Acessibilidade / Reportar erro

Tailoring Charge Carriers through Solid-State Mechanochemical Synthesis Bismuth Nanoparticles Directly onto BiVO4 Surface

Abstract

Metallic nanoparticles remain the focus of diverse studies across various fields due to their exceptional properties and applications in biology, sensing, energy, and (electro)catalysis. Typical syntheses of metallic nanoparticles are solution-based and require the use of stabilizing agents. Although these surface agents prevent agglomeration, they also hinder access to active sites on the nanoparticle surface. A promising alternative is mechanochemical synthesis, offering a solvent and surfactant-free approach, which yields uniform nanoparticles with clean surfaces. Here, we employ ball milling as a mechanochemical method to prepare metallic bismuth nanoparticles in a single step, using BiVO4 semiconductor both as a bismuth source and a support. Our results demonstrate that, in addition to the formation of nanoparticles by the partial reduction of BiVO4, there is an increase in charge carriers in the semiconductor structure as probed by electrochemical experiments. The final properties of Bi@BiVO4 material make it a relevant candidate for (electro)catalysis.

Keywords:
bismuth nanoparticles; semiconductors; mechanochemistry; charge carriers


Introduction

Nanomaterials, particularly metallic nanoparticles (NPs), continue to gain attention due to their properties and consequent applications in various fields such as biology, energy devices and catalysis, among others.11 da Silva, R. T. P.; Petri, M. V.; Valencia, E. Y.; Camargo, P. H. C.; de Torresi, S. I. C.; Spira, B.; Photodiagnosis Photodyn. Ther. 2020, 31, 101908. [Crossref]
Crossref...
, 22 Durán, N.; Marcato, P. D.; de Conti, R.; Alves, O. L.; Costa, F. T. M.; Brocchi, M.; J. Braz. Chem. Soc. 2010, 21, 949. [Crossref]
Crossref...
, 33 Abbasi-Ahd, A.; Shokoufi, N.; Adeleh, S.; Kargosha, K.; J. Braz. Chem. Soc. 2017, 28, 1843. [Crossref]
Crossref...
, 44 Stratakis, E.; Kymakis, E.; Materials Today 2013, 16, 133. [Crossref]
Crossref...
, 55 Miguel, V. M.; Rodrigues, M. P. S.; Braga, A. H.; de Torresi, S. I. C.; ACS Appl. Nano Mater. 2022, 5, 2943. [Crossref]
Crossref...
, 66 Rodrigues, M. P. S.; Dourado, A. H. B.; Cutolo, L. O.; Parreira, L. S.; Alves, T. V.; Slater, T. J. A.; Haigh, S. J.; Camargo, P. H. C.; Cordoba de Torresi, S. I.; ACS Catal. 2021, 11, 13543. [Crossref]
Crossref...
The small size of the NPs (1-100 nm) reflects in unique properties which include a high surface area, electrical conductivity, and physical chemical properties like modified optical, mechanical, and thermal stability.77 Yaqoob, A. A.; Ahmad, H.; Parveen, T.; Ahmad, A.; Oves, M.; Ismail, I. M. I.; Qari, H. A.; Umar, K.; Mohamad Ibrahim, M. N.; Front Chem. 2020, 8, 341. [Crossref]
Crossref...
Furthermore, the metal NPs have photoinduced properties due to the localized surface plasmon resonance (LSPR), in which, when the NPs are light-irradiated by an appropriated wavelength, typical in the UV-Vis range, the electromagnetic field induces the electron oscillation over the NPs surface, resulting in dipoles. As a result, local effects are observed, like increases in the local temperature, generation of hot carriers and near field enhancement.88 Rodrigues, M. P. S.; Dourado, A. H. B.; Sampaio de Oliveira-Filho, A. G.; de Lima Batista, A. P.; Feil, M.; Krischer, K.; Córdoba de Torresi, S. I.; ACS Catal. 2023, 15, 267. [Crossref]
Crossref...

Different ways to synthesize metal NPs have been studied for decades, such as electrodeposition, sputtering and solution-based methods, which include chemical reduction, sol-gel, and hydrothermal methods.99 Jamkhande, P. G.; Ghule, N. W.; Bamer, A. H.; Kalaskar, M. G.; J. Drug Delivery Sci. Technol. 2019, 55, 101174. [Crossref]
Crossref...
As the properties of the NPs are strongly affected by the shape and size, the selected method of synthesis can be very important, since the interaction of the metal ions and reducing agents, stabilizing agent, precursor and solvents can change the final structure of the NPs.

