DOI:
10.1039/D5RA06653K
(Paper)
RSC Adv., 2025,
15, 45038-45047
Fabrication of CuO/BiVO4 composites for enhanced visible-light-driven photocatalytic antibacterial activity
Received
4th September 2025
, Accepted 6th November 2025
First published on 18th November 2025
Abstract
In recent years, photocatalytic antibacterial technology has attracted wide attention due to its advantages of broad-spectrum antibacterial activity, high stability, safety and non-toxicity. Bismuth vanadate (BiVO4) reveals a strong response to visible light, however, the photocatalytic activity of pure BiVO4 remains unsatisfactory. In the present work, CuO/BiVO4 composites were successfully constructed, and the photocatalytic performance and antibacterial mechanism were systematically studied. The antibacterial results confirmed that the CuO/BiVO4 composites exhibited enhanced photocatalytic antibacterial activity against Escherichia coli (E. coli). With the CuO addition of 12.5%, the CuO/BiVO4 composites presented superior antibacterial performance, and the antibacterial rate reached 100% under visible light irradiation for 30 min, while the antibacterial rate for BiVO4 was less than 20% under the same conditions. In addition, CuO/BiVO4 composites displayed a long-term effect, and the antibacterial rate was kept at >90% after 5 cycles. The antibacterial mechanism was mainly from ROS oxidative damage, in which ·O2− played a major role in antibacterial activity.
1. Introduction
The proliferation of bacteria is becoming increasingly serious and constantly threatening human health and life, and therefore, it is urgent to address the bacterial contamination. Photocatalysis, as an economical, efficient, and environmentally friendly method, has attracted increasing attention in various research fields over the past decades, such as in photocatalytic degradation of pollutants, water splitting, and photocatalytic antibacterial applications. When a photocatalyst absorbs light, it generates e−–h+ pairs, which react with water and dissolved oxygen to form reactive oxygen species (ROS),1,2 such as hydroxyl radicals (·OH), singlet oxygen (1O2) and superoxide radicals (·O2−). These highly oxidative reactive species can disinfect pathogens by destroying essential macromolecules within bacteria.3 It has been reported that TiO2 exhibits photocatalytic activity against several types of microorganisms, including bacteria and viruses.4 However, TiO2 only harvests ultraviolet light (L <390 nm, accounting for only 4% of total sunlight) because of the large bandgap energy.5,6
BiVO4 is a non-toxic, corrosion-resistant, highly stable, and environmentally friendly photocatalyst with significant application prospects. Since Kudo et al. first reported the photocatalytic water splitting of BiVO4 under visible light in 1998,7 it has attracted much attention in the fields of photocatalytic degradation of organic pollutants, hydrogen evolution from water splitting, and antibacterial applications.8–14 There are three different crystal phases for BiVO4, tetragonal zircon (t-z), tetragonal scheelite (t-S), and monoclinic scheelite (m-S).15,16 Given that the monoclinic scheelite phase (m-BiVO4) is the most thermodynamically stable and has demonstrated the highest photocatalytic activity (e.g., for visible-light degradation of pollutants and water splitting), it has attracted widespread research interest. Consequently, we focus herein on the crystal structure, electronic structure, and optical characteristics of m-BiVO4 to elucidate their influence on its photoelectrochemical properties. The conduction band of monoclinic BiVO4 is mainly composed of V 3d orbitals, and the valence band is a hybrid of Bi 6s and O 2s orbitals. The hybrid orbitals can reduce the bandgap, making BiVO4 broader visible light absorption region and suitable energy level positions as a visible-light-driven photocatalyst. However, the rapid recombination of photogenerated charge carriers and poor surface adsorption capacity greatly limits the practical application of pure BiVO4.17–19
In order to improve the photocatalytic performance, heterojunctions have been constructed, such as Cu/BiVO4,20 Ag/BiVO4,21 WO3/BiVO4,22 V2O5/BiVO4,23 CeO2/BiVO4,24 Bi2O3/BiVO4,25 etc., which have demonstrated the effectiveness of nano-composite methods for the photocatalytic activity improvement. Jiang prepared CuO/BiVO4 composite photocatalyst and found that the composite material showed the best photocatalytic performance when the loading amount of CuO was 2 wt%, achieving a methylene blue (MB) degradation rate of 47% in 180 min.26 Zhang prepared a new type of CuO–BiVO4 heterojunction composite material (CuO–BiVO4/FACS) by metal–organic decomposition impregnation method, and found that the composite material showed high photocatalytic activity for the degradation of MB under visible light irradiation, with a degradation rate up to 92% in 180 min.27 Zhao prepared carbon-loaded CuO–BiVO4 (CuO–BiVO4@C) composite photocatalytic materials by hydrothermal method, and CuO–BiVO4@1.0C showed the highest photocatalytic degradation activity for MB, with a degradation rate up to 100% in 2.5 h.28 Therefore, incorporating CuO into BiVO4 can effectively enhance the photocatalytic activity of pure BiVO4. However, to our knowledge, most CuO–BiVO4 composite materials have been studied for pollutant degradation, and there is still little research in the field of antibacterial applications. And there is a lack of in-depth analysis regarding the antibacterial mechanism of the CuO–BiVO4 composite materials.
