S-anion effects on the improvement of adsorption capacity and performance for benzyl alcohol electro-oxidation catalysts

Yufeng ZHANG Haotian QI Jingya ZHONG Leiming LANG Guojun YUAN Siqi LU Haiying WANG Guangxiang LIU

Citation:  Yufeng ZHANG, Haotian QI, Jingya ZHONG, Leiming LANG, Guojun YUAN, Siqi LU, Haiying WANG, Guangxiang LIU. S-anion effects on the improvement of adsorption capacity and performance for benzyl alcohol electro-oxidation catalysts[J]. Chinese Journal of Inorganic Chemistry, 2025, 41(12): 2591-2600. doi: 10.11862/CJIC.20250282 shu

硫阴离子效应对苯甲醇电氧化催化剂的吸附能力优化及性能提升

    通讯作者: 卢思奇, siqi_lu@126.com
    王海英, wanghaiying@nju.edu.cn
    刘光祥, liugx@njxzc.edu.cn
  • 基金项目:

    国家自然科学基金 22405229

    配位化学全国重点实验室开放课题 SKLCC2514

摘要: 通过液相法制备了Ni2CoS4, 并将其应用于催化苯甲醇电氧化反应(BAOR)。结果表明, Ni2CoS4具有理想的催化活性[1.40 V(vs RHE)下的电流密度达到271 mA·cm-2]和长效稳定性。其中, 硫阴离子效应能够调节催化剂表面的电荷分布, 从而增强Co位点对OH-的额外吸附能力。结合表征和理论计算可以发现, 这一过程可以增加OH*中间体的浓度, 加速Ni位点的活化过程, 最终实现整体活性和稳定性的提升。

English

  • The electrochemical production of hydrogen fuel is currently considered an important method for addressing energy and climate issues[1-3]. However, traditional water electrolysis is limited by the relatively slow kinetics of the oxygen evolution reaction at the anode, and its overall efficiency in energy utilization still needs to be improved. By changing the anode substrate to aliphatic compounds, it is possible to significantly reduce energy consumption and synthesize more valuable organic products[4-8]. Among all organic substrates, the benzyl alcohol electro-oxidation reaction (BAOR) shows the advantages of low production cost and a wide range of raw material sources[9-10]. Besides, the oxidation product of benzyl alcohol, benzoic acid, is widely used in numerous fields, including medicine, food, and the chemical industry[11-13].

    Ni-based non-precious metal catalysts offer certain advantages, including low cost and strong tolerance to poisoning[14-15]. Meanwhile, Ni-based catalysts need to adsorb OH- at a higher potential to effectively catalyze the electrochemical oxidation of alcohols[16-17]. Therefore, reducing the activation potential and enhancing the overall activity are important means for developing Ni-based BAOR catalysts. Adding other cations is a common method to enhance the activity of Ni-based catalysts[9,17-20]. In particular, Ni-Co binary metal compounds are a common method for enhancing the performance of Ni-based electrocatalysts[21-23]. The Co site has a strong adsorption effect on OH-, which is conducive to the use of various anode electrocatalysts[23-24]. Therefore, adding Co cations is an important means for optimizing Ni-based anode catalysts.

    By taking advantage of the anion effect, the charge distribution on the catalyst surface can be adjusted, thereby further enhancing the strengthening function of Co[9,25-26]. Therefore, we introduced S anions into Ni-Co bimetallic compounds, enhanced the additional adsorption of Co, and resulted in the formation of a highly active Ni2CoS4 BAOR catalyst. The specific activity of the Ni2CoS4 BAOR catalyst was 271 mA·cm-2 at 1.40 V (vs RHE), 3.3 times of Ni2Co(OH)x (83 mA·cm-2), and 7.7 times of Ni(OH)2 (35 mA·cm-2). By S- anion effects, the charge distribution among different metals on the surface of the catalyst was re-adjusted, strengthening the positive charge distribution on the Co surface and enhancing its OH- adsorption ability. Finally, the overall catalytic performance of Ni2CoS4 for BAOR was improved.

    Cobalt chloride hexahydrate (CoCl2·6H2O, 99%, Macklin), nickel chloride hexahydrate (NiCl2·6H2O, 99%, Macklin), potassium hydroxide (KOH, 99.999%, Aladdin), thiourea (99%, Aladdin), urea (99%, Aladdin), benzyl alcohol (99.9%, Aladdin) were obtained commercially and used without further purification.

    The X‐ray diffraction (XRD) patterns were obtained on a Rigaku D/Max 2500 VB2+/PC X-ray powder diffractometer equipped with Cu radiation (λ=0.154 nm) operating at 40 kV and 40 mA with a 2θ range of 20° to 90°. The transmission electron microscopy (TEM) image was performed on a 2100F/F200X transmission electron microscope (200 kV). The valence states of the elements were recorded on a Thermo Fisher Escalab Xi+ X-ray photoelectron spectrometer (XPS) with a monochromatic Al X-ray source. The binding energies derived from XPS measurements were calibrated to the C1s at 284.8 eV. The morphologies were obtained on a Thermo Fisher Scientific Apreo 2C scanning electron microscope. The elemental contents were tested by an Agilent 5100 inductively coupled plasma optical emission spectrometer (ICP-OES).

