suggest that during the reaction process, electrons are transferred from the CB of Co3O4 to the VB of BiOBr, consistent with the electron transport pathway of the Z-scheme heterojunction.

    Figure 5

    Figure 5.  (a) Stability test of Co3O4/BiOB-0.8; (b) XRD patterns, (c) Bi4f, and (d) Co2p spectra of Co3O4/BiOB-0.8 before and after the stability test

    Fig. 6a illustrates the photoluminescence (PL) spectra of pure BiOBr and Co3O4/BiOBr heterojunctions with different ratios. The PL intensity decreased first and then increased with the increase of Co3O4 amount, suggesting that the recombination rate of electron-hole pairs in BiOBr is significantly suppressed through the incorporation of Co3O4[24]. From the figure, we can see that the most suitable ratio was 0.8∶1. Fig. 6b displays the time-resolved PL decay (TRPL) curves of Co3O4/BiOBr-0.8, pure BiOBr, and pure Co3O4. The TPRL curves can be analyzed using Eq.2:

    $ {\tau }_{a}=\frac{{A}_{1}{\tau }_{1}^{2}+{A}_{2}{\tau }_{2}^{2}}{{A}_{1}{\tau }_{1}+{A}_{2}{\tau }_{2}} $

    (2)

    Figure 6

    Figure 6.  (a) Steady PL spectra, (b) TRPL curves, (c) transient photocurrent density curves, and (d) EIS of pure Co3O4, pure BiOBr, and Co3O4/BiOBr-0.8

    where A1 (the amplitude coefficient of the short-lived component) and A2 (the amplitude coefficient of the long‑lived component) represent the corresponding amplitudes, and τ1 and τ2 denote the fluorescent lifetimes of fast decay and slow decay stages, respectively. For Co3O4/BiOBr‑0.8, τ1=0.83 ns and τ2=3.53 ns, which were longer than the values of pure Co3O4 (τ1=0.73 ns; τ2=2.53 ns) and BiOBr (τ1=0.82 ns; τ2=3.48 ns). Consequently, the average fluorescent lifetimes of Co3O4/BiOBr-0.8 (1.54 ns) were prolonged than pristine Co3O4 (1.37 ns) and BiOBr (1.30 ns) as calculated from Eq.2. The prolonged fluorescence lifetime further substantiates that the heterojunction structure can efficiently separate photogenerated electrons and holes into distinct phases, thereby significantly inhibiting the recombination of photogenerated carriers[25]. Fig.6c illustrates the transient photocurrent responses of pure BiOBr, pure Co3O4, and Co3O4/BiOBr-0.8 under intermittent light illumination. This method is embodied by using an electrochemical method to test the current density of the photosensitive semiconductor under the switching light conversion for assessing the separation efficiency of photogenerated electron-hole pairs. The photocurrent generated by Co3O4/BiOBr-0.8 was notably higher than that of pure BiOBr and pure Co3O4, indicating a significant enhancement and stability in the separation efficiency of photoinduced electron-hole pairs for Co3O4/BiOBr-0.8[26]. Fig.6d presents the electrochemical impedance spectra (EIS) of pure BiOBr, pure Co3O4, and Co3O4/BiOBr-0.8. The arc radius of Co3O4/BiOBr-0.8 was considerably smaller compared to pure BiOBr and Co3O4, suggesting superior interfacial charge transfer efficiency for Co3O4/BiOBr-0.8[27]. Consequently, the Z-scheme heterojunction of Co3O4/BiOBr demonstrates enhanced separation of electron‑hole pairs and efficient electron transport, which aligns well with the experimental findings.

