Recent progress of in situ/operando characterization techniques for electrocatalytic energy conversion reaction

Zhao Li Huimin Yang Wenjing Cheng Lin Tian

Citation:  Zhao Li, Huimin Yang, Wenjing Cheng, Lin Tian. Recent progress of in situ/operando characterization techniques for electrocatalytic energy conversion reaction[J]. Chinese Chemical Letters, 2024, 35(9): 109237. doi: 10.1016/j.cclet.2023.109237 shu

Recent progress of in situ/operando characterization techniques for electrocatalytic energy conversion reaction

English

  • The increasing demand in energy for the future development stimulated a great growth in research on pursuing renewable and carbon-neutral energy supplies [13]. Nowadays, many efficient conversion technologies, including fuel cells, water splitting, and electrocatalytic CO2 reduction, play a crucial role and appeal extensive interest as potential approaches to facilitate the energy conversion and supply highly valuable products [46]. As well known, catalysts play a crucial role in realizing efficient electrochemical reactions by significantly accelerating reaction kinetics [715]. Accordingly, the development of cost-effective and high-performance electrocatalysts is crucial for enhancing the energy conversion process. In essence, the catalysts used for electrochemical reactions should be highly efficient, stable, and economical [1618]. Despite significant progress in the rational design and fabrication of advanced electrocatalysts, a fundamental question remains unanswered, "what is the nature of active sites under working conditions?" To this end, precisely analyzing the evolution process and comprehensively studying the mechanism are significant for the design and exploration of expected electrocatalysts [19,20].

    A deep understanding on the reaction steps, reaction mechanisms, and adsorption energy of intermediates is still lacking, which require more endeavors devoted. In recent years, some advanced in situ characterization techniques have been developed to investigate the morphology, structure, composition, and chemical valences of the catalyst [2123]. Many advanced in situ technologies have been well developed for identifying the true reconstruction process of the catalysts, such as the in situ Fourier transform infrared spectroscopy [24,25]. Previous works have demonstrated that in situ characterization techniques play a pivotal role in deeply understanding the dynamic evolution of reaction sites and the specific reaction mechanisms [26,27]. As expected, in situ characterizations have emerged as promising techniques for researchers due to their significant achievements in investigating energy-related reactions [28].

    Aiming to describe the recent achievements in this interesting field, we have organized a comprehensive review to summarize the significant advancements in situ techniques for studying the reconstruction in the electrochemical reactions (Scheme 1). First, we manifest the fundamental principles of various in situ characterization technologies, including XRD, XAS, Raman, FTIR, and TEM. Then, a section on the applications of these in situ techniques used for investigating the dynamic evolution process and the nature of active sites of catalysts during electrochemical operation is also presented. At the end of this review, the future development and application of in situ technologies in this fascinating field are systematically discussed.

    Scheme 1

    Scheme 1.  Scheme of the in situ technologies and their applications in energy conversion.

    In situ characterization is a technique used to study the properties of materials and surfaces by observing and measuring the structure, morphology, composition and electrochemical properties of materials under in situ conditions [29].

    The basic principle is to place the material to be studied in an experimental device, and then observe and measure it by various means. These means may include optical microscopy, scanning electron microscopy, transmission electron microscopy, atomic force microscopy, Raman spectrometer, X-ray diffractometer, electrochemical workstation, etc. Through these techniques, the changes in structure, morphology, composition and electrochemical properties of materials can be observed and measured in real time [30].

    In situ characterization technology has been widely used in materials science, energy storage and conversion, catalyst research, biomedicine and other fields. For example, in the field of materials science, in situ characterization technology can be used to study the process of material phase transition, crystal growth, interface reaction, etc. In the field of energy-related application, it can be used to study the changes in the performance of energy storage and conversion devices such as batteries and fuel cells. In the research of catalyst, it can be used to study the activity and stability of catalyst. In the biomedical field, it can be used to study properties such as biocompatibility and drug release of biological materials. In short, in situ characterization technology provides an important avenue and method for the research of material science and related fields by observing and measuring the property changes of materials under experimental conditions.

    Electron microscopy, especially for TEM, is one of the most effective strategies for studying the microstructure and morphology of the materials [31,32]. The rapid development of TEM analysis techniques has boosted the explosion of nanoscience and nanotechnology. Particularly, in situ TEM is now attracting increasing interest as potential characterization technique for revealing the chemical and physical process dynamics at atomic resolutions [33]. For instance, Ma et al. [34] demonstrated the syntheses of free-supporting Pt and Pt-Ni nanowire-network electrocatalysts, and the formation mechanism was also investigated by an in situ TEM study. According to the in situ TEM images (Figs. 1a-c), it was observed that the Pt nanowires were formed through the oriented attachment of individual nanoparticles. Besides in situ TEM, a combination of other in situ electron microscopy is also demonstrated to be conducive to the investigation of the dynamic structural and morphological evolution during electrocatalysis.

    Figure 1

    Figure 1.  (a-c) In situ TEM observation of Pt nanowire growth at different stages. Reproduced with permission [34]. Copyright 2017, Wiley-VCH.

