Ru掺杂Co3O4/氧化还原石墨烯的制备及其电催化析氧性能
English
Ru-doped Co3O4/reduced graphene oxide: Preparation and electrocatalytic oxygen evolution property
-
Key words:
- metal-organic framework
- / graphene
- / electrocatalyst
- / oxygen evolution reaction
-
With the escalating gravity of the energy crisis, the search for novel green energy sources to mitigate reliance on fossil fuels stands as the quintessential path toward sustainable development[1-2]. Electrolytic water splitting, known for its simplicity and cost-efficiency in hydrogen production, has attracted significant attention. The electrolysis process primarily comprises two half-reactions: the anodic oxygen evolution reaction (OER) and the cathodic hydrogen evolution reaction (HER). However, the widespread implementation of electrolytic water technology is restricted due to the intricate four-electron transfer mechanism in OER[3-4]. Nanoalloy and carbon composite materials have emerged as promising candidates for OER application with their advantageous properties such as multifaceted electrocatalytic sites, excellent conductivity, high stability, and structural diversity[5].
Metal-organic frameworks (MOFs) represent a class of three-dimensional porous organic-inorganic hybrids crafted through the coordination of metal ions and organic ligands. Unlike traditional porous materials, MOFs exhibit remarkable advantages, such as structural diversity, facile functionalization and modification, and tunable pore sizes. Inorganic nanomaterials derived from MOFs provide an ideal platform for developing highly efficient catalyst systems, due to their well-defined structure, large specific surface area, and interconnected porosity[6].
Graphene, a two-dimensional carbon nanomaterial with a hexagonal honeycomb lattice structure composed of sp2 hybridized carbon atoms, has garnered significant attention owing to its exceptional specific surface area, unparalleled electronic properties, and robust electrochemical stability[7]. Despite these promising attributes, pristine graphene materials often encounter challenges related to limited specific capacitance and durability. 2-Methylimidazole, as a ligand for transition metals or metal clusters, can simultaneously connect metals and carbon nanomaterials to form a network crystalline porous structure, providing more active sites and controllable mesopores[8-9]. Therefore, the complex network structure formed by 2-methylimidazole, Co, and reduced graphene oxide (rGO) has a higher specific surface area and surface functionality. Co3O4/rGO derived from this template has more diverse valence states and higher catalytic activity.
Extensive research has revealed that the integration of graphene surfaces with MOFs can substantially enhance the capacitance and electrochemical stability of these materials. Improving the catalytic performance of MOF materials through doping with non-precious metals is also an advanced strategy for designing catalysts. Zhang et al.[10] constructed a two-dimensional cobalt-doped MnPSe3 nanosheets (CMPS), which served as an outstanding bifunctional catalyst for alkaline seawater splitting, offering the current density of 10 mA·cm-2 with applied overpotentials of 59 and 300 mV for HER and OER, respectively, while that of the MnPSe3 nanosheets (MPS) was 120 and 370 mV, respectively. Peng et al.[11] provided platinum-containing ruthenium oxide nanoparticles (Pt@RuOx NPs) to achieve excellent overall water-splitting performance in acidic electrolytes and Pt@RuOx NPs exhibited more desirable OER activity than RuOx NPs. Wang et al.[12] also reported Fe nanoparticles confined by multiple-heteroatom-doped carbon frameworks which are used for aqueous Zn-air battery driving CO2 electrolysis. Nowadays, a series of research shows that Ru-doped Co/C ternary composites can also exhibit impressive electrochemical performance in varying acidic and alkaline environments[13-14].
Herein, a one-step precipitation method was employed to synthesize MOFs/rGO binary composites (ZIF-67/rGO), using Co(NO3)2·6H2O as the metal source, 2-methylimidazole as the organic ligand, and rGO as the carbon support. Then Ru ions were introduced through ion exchange and annealing, producing porous Ru-doped Co3O4/rGO electrocatalysts (Ru-Co3O4/rGO). This approach significantly improves the electrocatalytic performance of Co-based MOFs, offering promising applications in various electrochemical reactions.
1. Experimental
1.1 Chemicals
Ruthenium(Ⅲ) chloride hydrate (RuCl3·3H2O) and cobalt nitrate hexahydrate (Co(NO3)2·6H2O, 99%) were purchased from Sinopharm Chemical Reagent Co., Ltd. 2-Methylimidazole (99%) were purchased from Aladdin Reagent Co., Ltd. rGO was purchased from Nanjing XFNANO Materials Tech Co., Ltd. Other reagents are purchased from commercial companies and used without further treatment.