The bottom-up mechanochemical synthesis of metal NPs has gained more attention in the last years because it is considered a green and scalable alternative route to produce metallic NPs.1010 Baláž, M.; Bedlovičová, Z.; Daneu, N.; Siksa, P.; Sokoli, L.; Tkáčiková, Ľ.; Salayová, A.; Džunda, R.; Kováčová, M.; Bureš, R.; Bujňáková, Z. L.; Nanomaterials 2021, 11, 1139. [Crossref]
Crossref...
, 1111 de Oliveira, P. F. M.; Torresi, R. M.; Emmerling, F.; Camargo, P. H. C.; J. Mater. Chem. A 2020, 8, 16114. [Crossref]
Crossref...
In terms of the synthesis of NPs, the attention given to mechanochemical means is attributed to the fact that it does not require the use of solvents or stabilizing agents, producing uniform NPs with a clean surface allowing to explore their properties for different applications.1212 da Silva, R. T. P.; Córdoba de Torresi, S. I.; de Oliveira, P. F. M.; Front Chem. 2022, 10, 836597. [Crossref]
Crossref...
In addition, the synthesis of NPs over solid supports can be performed in one-pot by the chemical reduction of the metal precursor, followed by the nucleation and growth on the surface of the support.

In the present work, we explore the bottom-up mechanochemical route using ball milling to synthesize metallic bismuth (Bi) NPs by chemical reduction. BiVO4 served as both support and Bi source, resulting in Bi@BiVO4 materials with modified electrochemical and optical properties. The final materials were structurally characterized by powder X-ray diffraction (PXRD), UV-Vis diffuse reflectance spectroscopy (UV-Vis DRS) and transmission electron microscopy (TEM). Electrochemical studies were also conducted and demonstrated the effect of the amount of the reducing agent used for BiNPs formation on the density of the charge carriers.

Experimental

Materials

All solvents were high-performance liquid chromatography (HPLC) grade. Bi(NO3)3.5H2O (99%), NH4VO3 (98%), AgNO3 (99%), and NaBH4 (98%) were purchased from Aldrich (St. Louis, USA). Milli-Q water was used throughout the experiments.

Synthesis of BiVO4

Initially, the BiVO4 was synthesized following similar procedures previously reported in the literature by a wet route.1313 Yuan, M.; Chen, J.; Bai, Y.; Liu, Z.; Zhang, J.; Zhao, T.; Shi, Q.; Li, S.; Wang, X.; Zhang, G.; Chem. Sci. 2021, 12, 6048. [Crossref]
Crossref...
, 1414 Yuan, M.; Chen, J.; Bai, Y.; Liu, Z.; Zhang, J.; Zhao, T.; Wang, Q.; Li, S.; He, H.; Zhang, G.; Angew. Chem., Int. Ed. 2021, 60, 10910. [Crossref]
Crossref...
Typically, 3.0 mmol NH4VO3 and 3.0 mmol Bi(NO3)3.5H2O were dissolved in 12 mL of hot water and 12 mL of ethyleneglicol, respectively. Afterwards, the NH4VO4 solution was slowly dropwised in Bi(NO3)3.5H2O solution and vigorously stirred for 2 h. Then, the mixture was kept in a furnace at 90 ºC overnight resulting in a gel, which was further calcinated at 550 ºC for 120 min (3 ºC min11 da Silva, R. T. P.; Petri, M. V.; Valencia, E. Y.; Camargo, P. H. C.; de Torresi, S. I. C.; Spira, B.; Photodiagnosis Photodyn. Ther. 2020, 31, 101908. [Crossref]
Crossref...
) to form BiVO4.

Synthesis of Bi@BiVO4

The Bi@BiVO4 structures were synthesized through a mechanochemical route. In a poly(methyl methacrylate) (PMMA) jar, the pristine BiVO4 and the reducing agent NaBH4 were milled in different m/m ratios in a vibratory ball mill (Pulverissette 23, St. Louis, USA) with a single zirconia ball (2.8 g, diameter ø = 10 mm) for 30 min at 50 Hz. The final powder was recovered and washed several times with water, ethanol, and acetone to remove any residual and dried under vacuum. The ratios chosen were 1, 2.5, 5 and 10% of NaBH4 in relation to the BiVO4 (m/m) to give 1-Bi@BiVO4, 2.5-Bi@BiVO4, 5-Bi@BiVO4 and 10-Bi@BiVO4, respectively. In addition, the pristine BiVO4 was also milled in the absence of a reducing agent, NaBH4, at the same conditions previously described giving M@BiVO4.

