Ultrafast synthesis of Mo2N with highly dispersed Ru for efficient alkaline hydrogen evolution

Xinyu Hou Xuelian Yu Meng Liu Hengxing Peng Lijuan Wu Libing Liao Guocheng Lv

Citation:  Xinyu Hou, Xuelian Yu, Meng Liu, Hengxing Peng, Lijuan Wu, Libing Liao, Guocheng Lv. Ultrafast synthesis of Mo2N with highly dispersed Ru for efficient alkaline hydrogen evolution[J]. Chinese Chemical Letters, 2025, 36(4): 109845. doi: 10.1016/j.cclet.2024.109845 shu

Ultrafast synthesis of Mo2N with highly dispersed Ru for efficient alkaline hydrogen evolution

English

  • H2, with its high energy density (142,351 kJ/kg) and environmentally friendly nature, is considered the most promising alternative energy for resolving the environmental crisis [1,2]. Taking into account sustainable development, electrocatalytic water splitting has been recognized as a green and efficient method for H2 production [3-6]. The alkaline hydrogen evolution reaction (HER) has gained recognition over the acidic HER due to its greater stability and cost-effectiveness, making it more suitable for large-scale applications [7].

    Platinum (Pt) is well-known as one of the most effective noble metal catalysts for HER, primarily because of its moderate Pt-H bond [8,9]. However, Pt-based catalysts are expensive and impractical for large-scale applications. Recently, due to the excellent performance and significantly lower cost (approximately Pt~4%), researchers considered ruthenium (Ru) as a promising electrocatalyst for HER [10]. While excessively strong OH adsorption by Ru mono-atoms can lead to desorption difficulties and reduce the HER reaction rate [11]. Therefore, it is crucial to identify a suitable carrier with a strong water dissociation capacity to enhance the HER catalytic activity [12-14].

    Transition metal nitrides (TMNs) are promising materials that can sustain HER activity at low noble metal loadings [15]. Studies have shown that the interaction between nitrogen (N) elements in TMNs with metals affects their electronic structures. The d orbitals of metal atoms will extend after being hybridized with the s and p orbitals of nitrogen atoms. This process gives them d-band characteristics similar to noble metals [16,17]. Compared to other carriers, such as transition metal carbides (TMCs), TMNs offer additional benefits by desorbing with extra nitrogen atoms as N2 during synthesis, leaving a clean TMN surface for direct interaction with supported metal [18]. However, for TMCs, excess carbon may accumulate on the surface, preventing direct contact with the active metal [19]. In a recent study, Turaczy's team experimentally determined the hydrogen binding energy (HBE) value of molybdenum nitride, demonstrating its potential as a support to reduce noble metal loadings in HER catalysts [18]. The introduction of nitrogen atoms can alter the charge density of Mo atoms, effectively adjusting the strength of metal (Mo-H) bonds and optimizing electrochemical activity [20-22].

    However, current synthesis methods of molybdenum nitrides under N2 or NH3 atmosphere typically involve high pressure and toxic gases (Table S1 in Supporting information) [23,24]. These methods often result in low yield and incomplete nitridation [25]. Besides, prolonged high-temperature reactions can lead to severe aggregation of Ru species [26,27]. Therefore, the rapid and safe synthesis of molybdenum nitride carriers with enhanced exposure and dispersion of Ru active sites remains a major challenge. High temperature shock (HTS) provides a simple and efficient strategy through rapid heating. Lattice defects generated by extremely rapid temperature changes can enhance the electronic polarization of elemental surfaces [28,29]. With this method, Hu's team successfully synthesized high-entropy alloys, which effectively addressed the aggregation of active sites and greatly simplified the synthesis path [30]. Similarly, niobium-based oxides have recently been synthesized through ultrafast carbon thermal shock (CTS) by Zhao et al., avoiding the need for time-consuming and energy-intensive annealing processes [31].

    Based on the above considerations, we proposed an ultrafast (1.67 s) method to prepare Mo2N and simultaneously evenly disperse ultrafine Ru sub-nanoparticles on the Mo2N carrier. Under vacuum conditions, precursor powder was completely nitrided to Mo2N by the rapid heating and cooling effects of HTS, and the interaction between Ru and Mo2N effectively regulated the charge redistribution. The synthesized Ru/Mo2N catalyst shows excellent HER performance under alkaline conditions (1 mol/L KOH). It exhibited a low overpotential of 66 mV vs. RHE for 10 mA/cm2; it also maintains a significant advantage at an overpotential of 200 mV vs. RHE for 100 mA/cm2, which is lower than commercial Pt/C (220 mV).