    The photocatalytic intermediates on Co3O4/BiOBr-0.8 were identified using in-situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS). As illustrated in Fig.7, the DRIFTS peaks of Co3O4/BiOBr-0.8 exhibited enhanced intensity with prolonged reaction time, indicating that CO2 molecules were effectively reduced under light irradiation. The characteristic peak at 1 065 cm-1 is attributed to intermediate CHO*, resulting from the reaction of intermediate CO* with H[28]. The peak at 1 156 cm-1 corresponds to intermediate CH3O*, while the peak at 1 395 cm-1 is ascribed to bending vibrations of intermediate CH3*[29]. The peaks at 1 362 and 1 517 cm-1 can be attributed to the vibrational modes associated with monodentate carbonates (m-CO32-). The peaks at 1 626 and 1 651 cm-1 are attributed to bidentate carbonate (b-CO32-) species[30]. The peaks at 1 424 and 1 457 cm-1 are attributed to bicarbonate (HCO3-) produced by co-adsorption of CO2 in H2O vapor[11]. Additionally, the intermediate COOH* species is confirmed by the peaks at 1 544, 1 742, and 1 772 cm-1, which is a crucial intermediate in the photoreduction from CO2 to CO[31]. The characteristic peaks observed at 1 915, 1 962, and 2 081 cm-1 indicate the formation of CO* species during the photocatalytic reduction of CO2[32]. These peaks can be ascribed to intermediates during the photoreduction from CO2 to CO and CH4 with water, the possible reaction pathway is illustrated in Fig.8.

    Figure 7

    Figure 7.  In-situ DRIFTS of Co3O4/BiOBr-0.8

    Figure 8

    Figure 8.  Schematic illustration of the CO2 photoreduction in the presence of the catalyst composites

    * denotes the adsorption state at the surface of the catalyst.

    In summary, the Co3O4/BiOBr heterojunction was synthesized via a one-step solvothermal method. The photocatalytic reduction activity of CO2 was comprehensively examined. It was observed that the Co3O4/BiOBr heterojunction exhibited substantially elevated productions of CO during the CO2 photoreduction compared to pure BiOBr and Co3O4. Specifically, the evolution rate of CO for the Co3O4/BiOBr-0.8 heterojunction was 6.3 times higher than that of pristine BiOBr. The experimental results demonstrate that the Co3O4/BiOBr heterojunction exhibits a Z-scheme band structure, thereby markedly improving light utilization, improving electron transport between the two phases, and restraining the recombination of photogenerated carriers. This structure not only maintains superior redox capabilities but also significantly enhances the selectivity for CO production. This study provides valuable insights into the synthesis of highly selective Z-scheme heterojunction photocatalysts for efficient CO2 photoreduction.


    Supporting information is available at http://www.wjhxxb.cn
    1. [1]

      LI X, JIANG H P, MA C, ZHU Z, SONG X H, WANG H Q, HUO P W, LI X Y. Local surface plasma resonance effect enhanced Z-scheme ZnO/Au/g-C3N4 film photocatalyst for reduction of CO2 to CO[J]. Appl. Catal. B‒Environ., 2021, 283: 119638 doi: 10.1016/j.apcatb.2020.119638

    2. [2]

      PATIAL S, KUMAR R, RAIZADA P, SINGH P, LE Q V, LICHTFOUSE E, LE TRI NGUYEN D, NGUYEN V H. Boosting light-driven CO2 reduction into solar fuels: Mainstream avenues for engineering ZnO-based photocatalysts[J]. Environ. Res., 2021, 197: 111134 doi: 10.1016/j.envres.2021.111134

    3. [3]

      ZHAO J L, MIAO Z R, ZHANG Y F, WEN G Y, LIU L H, WANG X X, CAO X Z, WANG B Y. Oxygen vacancy-rich hierarchical BiOBr hollow microspheres with dramatic CO2 photoreduction activity[J]. J. Colloid Interface Sci., 2021, 593: 231-243 doi: 10.1016/j.jcis.2021.02.117

    4. [4]

      REN X J, GAO M C, ZHANG Y F, ZHANG Z Z, CAO X Z, WANG B Y, WANG X X. Photocatalytic reduction of CO2 on BiOX: Effect of halogen element type and surface oxygen vacancy mediated mechanism[J]. Appl. Catal. B‒Environ., 2020, 274: 119063 doi: 10.1016/j.apcatb.2020.119063

    5. [5]

      MIAO Z R, WANG Q L, ZHANG Y F, MENG L P, WANG X X. In situ construction of S-scheme AgBr/BiOBr heterojunction with surface oxygen vacancy for boosting photocatalytic CO2 reduction with H2O[J]. Appl. Catal. B‒Environ. Energy, 2022, 301: 120802 doi: 10.1016/j.apcatb.2021.120802