    1.2.1   Synthesis of Ni2CoS4

    A mixture was prepared by combining 0.48 g NiCl2·6H2O, 0.24 g CoCl2·6H2O, 0.4 g of thiourea, and 28 mL of deionized water. Three pieces of Ni foams (1 cm×2.5 cm) were immersed in the mixture, which was then transferred into a 50 mL Teflon-lined stainless-steel autoclave. The sealed autoclave was heated at 120 ℃ for 8 h. After the reaction, the Ni foam substrates with deposited Ni2CoS4 were taken out, rinsed thoroughly, and dried, yielding the final Ni2CoS4 product.

    1.2.2   Syntheses of Ni2Co(OH)x and Ni(OH)2

    A homogeneous solution was prepared by dissolving 0.48 g NiCl2·6H2O, 0.24 g CoCl2·6H2O, and 0.6 g urea in 28 mL of deionized water. Three pieces of Ni foams (1 cm×2.5 cm) were immersed in the solution, which was then transferred into a 50 mL Teflon-lined stainless‐steel autoclave. The sealed autoclave was heated at 120 ℃ for 8 h. After the reaction, the Ni foam substrates were collected, washed with ethanol and deionized water, and dried at 50 ℃ for 2 h to obtain Ni2Co(OH)x. Using the same procedure but without the addition of CoCl2·6H2O, pure Ni(OH)2 was synthesized.

    BAOR tests were performed in a standard three-electrode configuration using a Bio-Logic workstation. The electrolyte consisted of a 1.0 mol·L-1 KOH solution containing 1.0 mol·L-1 benzyl alcohol. A graphite rod and an Ag/AgCl electrode were employed as the counter and reference electrodes, respectively. The working electrode was prepared by depositing the catalyst onto Ni foam, with a fixed geometric area of 1 cm×1 cm. All measurements were carried out at a scan rate of 50 mV·s-1. All the electrochemical reactions were carried out at 30 ℃ under a N2 atmosphere. And through rotor agitation, the transfer of substances on the catalyst surface was accelerated.

    We have employed the Vienna Ab Initio Package (VASP) to perform all the density functional theory (DFT) calculations within the generalized gradient approximation (GGA) using the Perdew‐Burke‐ Ernzerhof (PBE) formulation. We have chosen the projected augmented wave (PAW) potentials to describe the ionic cores and take valence electrons into account using a plane wave basis set with a kinetic energy cutoff of 500 eV. Partial occupancies of the Kohn-Sham orbitals were allowed using the Gaussian smearing method and a width of 0.05 eV. The electronic energy was considered self-consistent when the energy change was smaller than 10-5 eV. A geometry optimization was considered convergent when the force change was smaller than 0.5 eV·nm-1. Grimme′s DFT‐D3 methodology was used to describe the dispersion interactions. Adsorption energies (Eads) were determined using the formula Eads=Ead/sub-Ead-Esub, where Ead/sub represents the total energy of the optimized adsorbate/substrate system, Ead is the energy of the isolated adsorbate, and Esub is the energy of the pristine substrate.

    Ni2CoS4 was synthesized through the hydrothermal method. Ni foam, NiCl2·6H2O, CoCl2·6H2O, and thiourea were added to a Teflon vessel and reacted for 8 h. Thiourea can provide the S element and regulate the pH. After that, the Ni-Co bimetallic sulfides precipitated on the Ni foam. To better study the structure and composition of this material, we conducted a series of characterization experiments.

    Firstly, XRD characterization was conducted, as shown in Fig. 1a. The obtained catalysts matched well with the peaks of Ni2CoS4 (PDF No.24-0334), which could be preliminarily concluded that Ni2CoS4 was synthesized. To further confirm the results, we also conducted ICP-OES characterization, and the results are shown in Table S1 (Supporting information). The molar ratio of Ni and Co was approximately 2∶1, consistent with the XRD results.

    Figure 1

    Figure 1.  Characterization of Ni2CoS4: (a) XRD patterns; (b, c) SEM images; (d) elemental mappings of Ni, Co, and S on the surface of Ni foam; (e) HRTEM image; (f) elemental mappings of Ni, Co, and S in HRTEM image

    We also conducted the SEM characterization, and the results are shown in Fig. 1b and 1c. It could be observed that there were nanorods with abundant defects on the surface of the Ni foam, and the diameter was approximately 100 nm. The rich defect structure can provide more active sites and a larger specific area. We also conducted EDS (energy dispersive X-ray spectroscopy) mappings (Fig. 1d). It can be found that Ni, Co, and S were uniformly distributed on the surface of Ni foam. Then, we placed the Ni foam loaded with Ni2CoS4 in ethanol for ultrasonic treatment. The resulting solution was loaded on an ultrathin copper net and subjected to high-resolution TEM (HRTEM) characterization (Fig. 1e). The lattice fringe spacing was 0.28 nm, corresponding to the Ni2CoS4 (311) plane. In HRTEM image, the uniform distribution of Ni, Co, and S could be observed. This conclusion was consistent with previous characterizations of XRD and ICP-OES. By integrating the relevant characterization methods, it can be confirmed that the successful synthesis of Ni2CoS4.