    The specific synthesis method for Ru/Mo2N is provided in the supporting information. In the synthesis of precursors, melamine and cyanuric acid have the ability to undergo self-association through hydrogen bonding in solution, while Mo7O246− and Ru3+ can effectively anchor through the metal-N/O covalent bonding coordination effect [32,33]. After filtering and drying, the precursor powder was subjected to HTS for 1.67 s to form Ru/Mo2N. In terms of structural characterization, we used X-ray diffraction (XRD) for compositional analysis of Ru/Mo2N. In Fig. 1a, the diffraction peaks at 37.38°, 43.44°, 63.10°, and 75.95° are in good agreement with the characteristic crystal planes (111), (420), (311), and (200) of the cubic γ-Mo2N (JCPDS card No. 25-1366). The significant bulge around 23° may be attributed to non-crystalline carbon. Meanwhile, Raman spectroscopy shows two distinct peaks of carbon: the D band (1342 cm−1) and the G band (1586 cm−1) (Fig. 1b). The value of ID/IG (1.16) further indicates the presence of non-crystalline carbon and numerous lattice defects in Ru/Mo2N [34]. The morphology and structure of the prepared samples were investigated using scanning electron microscopy (SEM) and transmission electron microscopy (TEM). As shown in Fig. 1c, Ru/Mo2N exhibits a curled nanosheet structure with open and continuous channels, and the specific surface area is 13.85 m2/g. These channels may be derived from decomposable melamine and cyanuric acid, which are expected to enhance the accessibility of active sites [35,36]. Moreover, TEM images reveal a relatively high and uniform dispersion of Ru/Mo2N (Fig. 1d). Notably, the carbon cladding layers were visualized in the TEM image (Fig. 1e), corresponding to the previously mentioned amorphous carbon. These carbon layers provide structural support to the entire Ru/Mo2N particle and accelerate charge transfer. Furthermore, the lattice spacing of the Ru/Mo2N sample was calculated to be 0.240 nm, corresponding to the (111) crystal plane of γ-Mo2N. The structure of Ru/Mo2N and the highly homogeneous dispersion of Mo, C, N, and Ru were further demonstrated by high-resolution TEM (HR-TEM) and energy-dispersive X-ray spectroscopy (EDS) mapping (Figs. 1fj). Additionally, there are no obvious aggregates of Ru or its oxides on the entire nanosheet structure, indicating that Ru is dispersed as ultrafine sub-nanoparticles on the Mo2N matrix [2].

    Figure 1

    Figure 1.  Ru/Mo2N: (a) XRD pattern. (b) Raman spectra. (c, d) SEM images. (e) TEM image. (f) HR-TEM image and (g-j) EDS mapping images.

    In order to clarify the valence states of the chemical elements and their composition on the surface, X-ray photoelectron spectroscopy (XPS) was utilized. The XPS survey clearly shows signals of Mo, N, C, and Ru (Fig. 2a). The XPS spectra of Ru 3p (Fig. 2b) exhibits binding energies at 483.1 eV and 460.8 eV corresponding to the Ru 3p1/2 and Ru 3p3/2 peaks, respectively. Interestingly, the Ru 3p spectra of Ru/Mo2N are negatively shifted compared to RuO2 in other studies (486.3 eV and 464.1 eV), indicating an electron-enriched state in Ru species [37,38]. Fig. 2c shows N 1s XPS spectra in Ru/Mo2N with two peaks located at 401.9 eV and 398.5 eV, corresponding to graphitic N and pyridinic N, respectively. The higher content of pyridinic N typically exhibits stronger alkalinity and nucleophilicity, making it capable of coordinating with metals. A typical N-Mo peak appears at 395.4 eV, indicating the presence of nitrides. The two peaks mentioned above collectively validate the existence of Mo2N [39]. Fig. 2d displays the XPS spectrum of Mo 3d, showing three pairs of peaks corresponding to Mo2+, Mo4+, and Mo6+. The peaks at 227.5 eV and 230.7 eV are attributed to Mo2+, mainly derived from Mo2N. The peaks at 228.3 eV and 233.1 eV correspond to Mo4+, while the peaks at 231.8 eV and 234.9 eV correspond to Mo6+. The abundance of high-valence molybdenum attributes to the surface oxidation of Mo [40].