Nano CuO is a low-cost, highly reactive broad-spectrum antibacterial material with advantages such as good heat resistance and stability. As an inorganic antibacterial material, it has good antibacterial performance without drug resistance, therefore, copper-based antibacterial materials have broad research and practical application prospects. It has been reported that needle-shaped nano CuO has strong antibacterial effects on Escherichia coli (E. coli) and Staphylococcus aureus (S. aureus).29 Ran immobilized CuO/BiVO4 photocatalytic materials on cotton fabric through a polydopamine template, achieving efficient visible light-driven photocatalysis, antibacterial, and ultraviolet protection applications.30 In this paper, a series of CuO/BiVO4 composite materials with different amounts of CuO loading were prepared by the hydrothermal-impregnation method. CuO/BiVO4 composite materials enhanced the visible-light response and reactive oxygen species (ROS) generation of BiVO4 for the photocatalytic antibacterial applications. Furthermore, the analysis of the antibacterial contribution of the CuO/BiVO4 composite materials successfully confirmed that photogenerated ROS serve as the primary mechanism, with superoxide anions (·O2−) and hydroxyl radicals (·OH) identified as the dominant reactive species. The process of bacterial inactivation by ROS was also elucidated.
2. Materials and methods
2.1 Materials
All reagents used in the experiments were of analytical grade, and all experiments were conducted with deionized water. All instruments used in the experiments were sterilized with an autoclave before use. Bismuth nitrate pentahydrate (Bi(NO3)3·5H2O) was obtained from Xilong Scientific Co., Ltd. Ammonium metavanadate (NH4VO3) was obtained from Tianjin Damao Chemical Reagent Factory (China). Dilute nitric acid, ammonia water, sodium dihydrogen phosphate and potassium dihydrogen phosphate were obtained from Tianjin Damao Chemical Reagent Factory (China). All biological reagents were purchased from Beijing Aoboxing Biotechnology Co., Ltd (China).
2.2 Synthesis of CuO/BiVO4 composite materials
Synthesis of BiVO4: 17.46 g of Bi(NO3)3·5H2O were dissolved in 80 mL of 2 mol per L dilute HNO3 to obtain a colorless transparent solution A. Then, 4.212 g of NH4VO3 were dissolved in 220 mL of 2 mol per L dilute HNO3 to obtain a yellow solution B. Solution A and B were mixed uniformly to obtain a yellow transparent solution and were stirred continuously for 0.5 h, then the pH was adjusted to 2 with ammonia to obtain an orange precipitate. After continue stirred at a constant speed for 1 h, the precipitate was aged at room temperature for 2 h. Then, the supernatant was poured off and about 80 mL of the slurry retained was transferred into a 100 mL Teflon-lined stainless steel autoclave, and heated at 200 °C for 24 h. After cooled to room temperature, the precipitate was washed three times with deionized water, dried under vacuum at 80 °C overnight, and then calcined in a muffle furnace at 400 °C for 2 h.
CuO/BiVO4 composite materials were prepared by impregnation-calcination method. Firstly, Cu(NO3)2 was dispersed in 20 mL of deionized water, and then the as-prepared BiVO4 was added to obtain a suspension. The suspension was stirred in a water bath at 80 °C until all the water evaporated. Finally, the remained powder was placed in a muffle furnace and calcined at 300 °C for 1 h. A series of composite materials were obtained by changing the amount of CuO loaded. The prepared composite materials were named as CuO/BiVO4-x% (x is the mass fraction of CuO in the CuO/BiVO4 composite material; x = 10, 12.5, 16.7, and 20).
2.3 Characterization
XRD measurement was conducted by Rigaku Ultima IV X-ray diffractometer from Japan. Copper target Cu-Kα radiation was used with wavelength of 1.5406 Å, divergence slit of 0.19 mm, tube voltage of 40 kV, and tube current of 40 mA. The scanning range was 10–80°, and the continuous scanning speed was 5° min−1. The surface morphology of samples was observed by Zeiss Gemini 300 scanning electron microscope from Germany. Before test, the samples were sputter-coated with gold to enhance the conductivity. The acceleration voltage was 5 kV, and the test mode was secondary electrons. XPS was performed by Shimadzu/Krayos AXIS Ultra DLD X-ray photoelectron spectrometer from Japan. Al-Kα (hν = 1486.6 eV) was used as the radiation source, and the power and pass energy of the spectrometer analyzer were 150 W and 50 eV, respectively. The electron binding energy was corrected based on the C 1s (284.8 eV) of the sample. The ultraviolet-visible spectra of the samples were tested by Shimadzu UV-3600i Plus spectrometer from Japan with the integrating sphere test mode. The wavelength range was 200–800 nm, and the diffuse reflectance (reflectance R%) data mode was selected. The copper ion concentration in the solution was tested by Agilent ICP-OES 725 ES with radio frequency power of 1.20 kW, flow rate of 15.0 L min−1, auxiliary flow rate of 1.50 L min−1, nebulizer flow rate of 0.75 L min−1, sampling delay of 10 s, replicate reading time of 15 s, and the number of repetitions of 3.