    Ni(OH)2 and Ni2Co(OH)x were also synthesized as control samples. For Ni(OH)2, the XRD pattern was shown in Fig. S1, which matched well with Ni(OH)2 (PDF No.22-0444). We also conducted SEM images (Fig. S2), revealing the coverage of the nanosheet on the surface of the Ni foam. Ni and O were evenly distributed in the Ni foam as well. The XRD characterization of Ni2Co(OH)x (Fig. S3) showed that the peaks shifted compared to Ni(OH)2. The result of ICP-OES indicates that the molar ratio of Ni to Co was approximately 2∶1. SEM images (Fig. S4) revealed nanorods with numerous surface defects. EDS mappings in Fig. S4c demonstrated the uniform distribution of Ni, Co, and O. Through ICP-OES, XRD, and SEM, the synthesis of Ni-Co hydroxides (Ni2Co(OH)x) can be preliminarily inferred. To further clarify the composition of the material, we also conducted HRTEM characterization (Fig. S5). The lattice fringe spacing was 0.22 nm between the Ni(OH)2 (101) plane and the Co(OH)2 (101) plane. Furthermore, by using EDS mappings at high magnification, Ni and Co were uniformly distributed. These results verified the successful synthesis of Ni2Co(OH)x.

    To study the performance of Ni-based catalysts with different components for BAOR, we conducted electrochemical tests in an electrolyte solution with 1.0 mol·L-1 KOH and 0.1 mol·L-1 benzyl alcohol. The loading amounts of Ni2CoS4, Ni2Co(OH)x, and Ni(OH)2 were 10.9, 10.6, and 10.8 mg·cm-2, respectively. In a 1.0 mol·L-1 KOH solution without benzyl alcohol, it could be observed that the polarization curve exhibited a weak current at a low potential. When 0.1 mol·L-1 benzyl alcohol was added, a significant anode current could be observed at 1.30 V (vs RHE) (Fig. S6). It can be concluded that the obtained Ni2CoS4 showed ideal catalytic activity for BAOR, and side reactions could be ignored.

    We tested and compared the polarization curves of Ni2CoS4, Ni2Co(OH)x, and Ni(OH)2 in an electrolyte with 1.0 mol·L-1 KOH and 0.1 mol·L-1 benzyl alcohol (Fig. 2a). Compared with Ni2Co(OH)x and Ni(OH)2, Ni2CoS4 showed a higher current density at the same potential, indicating better activity. The addition of Co can enhance the BAOR activity of the Ni-based catalyst, but the performance improvement was not significant. In contrast, Ni2CoS4 exhibited obviously higher current density and a lower onset potential than the other two catalysts. The specific activities of Ni2CoS4, Ni2Co(OH)x, and Ni(OH)2 at 1.40 V (vs RHE) are summarized in Fig. 2b. The specific activity of Ni2Co(OH)x was 83 mA·cm-2, which was slightly higher than that of Ni(OH)2 (35 mA·cm-2). Notably, Ni2CoS4 could reach 271 mA·cm-2 at 1.40 V (vs RHE), which was 7.7 times that of Ni(OH)2.

    Figure 2

    Figure 2.  Electrochemical performance of Ni2CoS4, Ni2Co(OH)x, and Ni(OH)2 in 1.0 mol·L-1 KOH with 1.0 mol·L-1 benzyl alcohol: (a) polarization curves; (b) current densities at 1.40 V (vs RHE); (c) Nyquist plots; (d) Cdl; (e) chronopotentiometry curves at 50 mA·cm-2 after IR correction

    We also calculated the turnover frequency (TOF), as shown in Fig. S7a. It can be observed that at the same potential, Ni2Co(OH)x showed a higher TOF compared to Ni(OH)2, while the TOF of Ni2CoS4 showed a more significant improvement. To conduct a quantitative analysis, we also calculated the TOF of the three catalysts at 1.40 V (vs RHE) (Fig. S7b), and the TOF values of Ni2CoS4, Ni2Co(OH)x, and Ni(OH)2 were 0.31, 0.11, and 0.05 s-1, respectively. This result indicated that the active center of Ni2CoS4 had the highest intrinsic activity. By combining the linear sweep voltammetry (LSV) curve and the TOF calculation results, it can be observed that, on the one hand, the Co-cation effects can enhance the intrinsic activity of the Ni sites; on the other hand, the S-anion effects can further strengthen the cationic effect of Co, enhancing the optimization effect of Co on the charge distribution and activity of the catalyst, achieving further improvement in overall performance.