    XRD shows diffraction patterns for different crystal phases, which can thus be a fingerprint for the analytical task. As well known, the majority of XRD investigations of materials can be used to identify the crystal phases. Therefore, the collection of an XRD pattern can be a potential strategy to study the crystal structures by altering the scattering vector, and XRD is also appearing as a promising technique for analyzing the crystal structure of materials [35] More recently, in situ XRD techniques are also developed and used for investigating the crystal plane and structure of the catalysts during electrochemical operation, providing a facile method for monitoring the structure evolution of catalyst and identifying the nature of catalytic active site [3638]. As an illustration, Yuan et al. [39] have studied the structural reconstruction of Bi(OH)3 during electrocatalytic CO2 reduction reaction (CO2RR) through a combination of ex situ XRD and in situ Raman (Figs. 2a-e). Specifically, the sample underwent XRD characterization each 5 mins during the electrochemical activation process in CO2-saturated 0.1 mol/L KHCO3 at −1.0 VRHE. According to the ex situ XRD pattern, it is clearly observed that the peaks at 47.0°, 42.4°, and 32.9° can be indexed to the Bi(OH)3, while the other peak located at 48.7°, 45.9°, 44.6°, 39.6° and 37.9° are associated with the metallic Bi (Figs. 2f and g). The ex situ XRD pattern suggested that the transformation of Bi(OH)3 to metallic Bi during electrocatalytic CO2RR.

    Figure 2

    Figure 2.  (a) Illustration of the fabrication of Bi(OH)3 and the reconstruction to pits-Bi nanosheets. (b) TEM image, (c, d) STEM image and (e) AC-STEM image of the pits-Bi nanosheets. (f) Ex situ XRD and (g) Raman spectra for the in situ reconstruction process. Reproduced with permission [39]. Copyright 2022, Wiley-VCH.

    The principle of in situ Raman spectroscopy is to use the scattering phenomenon that the incident light frequency is obviously changed by material surface molecules [4043]. Operando Raman spectroscopy is also conducted to establish the evolution of electrocatalysts during reactions process and emerging as a potential approach to explore the structure-performance relationship of catalysts. For example, Wu et al. [44] have performed the in situ surface-enhanced Raman analysis to clarify the role of cation defect of the catalyst during electrochemical reaction. Specifically, they have synthesized the defective NiFe-LDH nanosheets via a facile aprotic-solvent-solvation-induced strategy. According to the Raman spectra, it is clearly observed that the as-generated cationic vacancy defects tend to exist as VM (M = Ni/Fe) (Figs. 3a and b). As the voltage increases, the crystalline Ni(OH)x on the surface of NiFe-LDH gradually transforms into a disordered state. When oxygen bubbles begin to evolve under sufficiently high voltage, local NiOOH species appear as residual products (Figs. 3c and d). Therefore, in situ Raman spectra are beneficial for understanding the reconstruction behaviors of the NiFe-LDH. Besides better understanding the surface reconstruction process, in situ Raman technology is also advantageous for providing insights into the adsorption and transfer of the reaction intermediates. Previous studies have reported that in situ Raman technology can be utilized to identify the electron transfer and hydrogen spillover effects, which are critical for comprehending the reaction mechanisms.

    Figure 3

    Figure 3.  In situ Raman spectroscopy tests for studying the transformation of (a) the pristine NiFe-LDH and (b) d-NiFe-LDH during OER. (c, d) The evolution of I528/I457 of NiFe-LDH and d-NiFe-LDH. Reproduced with permission [44]. Copyright 2021, Wiley-VCH.

    In situ FTIR, pioneered by Bewick et al. [24] has been utilized to acquire the molecular information regarding ionic and neutral adsorbates at the electrode as well as solution species during electrochemical reactions [45,46]. In situ FTIR is employed to characterize the adsorbates, molecules, and the intermediated species. Many researches employed in situ FTIR to study the electrocatalytic processes, including the study of the reaction mechanisms of organic molecules oxidation, CO2RR, and the corresponding dynamic process [47,48]. Sun et al. have used this technique to investigate the reaction mechanism and intermediated species of many electrocatalytic processes [49]. For instance, they have investigated the reaction mechanisms of the propene electrooxidation on Pd/C and PdO/C via in situ FTIR [50]. Based on the in situ attenuated total reflection FTIR (Figs. 4a and b), a novel reaction pathway was distinctly observed compared with unconventional thermocatalysis (Figs. 4c-f). This work highlights the significant role of in situ FTIR in elucidating the electrochemical reaction mechanisms involved in propene oxidation.

    Figure 4

    Figure 4.  In situ ATR-FTIR spectra of C3H6 oxidation on (a) PdO/C and (b) Pd/C at different potentials in a C3H6-saturated 0.1 mol/L HClO4 solution. Different propene adsorbed configurations (c) on PdO below 0.80 V, (d) above 0.8 V, (e) on Pd below 0.80 V, and (f) above 0.80 V. Reproduced with permission [50]. Copyright 2021, American Chemical Society.

    XAS is extensively employed to investigate the local electronic and atomic structure of catalysts [5153]. Compared with the XRD, XAS is sensitive to the electronic structure of materials, which can thus be utilized as advanced characterizations to measure the electronic properties of the catalysts [54]. Recently, XAS characterization techniques have been widely utilized in the electronic structure analysis of nanocatalysts and single-atom catalysts [55], which have gained widespread utilization and extensive investigation [56,57]. For example, Yang and coworkers explored the structural and chemical transformation of the AgCu nanocatalysts during the electrochemical CO2RR [58]. According to a systematic investigation, it has been revealed that the two types of catalysts undergo similar structural evolution under CO2RR conditions (Figs. 5a-d). In the case of intermixed AgCu catalysts, Cu may undergo leaching and subsequently become enriched on the surface of particles or recrystallize elsewhere as new particles (Fig. 5e).

