Synergistic Effect of Ta2O5/F−C Composites for Effective Electrosynthesis of Hydrogen Peroxide from O2 Reduction

Ke WANG Yong-Yu PANG Huan XIE Yuan SUN Guo-Liang CHAI

Citation:  Ke WANG, Yong-Yu PANG, Huan XIE, Yuan SUN, Guo-Liang CHAI. Synergistic Effect of Ta2O5/F−C Composites for Effective Electrosynthesis of Hydrogen Peroxide from O2 Reduction[J]. Chinese Journal of Structural Chemistry, 2021, 40(2): 225-232. doi: 10.14102/j.cnki.0254-5861.2011-2817 shu

Synergistic Effect of Ta2O5/F−C Composites for Effective Electrosynthesis of Hydrogen Peroxide from O2 Reduction

English

  • Hydrogen peroxide (H2O2) is one of the 100 most important chemicals in the world[1]. Because of its strong oxidative properties and the only by-product is water[2, 3], H2O2 as a green oxidant is often used in almost all chemical industry fields, such as medicine, chemical synthesis, environmental protection, paper bleaching, disinfection, water treatment and so on[2, 4-7]. So far, 95% of the total H2O2 productions annually comes from the anthraquinone process (AO), which contains a multi-step complex process such as hydrogenation over Pd catalyst and rapid oxidation by O2 to generate H2O2[3, 8-10]. However, such AO process needs complicated steps and several side reactions which cause unnecessary energy consumption[9, 11]. Secondly, the extremely poor stability of high-concentration H2O2 poses safety hazard and high transportation costs[12]. Therefore, the development of low-cost and low-concentration H2O2 synthesis methods has received increasing attention[5]. One alternative process is the direct synthesis of H2O2 from H2 and O2 with a noble metal alloy catalyst[13-15]. However, this process is not only inefficient but also has a risk of explosion due to the hydrogen-oxygen mixture[12]. Another alternative is photocatalytic production of hydrogen peroxide from water and dioxygen[16, 17]. Nevertheless, this method also has the disadvantages of slow kinetics and low efficiency[1, 2]. Recently, electrosynthesis of H2O2 via two-electron O2 reduction replacing the AO process has caused more and more attention because it is a green process that is efficient, energy-saving and safe[2, 18].

    Due to its low price and abundance, transition metal oxides (such as MnO2, Co3O4, Fe3O4, V2O5, Ta2O5 and so no) have been widely used as an alternative to precious metals in electrocatalytic oxygen reduction reaction (ORR) catalysts[19-24]. In general, metal oxides show limited ORR performance due to poor electronic conductivity and agglomeration issues[25].

    Ta2O5 is currently used as an efficient ORR catalyst due to its high oxygen reduction onset potential (∼0.95 V vs. SHE) and excellent stability under corrosive environments[26-28]. However, the limited electronic transport of Ta2O5 always reduces its kinetic reaction rate[29]. Recently, Ta2O5 was found to be an efficient catalyst for two electrons ORR to generate H2O2[24, 26]. Carbon supports (such as Vulcan XC72R (V), printex L6 (P)[22, 30], RGO[31], carbon black[32], etc.) show good conductivity and large surface areas have been proved to be helpful to improve the electronic transport and catalytic activity of metal oxides.

    Herein, a simple and effective strategy has been proposed to improve the catalytic activity and electronic conductivity of Ta2O5 for H2O2 electrosynthesis. The polyvinylidene fluoride (PVDF) has been carbonized to high surface area F doped porous carbon and coupled with Ta2O5 nanospheres by a simple calcination to generate Ta2O5/F−C composite. The Ta2O5/F−C composite catalyst presents an excellent performance for H2O2 generation with H2O2 selectivity and achieves more than 80% and negligible overpotential in 0.1 M KOH alkaline solution, which is greatly enhanced compared to the counterparts of F−C and Ta2O5. The enhanced performance was attributed to the synergistic effect between F−C and Ta2O5, which increases the conductivity of Ta2O5 and improves the selectivity of H2O2.

    Potassium hydroxide (98.5%) and tantalum powder (99.9%) were bought from Aladdin company. Hydrofluoric acid (40%), nitric acid (65~68%), ammonium hydroxide (25~28%), N-methylpyrrolidone (NMP, 99%), monopotassium phosphate (99.5%) isopropanol (99.7%), glucose (AR) and dipotassium phosphate (99%) were gotten from Sinopharm Chemical Reagent Co. Ltd. PVDF (HSV900) was bought from HeiFei-kejing. Nafion solution (5%) is gotten from Shanghai Hesen Electric Co. Ltd.