    Figure 2

    Figure 2.  The XPS spectra of Ru/Mo2N. (a) XPS survey spectra of Ru/Mo2N. (b) Ru 3p. (c) N 1s. (d) Mo 3d.

    The electrochemical HER performance of Ru/Mo2N was evaluated in 1 mol/L KOH electrolyte using a standard three-electrode system. All potentials are calculated with an iR correction and reported at a reversible hydrogen electrode (RHE) potential of 10 mA/cm2. For comparison, the prepared samples and commercial Pt/C electrocatalysts were all tested. The preparation method for Mo2N is similar to that of Ru/Mo2N, except that Ru is not added. Linear sweep voltammetry (LSV) curves and overpotential in Figs. 3a and b show that the required overpotentials at 10 mA/cm2 (η10) and 100 mA/cm2 (η100) for Ru/Mo2N are 66 mV and 200 mV, respectively, which are significantly lower than those of the contrast samples, Mo2N (η10 = 222 mV) and commercial Pt/C (η10 = 63 mV, η100 = 220 mV). These results demonstrate that the inclusion of a trace amount of Ru significantly improved the electrocatalytic performance. We also studied the performance at different pH values (Fig. S1 in Supporting information). Moreover, the Tafel slope of Ru/Mo2N in Fig. 3c is 57 mV/dec, which is lower than that of Mo2N (80 mV/dec), suggesting faster kinetics on Ru/Mo2N, believed to follow the Volmer-Heyrovsky mechanism [41,42]. In order to obtain a more precise study of Ru/Mo2N, further calculations were performed on the mass activity (MA) and turnover frequency (TOF) related to the noble metals (Pt or Ru). MA represents the active species generated by the catalyst per unit mass, while TOF denotes the rate of reactant transformations per unit time in a catalyzed reaction. Impressively, the Ru/Mo2N catalyst exhibits a significantly higher MA than commercial Pt/C (Fig. 3d), indicating a high utilization of noble metals. Additionally, it can be observed that at 50, 75, and 100 mV, the TOF of Ru/Mo2N is also superior to Pt/C (Fig. 3e). Then, we assessed the electrocatalytic durability of the Ru/Mo2N catalyst of η10 (Fig. 3f). After a 50-h prolonged test, the decay of the potential of the Ru/Mo2N catalyst can be neglected, demonstrating its superior stability during the HER catalytic process. A range of previously reported Mo-N-based catalyst performances are demonstrated in Fig. 3g.

    Figure 3

    Figure 3.  (a–c) LSV curves, overpotentials and Tafel plots of Mo2N, Ru/Mo2N and Pt/C in 1 mol/L KOH. (d, e) Comparison of MA and TOF between Ru/Mo2N and Pt/C. (f) Chronopotentiometric (CP) curve recorded at a constant current density of −10 mA/cm2. (g) The overpotential (η10) of various Mo-N-based catalysts.