    Figure 5

    Figure 5.  (a) Operando XANES of Cu K-edge of the I-AgCu as a function of the reaction time at −1.0 V vs. RHE in CO2-saturated 0.1 mol/L KHCO3. (b) XANES pre-edges of I-AgCu magnified from the dashed box region in (a) and the comparison with Cu and Cu2O references (dashed lines). (c, d) Operando XANES and EXAFS spectra of I-AgCu for CO2RR. (e) The in situ evolution of intermixed and phase-separated AgCu catalysts in electrochemical CO2RR. Reproduced with permission [58]. Copyright 2023, American Chemical Society.

    Fuel cells offer a green and high-efficiency technology for the conversion of chemical energy into electrical energy [59,60]. The development of high-performance ORR catalysts largely relies on a comprehensive understanding on the redox behavior of the active sites [6167]. Therefore, the in situ/operando measurement techniques are highly imperative [6871]. Recent works have confirmed the significant role of in situ techniques in enhancing our understanding on the redox behavior of metal sites during electrochemical ORR. For instance, Huang and colleagues reported the in situ synthesis of Ru-doped Pt3Co octahedral nanoparticles grown on carbon (Ru–Pt3Co/C) [72], which exhibited remarkable ORR activity up to 1.05 A/mgPt. Ru doping effectively mitigates Co corrosion, substantially improving the morphological stability of Ru-Pt3Co/C. In situ XAFS analysis revealed that Ru promoted the Pt atom inversion (Figs. 6a-d), thereby limiting over-reduction of Pt sites to enhance the stability of Ru-Pt3Co/C. Additionally, DFT calculations demonstrated that adding Ru atom accelerated the oxygen intermediate breach and desorption (Fig. 6e).

    Figure 6

    Figure 6.  In situ XAS tests. (a) Scheme of the electrochemical setup for the in situ XAS experiment. Pt L3-edge XANES spectra of (b) Ru-Pt3Co/C and (c) Pt3Co/C during the ORR. (d) Peak positions of Pt 4f7/2 XPS spectra for Ru-Pt3Co/C, Pt3Co/C, and commercial Pt/C. (e) The corresponding ORR mechanism of Ru-Pt3Co/C. Schematic illustration of the (f) SHINERS study of the ORR at Pt (311) surface and the (g) ORR mechanism. (a-e) Reproduced with permission [72]. Copyright 2021, American Chemical Society. (f, g) Reproduced with permission [74]. Copyright 2020, American Chemical Society.

    In situ characterization techniques can be employed not only to investigate the nature of active sites, but also to examine the intermediates during electrochemical ORR [73]. Dong et al. employed the shell-isolated nanoparticle-enhanced Raman spectroscopy (SHINERS) to study the ORR mechanism at Pt (211) and Pt (311) surfaces [74]. Through control and isotope substitution experiments, clear evidence of OH and OOH intermediates at Pt (311) and Pt (211) surfaces can be observed. It was concluded that the difference in OOH adsorption on high-index surfaces posed a significant impact on the electrocatalytic ORR performance (Figs. 6f and g). The research has enhanced the comprehension of ORR mechanism on high-index Pt (hkl) surface and highlighted the crucial significance of in situ techniques for a thorough understanding of the ORR mechanism.

    The electrooxidation of hydrogen, CO, or small organic molecules is also a crucial electrode reaction of fuel cells [75,76]. Appraising the reaction intermediates is of crucial importance to unveil the pathways and mechanisms. Taking electrocatalytic hydrogen oxidation reaction (HOR) as an example, the HOR may follow the Heyrovsky-Volmer or Tafel-Volmer mechanism in acidic solution, where the H+ can be generated on the catalyst surface from adsorbed hydrogen and then convert to a proton by releasing an electron [77,78]. The EOR process typically involves a dual-pathway mechanism, namely the C1 and C2 pathways. The former can generate adsorbed carbon monoxide (COad) and carbonate radical (CO32−) intermediates [7982]. In contrast, the C2 pathway is involved in the generation of acetate radical (CH3COO). Therefore, measuring and analyzing the intermediated species will be significant for deeply understanding on the reaction pathways and mechanisms. Huang and coworkers investigated the EOR mechanism on the Pd-Sb hexagonal nanoplate (Figs. 7a-d) [83]. According to the in situ ATR-SEIRAS spectra, it is clearly found that there are two forms of CO adsorption on Pd8Sb3 HPs/C surface (bridge-bonded CO (COB) and multiply-bonded CO (COM)), while Pd/C has three forms of CO adsorption (linear-bonded CO (COL), COB and COM) (Figs. 7e and f). The negligible COL on Pd8Sb3 HPs/C reveals that the linear CO adsorption capacity of Pd8Sb3 HPs/C is weakened due to the introduction of Sb, and the stronger COM and COB on Pd8Sb3 HPs/C suggest a more positive CO, indicating that Pd8Sb3 HPs/C can greatly facilitate the C2 pathway of the EOR.