    The Ta2O5 sphere precursor was prepared by a co-precipitation process in a typical previous synthesis procedure[33]. First, 11 mL HF, 22 mL HNO3 and 77 mL deionized water were mixed and then 1 g tantalum powder was carefully poured into the above solution. Ta powder is completely dissolved after six hours. Finally, a certain amount of ammonium hydroxide was injected to the solution and a white precipitation was generated. The white precipitates were separated, washed with deionized water for three times and then dried in an oven at 60 ℃ overnight to obtain the final products of Ta2O5 nanospheres.

    Ta2O5/F−C was prepared using one-step process. 0.1 g PVDF and 0.1 g Ta2O5 nanospheres were milled with a mortar for 30 minutes, then a certain amount of NMP as a binder was added during the milling process. Such viscous colloids were dried overnight in an oven at 60 ℃ and then calcinated in a tube furnace at 800 ℃ for 2 h with a heating rate of 5 ℃/min in N2. Then, they were cooled down to room temperature naturally. The black powders were generated and marked as Ta2O5/F−C. As a comparison, PVDF and Ta2O5 nanospheres with a mass ratio of 1:2 and 2:1 were also treated with the same process. The products were marked Ta2O5/F−0.5C and Ta2O5/F−2C, respectively. 0.1 g PVDF was calcinated in a tube furnace at 800 ℃ for 2 h at a heating rate of 5 ℃/min in N2 and marked as Ta2O5-800. 0.1 g Ta2O5 nanosphere calcinated in a tube furnace at 800 ℃ for 2 h with a heating rate of 5 ℃/min in N2 was marked as F−C.

    The transmission electron microscopy (TEM) images were obtained by using JEOL field-emission microscope (JEOL, Tokyo, Japan) operated at 200 kV. A field emission scanning electron microscope (SEM, FESEM, JSM6700F) equipped with an energy dispersive X-ray spectroscope (EDX, Oxford INCA) is used for acquired SEM images and collecting EDX data (Random five points). X-ray diffraction (XRD) patterns were captured by Miniflex 600 X-ray diffractometer under Cu radiation. X-ray photoelectron spectroscopy (XPS) was recorded on an ESCALAB 250Xi X-ray photoelectron spectrometer (Thermo Fisher) using monochromatized Al radiation (15 kV, 10 mA). The specific surface areas were measured using the Brunauer-Emmett-Teller (BET) model.

    All electrochemical measurements were performed by an electrochemical workstation (Autolab, PGSTAT302N). Linear sweep voltammetry (LSV) and cyclic voltammetry (CV) experiments were performed using a single cell on a three-electrode system including a platinum gauze as the counter electrode, a saturated Ag/AgCl electrode as a reference electrode and a rotating ring-disk electrode (RRDE, the areas of ring and desk electrodes are 0.07212 and 0.19625 cm2) as working electrode. The H2O2 selectivity was acquired by RRDE experiments in oxygen saturated electrolyte (two electrolytes with pH~8 (phosphate buffered saline) and pH~13 (0.1 M KOH)) at a scan rate of 10 mV/s). 6.0 mg of each catalyst was dispersed in 1.49 mL of water, 0.57 mL of isopropanol, and 0.04 mL of a Nafion solution (5%). The catalyst ink was sonicated for one hour and 16.5 µL of catalyst ink drop-casted onto the desk electrode and then dried at room temperature. As a comparison, CV and LSV experiments in nitrogen saturated electrolytes were also performed. Electrochemical impedance spectroscopy (EIS) was carried out on a CHI760E working station. Stability test was obtained by a chronoamperometry measurement using RRED under 0.5V vs. RHE with a rotation speed of 1600 rpm for 10 hours. All potentials in this work were converted to reversible hydrogen (RHE) by the following equation (1), H2O2 selectivity was calculated by equation (2)[34], and electronic transfer number was calculated by equation (3)[5].

    ${E}_{\mathrm{R}\mathrm{H}\mathrm{E}}={E}_{\mathrm{A}\mathrm{g}/\mathrm{A}\mathrm{g}\mathrm{C}\mathrm{l}}+0.197+0.0592\times \mathrm{p}\mathrm{H}$

    (1)

    ${\mathrm{H}}_{2}{\mathrm{O}}_{2}\left(\mathrm{\%}\right)=\frac{200\times {I}_{\mathrm{r}}}{\left(N\times {I}_{\mathrm{d}}\right)+{I}_{\mathrm{r}}}$

    (2)

    $n=\frac{4\times {I}_{\mathrm{d}}}{{I}_{\mathrm{d}}+{I}_{\mathrm{r}}/N}$

    (3)

    where Id is the disk current, Ir the ring current, and N the collection efficiency (0.249).