Characterization

Powder X-ray diffraction (PXRD) patterns were recorded in Shimadzu 7.000 Maxima diffractometer (Kyoto, Japan) with Cu Kα radiation (λ Cu Kα = 1.5418 Å for combined Kα1 and Kα2) in a Bragg-Brentano geometry. The transmission electron microscopy (TEM) and high-resolution TEM (HRTEM) images were recorded in JEOL JEM 2100 (Peabody, USA) operated at 200 kV. The samples were prepared by drop casting using 5 µL of the materials in isopropanol/water (1:1). For the TEM and HRTEM images the samples were placed in copper grids. The UV-Vis diffuse reflectance spectra (UV-Vis DRS) of the solid samples were recorded in a Shimadzu UV-2600 UV-Vis spectrophotometer (Kyoto, Japan).

Rietveld refinement

Microstructure analysis and Rietveld refinement of monoclinic BiVO4 and Bi hexagonal phases were conducted using Fullprof Suite software.1515 Fullprof Suite, v. 5.10; Institut Laue-Langevin, Grenoble, France, 2001. The positions of diffraction peaks were calibrated using high-purity silicon as an external standard under identical experimental conditions as the samples, i.e., with a scanning step of 0.3° (2θ) and a range of 10 to 80°. Rietveld refinement quality was assessed using standard agreement indices, including the residual factor (Rp), weighted residual factor (Rwp), and the goodness-of-fit factor (χ2). Additionally, microstructure assessment was performed using the Thomson-Cox-Hastings (TCH) method. This approach utilizes Debye-Scherrer geometry and incorporates variations in Gaussian and Lorentzian half-widths to accurately represent experimental data.

Electrochemical measurements

All the electrochemical experiments were carried out in a Autolab PGSTAT 302N (Uterecht, Netherlands) coupled with three electrodes cell and using 0.2 mol L−1 KHCO3 as supporting electrolyte and glassy carbon (area = 0.04 cm2) as working electrode. The glassy carbon electrodes were also cleaned electrochemically in H2SO4 0.5 mol L−1 solution. The materials were deposited over the glassy carbon electrodes according to the following steps: 10.3 mg of the materials were dispersed in 700 µL of ethanol, 270 µL of water and 30 µL of Nafion solution 0.05%. The dispersion was sonicated for 1 h and then 7 µL were loaded over the electrode surface. Ag/AgCl/Cl(sat) electrode was used as reference and a platinum wire as a counter-electrode. All the potentials were converted to reversible hydrogen electrode (RHE) through the following equation E(VvsRHE)=E(VvsAg/AgCl/Cl(sat)+0.0591×pH+0.197(pH=8.9in Ar in 0.2 mol L1KHCO3) and carried out at room temperature.

The extension of charge carriers on the final materials produced by the milling process in the presence of a NaBH4 as reducing agent was estimated by Mott-Schottky plots analysis based on the changes in the density of the charge carriers. Then, impedance experiments were carried out in the capacitive potential range and the plots were constructed by the following equation.1616 Gelderman, K.; Lee, L.; Donne, S. W.; J. Chem. Educ. 2007, 84, 685. [Crossref]
Crossref...

(1) 1 C 2 = 2 εε 0 A 2 eN d ( V V fb k B T e )

where C is the differential capacitance, ε dielectric constant of the semiconductor (86 = BiVO4),1717 Ye, K.-H.; Yu, X.; Qiu, Z.; Zhu, Y.; Lu, X.; Zhang, Y.; RSC Adv. 2015, 5, 34152. [Crossref]
Crossref...
ε0 vacuum electric permittivity, A area, e elemental charge, Nd density of the charge carriers, V applied potential, Vfb flatband potential, kB is the Boltzmann constant and T is the temperature.