    6. [6]

      CHENG C, HE B W, FAN J J, CHENG B, CAO S W, YU J G. An inorganic/organic S-scheme heterojunction H2-production photocatalyst and its charge transfer mechanism[J]. Adv. Mater., 2021, 33(22): 2100317 doi: 10.1002/adma.202100317

    7. [7]

      LI X, LIU C Y, WU D Y, LI J Z, HUO P W, WANG H Q. Improved charge transfer by size-dependent plasmonic Au on C3N4 for efficient photocatalytic oxidation of RhB and CO2 reduction[J]. Chin. J. Catal., 2019, 40(6): 928-939 doi: 10.1016/S1872-2067(19)63347-4

    8. [8]

      HIRAGOND C B, LEE J, KIM H, JUNG J W, CHO C H, IN S I. A novel N-doped graphene oxide enfolded reduced titania for highly stable and selective gas-phase photocatalytic CO2 reduction into CH4: An in-depth study on the interfacial charge transfer mechanism[J]. Chem. Eng. J., 2021, 416: 127978 doi: 10.1016/j.cej.2020.127978

    9. [9]

      XIONG J, LI X B, HUANG J T, GAO X M, CHEN Z, LIU J Y, LI H, KANG B B, YAO W Q, ZHU Y F. CN/rGO@BPQDs high-low junctions with stretching spatial charge separation ability for photocatalytic degradation and H2O2 production[J]. Appl. Catal. B‒Environ., 2020, 266: 118602 doi: 10.1016/j.apcatb.2020.118602

    10. [10]

      LIANG T, YU Z B, BIN Y J, ZHANG S M, WEI J W, LIU Y J, ZHU T T, FAN S Y, SHEN Y X, WANG S F, HOU Y P. Tungsten and oxygen dual vacancies regulation of the S-scheme ZnSe/ZnWO4 heterojunction with local polarization electric field for efficient CO2 photocatalytic reduction[J]. Chem. Eng. J., 2024, 479: 147942 doi: 10.1016/j.cej.2023.147942

    11. [11]

      ZHU Q L, HUANG W X, SHEN J H, JIANG H B, ZHU Y H, LI C Z. In-situ preparation of BiOBr/Bi-doped CsPbBr3 S-scheme heterojunction for efficient photocatalytic CO2 reduction[J]. Chem. Eng. J., 2024, 499: 156663 doi: 10.1016/j.cej.2024.156663

    12. [12]

      WANG G J, TANG Z W, WANG J, LV S S, XIANG Y J, LI F, LIU C. Energy band engineering of Bi2O2.33CdS direct Z-scheme heterojunction for enhanced photocatalytic reduction of CO2[J]. J. Mater. Sci. Technol., 2022, 111: 17-27 doi: 10.1016/j.jmst.2021.09.018

    13. [13]

      WANG J J, HUANG L, SUN B J, ZHANG H F, HOU D F, QIAO X Q, MA H J, LI D S. Efficient photothermal catalytic CO2 reduction over in situ construction ZnIn2S4@Ni(OH)2/NiO Z-scheme heterojunction[J]. Chem. Eng. J., 2024, 479: 147719 doi: 10.1016/j.cej.2023.147719

    14. [14]

      WANG L, CHEN R J, ZHANG Z Q, CHEN X R, DING J, ZHANG J F, WAN H, GUAN G F. Constructing direct Z-scheme heterojunction g-C3N5/BiOBr for efficient photocatalytic CO2 reduction with H2O[J]. J. Environ. Chem. Eng., 2023, 11(2): 109345 doi: 10.1016/j.jece.2023.109345

    15. [15]

      ZHANG B F, ZHANG M T, ZHANG L, BINGHANM P A, LI W, KUBUKI S. PVP surfactant-modified flower-like BiOBr with tunable bandgap structure for efficient photocatalytic decontamination of pollutants[J]. Appl. Surf. Sci., 2020, 530: 147233 doi: 10.1016/j.apsusc.2020.147233

    16. [16]