2.4 Photocatalytic activity test
2.4.1 Inactivation of E. coli. The antibacterial performance of the CuO/BiVO4 composite materials was qualitatively measured by the colony counting method. 300 W xenon lamp with an ultraviolet light cutoff filter (L/420 nm) was used as the light source in the photocatalytic antibacterial experiment. The system temperature was maintained at around 25 °C during the antibacterial experiment by water cooling. All instruments and solutions required for the experiment were sterilized with high pressure, and the ultraviolet lamp in the sterile operation table was kept on 30 min before the experiment started. Firstly, 0.03 mg of CuO/BiVO4 composite material and 30 mL of sterilized PBS buffer solution were added to a 250 mL beaker, followed by the addition of E. coli suspension (106 cfu mL−1). Before the photocatalytic antibacterial experiment, the CuO/BiVO4 composite material and bacterial suspension were stirred in the dark for 30 min. During the experiment, 0.1 mL of bacterial suspension was taken out every 10 min and transferred to the sterile operation table, diluted to an appropriate concentration, and then a suitable amount of suspension was dropped onto a Petri dish containing LB solid medium and spread.After incubating the Petri dishes upside down in a constant temperature incubator for 24 h, the number of living cells (cfu) was counted. Dark control (no light) and blank control (no photocatalyst) experiments, as well as copper ion control experiments, were also conducted. All experiments were performed in triplicate, and the average values were given. The antibacterial rate X was calculated by eqn (1).
| |
 | (1) |
In the antibacterial experiments,
N0 and
Nt represent the number of viable cells in the blank control without CuO/BiVO
4 and that after the photocatalytic reaction with the addition of CuO/BiVO
4, respectively.
2.4.2 Degradation of Rhodamine B. The photocatalytic degradation experiment of the CuO/BiVO4 composite materials was conducted under room temperature and visible light condition. A 300 W xenon lamp with an ultraviolet light cutoff filter (L/420 nm) was used as the light source. 100 mL of Rhodamine B solution (concentration of 10 mg L−1) was placed in a dry and clean beaker. The initial absorbance A0 was measured at wavelength of 554 nm. Then, the weighed CuO/BiVO4 composite materials were added to the beaker. After stirred in dark for 30 min, the absorbance was measured and recorded. Subsequently, the light source was turned on to initiate the photodegradation experiment. During the experiment, the absorbance of the solution was measured every 60 min. The absorbance recorded at different times was denoted as At. The degradation efficiency η was calculated by eqn (2):| |
 | (2) |
2.5 Determination of reactive oxygen species (ROS)
To analyze the free radicals, three reagents, superoxide dismutase (SOD), catalase (CAT), and D-mannitol were used to scavenge the ·O2−, H2O2, and ·OH radicals produced in the antibacterial process, respectively.31 E. coli bacterial solution (106 cfu mL−1) and CuO/BiVO4 composite materials (1000 µg mL−1) were placed into the beakers (Groups A, B, C), and then 300 µL of superoxide dismutase (SOD, 100 unit mL−1), 300 µL of catalase (CAT, 100 unit mL−1), and 300 µL of D-mannitol (10 mM) scavengers were added to the three beakers, respectively. The conditions and steps of the photocatalytic antibacterial experiment were the same as those in previous experiments. At the same time, control experiments with the scavengers alone in bacterial suspensions under visible light irradiation were also conducted. The influence of individual ROS species on the antibacterial performance of the CuO/BiVO4 composite material was analyzed by the colony counting method.
3. Results and discussion
3.1 Characterization of CuO/BiVO4 composite materials
To investigate the influence of CuO addition on the BiVO4 structure, the crystal structure of BiVO4 and CuO/BiVO4 composite materials was characterized by XRD. As shown in Fig. 1A, all the main characteristic peaks of pure BiVO4 matched with the monoclinic BiVO4 (m-BiVO4, JCPDS No. 14-0133), indicating the successful synthesis of monoclinic BiVO4. For the CuO/BiVO4 composite materials, the main characteristic peaks corresponded to those of pure BiVO4, and additional peaks observed at 35.6° and 38.7° were attributed to CuO (JCPDS No. 72-0629). The XRD diffraction peaks of pure CuO sample (Fig S1 in the SI) were consistent with those identified for CuO in the CuO/BiVO4 composite patterns presented in Fig. 1A, indicating the successful preparation of CuO/BiVO4 composite materials.