    Furthermore, to validate the advantages of Mo2N synthesized by HTS, Mo2N-F and Ru/Mo2N-F were prepared using the same raw materials but high-temperature tube furnace sintering method. Interestingly, the X-ray diffraction patterns of Mo2N-F and Ru/Mo2N-F (Fig. 4a), in addition to coinciding with some of the γ-Mo2N peaks and also show good correspondence with the characteristic crystal planes (101), (110), (103), and (112) of hexagonal β-Mo2C (JCPDS card No. 65–8766) at 2θ = 39.58°, 61.87°, 69.76°, and 75.19°, respectively. Meanwhile, the crystal plane spacing of Ru/Mo2N-F was calculated to be 0.228 nm and 0.236 nm based on the TEM images, corresponding to the (101) and (002) crystal faces of β-Mo2C, respectively (Fig. S2 in Supporting information). The above results indicate that incomplete nitridation exists in Mo2N-F and Ru/Mo2N-F synthesized by high-temperature tube furnace sintering. EDS confirmed the uniform distribution of Mo, C, N, and Ru elements throughout the Ru/Mo2N-F material (Figs. S3b-e in Supporting information). As for electrochemical HER performance, enormous distinctions were observed, LSV revealed that the overpotential (η10) values of Mo2N and Ru/Mo2N (222 mV, 66 mV) were significantly lower than those of Mo2N-F and Ru/Mo2N-F (287 mV, 148 mV), as illustrated in Fig. 4b. To obtain a more precise investigation of these distinctions, XPS was employed (Fig. 4c). Compared with Ru/Mo2N-F, the Ru 3p peak of Ru/Mo2N shows a slight negative shift, implying a higher degree of electron transfer after HTS [43]. Meanwhile, the N 1s spectra of Ru/Mo2N shows an obvious positive shift relative to that of Ru/Mo2N-F (Fig. S4 in Supporting information). The above results suggest that there is an extensive interaction between Mo2N and Ru species under HTS. To further compare the intrinsic HER activity of these samples, the electrochemical double-layer capacitance (Cdl) of the catalysts was determined by cyclic voltammetry (CV) as a function of scan rate (20–100 mV/s) (Fig. 4d) [44-46]. It is worth noting that the Cdl value of Ru/Mo2N was 6.82 mF/cm2, which was significantly higher than the Cdl value of Ru/Mo2N-F (1.39 mF/cm2). This also demonstrates that Ru/Mo2N prepared by HTS disperses Ru more evenly, which possesses more catalytically active sites and exhibits higher catalytic activity.

    Figure 4

    Figure 4.  (a) XRD patterns of Ru/Mo2N and Ru/Mo2N-F. (b) LSV curves of various samples in 1 mol/L KOH. (c) The XPS spectra of Ru 3p (Ru/Mo2N and Ru/Mo2N-F). (d) The electrochemical Cdl of Ru/Mo2N and Ru/Mo2N-F.

    In conclusion, we successfully developed a highly effective electrocatalyst, Ru/Mo2N, using a rapid heating method, resulting in homogeneously dispersed Ru on Mo2N. As evidenced by experimental analysis, we propose that the superior performance is attributed to a higher degree of electron transfer exhibited by Ru during the HTS, and a synergistic effect between highly dispersed Ru sub-nanoparticle sites and Mo2N, promoting abundant metal-support interactions in Ru/Mo2N. With an overpotential (η10) of 66 mV and outstanding catalytic activity, the prepared Ru/Mo2N catalyst exhibits its remarkable performance. The catalyst also demonstrates excellent long-term electrochemical stability in 1 mo/L KOH solution. This work provides a rational pathway for the preparation of Mo2N-based electrocatalysts.

    No conflict of interest exits in the submission of this manuscript, and manuscript is approved by all authors for publication. I would like to declare on behalf of my coauthors that the work described was original research that has not been published previously, and not under consideration for publication elsewhere, in whole or in part. All the authors listed have approved the manuscript that is enclosed.

    Xinyu Hou: Data curation, Methodology, Validation, Writing – original draft. Xuelian Yu: Supervision, Conceptualization, Formal analysis, Methodology. Meng Liu: Conceptualization, Data curation, Supervision, Validation. Hengxing Peng: Conceptualization. Lijuan Wu: Conceptualization, Software. Libing Liao: Methodology, Supervision. Guocheng Lv: Methodology, Resources, Supervision.

    This work was supported by the Beijing Natural Science Foundation (No. 2232061) and the National Natural Science Foundation of China (No. 42377227).

    Supplementary material associated with this article can be found, in the online version, at doi:10.1016/j.cclet.2024.109845.


    1. [1]

      J.Q. Chi, M. Yallg, Y.M. Chai, et al., J. Energy Chem. 48 (2022) 398–423.

    2. [2]

      X.L. Fan, C. Liu, M.Y. Wu, et al., Appl. Catal. B: Environ. 318 (2022) 121867.

    3. [3]

      Y.M. Liu, Chem. Bio. Engin. 3 (2007) 79–81.

    4. [4]

      B.H. Suryanto, Y. Wang, R.K. Hocking, et al., Nat. Commun. 10 (2019) 5599. doi: 10.1038/s41467-019-13415-8

    5. [5]

      S.D. Lu, B.Y. Lu, G.C. Tan, et al., Biosens. Bioelectron. 167 (2020) 112491.