    The Ta2O5/F−C catalyst was successfully synthesized by calcining PVDF and Ta2O5 spheres in N2 at 800 ℃. Figs. 1a and S1a show the SEM images with different magnification of the Ta2O5 spheres which have a smooth surface and a particle size of tens to hundreds of nanometers. For comparison, the SEM image of Ta2O5-800 (Ta2O5 spheres were calcined at 800 ℃) shown in Fig. S1b suggests that the morphology of Ta2O5 sphere maintained after calcination. Figs. 1b and S1c suggest that the F doped C (F−C) displayed as a block shape. Compared with the Ta2O5 sphere precursor, the rough surfaces of Ta2O5/F−C composite shown in Figs. 1c and S1d illustrate that the F doped C is coupled with Ta2O5 sphere. The TEM image of Ta2O5 spheres (Fig. 2a) further proves the Ta2O5 spheres are dense of several smaller nanospheres, 30~100 nm in size. In Fig. 2b, the TEM structure of F doped C indicates that F−C has a porous layered structure. The TEM image of Ta2O5/F−C (Fig. 2c) further suggests that the Ta2O5 sphere was coupled with a layer of F doped carbon. Fig. S2a~2c shows the results of five spots EDX data of Ta2O5/F−0.5C, Ta2O5/F−C and Ta2O5/F−2C, illustrating that the composite contains F, O, Ta, and C elements and the atomic ratios of C and Ta are 8:1, 15:1, 31:1, respectively. This corresponds to the feed ratio of Ta2O5 spheres and PVDF. Fig. S2d reveals the EDX data of Ta2O5-800 draw near to 2:5 which can further prove the presence of Ta2O5 phase. The EDX result of F−C (Fig. S2e) shows that the atom ratio of F element is 1.5%.

    Figure 1

    Figure 1.  SEM images of (a) Ta2O5 spheres, (b) F doped carbon and (c) Ta2O5/F−C

    Figure 2

    Figure 2.  TEM images of (a) Ta2O5 spheres, (b) F doped carbon and (c) Ta2O5/F−C

    The XRD (X-ray diffraction) patterns of Ta2O5 (Fig. S3) before calcination show two wide diffraction peaks, suggesting that Ta2O5 spheres synthesized at room temperature are amorphous phase[35]. The XRD pattern of F−C (in Fig. S3) shows a wide peak of graphite carbon. The X-ray diffraction (XRD) pattern of Ta2O5 and Ta2O5/F−C (Fig. 3) mainly demonstrate the orthorhombic pattern of Ta2O5 (ICSD No. 25-0922).

    Figure 3

    Figure 3.  XRD patterns of Ta2O5/F−C and Ta2O5-800

    The high resolution XPS spectrum for C1s is shown in Fig. 4a, which reveals that the fitting peaks at 284.5, 286.2, 288.2 and 289.8 eV are corresponding to C−C, C−O, C−F and O−C=O chemical bonds, respectively[36]. It confirmed that F atoms are incorporated into C matrix. The fitting peaks of O 1s spectrum (Fig. 4b) at 531.4, 530.1 and 533.1 eV are corresponding to C=O, O−Ta and O−C=O bonds, which demonstrates that O atoms are combined with C and Ta atoms[37, 38]. The binding energy of the F 1s peak is 688.9 eV, which corresponds to F−C bond, as shown in Fig. 4c. This further indicates that fluorine is successfully doped into carbon after PVDF carbonization[36, 37]. The Ta 4f7/2 and Ta 4f5/2 peaks in Fig. 4d are at 26.8 and 28.7 eV, which indicates the presence of Ta2O5[33].

    Figure 4

    Figure 4.  X-ray photoelectron spectra of (a) C 1s, (b) O 1s, (c) F 1s and (d) Ta 4f for Ta2O5/F−C, (e) Nitrogen adsorption-desorption isotherms for Ta2O5/F−C and (f) BJH desorption pore-size distribution of Ta2O5/F−C

    The nitrogen adsorption-desorption isotherms of Ta2O5/F−C (Fig. 4e) demonstrate extremely rapid nitrogen absorption at very low relative pressures, indicating the typical microporosity property of Ta2O5/F−C[39]. In Fig. 4f, the BJH desorption pore-size distribution suggests that the pore width distribution is 1.16 nm. The BET surface area of Ta2O5/F−C is 152.6 m2/g and the micropore area is 140 m2/g with a micropore volume of 0.075 cm3/g. However, Ta2O5-800 has no conspicuous hysteresis loop at high relative pressure compared to Ta2O5/F−C, as shown in Fig. S4. Its BET surface area is only 8 m2/g and the pore volume is insignificant. As a result, it can be deduced that the increased BET surface area and pore volume of Ta2O5/F−C mainly comes from piled pore and porous F−C. The increased BET surface area not only exposes more catalytically active sites, but also improves mass transfer. This is more conducive to the production of hydrogen peroxide[40].