    By comparing the electrochemical impedance spectroscopy (EIS) of the three catalysts (Fig. 2c), the kinetic processes of the three catalysts can be studied. The EIS test was set to a frequency of 105 Hz to 50 mHz at a potential of 1.40 V (vs RHE). The faster kinetic process corresponds to a smaller charge transfer resistance (Rct). The Rct value of Ni2CoS4 was only 1.34 Ω, smaller than those of Ni2Co(OH)x and Ni(OH)2, indicating a faster kinetic process. By analyzing the cyclic voltammogram (CV) curves obtained at different scan rates within the range of 1.20 to 1.40 V (vs RHE) (Fig. S8), the Cdl of the catalyst can be studied. A larger Cdl corresponds to a higher electrochemically active area. Based on the fitting results, the Cdl values for Ni2CoS4, Ni2Co(OH)x, and Ni(OH)2 were 23.1, 14.4, and 8.8 mF·cm-2, respectively (Fig. 2d). The multi-defect structure of Ni2CoS4 significantly increased the contact area between the catalyst and the reactants, which is beneficial for the catalytic process of BAOR.

    Stability is another crucial indicator for catalyst performance. At the current density of 50 mA·cm-2, we conducted a chronopotentiometry test to study the long-term stability of Ni2CoS4, Ni2Co(OH)x, and Ni(OH)2. As shown in Fig. 2e, during the 10‐hour test, both Ni2Co(OH)x and Ni(OH)2 exhibited significant performance degradation. However, the working potential of Ni2CoS4 remained consistently at around 1.34 V (vs RHE), with a reduction of only 10 mV, demonstrating significantly higher stability.

    Based on the relevant electrochemical test results, it can be concluded that the Co cation can optimize the activity of Ni-based BAOR catalyst to a certain extent, while the S anion can further enhance the performance.

    The alcohol electro-oxidation reaction catalyzed by Ni-based catalysts needs to be carried out under alkaline conditions[27-28]. The OH- absorbed on Ni-based active sites to form higher-valent Ni(Ⅲ) centers, which further catalyze the electrochemical oxidation of benzyl alcohol to benzoic acid[27]. The stronger OH-adsorption ability can help to improve the overall performance of Ni-based BAOR catalysts. By analyzing the Bode diagrams obtained from EIS at different potentials, the changes of oxidation species on the surface of Ni-based catalyst can be studied[18,29-30]. The phase angle change at lower frequencies often corresponds to the change of oxidation species on the surface. A lower phase angle variation potential is often associated with a lower adsorption potential and a lower catalyst activation potential[29,31]. The Bode diagrams of Ni2CoS4, Ni2Co(OH)x, and Ni(OH)2 catalysts are shown in Fig. 3, and the phase angle change potentials can be obtained. The Co cations reduced the phase angle change potential of the Ni-based catalysts. The introduction of S anions could further decrease the external electrical energy required for this process. This demonstrates that the synergistic anion-cation effects of S and Co can effectively reduce the energy required for the adsorption process, thereby enhancing the overall activity of the catalyst.

    Figure 3

    Figure 3.  Bode plots of (a) Ni2CoS4, (b) Ni2Co(OH)x, and (c) Ni(OH)2 in 1.0 mol·L-1 KOH with 1.0 mol·L-1 benzyl alcohol

    The density of states (DOS) of the three catalysts was calculated to study the effects of S and Co on the charge distribution (Fig. 4a-4c, and calculation models are shown in Fig. S9). Compared with Ni(OH)2 without Co, the total DOS of Ni-Co diohydroxide provided a higher electron cloud density near the Fermi level, which is beneficial for their adsorption ability. As for Ni2CoS4, due to the lattice expansion of S, the d-band center of Ni2CoS4 (-1.115 eV) was also closer to the Fermi level than that of Ni2Co(OH)x (-1.310 eV). This electron distribution facilitates the adsorption of OH* on the catalyst, thereby reducing the reaction energy barrier[9,32].

    Figure 4

    Figure 4.  DOS of (a) Ni(OH)2, (b) Ni2Co(OH)x, and (c) Ni2CoS4; (d) Calculated model for OH- adsorption of Ni2CoS4; (e) Calculated adsorption energy of OH- of Ni(OH)2, Ni2Co(OH)x, and Ni2CoS4

    In addition, we also calculated the adsorption energy of OH- on Ni2CoS4, Ni2Co(OH)x, and Ni(OH)2, and the calculation models are shown in Fig. 4d and S10. The adsorption energy of OH- in Ni2Co(OH)x was only -0.839 eV, lower than -0.398 eV of Ni(OH)2. Through calculation, it can be found that the Co sites showed a strong adsorption capacity for OH-, which can provide additional active centers and increase the concentration of OH* on the surface. Meanwhile, in Ni2CoS4, the adsorption energy of OH- on the Co sites was even lower (-1.724 eV). The redistribution of charges caused by the S enhanced the additional adsorption capacity of Co and accelerated the activation process. The results of the adsorption energy also confirm the calculations of DOS.

    We investigated the oxidation states and chemical environments through XPS. Fig. 5a shows the Ni2p spectra of Ni2CoS4, Ni2Co(OH)x, and Ni(OH)2. It can be observed that a clear Niδ+2p3/2 peak at 852.7 eV in Ni2CoS4[33-35], but could not be detected in Ni2Co(OH)x and Ni(OH)2. The peaks at 852.7 eV are mainly observed in Ni-based sulfides[17,35-37]. It might be due to the charge redistribution caused by the S-anion effect.