    6. [6]

      L.N. Chen, S.H. Wang, P.Y. Zhang, et al., Nano Energy 88 (2021) 106211. doi: 10.1016/j.nanoen.2021.106211

    7. [7]

      H.Q. Fu, M. Zhou, P.F. Liu, et al., J. Am. Chem. Soc. 144 (2022) 6028–6039. doi: 10.1021/jacs.2c01094

    8. [8]

      J.Y. Yu, A.Z. Wang, W.Q. Yu, et al., Appl. Catal. B: Environ. 277 (2020) 119236.

    9. [9]

      D.Y. Ren, G.W. Wang, L.Y. Li, et al., Chem. Eng. J. 454 (2023) 140557.

    10. [10]

      Y.N. Zhou, F.L. Wang, J. Nan, et al., Appl. Catal. B: Environ. 304 (2022) 120917.

    11. [11]

      J. Zhang, G. Chen, Q. Liu, et al., Angew. Chem. Int. Ed. 61 (2022) e20220948.

    12. [12]

      Z.X. Yang, Z.Z. Wang, J.T. Wan, et al., Environ. Sci. Technol. 56 (2022) 18008–18017. doi: 10.1021/acs.est.2c06571

    13. [13]

      C. Li, H. Jang, M.G. Kim, et al., Appl. Catal. B: Environ. 307 (2022) 121204. doi: 10.1016/j.apcatb.2022.121204

    14. [14]

      H. Song, M. Wu, Z. Tang, et al., Angew. Chem. Int. Ed. 60 (2021) 7234–7244. doi: 10.1002/anie.202017102

    15. [15]

      X. Shi, A.P. Wu, H.J. Yan, et al., J. Mater. Chem. A 6 (2018) 20100–20109. doi: 10.1039/c8ta07906d

    16. [16]

      C.G. Morales-Guio, L.A. Stern, X. Hu, Chem. Soc. Rev. 43 (2014) 6555–6569. doi: 10.1039/c3cs60468c.

    17. [17]

      X. Wang, J. He, B. Yu, et al., Appl. Catal. B: Environ. 258 (2019) 117996.

    18. [18]

      K.K. Turaczy, W.J. Liao, et al., ACS Catal. 13 (2023) 14268–14276. doi: 10.1021/acscatal.3c04063

    19. [19]

      Y.C. Kimmel, D.V. Esposito, R.W. Birkmire, et al., Int. J. Hydrogen Energy 37 (2012) 3019–3024.

    20. [20]

      Y.Y. Chen, Y. Zhang, J. W, et al., ACS Nano 10 (2016) 8851–8860. doi: 10.1021/acsnano.6b04725

    21. [21]

      D.J. Ham, J.S. Lee, Energies 2 (2009) 873–899. doi: 10.3390/en20400873

    22. [22]

      L. Zhe, M. Tahir, X.W. Lang, et al., J. Mater. Chem. A 5 (2017) 20932. doi: 10.1039/C7TA06981B

    23. [23]

      K. Ojha, S. Saha, S. Banerjee, A.K. Ganguli, et al., ACS Appl. Mater. Interfaces 9 (2017) 19455–19461. doi: 10.1021/acsami.6b10717

    24. [24]

      L.M. Dai, F.L. Yao, L. Yu, et al., Adv. Energy Mater. 12 (2022) 2200974.

    25. [25]

      J. Li, Y.N. Zhou, W.J. Tang, et al., Appl. Catal. B: Environ. 285 (2021) 119861.

    26. [26]

      Z. Wei, P. Grange, B. Delmon, Appl. Surf. Sci. 135 (1998) 107–114. doi: 10.3724/j.issn.1000-0518.1998.4.107

    27. [27]

      W.F. Chen, K. Sasaki, C. Ma, et al., Angew. Chem. Int. Ed. 51 (2012) 6131–6135. doi: 10.1002/anie.201200699

    28. [28]

      Z.X. Yang, Y. Li, X.T. Wang, et al., J. Environ. Sci. 142 (2024) 204–214. doi: 10.3390/agronomy14010204

    29. [29]

      Z. Zhao, J. Sun, X. Li, et al., Appl. Catal. B 340 (2024) 123277–123293. doi: 10.1016/j.apcatb.2023.123277

    30. [30]

      Y. Yao, Z. Huang, P. Xie, et al., Science 359 (2018) 1489–1494. doi: 10.1126/science.aan5412

    31. [31]

      Q.L. Wu, Y.H. Kang, G.H. Chen, et al., Adv. Funct. Mater. (2024) 2315248.