 |
| | Fig. 1 XRD patterns and XPS spectra of CuO/BiVO4 composite materials ((A) XRD; (B) total spectra; (C) V 2p; (D) Bi 4f; (E) Cu 2p; (F) O 1s). | |
Further analysis of the elemental valence states and chemical composition of the CuO/BiVO4 composite materials was conducted by XPS. As shown in Fig. 1B, signals of Cu, V, Bi, O, and C were present in the composite materials. Fig. 1C–F were the high-resolution XPS spectra of V 2p, Bi 4f, Cu 2p, and O 1s, respectively. The high-resolution XPS spectra of V and Bi revealed that BiVO4 and CuO/BiVO4 were characterized by spin–orbit doublets. In the V 2p region (Fig. 1C), the peaks at around 516.5 eV and 524.0 eV were assigned to V 2p3/2 and V 2p1/2, respectively, signifying the V5+ oxidation state.32 In the Bi 4f region (Fig. 1D), the peaks located at approximately 159.0 eV and 164.1 eV were attributed to Bi 4f7/2 and Bi 4f5/2, respectively, characteristic of Bi3+.34 The XPS results confirmed the consistent oxidation states of the metal cations in materials. XPS analysis revealed that the Bi and V peaks in the CuO/BiVO4 composite shifted towards lower binding energies compared to the pristine BiVO4 sample. The peak corresponding to the Cu 2p3/2 orbital in Fig. 1E was deconvoluted, and the results suggested that there were Cu2+ and Cu+ in the composite materials.33 In Fig. 1F, O 1s binding energy was located at the range of 526.0 eV to 535.0 eV. The deconvolution results indicated that there were three types of oxygen in the composite materials, lattice oxygen, oxygen vacancy and adsorbed oxygen. Furthermore, the deconvolution of the O 1s spectra showed a notable increase in the proportion of oxygen vacancies, suggesting a higher concentration of surface oxygen vacancies in the CuO/BiVO4 composite materials.
SEM was conducted to explore the morphology of pure BiVO4 and CuO/BiVO4 composite materials. As shown in Fig. 2A, pure BiVO4 exhibited a relatively regular dodecahedral structure, with the length of 1.2 to 2.4 µm. The presence of some irregular structures may be due to incomplete growth of material. Fig. 2B and S2 in the SI displayed the morphology of the samples with CuO content increasing from 10 to 20%. The composite materials exhibited a relatively regular spherical appearance, which might originate from the recrystallization of BiVO4 in the acidic Cu(NO3)2 solution (pH of 3–4). Furthermore, EDS mapping confirmed the uniform distribution of Bi, V, Cu, and O elements on the surface of the composite materials (Fig. 2C). In order to determine the internal information of the particles, cross section polisher was used to cut the CuO/BiVO4 for further analysis by EDS. The cross-sectional SEM image (Fig. 2D) of the CuO/BiVO4 confirmed that there was no distinct interface between CuO and BiVO4 within the composite. Meanwhile, the cross-sectional EDS mapping (Fig. 2E) verified the uniform distribution of the Bi, V, Cu, and O elements inside the CuO/BiVO4 composite. Additionally, small CuO nanoparticles could be observed on the surface of the CuO/BiVO4 composite materials by TEM image (inset of Fig. 2F). The HRTEM image in Fig. 2F provided detailed investigation of the material's microstructure. Two types of lattice fringes with spacings of 0.26 nm and 0.19 nm were clearly observed, which could be assigned to the (111) plane of CuO and the (060) plane of BiVO4, respectively.
 |
| | Fig. 2 SEM/TEM images and EDS of CuO/BiVO4 composite materials ((A) pure BiVO4; (B) CuO/BiVO4-12.5%; (C) EDS mapping results; (D) cross-sectional SEM; (E) cross-sectional EDS mapping; (F) TEM and HRTEM). | |
3.2 Photocatalytic antibacterial activity
The antibacterial performance of CuO/BiVO4 composite materials with CuO loading of 10%, 12.5%, 16.7%, and 20%, respectively, was evaluated by inactivation of E. coli under visible light. As shown in Fig. 3A, under visible light irradiation, CuO/BiVO4 composite materials exhibited good inactivation effects on E. coli, with antibacterial rate of 100% within 60 min. In contrast, the inactivation capability of pure BiVO4 was inferior, with an antibacterial rate of only 26% within 60 min. With the CuO content increased from 10% to 20%, the antibacterial rate of CuO/BiVO4 was 55%, 100%, 92%, and 61%, respectively, within 30 min. Therefore, 12.5% was selected as the optimal loading amount to prepare the CuO/BiVO4 composite materials. The antibacterial activity of CuO/BiVO4 in this study was comparable to that of recently reported BiVO4 based composite materials against E. coli, as shown in Table S1 in the SI. To exclude the influence of other factors on the antibacterial property, control experiments were conducted and the results were shown in Fig. 3B. In the dark control (with CuO/BiVO4) and blank control (light without CuO/BiVO4) experiments, the E. coli grew naturally, which indicated that the CuO/BiVO4 composite materials had no antibacterial ability in the absence of light, and the influence of visible light on bacteria was negligible. To exclude the influence of CuO on antibacterial performance, antibacterial tests were conducted on pure CuO. The results (Fig. S3 in the SI) showed that after 60 min of light exposure, the antibacterial rate of pure CuO was approximately 32%, indicating a negligible impact on antibacterial performance. Based on the ICP test, the concentration of Cu2+ released in solution was 1.35 mg L−1 after 60 min of illumination, and the inactivation effect of Cu2+ was insignificant.