Results and Discussion

The resulting Bi@BiVO4 materials prepared mechanochemically were initially characterized by PXRD, Figure 1. The Bragg reflections observed for the pristine BiVO4 are consistent with the PXRD pattern of monoclinic BiVO4 (JCPDS 98-010-0602). The milled material in the absence of NaBH4, M@BiVO4, also exhibits the same reflections observed for the pristine material, indicating no phase transition as a result of mechanical milling. For the materials milled in the presence of a reducing agent, NaBH4, the reflections relative to the BiVO4 are still maintained. However, the peak intensities are decreased probably due to the loss of crystallinity, resulted from the mechanical stress in the material structure, and the partial disorder caused by the inclusion of NaBH4 in the media. In addition, the PXRD patterns of the materials 5-Bi@BiVO4 and 10-Bi@BiVO4 presents a new reflection at 2θ = 27º (012), which matches with the metallic bismuth in a hexagonal crystalline structure (JCPDS 98-005-3797). Further less intense reflections at 2θ = 38º, 48º, 62º and 64º are also consistent with the metallic hexagonal phase of bismuth. The PXRD results clearly demonstrate that the formation of metallic bismuth is a result of the partial reduction of BiVO4 by NaBH4 resulting in Bi0 over a BiVO4 stressed structure. In fact, the formation of metallic bismuth starting from BiVO4 have already been reported in the literature,1414 Yuan, M.; Chen, J.; Bai, Y.; Liu, Z.; Zhang, J.; Zhao, T.; Wang, Q.; Li, S.; He, H.; Zhang, G.; Angew. Chem., Int. Ed. 2021, 60, 10910. [Crossref]
Crossref...
, 1818 de Brito, J. F.; Corradini, P. G.; Zanoni, M. V. B.; Marken, F.; Mascaro, L. H.; J. Alloys Compd. 2021, 851, 156912. [Crossref]
Crossref...
however, the authors reported the synthesis based on solution and temperature dependence. In our solid state mechanochemical approach, the use of solvents or higher temperatures are not necessary, producing the same results in Bi NPs in a well distribution.

Figure 1
(a) PXRD patterns of (from bottom to top) pristine BiVO4, M@BiVO4, 1-Bi@BiVO4, 2.5-Bi@BiVO4, 5-Bi@BiVO4 and 10-Bi@BiVO4 materials.

Rietveld refinement was carried out for the following samples, 5-Bi@BiVO4 and 10-Bi@BiVO4. The Rietveld refinement revealed that the fraction (in percentage) of metallic bismuth formed during the mechanochemical synthesis are 3.08 and 3.79% for 5-Bi@BiVO4 and 10-Bi@BiVO4 respectively (see Table S1 in Supplementary Information (SI) section). The data revealed that the fraction of metallic bismuth does not increase in the same step as the amount of NaBH4 is added in the synthesis.

The morphology of the materials was investigated by TEM, Figure 2. The pristine BiVO4 and the M@BiVO4 displayed well-defined lattice with an interplanar distance of 0.584 nm, corresponding to the (020) plane of monoclinic phase of BiVO4, the same phase identified in the PXRD patterns. Regarding the BiVO4 milled in the presence of different amounts of NaBH4, one can easily see a uniform distribution of spherical nanoparticles over the BiVO4 sites (Figures 2c-2f). The size of these nanoparticles ranges around 7 ± 1 nm, with an interplanar distance calculated of d = 0.328, which corresponds to the (012) planes of metallic hexagonal bismuth, also in agreement with the PXRD patterns. Interestingly, different from PXRD patterns of reduced samples, which only displayed the reflections of Bi(0) when higher amounts of NaBH4 (5 and 10%) were added, the TEM images show that already with 1% of the reducing agent, Bi(0) can be formed (Figure 2c). The apparent divergence between PXRD and TEM is most probably due to the low content of Bi(0) when NaBH4 is used below 5 wt.%, which makes it difficult to detect Bi(0) by PXRD in our conditions of acquisition. Additionally, a deep investigation of the size of nanoparticles by TEM images, revealed that in higher concentrations of NaBH4, i.e., 5 and 10%, another family of large Bi NPs are produced along with the small ones, with sizes around 24 nm for 5-Bi@BiVO4 and 50 nm for 10-BiVO4 (see Figures S1 and S2, SI section). This apparently intriguing result is a consequence of welding/coalesce of smaller NPs quickly formed at the beggining of the reaction, when there is more NaBH4 in the system. Then, the constant action of mechanical energy will induce the coalescence and welding of the small NPs, resulting in another family of particle sizes, in addition of that of 7 nm.1919 de Oliveira, P. F. M.; Torresi, R. M.; Emmerling, F.; Camargo, P. H. C.; J. Mater. Chem. A 2020, 8, 16114. [Crossref]
Crossref...