1.2 Synthesis of ZIF-67/rGO precursor
rGO (120 mg) and Co(NO3)2·6H2O (1 532 mg) were dispersed in a 120 mL methanol/ethanol (1∶1, V/V) mixed solution by ultrasound for 2 h and then stirred for 3 h. 2-Methylimidazole (1 729 mg) was added to the above solution, and the mixture was vigorously stirred for 20 min and let settle for 24 h. The product was washed multiple times with ethanol, and dried overnight in a 60 ℃ oven to obtain ZIF-67/rGO.
1.3 Synthesis of Ru-ZIF-67/rGO catalyst
ZIF-67/rGO (55 mg) was dispersed in 25 mL ethanol solution using ultrasound and 200 μL ethanol solution of RuCl3 was added to stir and disperse for 3 h. The product was washed multiple times with ethanol and dried overnight in a 60 ℃ oven to obtain Ru-ZIF-67/rGO.
1.4 Synthesis of Ru-Co3O4/rGO composite materials
Ru-ZIF-67/rGO was calcined at 350 ℃ in an N2 atmosphere for 2 h to prepare Ru-Co3O4/rGO composite materials. The preparation process is shown in Fig. 1. Firstly, rGO was used as the induction template, Co2+ was used as the metal source, and 2-methylimidazole was used as the organic ligand. Cobalt-based MOF material (ZIF-67) was grown on a rGO surface by rich oxygen-containing functional groups of rGO. The purified sample was dissolved and dispersed in ethanol, and RuCl3 was added as the ruthenium source. The aim was to fully contact the surface of the introduced ruthenium and ZIF-67/rGO for ion exchange, resulting in a uniform ruthenium-doped ternary composite material (Ru-ZIF-67/rGO). Subsequently, the material was used as a precursor and annealed in a nitrogen atmosphere to remove the organic components in Ru-ZIF-67/rGO, resulting in a high-performance catalyst Ru-Co3O4/rGO.
Figure 1
1.5 Structural characterization of catalysts
X-ray diffractometer (XRD, Cu Kα, λ=0.154 nm) was used under the operating voltage of 40 kV, operating current of 40 mA, scanning speed of 2 ℃·min-1, and scanning range of 5°-80°. The morphology of the samples was observed by the S-4800 scanning electron microscope (SEM) of Hitachi High-tech Co. Ltd. The surface element content and the presence form of main elements were analyzed by X-ray photoelectron spectrometer. The structural features and elemental composition of Ru-Co3O4/rGO were carefully observed and accurately determined through transmission electron microscopy (TEM, JEM-1400-plus, Japan) at an operating voltage of 100 kV. The metal components of the catalysts were determined by inductively coupled plasma-optical emission spectrometry (ICP-OES, OES5110, MS7700).
1.6 Performance test of electrocatalyst materials
At room temperature, an electrochemical workstation (CHI 660E, China) was used in a three-electrode system. The glassy carbon electrode loaded with samples was used as the working electrode, the graphite rod electrode was used as the opposing electrode, and the Hg/HgO electrode (internal infiltration 1 mol·L-1 KOH) was used as the reference electrode. All electrochemical measurements were carried out on the electrochemical workstation. To test the electrocatalytic performance of the catalyst and explore the mechanism of electrocatalysis, linear scan voltammetry (LSV) (a scan rate of 2 mV·s-1), cyclic voltammetry (CV) (scan rates ranging from 20 to 100 mV·s-1), constant voltage or current stability, electrochemical impedance (EIS) (an amplitude voltage of 5 mV), and other tests were usually required.
2. Results and discussion
2.1 Material characterization
2.1.1 XRD analysis
XRD was used to analyze the composition and structure of the synthesized material. As shown in Fig. 2, the XRD pattern of the ZIF-67/rGO precursor matches the reported diffraction peaks, confirming its successful synthesis[15-16]. The XRD patterns of Co3O4/rGO and Ru-Co3O4/rGO exhibited similar characteristics. Specifically, the peaks at 2θ of 37.0°, 44.9°, and 65.2° correspond to the (311), (400), and (440) planes of Co3O4, respectively, indicating the successful transformation of ZIF-67 into Co3O4 under high-temperature calcination. Introducing Ru to Co3O4/rGO did not significantly alter peak positions or intensities, indicating no new crystal phases or significant structural changes in the catalyst. This further confirms the successful doping of Ru into the Co3O4 lattice.