      ZHANG Q, YANG P J, ZHANG H X, ZHAO J H, SHI H, HUANG Y M, YANG H Q. Oxygen vacancies in Co3O4 promote CO2 photoreduction[J]. Appl. Catal. B‒Environ., 2022, 300: 120729 doi: 10.1016/j.apcatb.2021.120729

    17. [17]

      WU X L, NG Y H, WANG L, DU Y, DOU S X, AMAL R, SCOTT J. Improving the photo-oxidative capability of BiOBr via crystal facet engineering[J]. J. Mater. Chem. A, 2017, 5(17): 8117-8124 doi: 10.1039/C6TA10964K

    18. [18]

      SHEN M T, ZHU X Y, LIN L W, LI H, WANG Y N, LIANG Q, ZHOU M, LI Z Y, XU S. MOFs-derived S-scheme ZnO/BiOBr heterojunction with rich oxygen vacancy for boosting photocatalytic CO2 reduction[J]. Sep. Purif. Technol., 2025, 353: 128620 doi: 10.1016/j.seppur.2024.128620

    19. [19]

      LIU H B, QIU Y B, GAN W X, ZHUANG G X, CHEN F F, YANG C K, YU Y. MOF-derived Co3O4/ZrO2 mesoporous octahedrons with optimized charge transfer and intermediate conversion for efficient CO2 photoreduction[J]. Sci. China Mater., 2024, 67(2): 588-597 doi: 10.1007/s40843-023-2707-3

    20. [20]

      GUO J G, LIU Y, HAO Y J, LI Y L, WANG X J, LIU R H, LI F T. Comparison of importance between separation efficiency and valence band position: The case of heterostructured Bi3O4Br/α-Bi2O3 photocatalysts[J]. Appl. Catal. B‒Environ., 2018, 224: 841-853 doi: 10.1016/j.apcatb.2017.11.046

    21. [21]

      JIA T, WU J, SONG J, LIU Q Z, WANG J M, QI Y F, HE P, QI X M, YANG L T, ZHAO P C. In situ self-growing 3D hierarchical BiOBr/BiOIO3 Z-scheme heterojunction with rich oxygen vacancies and iodine ions as carriers transfer dual-channels for enhanced photocatalytic activity[J]. Chem. Eng. J., 2020, 396: 125258 doi: 10.1016/j.cej.2020.125258

    22. [22]

      FU J W, JIANG K X, QIU X Q, YU J G, LIU M. Product selectivity of photocatalytic CO2 reduction reactions[J]. Mater. Today, 2020, 32: 222-243 doi: 10.1016/j.mattod.2019.06.009

    23. [23]

      HUANG Q W, TIAN S Q, ZENG D W, WANG X X, SONG W L, LI Y Y, XIAO W, XIE C S. Enhanced photocatalytic activity of chemically bonded TiO2/graphene composites based on the effective interfacial charge transfer through the C—Ti bond[J]. ACS Catal., 2013, 3(7): 1477-1485 doi: 10.1021/cs400080w

    24. [24]

      JIA Y F, ZHANG W B, DO J Y, KANG M, LIU C L. Z-scheme SnFe2O4/α-Fe2O3 micro-octahedron with intimated interface for photocatalytic CO2 reduction[J]. Chem. Eng. J., 2020, 402: 126193 doi: 10.1016/j.cej.2020.126193

    25. [25]

      QIAN Z R, ZHANG L, ZHANG Y F, CUI H. Synergistically boosting of CO2 photoreduction over Bi/BiOBr nanostructure via in-situ formation of oxygen vacancy and metallic Bi[J]. Sep. Purif. Technol., 2023, 324: 124581 doi: 10.1016/j.seppur.2023.124581

    26. [26]

      BI W, HU Y J, JIANG N, ZHANG L, JIANG H, ZHAO X, WANG C Y, LI C Z. Ultra-fast construction of plaque-like Li2TiO3/TiO2 heterostructure for efficient gas-solid phase CO2 photoreduction[J]. Appl. Catal. B‒Environ., 2020, 269: 118810 doi: 10.1016/j.apcatb.2020.118810

    27. [27]