Figure 2
TEM images for (a) the BiVO4, M@BiVO4, 1-Bi@BiVO4, 2.5-Bi@BiVO4, 5-Bi@BiVO4 and 10-Bi@BiVO4 materials. Inset: detail of the size and planes of Bi metallic nanoparticles.

The optical properties of the BiVO4 and Bi@BiVO4 materials were investigated, Figure 3. The UV-Vis DRS for the pristine material presents higher percentual of reflectance between 500-800 nm; for the milled materials Bi@BiVO4, as the amount of NaBH4 increases, the absorption over this range is increased, Figure 3a. In order to investigated the effect associated to it, the band gap of the materials was calculated by Tauc Plot considering the Kubelka-Munk function applied to the UV-Vis DRS of solids (see SI section). The band gap achieved for the pristine BiVO4 is 2.50 eV, which agrees with other reports in the literature for the BiVO4 with values between 2.40 and 2.50 eV.2020 Chen, X.; Wu, Y.; Deng, M.; Shen, H.; Ding, J.; Wang, W.; ACS Omega 2022, 7, 17075. [Crossref]
Crossref...
, 2121 Pingmuang, K.; Chen, J.; Kangwansupamonkon, W.; Wallace, G. G.; Phanichphant, S.; Nattestad, A.; Sci. Rep. 2017, 7, 8929. [Crossref]
Crossref...
The milled materials, however, both in the presence and absence of reducing agent, no significant differences in the band gap energy could be observed, with values close to 2.53 eV. The values achieved reveal that the production of Bi NPs on the BiVO4 surface does not lead to significant changes in the energy of the band gap of the semiconductor. Otherwise, observations in terms of reflectance percentage and changes in the solid color, Figure 3c, were observed suggesting that other effects are related to the changes in the optical properties of Bi@BiVO4 materials. Then, the solids were suspended in acetonitrile and the UV-Vis spectra recorded in the transmission mode, Figure 3b. The pristine and the milled material in the absence of NaBH4 material do not present any bands between 400-800 nm. For the materials milled in the presence of a reducing agent, a band is observed in 460 nm and is related to the plasmon band of bismuth nanoparticles.2222 Singh, S. P.; Karmakar, B.; Mater. Chem. Phys. 2010, 119, 355. [Crossref]
Crossref...
For the materials milled in the presence of high amounts of NaBH4, i.e., 5 and 10%, a broad band between 500 and 800 nm is observed. The broad band absorption observed for the milled materials with higher amounts of NaBH4 can be a result of a production of large bismuth nanoparticles, as observed by TEM images and by the crystallite size obtained by Rietveld refinement. In fact, the plasmon bands of metallic nanoparticles are highly dependent of the shape and size. Some reports in literature2323 McMahon, J. M.; Schatz, G. C.; Gray, S. K.; Phys. Chem. Chem. Phys. 2013, 15, 5415. [Crossref]
Crossref...
, 2424 Wang, Z.; Jiang, C.; Huang, R.; Peng, H.; Tang, X.; J. Phys. Chem. C 2014, 118, 1155. [Crossref]
Crossref...
observed that, as the size of bismuth nanoparticles increase, there is a red-shift in the UV-Vis spectra of bismuth plasmon band.

Figure 3
(a) UV-Vis DRS, (b) UV-Vis in transmission mode and (c) color changes of pristine BiVO4, M@BiVO4, 1-Bi@BiVO4, 2.5-Bi@BiVO4, 5-Bi@BiVO4 and 10-Bi@BiVO4 materials.

The milling process in the presence of a reducing agent can result not only in the formation of metallic nanoparticles as reported here, but also in defects in the structure of semiconductors such as oxygen vacancies,2525 Zou, W.; Dong, J.; Ji, M.; Wang, B.; Li, Y.; Yin, S.; Li, H.; Xia, J.; ACS Appl. Nano Mater. 2023, 6, 4309. [Crossref]
Crossref...
that can be valuable for some applications, especially photocatalytic ones.2626 Hao, L.; Yan, J.; Guan, S.; Cheng, L.; Zhao, Q.; Zhu, Z.; Wang, Y.; Lu, Y.; Liu, J.; Appl. Surf. Sci. 2019, 466, 490. [Crossref]
Crossref...
The production of these species directly affects the density of charge carriers. Electrochemical measurements were then used to evaluate and estimate the changes in the density of the charge after the milling process, by employing the Mott-Schottky equation, Figure 4.