 |
| | Fig. 3 Antibacterial test and control experiment results of CuO/BiVO4 composite materials ((A) antibacterial test results; (B) control experiment results; (C) schematic diagram of antibacterial results). | |
Fluorescence microscopy can clearly observe the apoptosis of E. coli during the experimental process. According to the reagent manual, live bacterial cells stained with acridine orange (AO) reagent exhibit green fluorescence, while damaged or dead bacterial cells stained with propidium iodide (PI) reagent exhibit red fluorescence. As shown in Fig. 4, bacteria and composite material kept in dark for 30 min showed high-intensity green fluorescence, which was the same as the negative control (E. coli growing naturally), indicating that the dark adsorption process at the beginning of the experiment did not affect the growth of bacteria. While, all bacteria exhibited red fluorescence after 30 min of illumination, indicating that the composite material has completely inactivated the E. coli within the system.
 |
| | Fig. 4 Fluorescence micrograph of CuO/BiVO4-12.5% for E. coli inactivation (A) negative control (B) dark for 30 min (C) light for 30 min (D) positive control. | |
The cyclic and long-term antibacterial performance of the CuO/BiVO4-12.5% composite material against E. coli was presented in Fig. 5. After 5 cycles of antibacterial experiments, the antibacterial rate remained >90%, indicating that the composite material maintained good antibacterial activity and demonstrated excellent reusability and stability. The antibacterial rate of the composite material stored for 180 days was kept >90%, showing good long-term effectiveness. In summary, the CuO/BiVO4 composite material exhibited good reusability and stability, making it a promising visible light photocatalytic composite with practical application prospects.
 |
| | Fig. 5 Cyclic and long-term experiments. | |
To further explore the photocatalytic activity of the CuO/BiVO4 composite material, Rhodamine B was used as a model dye for degradation experiments. The initial concentration of Rhodamine B was 10 mg L−1 and the concentration of composite material was 1.0 g L−1. As shown in Fig. 6A, CuO/BiVO4 composite materials demonstrated good photocatalytic degradation performance under visible light irradiation, with much higher degradation rate than that of pure BiVO4. And the CuO/BiVO4-12.5% composite material showed the best photocatalytic degradation performance, achieving a degradation rate of 70% within 240 min, while the photocatalytic degradation performance of pure BiVO4 was relatively inferior, with a degradation rate of only 19%. The results indicated that the CuO loading could significantly enhance the photocatalytic activity of BiVO4. Fig. 6B is the first-order kinetic fitting curves for the degradation of Rhodamine B by the CuO/BiVO4 composite materials, and the R2 value (as shown in Table S2 in the SI) was close to 1, indicating a high degree of fitting. The slope of CuO/BiVO4-12.5% composite material was the largest, indicating the largest reaction constant and the highest photocatalytic activity. The dye degradation results further indicated that the significantly enhanced photocatalytic performance of CuO/BiVO4 composite material was originated from CuO loading.
 |
| | Fig. 6 Photocatalytic degradation of Rhodamine B under visible light irradiation: (A) photodegradation rate; (B) dynamics curve. | |
UV-Vis diffuse reflectance spectroscopy (DRS) was used to study the optical absorption properties of the prepared materials. As shown in Fig. 7A, the visible light absorption edge of BiVO4 was at 562 nm. In contrast, the visible light absorption edge of CuO/BiVO4 extended to approximately 611 nm, with obvious absorption of visible light. The results demonstrated that CuO/BiVO4 possessed a broader visible light response range and higher visible light utilization efficiency than BiVO4. Electrochemical measurements were performed to evaluate the photocurrent density and charge transfer efficiency of CuO/BiVO4. As shown in Fig. 7B, the CuO/BiVO4 composite exhibited a significantly higher transient photocurrent density than pure BiVO4. The enhanced photocurrent response indicated a more efficient separation of photogenerated electron–hole pairs. To probe the interfacial charge mobility within the photocatalyst, electrochemical impedance spectroscopy (EIS) was employed (Fig. 7C). The Nyquist plot of the CuO/BiVO4 composite showed a markedly smaller arc radius compared to that of pure BiVO4, suggesting a lower charge transfer resistance. The reduction in resistance was conducive to more efficient charge migration, which was consistent with the superior photoresponse performance observed for the CuO/BiVO4 composite materials. The Mott–Schottky measurements (Fig. 7D) identified CuO as p-type and BiVO4 as n-type semiconductors. The results further confirmed the superior photoelectrochemical performance of CuO/BiVO4 composite materials compared with BiVO4.