Figure 2
2.1.2 Morphology analysis
SEM was employed to examine the morphological characteristics of the synthesized samples. As depicted in Fig. 3a and 3b, rGO has a lamellar structure. Fig. 3c and 3d showed the ZIF-67 particles exhibited a uniform distribution with relatively consistent sizes. Notably, a significant portion of these ZIF-67 particles was observed to adhere to the folds of rGO indicating good dispersion and compatibility. Fig. 3e reveals that the introduction of a small quantity of Ru3+ ions did not appreciably alter the original morphology of the material. Subsequent high-temperature calcination of the ZIF-67/rGO precursor led to the formation of Co3O4/rGO, as shown in Fig. 3f. It is evident that most of the spherical ZIF-67 particles had undergone collapse and decomposition, while the original sheet-like structure of rGO remained intact, albeit in a stacked configuration. This suggests that the calcination process primarily affects the ZIF-67 component while preserving the overall morphology of the rGO carrier. Fig. 4a and 4b show the SEM images of Ru-ZIF-67/rGO after high-temperature calcination. The morphology observed is consistent with that of Co3O4/rGO, indicating that the addition of Ru3+ ions did not significantly influence the morphological features of the original material. This observation further validates the successful doping of Ru into the Co3O4 lattice while maintaining the overall structure. In Fig. 4c and 4d, Ru-Co3O4/rGO was still covered with many nanoparticles (ca. 10 nm), consistent with the SEM results.
Figure 3
Figure 4
2.1.3 XPS analysis
XPS analysis was conducted to elucidate the elemental composition and chemical states of the Ru-Co3O4/rGO. As depicted in Fig. 5a, the C1s XPS spectrum exhibits three characteristic peaks, corresponding to C=C (284.2 eV), C—O (285.1 eV), and C=O (288.2 eV), respectively. The high-resolution Co2p XPS spectrum (Fig. 5b) reveals characteristic peaks attributed to Co3+ at 780.1 and 795.3 eV, corresponding to Co2p3/2 and Co2p1/2, respectively. Additionally, peaks at 782.1 and 797.2 eV indicate the presence of Co2+, confirming the mixed valence state of Co in the catalyst[12]. Satellite peaks were also observed near 788.1 and 804.2 eV, further validating the XPS analysis. The high-resolution O1s XPS spectrum (Fig. 5c) exhibited three distinct peaks. These peaks can be attributed to M—O bonds (M=Co, Ru, 529.7 eV), oxygen vacancies, adsorbed hydroxyl groups (—OH, 531.3 eV), and H2O (532.7 eV) on the catalyst surface. The presence of these oxygen species suggests the diverse chemical environments and potential catalytic activity of the material[15-16]. Finally, Fig. 5d demonstrates that Ru0 was detected in the Ru-Co3O4/rGO catalyst, and the corresponding peaks of Ru3p were 463.8 and 485.5 eV[17]. The distinct peaks corresponding to the Ru element confirm its incorporation into the Ru-Co3O4/rGO material, further validating the XPS analysis and compositional characterization. The loading (mass fraction) of Ru and Co in the Ru-Co3O4/rGO catalyst was 0.441% and 24.6% respectively, which was determined by ICP-OES.
Figure 5
2.2 Electrocatalytic performance
Fig. 6a illustrates the LSV curves recorded for various catalysts during OER. The Ru-Co3O4/rGO catalyst exhibited an overpotential of 363.5 mV at a current density of 50 mA·cm·2, which was notably lower than the 402.3 mV observed for Co3O4/rGO without Ru doping. This significant reduction in overpotential indicates that Ru doping effectively enhances the OER performance of the composite material, leading to a higher catalytic efficiency. In practical applications, the application of a higher overpotential is often necessary to achieve a desired increase in current density. Ideally, a lower overpotential resulted in a faster increase in current density. The correlation between current density and overpotential can be mathematically described using the Butler-Volmer equation, which enables the calculation of the Tafel slope (b) for the OER from the LSV curve. A lower Tafel slope signifies a reduced overpotential requirement for the catalytic reaction to achieve a given current density[18-19]. As expected, Ru-Co3O4/rGO exhibited a low overpotential of 363.5 mV at a turnover frequency (TOF) of 0.000 12 s-1, which suggests that Ru-Co3O4/rGO possessed excellent intrinsic electrocatalytic OER activity. As depicted in Fig. 6b, the Tafel slope of Ru-Co3O4/rGO was 103 mV·dec-1, significantly lower than that of Co3O4/rGO (140 mV·dec-1). This lower Tafel slope suggests that Ru-Co3O4/rGO possessed higher catalytic activity and faster electrode reaction kinetic rates, further validating the beneficial effects of Ru doping on the OER performance of the composite material.