Figure 4
Mott-Schottky plots for (a) BiVO4 pristine and 1-Bi@BiVO4, (b) charge carriers’ density obtained for all materials by Mott-Schottky plots at 100 Hz in 0.2 mol L−1 KHCO3.

The Mott-Schottky analysis was conducted in the capacitive range of potential and exhibits a linear region with a positive slope for all the materials. The positive slope is a typical behavior of n-type semiconductors and can be used to estimate the density of charge carriers (Nd) applying the Mott-Schottky equation.2727 Hankin, A.; Bedoya-Lora, F. E.; Alexander, J. C.; Regoutz, A.; Kelsall, G. H.; J. Mater. Chem. A 2019, 7, 26162. [Crossref]
Crossref...
Regarding BiVO4 and M@BiVO4, the estimated Nd values have the same magnitude with values of 6.2 × 1020 and 3.0 × 1020 cm−3, respectively. For the materials milled with 1 and 2.5 wt.% of NaBH4, the Nd values also falls in the same magnitude when compared to the pristine BiVO4 and M@BiVO4 with values of 7.7 × 1020 and 9.7 × 1020 cm−3 respectively for 1-Bi@BiVO4 and 2.5-Bi@BiVO4. Despite being of similar order, one can already observe a small increase in the number of charge carriers, when comparing the pristine and the milled systems in the presence of a reducing agent. The difference becomes clearer when higher amounts of reducing agent (5 and 10 wt.% of NaBH4) is employed (Figure 4). The density of charge carriers increases significantly in terms of magnitude presenting values of 2.0 × 1021 and 6.1 × 1022 cm3 for 5-Bi@BiVO4 and 10-Bi@BiVO4, respectively. Therefore, concerning the materials milled with the reducing agent, it is possible to observe a trend. The higher the amount of NaBH4, the higher the density of charge carriers, with a more pronounced difference for higher concentrations of reducing agent.

One can attribute the increase of the charge carriers to the reduction of part of Bi3+ from BiVO4 into Bi0 producing metallic Bi NPs. Nonetheless, it is not possible to rule out the possible formation of oxygen vacancies during the milling process in the presence of the reducing agent. There are some reports that observe the formation of oxygen vacancies during the milling process. For instance, oxygen vacancies were formed in CuO structure when milled with Mg as reducing agent, according to Chen et al.2828 Chen, H.; Zhou, S.; Han, Z.; Jiang, Y.; Yu, H.; Zhou, X.; Jiang, R.; Liu, X.; Li, X.; J. Mater. Chem. A 2016, 4, 17207. [Crossref]
Crossref...
Another interesting example is related to the chemical reduction of Au3+ over MoO3 in the presence of NaBH4, resulting in the formation of metallic AuNPs over a defective MoO3-x structure.2929 Quiroz, J.; de Oliveira, P. F. M.; Shetty, S.; Oropeza, F. E.; de la Peña O’Shea, V. A.; Rodrigues, L. C. V.; de S. Rodrigues, M. P.; Torresi, R. M.; Emmerling, F.; Camargo, P. H. C.; ACS Sustainable Chem. Eng. 2021, 9, 9750. [Crossref]
Crossref...
However, when compared to MoO3-x case, for Bi@BiVO4 there is no meaningful change in the band gap.

The increase in the density of charge carriers in the semiconductor structure driven by the formation of NPs and defects through mechanochemistry can be advantageous for the development of materials based on semiconductors. This strategy has potential to be beneficial in different fields, such as catalysis enhanced by localized surface plasmon resonance.

Conclusions

The pristine BiVO4 semiconductor was synthesized through a wet route and submitted to the milling process in different ratios of a reducing agent. In a one-pot synthesis, the mechanochemical approach could generate metallic Bi NPs over BiVO4, without external heating or use of surface agents and solvents. PXRD and TEM confirmed that the milling process induces the formation of metallic Bi nanoparticles well distributed onto BiVO4 surface. Although no significant changes were observed in the band gap of Bi@BiVO4 materials, the changes of color and UV-Vis absorption spectra revealed red-shifted bands in the presence of higher amounts of NaBH4 due to the formation of large Bi NPs. Besides no significant changes in the amount of metallic Bi produced in higher amounts of NaBH4, the electrochemical measurements through Mott-Schottky analysis indicate that the number of charge carriers increases according to the amount of reducing agent used. The mechanochemical synthesis of Bi@BiVO4, displaying both nanoparticles and semiconductor properties, can be used to tune the density of charge carriers for different applications, especially catalytic ones. Furthermore, the mechanochemical synthesis approach can amplify the use of solid-state ball milling for the preparation of multicomponent advanced materials for technological applications.