    Figure 7

    Figure 7.  (a, b) HAADF-STEM images and the (c) elemental mapping images of the Pd8Sb3 HPs. (d) CV curves of the Pd8Sb3 HPs/C, Pd/C, and Pt/C for EOR. In situ ATR-SEIRAS spectra of (e) Pd8Sb3 HPs/C and (f) Pd/C during CV cycling. (g) Atomic-resolution HAADF-STEM image of the 22% YOx/MoOx-Pt nanowires. (h) Pt mass-normalized MOR, (i) comparison on the specific activities and mass activities of 22% YOx/MoOx-Pt nanowires and other references. (j) In situ FTIR spectra recorded from 0.15 V to 1.20 V and (k) magnified in situ FTIR spectra recorded from 0.15 V to 0.60 V versus RHE during MOR on 22% YOx/MoOx-Pt. (l, m) In situ FTIR spectra during MOR. (a-f) Reproduced with permission [83]. Copyright 2022, Wiley-VCH. (g-m) Reproduced with permission [84]. Copyright 2021, Wiley-VCH.

    Guo et al. employed the in situ FTIR test to obtain in-depth understanding on the improvement MOF performance on the YOx/MoOx-Pt catalyst (Figs. 7g-i) [84]. In situ FTIR spectra showed several bands between 2500 cm−1 and 1000 cm−1, where the downward bands at 2341 cm−1 was assigned to the asymmetric stretching vibration of CO2, implying the complete oxidation of methanol (Figs. 7j-m). It was also demonstrated that MoOx and YOx contributed to the enhanced anti-CO poisoning capability of Pt nanowires. Moreover, DFT calculations further revealed that surface Y and Mo atoms, with enabled oxidation states, facilitated the binding of COOH* to the surface via both C and O atoms. This effectively reduced the free energy barriers for CO* oxidation into COOH*. As a result, a mass activity of 2.10 A/mgPt and a high specific activity of 3.35 mA/cm2 were achieved for the optimized 22% YOx/MoOx-Pt.

    OER plays a critical role in electrochemical water splitting as it involves a four-electron transfer process, which directly determines the energy barrier and reaction efficiency of the overall reaction [8590]. Therefore, significant efforts have been dedicated to the development and production of more efficient electrocatalysts to enhance the reaction kinetics of OER [9092]. It is demonstrated that the in situ characterizations play a key role in identifying the nature of true active sites of catalysts during OER [9395]. For instance, Zheng et al. [96] endeavored to explore the electrochemical reconstruction of MOFs and elucidate the function of in situ generated species during electrochemical OER (Figs. 8a and b). Dramatic morphological changes that expose electron-accessible Co sites are clearly observed (Figs. 8c-e). Subsequent OER experiments suggest that the dominating active sites are CoOOH species produced from α/β-Co(OH)2.

    Figure 8

    Figure 8.  (a) In situ UV–vis and (b) Raman devices. (c) In situ UV–vis spectroelectrochemical study between 0.925 V and 1.705 V. (d) In situ Raman spectroelectrochemical study between 0.925 V and 1.585 V. (e) Scheme for the transformation of ZIF-67 to α-Co(OH)2 and β-Co(OH)2. Reproduced with permission [96]. Copyright 2019, American Chemical Society.

    Electrocatalytic HER offers a promising strategy to provide clean energy carriers to alleviate the energy crisis [97101]. According to the Sabatier's principle and volcano plot of hydrogen intermediate adsorption (∆GH), the best HER catalyst should possess the close-to-zero ∆GH [102,103]. To this end, designing the synthesizing the advanced HER electrocatalysts with close-to-zero ∆GH is highly favorable for promoting the hydrogen production [104]. However, it is also demonstrated that the hydrogen spillover effect and surface reconstruction of catalyst during electrochemical reactions are also crucial for affecting the HER performance [105108]. Operating in situ characterization tests is imperative to understand the reaction mechanisms and reconstruction process of catalyst, which will provide guidance for future rational design and fabrication of more excellent HER catalysts. Using Wei's work as an example, they have reported the preparation of Pt/TiO2 catalyst with abundant deficient oxygen vacancies (VO-deficient Pt/TiO2) and rich oxygen vacancies (VO-rich Pt/TiO2) (Figs. 9a-c) [109]. Remarkably, it is measured that the VO-rich Pt/TiO2 has a mass activity of 45.28 A/mgPt at −0.1 V vs. RHE, which is 58.8 and 16.7 times higher than those of commercial Pt/C and VO-deficient Pt/TiO2, respectively (Figs. 9d and e). In order to uncover the outstanding electrocatalytic HER performance, they have operated the in situ Raman tests and DFT calculations. Based on the in situ Raman spectra, it is obviously observed the H* transfer from Pt to TiO2 support. In contrast, no obvious peak related with Eg(1) are observed during the catalytic process due to the absence of H* (Fig. 9f). DFT calculations have demonstrated that TiO2 with abundant oxygen vacancies can facilitate both reversed charge transfer and hydrogen spillover from Pt to the TiO2 support, thereby contributing to the improved electrocatalytic performance for HER (Figs. 9g-k).

    Figure 9

    Figure 9.  (a) Illustration of the synthesis procedures of the VO-rich Pt/TiO2 and VO-deficient Pt/TiO2 catalysts. (b) ESR spectra and (c) XPS profiles. (d) HER polarization curves and (e) corresponding mass activity at the overpotential of 100 mV. (f) DFT calculation of the free energy. (g) In situ Raman spectra of VO-rich Pt/TiO2 at −0.1 V vs. RHE with different electrolysis durations. (h) Magnified view of the Eg(1) mode. (i–k) In situ Raman spectra of VO-rich Pt/TiO2 (i), VO-deficient Pt/TiO2 (j), and TiO2 (k) at a potential of −0.2 V vs. RHE. Reproduced with permission [109]. Copyright 2021, Wiley-VCH.