Figure 6
The OER performances of Ru-Co3O4/rGO with different Ru doping amounts (The volume of RuCl3 solution added was 100, 200, and 300 μL, respectively.) were tested. By comparing LSV curves, we found that when the volume of RuCl3 solution was 200 μL, the overpotential was much lower than that of undoped Ru samples and samples with the volume of RuCl3 solution of 100 and 300 μL (Fig. 7a). EIS (Fig. 7b) was used to characterize in surface charge transfer impedance of electrode during OER. Among them, Ru-Co3O4/rGO had the smallest radius of curvature, indicating the surface of Ru-Co3O4/rGO had the smallest charge transfer impedance, which is beneficial to improve the OER activity of the catalyst.
Figure 7
To quantitatively compare the electrochemical active surface areas (ECSAs) of Co3O4/rGO and Ru-Co3O4/rGO, cyclic voltammetry (CV) measurements were performed at varying scan rates (20, 40, 60, 80, and 100 mV·s·-1) to generate the CV curves depicted in Fig. 8a and 8b. The observed linear relationship between current density and scan rate provided insights into the electroactive surface during the catalytic process. Specifically, the double-layer capacitance (Cdl) was examined at the liquid/solid interface in a 1 mol·L-1 KOH electrolyte solution to study the electroactive surface[16-17]. Based on the CV curves of Co3O4/rGO and Ru-Co3O4/rGO at different scanning rates, the relationship between capacitive current and scanning rate was derived (Fig. 8c). The calculated value of Ru-Co3O4/rGO, based on the slope of the fitting line, was 6.5 mF·cm2, exceeding the corresponding value of 2.1 mF·cm2 for Co3O4/rGO. This significant increase indicates that Ru-Co3O4/rGO possesses a larger electrochemical surface area, enabling the exposure of more active sites and thereby enhancing the catalytic activity for OER[20-21]. The durability of Ru-Co3O4/rGO was assessed through a stability test presented in Fig. 8d, conducted at a potential of 1.52 V. The fluctuations observed in the current density-time (i-t) curve reflect the dynamic nature of the OER process at this potential. Initially, the generated oxygen bubbles adsorb onto the catalyst layer, reducing the contact area between the electrode and the electrolyte and decreasing current density. However, as the bubbles accumulated and eventually detached from the electrode surface, the contact area was restored, increasing current density. This cycle of oxygen generation and removal from the catalyst layer accounts for the jitter in the i-t curve. After 24 h of constant i-t testing, the current density of Ru-Co3O4/rGO decreased from the initial ca. 4 to 2 mA·cm-2. As shown in Fig. 9, After OER, the morphology of Ru-Co3O4/rGO was not significantly changed, as observed by SEM, compared to the original sample. However, the crystal shape was altered considerably, as observed by XRD. This change may be due to a small amount of Co2+ being converted to Co(OH)2 and CoOOH during the OER process.
Figure 8
Figure 9
3. Conclusions
In summary, a ZIF-67/rGO layered composite was constructed, and Ru-Co3O4/rGO catalysts were prepared using a high-temperature thermal method. To enhance the OER performance of the catalyst, Ru doped precursor was calcined to obtain Ru-Co3O4/rGO catalyst. Characterization analysis revealed that a small amount of Ru doping did not affect the nucleation and growth of the ZIF-67 precursor. Adding Ru resulted in similar or slightly enhanced OER performance for the Ru-Co3O4/rGO compared to Co3O4/rGO. The Ru-Co3O4/rGO catalyst has highly dispersed active sites, rich pore structure advantages, and excellent conductivity. It showed excellent OER catalytic activity in a saturated 1 mol·L-1 KOH electrolyte, demonstrating its great potential as a fuel cell cathode catalyst. Against the backdrop of the energy crisis and environmental pollution, this study fully integrated the advantages of MOFs and graphene materials, opening a new approach to designing efficient electrocatalysts. Based on the diversity and flexibility of MOF material structure and composition, this research approach has great potential for efficient catalyst design. It can be widely applied to optimize the performance of catalysts in different fields and catalytic systems.