    The electrochemical performance measurement of the catalysts was based on rotating ring-disk electrode (RRDE) with a rotating speed of 1600 rpm in a three-electrode system. To better evaluate the ORR activity, the ring electrode is made of Pt to oxidize H2O2 under a ring potential[41]. Fig. 5a and 5b reveals the cyclic voltammetry (CV) curves and the linear sweep voltammetry (LSV) profiles of the Ta2O5/F−C composites in N2 and O2 saturated alkaline electrolytes (0.1 M KOH solution). A remarkable increase of both the ring and disk currents in the presence of O2 respect to the N2 background was observed. The CV curve shows a flat slope in the N2 saturated solution and an obvious reduction peak under O2 saturated electrolyte, which indicates the reliability of oxygen reduction performance of Ta2O5/F−C.

    Figure 5

    Figure 5.  (a) CV and (b) LSV curves of Ta2O5/F−C in N2 and O2 saturated 0.1 M KOH solution

    Fig. 6a shows the LSV curves of Ta2O5/F−C, Ta2O5-800 and F−C in 0.1 M KOH electrolyte. Obviously, the ring current of Ta2O5/F−C is the largest among all the three catalysts. Although Ta2O5-800 has a favorable H2O2 selectivity, the oxygen reduction reactivity of Ta2O5-800 is poor. As shown in Fig. 6a and 6b, H2O2 selectivity of Ta2O5-800 can reach nearly 70% but the limiting current density is just less than 1 mA/cm2 and the onset potential at 0.64 V vs. RHE (defined as the potential at which a current density of 0.1 mA/cm2 is achieved). The poor activity of Ta2O5-800 is ascribed to the lower electronic conductivity of Ta2O5[25]. As shown in Fig. 6b, the F−C has a poor H2O2 selectivity about 55%, which is far less than the Ta2O5/F−C (80.8%). The resistance measurement further confirms the enhanced conductivity of the Ta2O5/F−C electrode. The high frequency range of EIS (Fig. S5) illustrated charge transfer ability for Ta2O5/F−C and Ta2O5-800. EIS in high frequency region (100 kHz to 10 Hz) was fitted with an equivalent circuit diagram to get charge transfer resistance. Obviously, the charge transfer resistance of Ta2O5-800 is two orders of magnitude larger than Ta2O5/F−C, which indicates the introduction of F−C improved the conductivity of Ta2O5 successfully.

    Figure 6

    Figure 6.  (a) LSV curves of Ta2O5/F−C, Ta2O5-800 and F−C (scan rate of 10 mV/s with a rotation speed of 1600 rpm). (b) H2O2 selectivity of Ta2O5/F−C, Ta2O5-800 and F−C in O2 saturated 0.1 M KOH electrolyte. (c) Tafel slopes for corresponding catalysts in O2 saturated 0.1 M KOH electrolyte. (d) Stability test for Ta2O5/F−C. All plots were subtracted from the nitrogen background current and the pH of test condition is 13