    Electrocatalytic CO2RR is a promising renewable technology for the conversion of greenhouse gas into value-added liquid fuels and chemicals [110,111]. Cu, Pd, Pb, and Bi-based catalysts are established as the predominant candidate for selective CO2 electroreduction to multicarbon products [111113]. However, the comprehension of the nature of active sites during reaction conditions remains limited, particularly for some electrocatalysts with complicated components. To this end, in situ characterization technologies are emerging as promising candidates for addressing these issues. Yang and colleagues emphasized the importance of understanding and regulating the dynamic characteristics of nanocatalysts to attain optimized CO2 electrolysis [114]. Specifically, they have systematically investigated the dynamic aspects of both the active site and its surrounding reaction microenvironment by analyzing the structural transformation via in situ/operando characterizations [115]. More recently, they also identified the nature of reaction sites of the high-activity Cu nanocatalysts through time-resolved operando techniques (Figs. 10a and b). After a systematic investigation, it is clearly found that an ensemble of monodisperse Cu nanoparticles would undergo a structural transformation process. According to the operando EC-STEM image, 4D-STEM image and XANES spectra, it is uncovered that the Cu nanograins underwent a typical structural reconstruction (Figs. 10c-m). As a result, a 7 nm Cu nanoparticle ensemble, with a unity fraction of active Cu nanograins, exhibits sixfold higher C2+ selectivity than the 18 nm counterpart with one-third of active Cu nanograins (Figs. 10n-p).

    Figure 10

    Figure 10.  (a) Operando/in situ methods used to reveal the dynamic changes during electrocatalytic CO2RR. (b) Scheme of operando EC-STEM and 4D-STEM under CO2RR-relevant conditions. (c-m) EC-STEM images of the initial growth of Cu nanocatalysts after a single negative-direction LSV scan, from 0.4 V to 0 V. (n) Quantitative analysis of relative fraction of metallic Cu. Operando XANES (o) and corresponding quantitative analysis (p) of 18 nm nanoparticles at −0.8 V. Reproduced with permission [115]. Copyright 2023, Nature Publishing Group.

    In summary, we have focused on the in situ characterizations utilized for identifying the true active sites of catalyst and better understanding the reaction mechanism of electrocatalytic reactions. The aim of this review is to clarify the great significance of the in situ characterization techniques on various electrocatalytic reactions. By systematically discussing the fundamentals of various in situ technologies, such as in situ electron microscopy, in situ XRD, in situ FTIR, in situ Raman, in situ XAS, we hope to provide readers with insightful ideas about the development of in situ technologies. Subsequently, the applications of various in situ technologies used for identifying the nature of reaction sites and the surface adsorption of intermediated species are also discussed in detail, which will benefit for the development of advanced electrocatalysts. Despite the significant progress achieved in this interesting field, a comprehensive visualization is still far from being attained. More efforts need to be devoted and some challenges must also be overcome.

    Notably, many factors can govern the electrocatalytic performance of the catalysts, including the oxidation state, local electronic structure. Refining the monitoring of electrocatalytic processes through a single in situ technique and interpretation of mechanisms may result in an incomplete assessment. Therefore, it is of utmost importance to integrate multiple in situ analytical techniques to provide a comprehensive insight into the underlying mechanisms of electrochemical reactions.

    In addition, it should be noted that a significant disparity exists between the in situ employed devices and actual energy devices. Currently, most researches are focused on the half-reaction of energy conversion devices, such as detecting the intermediates of small molecules in fuel cells. To this end, optimizing the electrochemical performance of in situ cells related to real energy devices should be the future direction.

    Another important challenge that should be overcome is to establish a reconstruction-performance relationship. Previous works have demonstrated that the in situ reconstruction of some catalysts can yield highly active species to further enhance the electrocatalytic performance, while some are hindering the surface reconstruction to achieve high electrocatalytic performance. For example, the surface reconstruction of Ni(OH)2 can yield highly active NiOOH to boost electrocatalytic OER, while some works are hampering the surface reconstruction of Ni(OH)2 by incorporating the Fe3+. It is still unclear whether surface reconstruction can promote the electrocatalytic performance or not.

    The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

    This work was financed by the National Natural Science Foundation of China (No. 22202169), Xuzhou Science and Technology Plan Project of China (No. KC21294).


    1. [1]

      Y. Yue, P. Cai, K. Xu, et al., J. Am. Chem. Soc. 143 (2021) 18052–18060. doi: 10.1021/jacs.1c06238

    2. [2]

      Y.R. Hong, S. Dutta, S.W. Jang, et al., J. Am. Chem. Soc. 144 (2022) 9033–9043. doi: 10.1021/jacs.2c01589

    3. [3]

      J. Qin, H. Liu, P. Zou, et al., J. Am. Chem. Soc. 144 (2022) 2197–2207. doi: 10.1021/jacs.1c11331

    4. [4]

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

    5. [5]

      J. Yang, H. Qi, A. Li, et al., J. Am. Chem. Soc. 144 (2022) 12062–12071. doi: 10.1021/jacs.2c02262

    6. [6]

      J. Liu, W. Niu, G. Liu, et al., J. Am. Chem. Soc. 143 (2021) 4387–4396. doi: 10.1021/jacs.1c00612

    7. [7]