-
-
[1]
HAN S M, MA Y, YUN Q B, WANG A L, ZHU Q S, ZHANG H, HE C H, XIA J, MENG X M, GAO L, CAO W B, LU Q P. The synergy of tensile strain and ligand effect in PtBi nanorings for boosting electrocatalytic alcohol oxidation[J]. Adv. Funct. Mater., 2022, 32(48): 2208760. doi: 10.1002/adfm.202208760
-
[2]
YU J, HE Q J, YANG G M, ZHOU W, SHAO Z P, NI M. Recent advances and perspective in ruthenium-based materials for electrochemical water splitting[J]. ACS Catal., 2019, 9(11): 9973-10011. doi: 10.1021/acscatal.9b02457
-
[3]
JIANG Y, XIE M, WU F, YE Z Q, ZHANG Y X, WANG Z H, ZHOU Y Z, LI L, CHEN R J. Cobalt selenide hollow polyhedron encapsulated in graphene for high-performance lithium/sodium storage[J]. Small, 2021, 17(40): 2102893. doi: 10.1002/smll.202102893
-
[4]
TAMILARASI B, JITHUL K, PANDEY J. Non-noble metal-based electro-catalyst for the oxygen evolution reaction (OER): Towards an active stable electro-catalyst for PEM water electrolysis[J]. Int. J. Hydrog. Energy, 2024, 58: 556-582. doi: 10.1016/j.ijhydene.2024.01.222
-
[5]
JIANG J, JIN B, MENG L. Research progress on nonprecious metal catalysts based on electrocatalytic oxygen evolution reaction[J]. J. Compos. Mater., 2023, 40(3): 1365-1380.
-
[6]
李京修, 赵媛, 薛建军, 何娉婷, 王玲. ZIF67衍生硫化钴/多孔碳复合催化剂的制备及其电催化性能[J]. 无机化学学报, 2019,35,(8): 1363-1370. LI J X, ZHAO Y, XUE J J, HE P T, WANG L. Preparation and electrocatalytic property of cobalt sulfide/porous carbon composite catalyst derived from ZIF67[J]. Chinese J. Inorg. Chem., 2019, 35(8): 1363-1370.
-
[7]
ZHOU G Y, ZHANG S, ZHU Y F, LI J, SUN K, PANG H, ZHANG M Y, TANG Y W, XU L. Manipulating the rectifying contact between ultrafine Ru nanoclusters and N-doped carbon nanofibers for high-efficiency pH-universal electrocatalytic hydrogen evolution[J]. Small, 2023, 19: 2206781. doi: 10.1002/smll.202206781
-
[8]
何应, 张宇, 何清, 刘辉, 李亮. 用于高性能超级电容器的金属有机骨架衍生Co(OH)2/C-N@GP纳米复合材料[J]. 无机化学学报, 2023,39,(12): 2432-2440. doi: 10.11862/CJIC.2023.192HE Y, ZHANG Y, HE Q, LIU H, LI L. Metal-organic-frameworks-derived Co(OH)2/nitrogen-doped carbon graphene nanocomposites for high-performance supercapacitors[J]. Chinese J. Inorg. Chem., 2023, 39(12): 2432-2440. doi: 10.11862/CJIC.2023.192
-
[9]
刘逸帆, 张沾, 朱荣妹, 邱紫铭, 庞欢. 三维花状铜基复合物及其低温煅烧衍生物用于高效析氧反应[J]. 无机化学学报, 2024,40,(5): 979-990. LIU Y F, ZHANG Z, ZHU R M, QIU Z M, PANG H. A three-dimensional flower-like Cu-based composite and its low-temperature calcination derivatives for efficient oxygen evolution reaction[J]. Chinese J. Inorg. Chem., 2024, 40(5): 979-990.