  • This paper is dedicated to our colleague, friend, excellent scientist and inspiring professor Oswaldo Alves. His legacy will always remain in the Brazilian chemical community.

Supplementary Information

Supplementary data are available free of charge at http://jbcs.sbq.org.br as PDF file.PDF

Acknowledgments

The authors are thankful to Fundação de Amparo à Pesquisa do Estado de São Paulo (FAPESP, 20/15230-5, 21/00675-4, 22/04604-7, 23/01223-5), Conselho Nacional de Desenvolvimento Científico e Tecnológico (303269/2021-9). PFMO also acknowledge FAPESP for funding (FAPESP 2020/14955-6). The authors would also thankfully acknowledge the support of the RCGI-Research Centre for Greenhouse Gas Innovation, hosted by the University of São Paulo (USP) and sponsored by FAPESP and Shell Brasil, and the strategic importance of the support given by ANP (Brazil’s National Oil, Natural Gas, and Biofuels Agency) through the R&D levy regulation.

References

  • 1
    da Silva, R. T. P.; Petri, M. V.; Valencia, E. Y.; Camargo, P. H. C.; de Torresi, S. I. C.; Spira, B.; Photodiagnosis Photodyn. Ther. 2020, 31, 101908. [Crossref]
    » Crossref
  • 2
    Durán, N.; Marcato, P. D.; de Conti, R.; Alves, O. L.; Costa, F. T. M.; Brocchi, M.; J. Braz. Chem. Soc. 2010, 21, 949. [Crossref]
    » Crossref
  • 3
    Abbasi-Ahd, A.; Shokoufi, N.; Adeleh, S.; Kargosha, K.; J. Braz. Chem. Soc. 2017, 28, 1843. [Crossref]
    » Crossref
  • 4
    Stratakis, E.; Kymakis, E.; Materials Today 2013, 16, 133. [Crossref]
    » Crossref
  • 5
    Miguel, V. M.; Rodrigues, M. P. S.; Braga, A. H.; de Torresi, S. I. C.; ACS Appl. Nano Mater. 2022, 5, 2943. [Crossref]
    » Crossref
  • 6
    Rodrigues, M. P. S.; Dourado, A. H. B.; Cutolo, L. O.; Parreira, L. S.; Alves, T. V.; Slater, T. J. A.; Haigh, S. J.; Camargo, P. H. C.; Cordoba de Torresi, S. I.; ACS Catal. 2021, 11, 13543. [Crossref]
    » Crossref
  • 7
    Yaqoob, A. A.; Ahmad, H.; Parveen, T.; Ahmad, A.; Oves, M.; Ismail, I. M. I.; Qari, H. A.; Umar, K.; Mohamad Ibrahim, M. N.; Front Chem. 2020, 8, 341. [Crossref]
    » Crossref
  • 8
    Rodrigues, M. P. S.; Dourado, A. H. B.; Sampaio de Oliveira-Filho, A. G.; de Lima Batista, A. P.; Feil, M.; Krischer, K.; Córdoba de Torresi, S. I.; ACS Catal. 2023, 15, 267. [Crossref]
    » Crossref
  • 9
    Jamkhande, P. G.; Ghule, N. W.; Bamer, A. H.; Kalaskar, M. G.; J. Drug Delivery Sci. Technol. 2019, 55, 101174. [Crossref]
    » Crossref
  • 10
    Baláž, M.; Bedlovičová, Z.; Daneu, N.; Siksa, P.; Sokoli, L.; Tkáčiková, Ľ.; Salayová, A.; Džunda, R.; Kováčová, M.; Bureš, R.; Bujňáková, Z. L.; Nanomaterials 2021, 11, 1139. [Crossref]
    » Crossref
  • 11
    de Oliveira, P. F. M.; Torresi, R. M.; Emmerling, F.; Camargo, P. H. C.; J. Mater. Chem. A 2020, 8, 16114. [Crossref]
    » Crossref
  • 12
    da Silva, R. T. P.; Córdoba de Torresi, S. I.; de Oliveira, P. F. M.; Front Chem. 2022, 10, 836597. [Crossref]