    Figure 5

    Figure 5.  High-resolution XPS spectra of (a) Ni2p, (b) Co2p, and (c) S2p in Ni-based catalysts; High-resolution XPS spectra of (d) Ni2p, (e) Co2p, and (f) S2p in Ni2CoS4 before and after stability test

    Furthermore, we also analyzed the valence states of Co in Ni2CoS4 and Ni2Co(OH)x. As shown in Fig. 5b, for Ni2CoS4, the peak at 780.8 eV can be attributed to Co3+2p3/2, and the characteristic peak of Co2+2p3/2 was located at 782.5 eV[38-40]. In Ni2CoS4, the content (atomic fraction) of Co3+ was 57%, and that of Co2+ was 43%. The content of Co3+ in Ni2CoS4 was higher than that in Ni2Co(OH)x (44%). This indicates that the oxidation state of the Co sites in Ni2CoS4 was more positive and conducive to the adsorption of OH-. Besides, the peaks in Fig. 5c near 162 eV correspond to S—M, where M=Co and Ni[30,33].

    Compared to the XPS characterization results before and after the stability test, we investigated the changes in the valence states of elements on the catalyst surface after a long period of testing. The oxidation state of Ni slightly increased after the reaction, with a positive shift of 0.3 eV. Some Ni sites were partially oxidized, and Niδ+ gradually disappeared (Fig. 5d). Meanwhile, the chemical valence state of Co slightly decreased, and the content (atomic fraction) of Co3+ changed to 54% (Fig. 5e). The S underwent a more significant oxidation, with the content of S—M decreasing and the content of S—O increasing (Fig. 5c, 5f). The changes in the oxidation states of S and Co were likely to be the cause of catalyst deactivation.

    Furthermore, SEM images shown in Fig. S11 after the stability test, it can be observed that after a long period of testing, Ni2CoS4 still maintained a well‐ dispersed nanorod structure, and the microscopic morphology changes were not significant. The XRD patterns of Ni2CoS4 after BAOR (Fig. S12) indicated that the overall composition of Ni2CoS4 was basically consistent with Ni2CoS4 (PDF No.24-0334), and the material composition changes were not obvious. The influences of morphology and phase transformation on the stability of the catalyst were not significant.

    By combining with the adsorption energy calculation, it can be found that the S-anionic effects provide more positive charge centers for Co, thereby effectively enhancing the additional adsorption capacity of Co for OH-. This achieves the enrichment of OH* species at a lower potential, which is beneficial to the overall improvement of activity.

    Combining electrochemical tests, theoretical calculations, and material characterization, we can observe that in the Ni-based BAOR catalyst, the Co cations can utilize their strong adsorption ability for OH- to provide abundant surface species and a lower onset potential for the activation of the Ni active centers. The S-anion effects can achieve the redistribution of the charge, further enhance the additional adsorption capacity of Co sites, and significantly improve the activity of the Ni-based catalyst. By combining the dual regulation of Co and S, an efficient Ni2CoS4 catalyst for BAOR was ultimately obtained.


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

      YIN C, WANG S L, YANG F L, FENG L G. Spontaneous charge redistribution with diverse nucleophilic and electrophilic sites in NiTe/Mo6Te8 for urea-assisted water electrolysis[J]. J. Eenergy Chem., 2025, 106: 340-350 doi: 10.1016/j.jechem.2025.02.042

    2. [2]

      WANG K L, LIU P C, WANG M Z, WEI T R, LU J T, ZHAO X L, JIANG Z Y, YUAN Z M, LIU X J, HE J. Modulating d-d orbitals coupling in PtPdCu medium-entropy alloy aerogels to boost pH-general methanol electrooxidation performance[J]. Chin. Chem. Lett., 2025, 36(4): 110532 doi: 10.1016/j.cclet.2024.110532

    3. [3]

      LI W S, NIE R R, SONG Y, NI L J, WU D, WU G G, CHU R Z, MENG X L. Microwave-driven ethanol steam reforming for low‐ temperature H2 production over the carbon nanotubes supported NiFe-based catalysts[J]. Int. J. Hydrog. Energy, 2025, 101: 490-503 doi: 10.1016/j.ijhydene.2024.12.414

    4. [4]

      XIANG C, LING Y Y, ZHOU Z T, ZHU X Y, XUE F, FENG Z J, WANG Y W, CHENG X Y, WANG M F, CHENG X M. Efficient synergism of concentric ring structures and carbon dots for enhanced methanol electro-oxidation[J]. RSC Adv., 2024, 14(41): 30091-30101 doi: 10.1039/D4RA04685D

    5. [5]

      HOU Y C, DING M W, LIU S K, WU S K, LIN Y C. Ni-substituted LaMnO3 perovskites for ethanol oxidation[J]. RSC Adv., 2014, 4(11): 5329-5339 doi: 10.1039/c3ra46323k

    6. [6]