    32. [32]

      F. Chen, X.L. Wu, C. Shi, et al., Adv. Funct. Mater. 31 (2021) 2007877. doi: 10.1002/adfm.202007877

    33. [33]

      B.X. Zhou, S.S. Ding, K.X. Yang, et al., Adv. Funct. Mater. 31 (2020) 2009230. doi: 10.1002/adfm.202009230

    34. [34]

      J. Jia, T.L. Xiong, L.L. Zhao, et al., ACS Nano 11 (2017) 12509–12518. doi: 10.1021/acsnano.7b06607

    35. [35]

      C. Tang, H. Zhang, K. Xu, et al., J. Mater. Chem. A 7 (2019) 18030–18038. doi: 10.1039/c9ta04374h

    36. [36]

      L.N. Zhang, Z.L. Lang, Y.H. Wang, et al., Science 12 (2019) 2569–2580. doi: 10.1039/c9ee01647c

    37. [37]

      P. Li, M. Wang, X.L. Duan, et al., Nat. Commun. 10 (2019) 1711. doi: 10.1007/s00192-018-3796-y

    38. [38]

      Y. Hu, G. Luo, L. Wang, et al., Adv. Energy Mater. 11 (2021) 2002816. doi: 10.1002/aenm.202002816

    39. [39]

      A.C. Mtukula, X.J. Boand, L.P. Guo, J. Alloys Compd. 692 (2017) 614–621. doi: 10.1016/j.jallcom.2016.09.079

    40. [40]

      Y.P. Zhu, G. Chen, X.M. Xu, et al., ACS Catal. 7 (2017) 3540–3547. doi: 10.1021/acscatal.7b00120

    41. [41]

      M. Zeng, Y. Li, Mater. Chem. A 3 (2015) 14942–14962. doi: 10.1039/C5TA02974K

    42. [42]

      M. Yao, H. Hu, B. Sun, et al., Small 15 (2019) 1905201. doi: 10.1002/smll.201905201

    43. [43]

      T.T. Chao, W.B. Xie, Y.M. Hu, et al., Energy Environ. Sci. 17 (2024) 1397–1406. doi: 10.1039/d3ee02760k

    44. [44]

      J.S. Li, Y. Wang, C.H. Liu, et al., Nat. Commun. 7 (2016) 11204. doi: 10.1038/ncomms11204

    45. [45]

      K. Qu, Y. Zheng, X. Zhang, et al., ACS Nano 11 (2017) 7293–7300. doi: 10.1021/acsnano.7b03290

    46. [46]

      C. Lu, D. Tranca, J. Zhang, et al., ACS Nano 11 (2017) 3933–3942. doi: 10.1021/acsnano.7b00365

  • Figure 1  Ru/Mo2N: (a) XRD pattern. (b) Raman spectra. (c, d) SEM images. (e) TEM image. (f) HR-TEM image and (g-j) EDS mapping images.

    Figure 2  The XPS spectra of Ru/Mo2N. (a) XPS survey spectra of Ru/Mo2N. (b) Ru 3p. (c) N 1s. (d) Mo 3d.

    Figure 3  (a–c) LSV curves, overpotentials and Tafel plots of Mo2N, Ru/Mo2N and Pt/C in 1 mol/L KOH. (d, e) Comparison of MA and TOF between Ru/Mo2N and Pt/C. (f) Chronopotentiometric (CP) curve recorded at a constant current density of −10 mA/cm2. (g) The overpotential (η10) of various Mo-N-based catalysts.

    Figure 4  (a) XRD patterns of Ru/Mo2N and Ru/Mo2N-F. (b) LSV curves of various samples in 1 mol/L KOH. (c) The XPS spectra of Ru 3p (Ru/Mo2N and Ru/Mo2N-F). (d) The electrochemical Cdl of Ru/Mo2N and Ru/Mo2N-F.

  • 加载中
计量
  • PDF下载量:  1
  • 文章访问数:  91
  • HTML全文浏览量:  4
文章相关
  • 发布日期:  2025-04-15
  • 收稿日期:  2024-03-03
  • 接受日期:  2024-03-29
  • 修回日期:  2024-03-21
  • 网络出版日期:  2024-03-30
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