    » Crossref
  • 13
    Yuan, M.; Chen, J.; Bai, Y.; Liu, Z.; Zhang, J.; Zhao, T.; Shi, Q.; Li, S.; Wang, X.; Zhang, G.; Chem. Sci. 2021, 12, 6048. [Crossref]
    » Crossref
  • 14
    Yuan, M.; Chen, J.; Bai, Y.; Liu, Z.; Zhang, J.; Zhao, T.; Wang, Q.; Li, S.; He, H.; Zhang, G.; Angew. Chem., Int. Ed. 2021, 60, 10910. [Crossref]
    » Crossref
  • 15
    Fullprof Suite, v. 5.10; Institut Laue-Langevin, Grenoble, France, 2001.
  • 16
    Gelderman, K.; Lee, L.; Donne, S. W.; J. Chem. Educ. 2007, 84, 685. [Crossref]
    » Crossref
  • 17
    Ye, K.-H.; Yu, X.; Qiu, Z.; Zhu, Y.; Lu, X.; Zhang, Y.; RSC Adv. 2015, 5, 34152. [Crossref]
    » Crossref
  • 18
    de Brito, J. F.; Corradini, P. G.; Zanoni, M. V. B.; Marken, F.; Mascaro, L. H.; J. Alloys Compd. 2021, 851, 156912. [Crossref]
    » Crossref
  • 19
    de Oliveira, P. F. M.; Torresi, R. M.; Emmerling, F.; Camargo, P. H. C.; J. Mater. Chem. A 2020, 8, 16114. [Crossref]
    » Crossref
  • 20
    Chen, X.; Wu, Y.; Deng, M.; Shen, H.; Ding, J.; Wang, W.; ACS Omega 2022, 7, 17075. [Crossref]
    » Crossref
  • 21
    Pingmuang, K.; Chen, J.; Kangwansupamonkon, W.; Wallace, G. G.; Phanichphant, S.; Nattestad, A.; Sci. Rep. 2017, 7, 8929. [Crossref]
    » Crossref
  • 22
    Singh, S. P.; Karmakar, B.; Mater. Chem. Phys. 2010, 119, 355. [Crossref]
    » Crossref
  • 23
    McMahon, J. M.; Schatz, G. C.; Gray, S. K.; Phys. Chem. Chem. Phys. 2013, 15, 5415. [Crossref]
    » Crossref
  • 24
    Wang, Z.; Jiang, C.; Huang, R.; Peng, H.; Tang, X.; J. Phys. Chem. C 2014, 118, 1155. [Crossref]
    » Crossref
  • 25
    Zou, W.; Dong, J.; Ji, M.; Wang, B.; Li, Y.; Yin, S.; Li, H.; Xia, J.; ACS Appl. Nano Mater. 2023, 6, 4309. [Crossref]
    » Crossref
  • 26
    Hao, L.; Yan, J.; Guan, S.; Cheng, L.; Zhao, Q.; Zhu, Z.; Wang, Y.; Lu, Y.; Liu, J.; Appl. Surf. Sci. 2019, 466, 490. [Crossref]
    » Crossref
  • 27
    Hankin, A.; Bedoya-Lora, F. E.; Alexander, J. C.; Regoutz, A.; Kelsall, G. H.; J. Mater. Chem. A 2019, 7, 26162. [Crossref]
    » Crossref
  • 28
    Chen, H.; Zhou, S.; Han, Z.; Jiang, Y.; Yu, H.; Zhou, X.; Jiang, R.; Liu, X.; Li, X.; J. Mater. Chem. A 2016, 4, 17207. [Crossref]
    » Crossref
  • 29
    Quiroz, J.; de Oliveira, P. F. M.; Shetty, S.; Oropeza, F. E.; de la Peña O’Shea, V. A.; Rodrigues, L. C. V.; de S. Rodrigues, M. P.; Torresi, R. M.; Emmerling, F.; Camargo, P. H. C.; ACS Sustainable Chem. Eng. 2021, 9, 9750. [Crossref]
    » Crossref

Edited by

Editor handled this article: André Galembeck (Guest)

Publication Dates

  • Publication in this collection
    22 July 2024
  • Date of issue
    2024

History

  • Received
    30 Jan 2024
  • Accepted
    18 June 2024
Sociedade Brasileira de Química Instituto de Química - UNICAMP, Caixa Postal 6154, 13083-970 Campinas SP - Brazil, Tel./FAX.: +55 19 3521-3151 - São Paulo - SP - Brazil
E-mail: office@jbcs.sbq.org.br