      LIU S L, ZHOU L L, ZHANG W J, JIN J Y, MU X Q, ZHANG S D, CHEN C Y, MU S C. Stabilizing sulfur vacancy defects by performing "click" chemistry of ultrafine palladium to trigger a high-efficiency hydrogen evolution of MoS2[J]. Nanoscale, 2020, 12(18): 9943-9949 doi: 10.1039/D0NR01693D

    7. [7]

      ZHENG Y, WAN X J, CHENG X, CHENG K, DAI Z F, LIU Z H. Advanced catalytic materials for ethanol oxidation in direct ethanol fuel cells[J]. Catalysts, 2020, 10(2): 22-24

    8. [8]

      HEDAYATINASAB F, GHAFFARINEJAD A. Enhanced ethanol oxidation using nickel-tungsten alloy: Durable catalysts for direct ethanol fuel cells[J]. J Environ. Chem. Eng., 2025, 13(1): 115228 doi: 10.1016/j.jece.2024.115228

    9. [9]

      WEN H, DUAN W, GUO L, WANG Q, FU X, WANG Y H, LI R, JIN B B, DU R, YANG C M, WANG D J. Directing charge transfer in a chemical-bonded Ni/Cd0.7Mn0.3S Schottky heterojunction for selective photocatalytic oxidation of benzyl alcohol structural organic platform molecules coupled with hydrogen evolution reaction[J]. Appl. Catal. B‒Environ. Energy, 2024, 345: 123641 doi: 10.1016/j.apcatb.2023.123641

    10. [10]

      SU W L, ZHENG X Z, XIONG W, OUYANG Y, ZHANG Z, ZENG W W, DUAN H T, CHEN X Y, SU P Y, SUN Z M, YUAN M W. Open active sites in Ni-based MOF with high oxidation states for electrooxidation of benzyl alcohol[J]. Inorg. Chem., 2024, 63(27): 12572-12581 doi: 10.1021/acs.inorgchem.4c01507

    11. [11]

      ZHANG Y F, LIU D, WU H, YANG Z Y, XIA Z X, WU Q H, YU S S, WANG H Y, LANG L M, LIU G X. The metal-support interactions of Cr(OH)3 enhance the performance of supported Au-based benzyl alcohol electrooxidation catalysts[J]. Chem. Commun., 2025, 61(16): 3379-3382 doi: 10.1039/D5CC00245A

    12. [12]

      FENG D, DONG Y, ZHANG L, GE X, ZHANG W, DAI S, QIAO Z A. holey lamellar high-entropy oxide as an ultra-high-activity heterogeneous catalyst for solvent-free aerobic oxidation of benzyl alcohol[J]. Angew. Chem. -Int. Edit., 2020, 59(44): 19503-19509 doi: 10.1002/anie.202004892

    13. [13]

      BABAR N U, JOYA K S, TAYYAB M A, ASHIQ M N, SOHAIL M. Highly sensitive and selective detection of arsenic using electrogenerated nanotextured gold assemblage[J]. ACS Omega, 2019, 4(9): 13645-13657 doi: 10.1021/acsomega.9b00807

    14. [14]

      SULTAN M A, HASSAN H B, HASSAN S S, ISMAIL K M. Enhanced electrocatalytic alcohol oxidation with Ni-MOF for direct alcohol fuel cell applications[J]. Int. J. Hydrog. Energy, 2025, 100: 528-547 doi: 10.1016/j.ijhydene.2024.12.335

    15. [15]

      LI X L, CHEN M C, YE Y T, CHEN C J, LI Z L, ZHOU Y F, CHEN J, XIE F Y, JIN Y S, WANG N, MENG H. Electronic structure modulation of nickel sites by cationic heterostructures to optimize ethanol electrooxidation activity in alkaline solution[J]. Small, 2023, 19(18): 2207086 doi: 10.1002/smll.202207086

    16. [16]

      DU R X, WU B, TAN W L, LEI Y C, TU Y C, MEBRAHTU C, CHEN Z H, LI S L, ZHOU Z H, TANG Z C, CHEN H H, CHEN S M, CHEN L, WANG J J, SHI X F, YE Y F, WANG D S, PALKOVITS R, ZHAO W, ZENG F. Metallic Ni as electron acceptor modulates the redox of catalytic centers at activated Ni0/Ni(OH)2 heterojunctions for efficient ethanol electrooxidation[J]. Angew. Chem. ‒Int. Edit., 2025: e202510285

    17. [17]

      TAO J G L, CHEN J X, ZHAO B, FENG R F, SHAKOURI M, CHEN F. Ni3C/Ni3S2 heterojunction electrocatalyst for efficient methanol oxidation via dual anion co-modulation strategy[J]. Small, 2024, 20(46): 2402492 doi: 10.1002/smll.202402492

    18. [18]

      HUANG L, LIN X C, ZHANG K, ZHANG J, WANG C H, QU S J, WANG Y G. Extraordinary d-d hybridization in Co(Cu)0.5OxHy microcubes facilitates PhCH2O*-Co(Ⅳ) coupling for benzyl alcohol electrooxidation[J]. Appl. Catal. B‒Environ. Energy, 2024, 346: 123739 doi: 10.1016/j.apcatb.2024.123739