    The LSV profiles of Ta2O5/F−0.5C, Ta2O5/F−C, and Ta2O5/F−2C that contain different ratios of F are shown in Fig. S6a. The Ta2O5/F−C displayed the highest disk current and ring current among the three catalysts, demonstrating that Ta2O5/F−C shows an excellent activity for two-electron ORR that is better than Ta2O5/F−C−0.5 and Ta2O5/F−C−2. The Ta2O5/F−C exhibits a high selectivity of H2O2 of 80.8% at 0.49 V vs. RHE, and the selectivity is close to 80% within a wide potential range of 0.25~0.55 V. This catalyst also displays an excellent oxygen reduction activity with the onset potential of 0.78 V vs. RHE, which shows almost no overpotential. These efficient activity of Ta2O5/F−C surpass those of most other carbon support metal oxides in alkaline media (Table S1). The different selectivity shown in Fig. S6b can illustrate the feed ratio of C and Ta also play a nonnegligible role in the activity of such composites and the feed ratio of 1:1 is optimal. In Fig. 6a, F−C shows the largest disk current and the most positive onset potential among all catalysts, but the ring current is lower than Ta2O5/F−C. This unusual performance can be explained that the F−C tends to take four-electron ORR pathway to generate H2O. The H2O2 selectivity of Ta2O5/F−C is better than that of F−C and Ta2O5, which demonstrates that F−C and Ta2O5 propose a synergistic catalytic effect. Both the ring and disk currents were also obtained by RRDE in a neutral condition (0.1 M PBS solution with PH = 8) for Ta2O5/F−C and F−C (Fig. S7a). In Fig. S7b, Ta2O5/F−C shows a considerable performance for H2O2 production with the selectivity of 68.5%, compared with the selectivity of 43.6% for F−C. Figs. 6c and S8 are the Tafel slopes for all the catalysts in different electrolytes. Obviously, Ta2O5/F−C has the lowest Tafel slope in alkaline electrolytes, so it shows the best kinetic rate than other catalysts. To evaluate its stability, a constant potential hold stability test at 0.5 V vs. RHE was performed by rotating ring-disk electrode at 1600 rpm. Fig. 6d shows that the ring and disk current density is rather stable during 10 h electrolysis for Ta2O5/F−C. Fig. S9 shows the H2O2 selectivity and electron transfer number (ETN) of the catalysts after 10 h electrolysis. The H2O2 selectivity kept around 80% and ETN is around 2.3 after electrolysis for 10 h.

    In summary, a synergistic effect is proposed to enhance the performance of Ta2O5/F−C composite catalysts for efficient H2O2 electrosynthesis. The electrical conductivity of Ta2O5 was significantly increased by forming Ta2O5/F−C interface. The Ta2O5/F−C catalyst shows an excellent selectivity more than 80% and long-term stability for H2O2 production. The greatly enhanced performance of Ta2O5/F−C compared to F−C (selectivity of 59%) Ta2O5-800 (current density of 0.85 mA/cm2) counterparts were ascribed to the synergistic effect between F−C and Ta2O5. This work could provide a new way to rational design of carbon-supported metal oxide catalysts for H2O2 electrosynthesis.


    1. [1]

      Zhou, W.; Meng, X.; Gao, J.; Alshawabkeh, A. N. Hydrogen peroxide generation from O2 electroreduction for environmental remediation: a state-of-the-art review. Chemosphere 2019, 225, 588−607. doi: 10.1016/j.chemosphere.2019.03.042

    2. [2]

      Campos-Martin, J. M.; Blanco-Brieva, G.; Fierro, J. L. Hydrogen peroxidesynthesis: an outlook beyond the anthraquinone process. Angew. Chem. Int. Ed. 2006, 45, 6962−6984. doi: 10.1002/anie.200503779

    3. [3]

      Yi, Y.; Wang, L.; Li, G.; Guo, H. A review on research progress in the direct synthesis of hydrogen peroxide from hydrogen and oxygen: noble-metal catalytic method, fuel-cell method and plasma method. Catal. Sci. Technol. 2016, 6, 1593−1610. doi: 10.1039/C5CY01567G

    4. [4]

      Drogui, P.; Elmaleh, S.; Rumeau, M.; Bernard, C.; Rambaud, A. Hydrogen peroxide production by water electrolysis: application to disinfection. J. Appl. Electrochem. 2001, 31, 877−882. doi: 10.1023/A:1017588221369

    5. [5]

      Jiang, Y.; Ni, P.; Chen, C.; Lu, Y.; Yang, P.; Kong, B.; Fisher, A.; Wang, X. Selective electrochemical H2O2 production through two-electron oxygen electrochemistry. Adv. Energy Mater. 2018, 8, 180190−9.

    6. [6]

      Lane, B. S.; Burgess, K. Metal-catalyzed epoxidations of alkenes with hydrogen peroxide. Chem. Rev. 2003, 103, 2457−2474. doi: 10.1021/cr020471z

    7. [7]

      Pan, Z.; Wang, K.; Wang, Y.; Tsiakaras, P.; Song, S. In-situ electrosynthesis of hydrogen peroxide and wastewater treatment application: a novel strategy for graphite felt activation. Appl. Catal. B-Environ. 2018, 237, 392−400. doi: 10.1016/j.apcatb.2018.05.079

    8. [8]

      García-Serna, J.; Moreno, T.; Biasi, P.; Cocero, M. J.; Mikkola, J. P.; Salmi, T. O. Engineering in direct synthesis of hydrogen peroxide: targets, reactors and guidelines for operational conditions. Green Chem. 2014, 16, 2320−2343. doi: 10.1039/c3gc41600c

    9. [9]

      Melchionna, M.; Fornasiero, P.; Prato, M. The rise of hydrogen peroxide as the main product by metal-free catalysis in oxygen reductions. Adv.  Mater. 2019, 31, 18029−20.