 |
| | Fig. 7 Characterization of the properties of the CuO/BiVO4 composite: (A) UV-Vis diffuse reflectance spectra (DRS), (B) photocurrent density, (C) EIS Nyquist plots, (D) Mott–Schottky plots. | |
3.3 Photocatalytic antibacterial mechanism
Currently, there are three main antibacterial mechanisms for inorganic photocatalytic materials: ROS oxidative damage mechanism, ion leaching mechanism, and mechanical damage mechanism.35 Herein, the photocatalytic antibacterial mechanism of the CuO/BiVO4 composite material was explored. Firstly, the dialysis tube experiment was determined. When the CuO/BiVO4 composite material was placed inside the dialysis tube, ROS and leached ions could diffuse into the bacterial solution through the dialysis membrane, while the CuO/BiVO4 was confined within the dialysis without direct contact with the bacteria, thus preventing mechanical damage of bacteria. As shown in Fig. 8A, the antibacterial rate of CuO/BiVO4 was 100% without dialysis tube. However, when the composite material was confined within the dialysis tube, the antibacterial rate reduced to 77.8%. The result indicated that the ROS oxidative and the ion leaching damage mechanism showed main contribution to the photocatalytic antibacterial mechanism of CuO/BiVO4. In order to estimate the effect of ROS, glutathione (GSH, ROS scavenger) was added to the system, and the results were shown in Fig. 8B. After the addition of GSH, the antibacterial rate decreased to 29.7%, which indicated that the ROS oxidation damage contributed 70.3% to the antibacterial mechanism of the composite material. Therefore, the antibacterial contribution of Cu2+ was ca. 7.5%, consistent with the insignificant effect of Cu2+ in Fig. 3B. Based on the above results, ROS oxidative damage is the primary mechanism of CuO/BiVO4 composite material, with the mechanical damage and leaching of Cu2+ playing a synergistic role (Fig. 8D).
 |
| | Fig. 8 Antibacterial activity ((A) dialysis tube experiment; (B) GSH experiment; (C) single ROS scavenger experiment) and the contribution of each antibacterial mechanism. | |
To further clarify the contribution of specific free radicals to the antibacterial activity of the CuO/BiVO4 composite material, single ROS scavenging test was conducted. Superoxide dismutase (SOD), catalase (CAT), and D-mannitol can respectively scavenge ·O2−, H2O2, and ·OH, and the three scavengers are specific and do not possess antibacterial capabilities themselves.36 As shown in Fig. 8C, after the addition of SOD, CAT, and D-mannitol, the antibacterial rates of CuO/BiVO4 decreased to 63.8%, 87.2%, and 88.8%, respectively. The results indicated that during the antibacterial process of the CuO/BiVO4 composite material, ·O2− played a major role, while H2O2 and ·OH played a secondary role. The EPR results further confirmed that CuO/BiVO4 could produce more ·O2− compared to pure BiVO4 under light excitation (Fig. S4 in the SI).
According to the antibacterial results and photoelectrochemical test, we propose a photocatalytic antibacterial mechanism of CuO/BiVO4 composite material (Scheme 1). The detailed calculation of CB, VB, and bandgap for BiVO4 was shown in the SI (Fig. S5). Under visible light irradiation, photogenerated electrons (e−) on the conduction band of the p-type CuO are transferred to the conduction band of the n-type BiVO4, meanwhile, holes (h+) on the valence band of the n-type BiVO4 move to the valence band of the p-type CuO. In this way, the photogenerated e−–h+ pairs can be quickly and effectively separated, thus significantly reducing the recombination of e−–h+ pairs and exhibiting improved photocatalytic performance. Subsequently, the photogenerated electrons (e−) and holes (h+) migrated to the surface and underwent redox reactions with adsorbed O2 and H2O molecules, generating reactive oxygen species (ROS) with strong oxidizing capabilities, such as superoxide radicals (·O2−) and hydroxyl radicals (·OH). These reactive substances will undergo oxidation reactions with proteins, nucleic acids, and cell membranes in bacterial cells, disrupting the normal growth and reproduction of bacteria and finally killing the bacteria.
 |
| | Scheme 1 Photocatalytic antibacterial mechanism of the CuO/BiVO4 composite material under visible light irradiation. | |
4. Conclusion
In summary, the CuO/BiVO4 composite photocatalyst was successfully prepared by hydrothermal-impregnation method. CuO loading onto BiVO4 was beneficial for the separation and transfer of photogenerated electrons (e−) and holes (h+), consequently, the photocatalytic antibacterial activity was significantly enhanced compared to pure BiVO4. The CuO/BiVO4-12.5% composite demonstrated the best bactericidal activity, achieving a 100% antibacterial rate under visible light irradiation within 30 min. The outstanding antibacterial performance is mainly originated from ROS oxidative damage, in which ·O2− played a major role in antibacterial activity. The CuO/BiVO4 composite material also exhibited good stability for repeated use. The performance advantages and scalable synthesis methods of CuO/BiVO4 composites make them highly promising for antibacterial applications.
Conflicts of interest
There are no conflicts to declare.