-
[10]
ZHANG H, MENG G, WEI T R, DING J Y, LIU Q, LUO J, LIU X J. Co-doping promotes the alkaline overall seawater electrolysis performance over MnPSe3 nanosheets[J]. Chem. Commun., 2023, 59: 12144-12147. doi: 10.1039/D3CC03434H
-
[11]
PENG Z M, ZHANG G, QI G C, ZHANG H, LIU Q, HU G Z, LUO J, LIU X J. Nanostructured Pt@RuOx catalyst for boosting overall acidic seawater splitting[J]. Chinese J. Struct. Chem., 2024, 43: 100191. doi: 10.1016/j.cjsc.2023.100191
-
[12]
WANG T W, ZHANG Q, LIAN K, QI G C, LIU Q, FENG L, HU G Z, LUO J, LIU X J. Fe nanoparticles confined by multiple-heteroatom-doped carbon frameworks for aqueous Zn-air battery driving CO2 electrolysis[J]. J. Colloid Interface Sci., 2024, 655: 176-186. doi: 10.1016/j.jcis.2023.10.157
-
[13]
XU Y Y, DENG P L, CHEN G D, CHEN J X, YAN Y, QI K, LIU H F, XIA B Y. 2D nitrogen-doped carbon nanotubes/graphene hybrid as bifunctional oxygen electrocatalyst for long-life rechargeable Zn-air batteries[J]. Adv. Funct. Mater., 2020, 30(6): 1906081. doi: 10.1002/adfm.201906081
-
[14]
ZHOU G Y, WEI C, LI Z J, HE B, LIU Z Y, LI J. Fe doping and interface engineering-induced dual electronic regulation of CoSe2/Co9S8 nanorod arrays for enhanced electrochemical oxygen evolution[J]. Inorg. Chem. Front., 2024, 11: 164-171.
-
[15]
FENG T L, YU G T, TAO S Y, ZHU S J, KU R Q, ZHANG R, ZENG Q S, YANG M X, CHEN Y X, CHEN W H, CHEN W, YANG B. A highly efficient overall water-splitting ruthenium-cobalt alloy electrocatalyst across a wide pH range via electronic coupling with carbon dots[J]. J. Mater. Chem. A, 2020, 8: 9638-9645. doi: 10.1039/D0TA02496A
-
[16]
XU H T, ZHOU X Y, LIN X R, WU Y H, QIU H J. Electronic interaction between in situ formed RuO2 clusters and a nanoporous Zn3V3O8 support and its use in the oxygen evolution reaction[J]. ACS Appl. Mater. Interfaces, 2021, 13(46): 54951-54958. doi: 10.1021/acsami.1c15119
-
[17]
李英杰, 王鑫, 周昱成. NiFe水滑石纳米片阵列负载Ru用于提升碱性电催化析氢和析氧性能[J]. 无机化学学报, 2023,39,(10): 1905-1913. LI Y J, WANG X, ZHOU Y C. Ru loaded on NiFe layered double hydroxide nanosheet arrays for boosting alkaline electrocatalytic hydrogen evolution and oxygen evolution abilities[J]. Chinese J. Inorg. Chem., 2023, 39(10): 1905-1913.
-
[18]
李航洋, 周凌, 韩博鸣, 谷向阳, 叶晴岚, 许雪棠, 王凡, 李斌. CoFeOx/MoO2准平行纳米阵列电极的析氢性能[J]. 无机化学学报, 2023,39,(7): 1275-1286. LI H Y, ZHOU L, HAN B M, GU X Y, YE Q L, XU X T, WANG F, LI B. Hydrogen evolution performance of CoFeOx/MoO2 quasi-parallel nanoarray electrode[J]. Chinese J. Inorg. Chem., 2023, 39(7): 1275-1286.
-
[19]
YAN C S, CHEN G, ZHOU X, SUN J X, LV C D. Template-based engineering of carbon-doped Co3O4 hollow nanofibers as anode materials for lithium-ion batteries[J]. Adv. Funct. Mater., 2016, 26: 1428-1436. doi: 10.1002/adfm.201504695
-
[20]
MA L B, SHEN X P, ZHOU H, ZHU G X, CHEN K M. CoP nanoparticles deposited on reduced graphene oxide sheets as an active electrocatalyst for the hydrogen evolution reaction[J]. J. Mater. Chem. A, 2015, 3: 5337-5343. doi: 10.1039/C4TA06458E
-
[21]
HE B, SONG J J, LI H Q, LI Y B, TANG Y W, SU Z, HAO Q L. Construction of hollow MoS2@N-C nanotubes for lithium-ion batteries[J]. Vacuum, 2022, 205: 111386. doi: 10.1016/j.vacuum.2022.111386
-
[1]
-
计量
- PDF下载量: 0
- 文章访问数: 5
- HTML全文浏览量: 0