    10. [10]

      Yang, S.; Verdaguer-Casadevall, A.; Arnarson, L.; Silvioli, L.; Čolić, V.; Frydendal, R.; Rossmeisl, J.; Chorkendorff, I.; Stephens, I. E. Toward the decentralized electrochemical production of H2O2: a focus on the catalysis. Acs Catal. 2018, 8, 4064−4081. doi: 10.1021/acscatal.8b00217

    11. [11]

      Perry, S. C.; Pangotra, D.; Vieira, L.; Csepei, L. I.; Sieber, V.; Wang, L.; de León, C. P.; Walsh, F. C. Electrochemical synthesis of hydrogen peroxide from water and oxygen. Nat. Rev. Chem. 2019, 3, 442−458. doi: 10.1038/s41570-019-0110-6

    12. [12]

      Zhang, J.; Zhang, H.; Cheng, M. J.; Lu, Q. Tailoring the electrochemical production of H2O2: strategies for the rational design of high-performance electrocatalysts. Small 2019, 190284−5.

    13. [13]

      Blanco-Brieva, G.; de Frutos Escrig, M. P.; Campos-Martin, J. M.; Fierro, J. L. Direct synthesis of hydrogen peroxide on palladium catalyst supported on sulfonic acid-functionalized silica. Green Chem. 2010, 12, 1163−1166. doi: 10.1039/c003700a

    14. [14]

      Pritchard, J. C.; He, Q.; Ntainjua, E. N.; Piccinini, M.; Edwards, J. K.; Herzing, A. A.; Carley, A. F.; Moulijn, J. A.; Kiely, C. J.; Hutchings, G. J. The effect of catalyst preparation method on the performance of supported Au–Pd catalysts for the direct synthesis of hydrogen peroxide. Green Chem. 2010, 12, 915−921. doi: 10.1039/b924472g

    15. [15]

      Xia, C.; Xia, Y.; Zhu, P.; Fan, L.; Wang, H. Direct electrosynthesis of pure aqueous H2O2 solutions up to 20% by weight using a solid electrolyte. Science 2019, 366, 226−231. doi: 10.1126/science.aay1844

    16. [16]

      Isaka, Y.; Oyama, K.; Yamada, Y.; Suenobu, T.; Fukuzumi, S. Photocatalytic production of hydrogen peroxide from water and dioxygen using cyano-bridged polynuclear transition metal complexes as water oxidation catalysts. Catal. Sci. Technol. 2016, 6, 681−684. doi: 10.1039/C5CY01845E

    17. [17]

      Weng, B.; Wu, J.; Zhang, N.; Xu, Y. J. Observing the role of graphene in boosting the two-electron reduction of oxygen in graphene-WO3 nanorod photocatalysts. Langmuir 2014, 30, 5574−5584. doi: 10.1021/la4048566

    18. [18]

      Siahrostami, S.; Verdaguer-Casadevall, A.; Karamad, M.; Deiana, D.; Malacrida, P.; Wickman, B.; Escudero-Escribano, M.; Paoli, E. A.; Frydendal, R.; Hansen, T. W. Enabling direct H2O2 production through rational electrocatalyst design. Nat. mater. 2013, 12, 1137−1143. doi: 10.1038/nmat3795

    19. [19]

      Aveiro, L. R.; da Silva, A. G.; Antonin, V. S.; Candido, E. G.; Parreira, L. S.; Geonmonond, R. S.; de Freitas, I. C.; Lanza, M. R.; Camargo, P. H.; Santos, M. C. Carbon-supported MnO2 nanoflowers: introducing oxygen vacancies for optimized volcano-type eelectrocatalytic activities towards H2O2 generation. Electrochim. Acta 2018, 268, 101−110. doi: 10.1016/j.electacta.2018.02.077

    20. [20]

      Cheng, F.; Su, Y.; Liang, J.; Tao, Z.; Chen, J. MnO2-based nanostructures as catalysts for electrochemical oxygen reduction in alkaline media. Chem. Mater. 2010, 22, 898−905. doi: 10.1021/cm901698s

    21. [21]

      Liang, Y.; Li, Y.; Wang, H.; Zhou, J.; Wang, J.; Regier, T.; Dai, H. Co3O4 nanocrystals on graphene as a synergistic catalyst for oxygen reduction reaction. Nat. Mater. 2011, 10, 780−786. doi: 10.1038/nmat3087

    22. [22]