Data availability
All data supporting this study are provided in the manuscript and supplementary information (SI) is available from the corresponding author upon reasonable request. Supplementary information: XRD pattern and antibacterial performance of CuO, the SEM images of CuO/BiVO4 composite materials, the EPR spectra of ·O2−, the band gap and VB/CB potential calculation, and slope and R2 value of dynamic fitting curve. See DOI: https://doi.org/10.1039/d5ra06653k.
Acknowledgements
This research was financially supported by the LiaoNing Revitalization Talents Program (XLYC1907137), and Fundamental Research Funds for the Central Universities (No. 3132023168, 3132023518).
References
- Z. Han, N. Wang and H. Fan, et al., Ag nanoparticles loaded on porous graphitic carbon nitride with enhanced photocatalytic activity for degradation of phenol, Solid State Sci., 2017, 65, 110–115 CrossRef.
- C. Liu, D. Kong and P.-C. Hsu, et al., Rapid water disinfection using vertically aligned MoS2 nanofilms and visible light, Nat. Nanotechnol., 2016, 11(12), 1098–1104 CrossRef PubMed.
- H. Yan, R. Wang and R. Liu, et al., Recyclable and reusable direct Z-scheme heterojunction CeO2/TiO2 nanotube arrays for photocatalytic water disinfection, Appl. Catal., B, 2021, 291, 120096 CrossRef.
- J. He, Z. Zheng and I. M. C. Lo, Different responses of gram-negative and gram-positive bacteria to photocatalytic disinfection using solar-light-driven magnetic TiO2-based material, and disinfection of real sewage, Water Res., 2021, 207, 117816 CrossRef PubMed.
- A. Kubacka, M. Fernandez-Garcia and G. Colon, Advanced nanoarchitectures for solar photocatalytic applications, Chem. Rev., 2012, 112(3), 1555–1614 CrossRef PubMed.
- K. Nakata and A. Fujishima, TiO2 photocatalysis: Design and applications, J. Photochem. Photobiol., C, 2012, 13(3), 169–189 CrossRef.
- J. Yu and A. Kudo, Effects of structural variation on the photocatalytic performance of hydrothermally synthesized BiVO4, Adv. Funct. Mater., 2006, 16(16), 2163–2169 CrossRef.
- L.-C. Chen, G.-T. Pan and T. C.-K. Yang, et al., In situ DRIFT and kinetic studies of photocatalytic degradation on benzene vapor with visible-light-driven silver vanadates, J. Hazard. Mater., 2010, 178(1–3), 644–651 CrossRef PubMed.
- P. Ju, P. Wang and B. Li, et al., A novel calcined Bi2WO6/BiVO4 heterojunction photocatalyst with highly enhanced photocatalytic activity, Chem. Eng. J., 2014, 236, 430–437 CrossRef CAS.
- P. Li, X. Chen and H. He, et al., Polyhedral 30-faceted BiVO4 microcrystals predominantly enclosed by high-index planes promoting photocatalytic water-splitting activity, Adv. Mater., 2018, 30(1), 1703119 CrossRef.
- K.-J. Shieh, M. Li and Y.-H. Lee, et al., Antibacterial performance of photocatalyst thin film fabricated by defection effect in visible light, Nanomed. Nanotechnol. Biol. Med., 2006, 2(2), 121–126 CrossRef CAS PubMed.
- T. Tachikawa, T. Ochi and Y. J. A. C. Kobori, Crystal-face-dependent charge dynamics on a BiVO4 photocatalyst revealed by single-particle spectroelectrochemistry, ACS Catal., 2016, 6(4), 2250–2256 CrossRef CAS.
- Y. Zhao, R. Li and L. Mu, et al., Significance of crystal morphology controlling in semiconductor-based photocatalysis: a case study on BiVO4 photocatalyst, Cryst. Growth Des., 2017, 17(6), 2923–2928 CrossRef CAS.
- J. Zhu, F. Fan and R. Chen, et al., Direct imaging of highly anisotropic photogenerated charge separations on different facets of a single BiVO4 photocatalyst, Angew. Chem., 2015, 127(31), 9239–9242 CrossRef.
- S. Chen, D. Huang and P. Xu, et al., Facet-engineered surface and interface design of monoclinic scheelite bismuth vanadate for enhanced photocatalytic performance, ACS Catal., 2019, 10(2), 1024–1059 CrossRef.
- X. Feng, X. Zhao and L. Chen, BiVO4/BiO0.67F1.66 heterojunction enhanced charge carrier separation to boost photocatalytic activity, J. Nanopart. Res., 2019, 21, 1–9 CrossRef CAS.
- D. Ke, T. Peng and L. Ma, Effects of hydrothermal temperature on the microstructures of BiVO4 and its photocatalytic O2 evolution activity under visible light, Inorg. Chem., 2009, 48, 4685–4691 CrossRef CAS PubMed.
- Y. K. Kho, W. Y. Teoh and A. Iwase, et al., Flame preparation of visible-light-responsive BiVO4 oxygen evolution photocatalysts with subsequent activation via aqueous route, ACS Appl. Mater. Interfaces, 2011, 3(6), 1997–2004 CrossRef CAS PubMed.