    19. [19]

      WU Y, GUO X L, CHEN H G, XIN Y C, DONG X A, HU X L, XIA L, YU P. Molybdenum triggers the bifunctional mechanism of oxygen evolution reaction of Fe34-xNi25Co25MoxB8P8 amorphous alloy with boosted catalytic activity[J]. J. Electroanal. Chem., 2024, 972: 118612 doi: 10.1016/j.jelechem.2024.118612

    20. [20]

      WANG S W, GENG Z, BI S H, WANG Y W, GAO Z J, JIN L M, ZHANG C M. Recent advances and future prospects on Ni3S2-based electrocatalysts for efficient alkaline water electrolysis[J]. Green Energy Environ., 2024, 9(4): 659-683 doi: 10.1016/j.gee.2023.02.011

    21. [21]

      ALDIES A M, AWAD S. Distinct influence of Cd in the electrocatalyst of Ni-Co-Cd/CNFs nanoparticles as a catalyst in direct alcohol fuel cells (DAFCs)[J]. Solid State Ionics, 2025, 423: 116846 doi: 10.1016/j.ssi.2025.116846

    22. [22]

      WAN Z M, WANG L Q, ZHOU Y H, XU S Y, ZHANG J, CHEN X, LI S, OU C J, KONG X Z. A frogspawn inspired twin Mo2C/Ni composite with a conductive fibrous network as a robust bifunctional catalyst for advanced anion exchange membrane electrolyzers[J]. Nanoscale, 2024, 16(11): 5845-5854 doi: 10.1039/D3NR06242B

    23. [23]

      CHEN D X, DING Y X, CAO X, SUN L C. Heterostructured Ni-Co electrocatalyst with enhanced interfacial charge transfer for efficient biomass upgrading[J]. Appl. Catal. B‒Environ. Energy, 2025, 378: 125539 doi: 10.1016/j.apcatb.2025.125539

    24. [24]

      GU X Y, JING H Y, MU X Q, YANG H, ZHOU Q, YAN S L, LIU S L, CHEN C Y. La-triggered synthesis of oxygen vacancy-modified cobalt oxide nanosheets for highly efficient oxygen evolution in alkaline media[J]. J. Alloy. Compd., 2020, 814: 152274 doi: 10.1016/j.jallcom.2019.152274

    25. [25]

      LI R C, KUANG P Y, WANG L X, TANG H L, YU J G. Engineering 2D NiO/Ni3S2 heterointerface electrocatalyst for highly efficient hydrogen production coupled with benzyl alcohol oxidation[J]. Chem. Eng. J., 2022, 431: 134137 doi: 10.1016/j.cej.2021.134137

    26. [26]

      WU T H, LIN Y C, HOU B W, LIANG W Y. Nanostructured β-NiS catalyst for enhanced and stable electro-oxidation of urea[J]. Catalysts, 2020, 10(11): 1280-1291 doi: 10.3390/catal10111280

    27. [27]

      MANSOR M, BUDIMAN S N, ZAINOODIN A M, KHAIRUNNISA M P, YAMANAKA S, JUSOH N W C, LIZA S. Candle soot as a novel support for nickel nanoparticles in the electrocatalytic ethanol oxidation[J]. Nanomaterials, 2024, 14(12): 1042 doi: 10.3390/nano14121042

    28. [28]

      HUANG H F, DENG X L, YAN L Q, WEI G, ZHOU W Z, LIANG X Q, GUO J. One-step synthesis of self-supported Ni3S2/NiS composite film on Ni foam by electrodeposition for high-performance supercapacitors[J]. Nanomaterials, 2019, 9(12): 1718 doi: 10.3390/nano9121718

    29. [29]

      CHEN W, SHI J Q, WU Y D, JIANG Y M, HUANG Y C, ZHOU W, LIU J L, DONG C L, ZOU Y Q, WANG S Y. Vacancy‐induced catalytic mechanism for alcohol electrooxidation on nickel‐based electrocatalyst[J]. Angew. Chem. ‒Int. Edit., 2023, 63(4): e202316449

    30. [30]

      ZHANG N, HU Y, AN L, LI Q Y, YIN J, LI J Y, YANG R, LU M, ZHANG S, XI P X, YAN C H. Surface activation and Ni‐S stabilization in NiO/NiS2 for efficient oxygen evolution reaction[J]. Angew. Chem. ‒Int. Edit., 2022, 61(35): e202207217 doi: 10.1002/anie.202207217

    31. [31]

      YAN Y D, WANG R Y, ZHENG Q, ZHONG J Y, HAO W C, YAN S C, ZOU Z G. Nonredox trivalent nickel catalyzing nucleophilic electrooxidation of organics[J]. Nat. Commun., 2023, 14(1): 7987-7997 doi: 10.1038/s41467-023-43649-6

    32. [32]