      Moraes, A.; Assumpção, M.; Papai, R.; Gaubeur, I.; Rocha, R. D. S.; Reis, R.; Calegaro, M. L.; Lanza, M. R. D. V.; Santos, M. C. D. Use of a vanadium nanostructured material for hydrogen peroxide electrogeneration. J. Electroanal. Chem. 2014, 719, 127−132. doi: 10.1016/j.jelechem.2014.02.009

    23. [23]

      Ye, Y.; Kuai, L.; Geng, B. A template-free route to a Fe3O4-Co3O4 yolk-shell nanostructure as a noble-metal free electrocatalyst for ORR in alkaline media. J. Mater. Chem. 2012, 22, 19132−19138. doi: 10.1039/c2jm33893a

    24. [24]

      Xu, A.; Han, W.; Li, J.; Sun, X.; Shen, J.; Wang, L. Electrogeneration of hydrogen peroxide using Ti/IrO2-Ta2O5 anode in dual tubular membranes electro-Fenton reactor for the degradation of tricyclazole without aeration. Chem. Eng. J. 2016, 295, 152−159. doi: 10.1016/j.cej.2016.03.001

    25. [25]

      Barros, W. R.; Wei, Q.; Zhang, G.; Sun, S.; Lanza, M. R.; Tavares, A. C. Oxygen reduction to hydrogen peroxide on Fe3O4 nanoparticles supported on printex carbon and graphene. Electrochim. Acta 2015, 162, 263−270. doi: 10.1016/j.electacta.2015.02.175

    26. [26]

      Carneiro, J. F.; Rocha, R. S.; Hammer, P.; Bertazzoli, R.; Lanza, M. Hydrogen peroxide electrogeneration in gas diffusion electrode nanostructured with Ta2O5. Appl. Catal. A-Gen. 2016, 517, 161−167. doi: 10.1016/j.apcata.2016.03.013

    27. [27]

      Oh, T.; Kim, J. Y.; Shin, Y.; Engelhard, M.; Weil, K. S. Effects of tungsten oxide addition on the electrochemical performance of nanoscale tantalum oxide-based electrocatalysts for proton exchange membrane (pem) fuel cells. J. Power Sources 2011, 196, 6099−6103. doi: 10.1016/j.jpowsour.2011.03.058

    28. [28]

      Chisaka, M.; Ishihara, A.; Suito, K.; Ota, K. I.; Muramoto, H. Oxygen reduction reaction activity of nitrogen-doped titanium oxide in acid media. Electrochim. Acta 2013, 88, 697−707. doi: 10.1016/j.electacta.2012.10.137

    29. [29]

      Kim, J. Y.; Oh, T. K.; Shin, Y.; Bonnett, J.; Weil, K. S. A novel non-platinum group electrocatalyst for pem fuel cell application. Int. J. Hydrogen Energy 2011, 36, 4557−4564. doi: 10.1016/j.ijhydene.2010.05.016

    30. [30]

      Assumpção, M. H. M. T.; Moraes, A.; De Souza, R.; Calegaro, M.; Lanza, M.; Leite, E.; Cordeiro, M.; Hammer, P.; Santos, M. C. D. Influence of the preparation method and the support on H2O2 electrogeneration using cerium oxide nanoparticles. Electrochim. Acta 2013, 111, 339−343. doi: 10.1016/j.electacta.2013.07.187

    31. [31]

      Carneiro, J. F.; Paulo, M. J.; Siaj, M.; Tavares, A. C.; Lanza, M. R. Nb2O5 nanoparticles supported on reduced graphene oxide sheets as electrocatalyst for the H2O2 electrogeneration. J. Catal. 2015, 332, 51−61. doi: 10.1016/j.jcat.2015.08.027

    32. [32]

      Carneiro, J. F.; Trevelin, L. C.; Lima, A. S.; Meloni, G. N.; Bertotti, M.; Hammer, P.; Bertazzoli, R.; Lanza, M. R. Synthesis and characterization of ZrO2/C as electrocatalyst for oxygen reduction to H2O2. Electrocatalysis 2017, 8, 189−195. doi: 10.1007/s12678-017-0355-0

    33. [33]

      Li, Z.; Liu, J.; Li, J.; Shen, J. Template free synthesis of crystallized nanoporous F-Ta2O5 spheres for effective photocatalytic hydrogen production. Nanoscale 2012, 4, 3867−70. doi: 10.1039/c2nr30721a

    34. [34]