- A. Martínez-De La Cruz, U. M. García-Pérez and S. Sepúlveda-Guzmán, Characterization of the visible-light-driven BiVO4 photocatalyst synthesized via a polymer-assisted hydrothermal method, Res. Chem. Intermed., 2013, 39, 881–894 CrossRef.
- M. Wang, P. Guo and T. Chai, et al., Effects of Cu dopants on the structures and photocatalytic performance of cocoon-like Cu-BiVO4 prepared via ethylene glycol solvothermal method, J. Alloys Compd., 2017, 691, 8–14 CrossRef CAS.
- A. Zhang and J. Zhang, Synthesis and characterization of Ag/BiVO4 composite photocatalyst, Appl. Surf. Sci., 2010, 256(10), 3224–3227 CrossRef CAS.
- J. Su, L. Guo and N. Bao, et al., Nanostructured WO3/BiVO4 heterojunction films for efficient photoelectrochemical water splitting, Nano Lett., 2011, 11(5), 1928–1933 CrossRef CAS.
- H. Jiang, M. Nagai and K. Kobayashi, et al., Enhanced photocatalytic activity for degradation of methylene blue over V2O5/BiVO4 composite, J. Alloys Compd., 2009, 479(1–2), 821–827 CrossRef CAS.
- N. Wetchakun, S. Chaiwichain and B. Inceesungvorn, et al., BiVO4/CeO2 nanocomposites with high visible-light-induced photocatalytic activity, ACS Appl. Mater. Interfaces, 2012, 4(7), 3718–3723 CrossRef CAS PubMed.
- J. Sun, X. Li and Q. Zhao, et al., Construction of pn heterojunction β-Bi2O3/BiVO4 nanocomposite with improved photoinduced charge transfer property and enhanced activity in degradation of ortho-dichlorobenzene, Appl. Catal., B, 2017, 219, 259–268 CrossRef CAS.
- H.-q. Jiang, H. Endo and H. Natori, et al., Fabrication and efficient photocatalytic degradation of methylene blue over CuO/BiVO4 composite under visible-light irradiation[J], Mater. Res. Bull., 2009, 44(3), 700–706 CrossRef CAS.
- J. Zhang, H. Cui and B. Wang, et al., Preparation and characterization of fly ash cenospheres supported CuO–BiVO4 heterojunction composite, Appl. Surf. Sci., 2014, 300, 51–57 CrossRef CAS.
- W. Zhao, Y. Wang and Y. Yang, et al., Carbon spheres supported visible-light-driven CuO-BiVO4 heterojunction: preparation, characterization, and photocatalytic properties, Appl. Catal., B, 2012, 115, 90–99 CrossRef.
- P. Pandey, S. Merwyn and G. Agarwal, et al., Electrochemical synthesis of multi-armed CuO nanoparticles and their remarkable bactericidal potential against waterborne bacteria, J. Nanopart. Res., 2012, 14, 1–13 CrossRef.
- J. Ran, H. Chen and X. Bai, et al., Immobilizing CuO/BiVO4 nanocomposite on PDA-templated cotton fabric for visible light photocatalysis, antimicrobial activity and UV protection, Appl. Surf. Sci., 2019, 493, 1167–1176 CrossRef CAS.
- S. Yang, Y. Nie and B. Zhang, et al., Construction of Er-doped ZnO/SiO2 composites with enhanced antimicrobial properties and analysis of antibacterial mechanism, Ceram. Int., 2020, 46(13), 20932–20942 CrossRef CAS.
- Y. Lin, C. Lu and C. Wei, Microstructure and photocatalytic performance of BiVO4 prepared by hydrothermal method, J. Alloys Compd., 2019, 781, 56–63 CrossRef CAS.
- M. F. R. Samsudin, S. Sufian and B. H. Hameed, Epigrammatic progress and perspective on the photocatalytic properties of BiVO4-based photocatalyst in photocatalytic water treatment technology: A review, J. Mol. Liq., 2018, 268, 438–459 CrossRef CAS.
- Y. Wang, X. Li and N. Wang, et al., Controllable synthesis of ZnO nanoflowers and their morphology-dependent photocatalytic activities, Sep. Purif. Technol., 2008, 62(3), 727–732 CrossRef CAS.
- K. Wang, M. Lv and T. Si, et al., Mechanism analysis of surface structure-regulated Cu2O in photocatalytic antibacterial process, J. Hazard. Mater., 2024, 461, 132479 CrossRef CAS.
- H. Ma, X. Yang and X. Tang, et al., Self-assembled Co-doped β-Bi2O3 flower-like structure for enhanced photocatalytic antibacterial effect under visible light, Appl. Surf. Sci., 2022, 572, 151348 CrossRef CAS.
|
| This journal is © The Royal Society of Chemistry 2025 |
Click here to see how this site uses Cookies. View our privacy policy here.