      WU Q K, WANG S R, GUO J H, FENG X Q, LI H, LV S S, ZHOU Y, CHEN Z. Insight into sulfur and iron effect of binary nickel-iron sulfide on oxygen evolution reaction[J]. Nano Res., 2021, 15(3): 1901-1908

    33. [33]

      VO T G, TRAN G S, CHIANG C L, LIN Y G, CHANG H E, KUO H H, CHIANG C Y, HSU Y J. Au@NiSx yolk@shell nanostructures as dual-functional electrocatalysts for concomitant production of value-added tartronic acid and hydrogen fuel[J]. Adv. Funct. Mater., 2023, 33(4): 2209386 doi: 10.1002/adfm.202209386

    34. [34]

      XUE Z Q, LI X, LIU Q L, CAI M K, LIU K, LIU M, KE Z F, LIU X L, LI G Q. Interfacial electronic structure modulation of NiTe nanoarrays with NiS nanodots facilitates electrocatalytic oxygen evolution[J]. Adv. Mater., 2019, 31(21): 1900430 doi: 10.1002/adma.201900430

    35. [35]

      LI D R, WAN W J, WANG Z W, WU H Y, WU S X, JIANG T, CAI G X, JIANG C Z, REN F. Self-derivation and surface reconstruction of Fe‐doped Ni3S2 electrode realizing high-efficient and stable overall water and urea electrolysis[J]. Adv. Energy. Mater., 2022, 12(39): 2201913 doi: 10.1002/aenm.202201913

    36. [36]

      GULTOM N S, LI C H, KUO D H, ABDULLAH H. Single-step synthesis of Fe-doped Ni3S2/FeS2 nanocomposites for highly efficient oxygen evolution reaction[J]. ACS Appl. Mater. Interfaces, 2022, 14(35): 39917-39926 doi: 10.1021/acsami.2c08246

    37. [37]

      XU Y Z, YUAN C Z, CHEN X P. Co-doped NiSe nanowires on nickel foam via a cation exchange approach as efficient electrocatalyst for enhanced oxygen evolution reaction[J]. RSC Adv., 2016, 6(108): 106832-106836 doi: 10.1039/C6RA23580H

    38. [38]

      KANSAL S, R R, ANSHU S, PRIYA S, MANDAL D, SRIVASTAVA A K, CHANDRA A. Lattice strain-induced d-band modulation in nanosheets of CuxNiCo-layered double hydroxides for enhanced water electrolysis[J]. ACS Sustain. Chem. Eng., 2024, 12(36): 13511-13524 doi: 10.1021/acssuschemeng.4c03545

    39. [39]

      MA G Y, YE J T, QIN M Y, SUN T Y, TAN W X, FAN Z H, HUANG L F, XIN X. Mn-doped NiCoP nanopin arrays as high- performance bifunctional electrocatalysts for sustainable hydrogen production via overall water splitting[J]. Nano Energy, 2023, 115: 108679 doi: 10.1016/j.nanoen.2023.108679

    40. [40]

      WANG Y, HU T, LIU Q, ZHANG L. CoMn2O4 embedded in MnOOH nanorods as a bifunctional catalyst for oxygen reduction and oxygen evolution reactions[J]. Chem. Commun., 2018, 54(32): 4005-4008 doi: 10.1039/C8CC00870A

  • Figure 1  Characterization of Ni2CoS4: (a) XRD patterns; (b, c) SEM images; (d) elemental mappings of Ni, Co, and S on the surface of Ni foam; (e) HRTEM image; (f) elemental mappings of Ni, Co, and S in HRTEM image

    Figure 2  Electrochemical performance of Ni2CoS4, Ni2Co(OH)x, and Ni(OH)2 in 1.0 mol·L-1 KOH with 1.0 mol·L-1 benzyl alcohol: (a) polarization curves; (b) current densities at 1.40 V (vs RHE); (c) Nyquist plots; (d) Cdl; (e) chronopotentiometry curves at 50 mA·cm-2 after IR correction

    Figure 3  Bode plots of (a) Ni2CoS4, (b) Ni2Co(OH)x, and (c) Ni(OH)2 in 1.0 mol·L-1 KOH with 1.0 mol·L-1 benzyl alcohol

    Figure 4  DOS of (a) Ni(OH)2, (b) Ni2Co(OH)x, and (c) Ni2CoS4; (d) Calculated model for OH- adsorption of Ni2CoS4; (e) Calculated adsorption energy of OH- of Ni(OH)2, Ni2Co(OH)x, and Ni2CoS4

    Figure 5  High-resolution XPS spectra of (a) Ni2p, (b) Co2p, and (c) S2p in Ni-based catalysts; High-resolution XPS spectra of (d) Ni2p, (e) Co2p, and (f) S2p in Ni2CoS4 before and after stability test

  • 加载中
计量
  • PDF下载量:  0
  • 文章访问数:  28
  • HTML全文浏览量:  4
文章相关
  • 发布日期:  2025-12-10
  • 收稿日期:  2025-09-08
  • 修回日期:  2025-10-10
通讯作者: 陈斌, bchen63@163.com
  • 1. 

    沈阳化工大学材料科学与工程学院 沈阳 110142

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

/

返回文章