      LIU T X, YANG K C, GONG H M, JIN Z L. Visible-light driven S-scheme Mn0.2Cd0.8S/CoTiO3 heterojunction for photocatalytic hydrogen evolution[J]. Renew. Energy, 2021, 173: 389-400 doi: 10.1016/j.renene.2021.03.146

    28. [28]

      GUAN C S, HOU T, NIE W Y, ZHANG Q, DUAN L B, ZHAO X R. Enhanced photocatalytic reduction of CO2 on BiOBr under synergistic effect of Zn doping and induced oxygen vacancy generation[J]. J. Colloid Interface Sci., 2023, 633: 177-88 doi: 10.1016/j.jcis.2022.11.106

    29. [29]

      NI M M, ZHU Y J, GUO C F, CHEN D L, NING J Q, ZHONG Y J, HU Y. Efficient visible-light-driven CO2 methanation with self- regenerated oxygen vacancies in Co3O4/NiCo2O4 hetero-nanocages: Vacancy-mediated selective photocatalysis[J]. ACS Catal., 2023, 13(4): 2502-2512 doi: 10.1021/acscatal.2c05577

    30. [30]

      FU H, LIU X L, WU Y Q, ZHANG Q Q, WANG Z Y, ZHENG Z K, CHENG H F, LIU Y Y, DAI Y, HUANG B B, WANG P. Construction of a bismuth-based perovskite direct Z-scheme heterojunction Au‑Cs3Bi2Br9/V2O5 for efficient photocatalytic CO2 reduction[J]. Appl. Surf. Sci., 2023, 622: 156964 doi: 10.1016/j.apsusc.2023.156964

    31. [31]

      XU S S, JIANG G C, ZHANG H K, GAO C Y, CHEN Z H, LIU Z H X, WANG J, DU J, CAI B, LI Z. Boosting photocatalytic CO2 methanation through interface fusion over CdS quantum dot aerogels[J]. Small, 2024, 20(38): 2400769

    32. [32]

      SONG Y X, LI X M, LI H, WANG L, XIAO S N, FEI H H, LI G S, SONG X L. Boosted photocatalytic CO2 conversion of a Cs2AgBiBr6@Co3O4 composite with high activity and selectivity under low-concentration CO2 and natural sunlight[J]. Appl. Catal. B‒Environ. Energy, 2025, 363: 124816 doi: 10.1016/j.apcatb.2024.124816

  • Figure 1  (a, b) SEM images, (c) TEM image, (d) HRTEM image, and (e, f) elemental mapping images of Co3O4/BiOBr-0.8

    Figure 2  (a) XRD patterns of Co3O4/BiOBr; (b) Br3d, (c) Bi4f, and (d) O1s XPS spectra of Co3O4, BiOBr, and Co3O4/BiOBr-0.8

    Figure 3  (a) UV-Vis DRS of different catalysts; (b) Band gaps, (c) Mott-Schottky curves, and (d) VB-XPS spectra of BiOBr and Co3O4

    Figure 4  (a) Productions of CO on various catalysts; (b) Productions of CO on Co3O4/BiOBr-0.8 with different reaction conditions; (c) Productions of CO on mechanical mixture and Co3O4/BiOBr-0.8; (d) Productions of CO on Co3O4/BiOBr-0.8 and BiOBr with different irradiation times

    Figure 5  (a) Stability test of Co3O4/BiOB-0.8; (b) XRD patterns, (c) Bi4f, and (d) Co2p spectra of Co3O4/BiOB-0.8 before and after the stability test

    Figure 6  (a) Steady PL spectra, (b) TRPL curves, (c) transient photocurrent density curves, and (d) EIS of pure Co3O4, pure BiOBr, and Co3O4/BiOBr-0.8

    Figure 7  In-situ DRIFTS of Co3O4/BiOBr-0.8

    Figure 8  Schematic illustration of the CO2 photoreduction in the presence of the catalyst composites

    * denotes the adsorption state at the surface of the catalyst.

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  132
  • HTML全文浏览量:  9
文章相关
  • 发布日期:  2025-09-10
  • 收稿日期:  2025-03-11
  • 修回日期:  2025-07-10
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

  1. 本站搜索
  2. 百度学术搜索
  3. 万方数据库搜索
  4. CNKI搜索

/

返回文章