      L. Zhang, J. Liang, L. Yue, et al., Nano Res. Energy 1 (2022) 9120028. doi: 10.26599/NRE.2022.9120028

    8. [8]

      L. Zhang, L. Li, J. Liang, et al., Inorg. Chem. Front. 10 (2023) 2766–2775. doi: 10.1039/d3qi00341h

    9. [9]

      T. Xu, D. Ma, T. Li, et al., Chem. Commun. 56 (2020) 14031–14034. doi: 10.1039/d0cc05917j

    10. [10]

      L. Ouyang, X. He, Y. Sun, et al., Inorg. Chem. Front. 9 (2022) 6602–6607. doi: 10.1039/d2qi02205b

    11. [11]

      L. Tian, Y. Liu, C. He, et al., Chem. Rec. 23 (2023) e202200213. doi: 10.1002/tcr.202200213

    12. [12]

      L. Tian, Z. Huang, X. Lu, et al., Inorg. Chem. 62 (2023) 1659–1666. doi: 10.1021/acs.inorgchem.2c04092

    13. [13]

      X. Lu, T. Wang, M. Cao, et al., Int. J. Hydrogen Energy 48 (2023) 34740–34749. doi: 10.1016/j.ijhydene.2023.04.257

    14. [14]

      X. Lu, M. Du, T. Wang, et al., Int. J. Hydrogen Energy 48 (2023) 34009–34017. doi: 10.1016/j.ijhydene.2023.05.105

    15. [15]

      L. Tian, X. Pang, H. Xu, et al., Inorg. Chem. 61 (2022) 16944–16951. doi: 10.1021/acs.inorgchem.2c03060

    16. [16]

      H. Xu, L. Yang, K. Wang, et al., Inorg. Chem. 62 (2023) 11271–11277. doi: 10.1021/acs.inorgchem.3c01701

    17. [17]

      H. Xu, K. Wang, L. Jin, et al., J. Colloid Interface Sci. 650 (2023) 1500–1508. doi: 10.1016/j.jcis.2023.07.109

    18. [18]

      H. Xu, K. Wang, G. He, et al., J. Mater. Chem. A 11 (2023) 17609–17615. doi: 10.1039/d3ta01766d

    19. [19]

      P. Geng, L. Wang, M. Du, et al., Adv. Mater. 34 (2022) e2107836. doi: 10.1002/adma.202107836

    20. [20]

      M. Du, P. Geng, C. Pei, et al., Angew. Chem. Int. Ed. 61 (2022) e202209350. doi: 10.1002/anie.202209350

    21. [21]

      D. Friebel, V. Viswanathan, D.J. Miller, et al., J. Am. Chem. Soc. 134 (2012) 9664–9671. doi: 10.1021/ja3003765

    22. [22]

      L. Huang, J.Y. Sun, S.H. Cao, et al., ACS Catal. 6 (2016) 7686–7695. doi: 10.1021/acscatal.6b02097

    23. [23]

      Z. Zhao, Z. Yuan, Z. Fang, et al., Adv. Sci. 5 (2018) e1800760. doi: 10.1002/advs.201800760

    24. [24]

      J.Y. Ye, Y.X. Jiang, T. Sheng, et al., Nano Energy 29 (2016) 414–427. doi: 10.1016/j.nanoen.2016.06.023

    25. [25]

      X. Hu, Z. Xiao, W. Wang, et al., J. Am. Chem. Soc. 145 (2023) 15109–15117. doi: 10.1021/jacs.3c00262

    26. [26]

      X. Chen, W. Li, G. Zhang, et al., Mater. Today Chem. 23 (2022) 100705. doi: 10.1016/j.mtchem.2021.100705

    27. [27]

      H. Zhou, S. Zheng, X. Guo, et al., J. Colloid Interface Sci. 628 (2022) 24–32. doi: 10.1016/j.jcis.2022.08.043

    28. [28]

      Y. Wang, Y. Li, T. Heine, J. Am. Chem. Soc. 140 (2018) 12732–12735. doi: 10.1021/jacs.8b08682

    29. [29]

      J. Liu, L. Guo, Matter 4 (2021) 2850–2873. doi: 10.1016/j.matt.2021.05.025

    30. [30]

      J. Zhao, J. Lian, Z. Zhao, Nano Micro Lett. 15 (2022) 19.

    31. [31]

      S. Liu, Y. Zhang, X. Mao, Energy Environ. Sci. 15 (2022) 1672–1681. doi: 10.1039/d1ee03687d

    32. [32]

      M. Wang, M. Wang, C. Zhan, et al., J. Mater. Chem. A 10 (2022) 18972–18977. doi: 10.1039/d2ta04882e

    33. [33]

      C. Zhang, K.L. Firestein, J.F.S. Fernando, et al., Adv. Mater. 32 (2020) e1904094. doi: 10.1002/adma.201904094

    34. [34]

      Y. Ma, W. Gao, H. Shan, et al., Adv. Mater. 29 (2017) e1703460. doi: 10.1002/adma.201703460

    35. [35]

      S. Tang, Y. Zhou, X. Lu, et al., J. Alloys Compds. 924 (2022) 166415. doi: 10.1016/j.jallcom.2022.166415

    36. [36]

      S. Liu, H. Ji, M. Wang, et al., ACS Appl. Mater. Interfaces 12 (2020) 15298–15304. doi: 10.1021/acsami.0c01940

    37. [37]