      Lu, Z.; Chen, G.; Siahrostami, S.; Chen, Z.; Liu, K.; Xie, J.; Liao, L.; Wu, T.; Lin, D.; Liu, Y.; Jaramillo, T. F.; Nørskov, J. K.; Cui, Y. High-efficiency oxygen reduction to hydrogen peroxide catalysed by oxidized carbon materials. Nat. Catal. 2018, 1, 156−162. doi: 10.1038/s41929-017-0017-x

    35. [35]

      Yuan, D.; Wei, Z.; Han, P.; Yang, C.; Huang, L.; Gu, Z.; Ding, Y.; Ma, J.; Zheng, G. Electron distribution tuning of fluorine-doped carbon for ammonia electrosynthesis. J. Mater. Chem. A 2019, 7, 16979−16983. doi: 10.1039/C9TA04141A

    36. [36]

      Guo, L.; He, H.; Ren, Y.; Wang, C.; Li, M. Core-shell SiO@ F-doped C composites with interspaces and voids as anodes for high-performance lithium-ion batteries. Chem. Eng. J. 2018, 335, 32−40. doi: 10.1016/j.cej.2017.10.145

    37. [37]

      Li, Z.; Xu, J.; Wang, J.; Niu, D.; Hu, S.; Zhang, X. Well-dispersed amorphous Ta2O5 chemically grafted onto multi-walled carbon nanotubes for high-performance lithium sulfur battery. Int. J. Electrochem. Sci 2019, 14, 6628−6642.

    38. [38]

      Yu, X.; Li, W.; Li, Z.; Liu, J.; Hu, P. Defect engineered Ta2O5 nanorod: one-pot synthesis, visible-light driven hydrogen generation and mechanism. Appl. Catal. B-Environ. 2017, 217, 48−56. doi: 10.1016/j.apcatb.2017.05.024

    39. [39]

      Zhuang, Y.; Seong, J. G.; Do, Y. S.; Jo, H. J.; Cui, Z.; Lee, J.; Lee, Y. M.; Guiver, M. D. Intrinsically microporous soluble polyimides incorporating Tröger's base for membrane gas separation. Macromolecules 2014, 47, 3254−3262. doi: 10.1021/ma5007073

    40. [40]

      Sun, Y.; Sinev, I.; Ju, W.; Bergmann, A.; Dresp, S.; Kühl, S.; Spoeri, C.; Schmies, H.; Wang, H.; Bernsmeier, D.; Paul, B.; Schmack, R.; Kraehnert, R.; Cuenya, B. R.; Strasser P. Efficient electrochemical hydrogen peroxide production from molecular oxygen on nitrogen-doped mesoporous carbon catalysts. ACS Catal. 2018, 8, 2844−2856. doi: 10.1021/acscatal.7b03464

    41. [41]

      Chen, S.; Chen, Z.; Siahrostami, S.; Higgins, D.; Nordlund, D.; Sokaras, D.; Kim, T. R.; Liu, Y.; Yan, X.; Nilsson, E. Designing boron nitride islands in carbon materials for efficient electrochemical synthesis of hydrogen peroxide. J. Am. Chem. Soc. 2018, 140, 7851−7859. doi: 10.1021/jacs.8b02798

  • Figure 1  SEM images of (a) Ta2O5 spheres, (b) F doped carbon and (c) Ta2O5/F−C

    Figure 2  TEM images of (a) Ta2O5 spheres, (b) F doped carbon and (c) Ta2O5/F−C

    Figure 3  XRD patterns of Ta2O5/F−C and Ta2O5-800

    Figure 4  X-ray photoelectron spectra of (a) C 1s, (b) O 1s, (c) F 1s and (d) Ta 4f for Ta2O5/F−C, (e) Nitrogen adsorption-desorption isotherms for Ta2O5/F−C and (f) BJH desorption pore-size distribution of Ta2O5/F−C

    Figure 5  (a) CV and (b) LSV curves of Ta2O5/F−C in N2 and O2 saturated 0.1 M KOH solution

    Figure 6  (a) LSV curves of Ta2O5/F−C, Ta2O5-800 and F−C (scan rate of 10 mV/s with a rotation speed of 1600 rpm). (b) H2O2 selectivity of Ta2O5/F−C, Ta2O5-800 and F−C in O2 saturated 0.1 M KOH electrolyte. (c) Tafel slopes for corresponding catalysts in O2 saturated 0.1 M KOH electrolyte. (d) Stability test for Ta2O5/F−C. All plots were subtracted from the nitrogen background current and the pH of test condition is 13

  • 加载中
计量
  • PDF下载量:  1
  • 文章访问数:  459
  • HTML全文浏览量:  20
文章相关
  • 发布日期:  2021-02-01
  • 收稿日期:  2020-03-23
  • 接受日期:  2020-04-20
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