      Y. Qiu, L. Xin, Y. Li, et al., J. Am. Chem. Soc. 140 (2018) 16580–16588. doi: 10.1021/jacs.8b08356

    38. [38]

      Z. Yan, H. Sun, X. Chen, et al., Nat. Commun. 9 (2018) 2373. doi: 10.1038/s41467-018-04788-3

    39. [39]

      Y. Yuan, Q. Wang, Y. Qiao, et al., Adv. Energy Mater. 12 (2022) e202200970.

    40. [40]

      S.K. Geng, Y. Zheng, S.Q. Li, et al., Nat. Energy 6 (2021) 904–912. doi: 10.1038/s41560-021-00899-2

    41. [41]

      Y. Bai, Y. Wu, X. Zhou, et al., Nat. Commun. 13 (2022) 6094. doi: 10.1038/s41467-022-33846-0

    42. [42]

      S.S. Mishra, P. Kumbhakar, S. Nellaiappan, et al., Energy Technol. 11 (2022) e202200860.

    43. [43]

      X. Zhang, H. Yi, M. Jin, et al., Small 18 (2022) e2203710. doi: 10.1002/smll.202203710

    44. [44]

      Y.J. Wu, J. Yang, T.X. Tu, et al., Angew. Chem. Int. Ed. 60 (2021) 26829–26836. doi: 10.1002/anie.202112447

    45. [45]

      Y. Wang, M. Zheng, Y. Li, et al., Angew. Chem. Int. Ed. 61 (2022) e202115735. doi: 10.1002/anie.202115735

    46. [46]

      Y. Wang, M. Zheng, H. Sun, et al., Appl. Catal. B: Environ. 253 (2019) 11–20. doi: 10.1016/j.apcatb.2019.04.022

    47. [47]

      Z.Y. Wu, Y.Q. Lu, J.T. Li, et al., ACS Omega 6 (2021) 27335–27350. doi: 10.1021/acsomega.1c04226

    48. [48]

      J.X. Tang, L.P. Xiao, C. Xiao, et al., J. Mater. Chem. A 9 (2021) 11049–11055. doi: 10.1039/d1ta00663k

    49. [49]

      J.X. Tang, N. Tian, L.P. Xiao, et al., J. Mater. Chem. A 10 (2022) 10902–10908. doi: 10.1039/d2ta01011a

    50. [50]

      X.C. Liu, T. Wang, Z.M. Zhang, et al., J. Am. Chem. Soc. 144 (2022) 20895–20902. doi: 10.1021/jacs.2c09105

    51. [51]

      S. Liu, H. Zhang, H. Yu, et al., Small 19 (2023) e2300388. doi: 10.1002/smll.202300388

    52. [52]

      E. Marelli, J. Gazquez, E. Poghosyan, et al., Angew. Chem. Int. Ed. 60 (2021) 14609–14619. doi: 10.1002/anie.202103151

    53. [53]

      J. Wang, C. Cheng, B. Huang, et al., Nano Lett. 21 (2021) 980–987. doi: 10.1021/acs.nanolett.0c04004

    54. [54]

      Q. Wang, X. Huang, Z.L. Zhao, et al., J. Am. Chem. Soc. 142 (2020) 7425–7433. doi: 10.1021/jacs.9b12642

    55. [55]

      M. Xie, S. Tang, B. Zhang, et al., Mater. Horiz. 10 (2023) 407–431. doi: 10.1039/d2mh01213h

    56. [56]

      Y. Zhu, T.R. Kuo, Y.H. Li, et al., Energy Environ. Sci. 14 (2021) 1928–1958. doi: 10.1039/d0ee03903a

    57. [57]

      M. Wang, L. Arnadottir, Z.J. Xu, et al., Nano Micro Lett. 11 (2019) 47. doi: 10.1117/12.2524195

    58. [58]

      P.C. Chen, C. Chen, Y. Yang, et al., J. Am. Chem. Soc. 145 (2023) 10116–10125. doi: 10.1021/jacs.3c00467

    59. [59]

      X. Chu, J. Li, W. Qian, et al., Chem. Rec. 23 (2023) e202200222. doi: 10.1002/tcr.202200222

    60. [60]

      X. Chu, L. Wang, J. Li, et al., Chem. Rec. 23 (2023) e202300013. doi: 10.1002/tcr.202300013

    61. [61]

      Z. Wang, S. Xu, M. Li, et al., Chem. Commun. 59 (2023) 4511–4514. doi: 10.1039/d3cc00221g

    62. [62]

      Y. Sheng, Y. Guo, H. Yu, et al., Small 19 (2023) e2207305. doi: 10.1002/smll.202207305

    63. [63]

      C. Xu, C. Guo, J. Liu, et al., Small 19 (2023) e2207675. doi: 10.1002/smll.202207675

    64. [64]

      Y. He, Y. Jia, B. Yu, et al., Small 19 (2023) e2206478. doi: 10.1002/smll.202206478

    65. [65]

      L. Li, N. Li, J. Xia, et al., J. Mater. Chem. A 11 (2023) 2291–2301. doi: 10.1039/d2ta08808h

    66. [66]

      L. Li, N. Li, J.W. Xia, et al., Nano Res 16 (2023) 9416–9425. doi: 10.1007/s12274-023-5625-y

    67. [67]

      D. Deng, S. Wu, H. Li, et al., Small 19 (2023) e2205469. doi: 10.1002/smll.202205469

    68. [68]

      H. Lu, Y. Jiang, G. Xiao, et al., J. Colloid Interface Sci. 616 (2022) 539–547. doi: 10.1016/j.jcis.2022.02.106

    69. [69]

      L. Tao, M. Sun, Y. Zhou, et al., J. Am. Chem. Soc. 144 (2022) 10582–10590. doi: 10.1021/jacs.2c03544

    70. [70]

      W. Deeloed, T. Priamushko, J. Cizek, et al., ACS Appl. Mater. Interfaces 14 (2022) 23307–23321. doi: 10.1021/acsami.2c00254

    71. [71]

      S. Kim, S. Ji, H. Yang, et al., Appl. Catal. B: Environ. 310 (2022) 121361. doi: 10.1016/j.apcatb.2022.121361

    72. [72]

      Y. Zhu, J. Peng, X. Zhu, et al., Nano Lett. 21 (2021) 6625–6632. doi: 10.1021/acs.nanolett.1c02064

    73. [73]

      L. Cao, X. Liu, X. Shen, et al., Acc. Chem. Res. 55 (2022) 2594–2603. doi: 10.1021/acs.accounts.2c00212

    74. [74]

      J.C. Dong, M. Su, V. Briega-Martos, et al., J. Am. Chem. Soc. 142 (2020) 715–719. doi: 10.1021/jacs.9b12803

    75. [75]

      H. Xu, B. Huang, Y. Zhao, et al., Inorg. Chem. 61 (2022) 4533–4540. doi: 10.1021/acs.inorgchem.2c00296

    76. [76]

      X. Chu, K. Wang, W. Qian, et al., Coord. Chem. Rev. 477 (2023) 214952. doi: 10.1016/j.ccr.2022.214952

    77. [77]

      C. Zhan, L. Bu, H. Sun, et al., Angew. Chem. Int. Ed. 62 (2023) e202213783. doi: 10.1002/anie.202213783

    78. [78]

      V.H. Do, P. Prabhu, V. Jose, et al., Adv. Mater. 35 (2023) e2208860. doi: 10.1002/adma.202208860

    79. [79]

      M. Zhou, J. Liu, C. Ling, et al., Adv. Mater. 34 (2022) e2106115. doi: 10.1002/adma.202106115

    80. [80]

      C. Liu, Y. Shen, J. Zhang, et al., Adv. Energy Mater. 12 (2022) e202103505.

    81. [81]

      W. Chen, S. Luo, M. Sun, et al., Adv. Mater. 34 (2022) e2206276. doi: 10.1002/adma.202206276

    82. [82]

      G. Liu, W. Zhou, Y. Ji, et al., J. Am. Chem. Soc. 143 (2021) 11262–11270. doi: 10.1021/jacs.1c05856

    83. [83]

      Y. Zhang, X. Liu, T. Liu, et al., Adv. Mater. 34 (2022) e2202333. doi: 10.1002/adma.202202333

    84. [84]

      M. Li, Z. Zhao, W. Zhang, et al., Adv. Mater. 33 (2021) e2103762. doi: 10.1002/adma.202103762

    85. [85]

      H. Xu, J. Yuan, G. He, et al., Coord. Chem. Rev. 475 (2023) 214869. doi: 10.1016/j.ccr.2022.214869

    86. [86]

      H. Xu, Y. Zhao, Q. Wang, et al., Coord. Chem. Rev. 451 (2022) 214261. doi: 10.1016/j.ccr.2021.214261

    87. [87]

      H. Xu, Y. Zhao, G. He, et al., Int. J. Hydrogen Energy 47 (2022) 14257–14279. doi: 10.1016/j.ijhydene.2022.02.152

    88. [88]

      H. Xu, C. Wang, G. He, et al., Inorg. Chem. 61 (2022) 14224–14232. doi: 10.1021/acs.inorgchem.2c02666

    89. [89]

      H. Xu, C. Wang, G. He, et al., Dalton Trans. 52 (2023) 8466–8472. doi: 10.1039/d3dt01228j

    90. [90]

      H. Xu, C. Wang, B. Huang, et al., Inorg. Chem. Front. 10 (2023) 2067–2074. doi: 10.1039/d3qi00138e

    91. [91]

      C. Wang, D. Liu, K. Zhang, et al., ACS Appl. Mater. Interfaces 14 (2022) 38669–38676. doi: 10.1021/acsami.2c07792

    92. [92]

      D. Liu, H. Xu, C. Wang, et al., J. Mater. Chem. A 9 (2021) 24670–24676. doi: 10.1039/d1ta06438j

    93. [93]

      Q. Niu, M. Yang, D. Luan, et al., Angew. Chem. Int. Ed. 61 (2022) e202213049. doi: 10.1002/anie.202213049

    94. [94]

      X. Zheng, J. Yang, Z. Xu, et al., Angew. Chem. Int. Ed. 61 (2022) e202205946. doi: 10.1002/anie.202205946

    95. [95]

      X. Li, K. Zheng, J. Zhang, et al., ACS Omega 7 (2022) 12430–12441. doi: 10.1021/acsomega.2c01423

    96. [96]

      W. Zheng, M. Liu, L.Y.S. Lee, ACS Catal. 10 (2019) 81–92. doi: 10.3390/s20010081

    97. [97]

      H. Xu, J. Li, X. Chu, Chem. Rec. 23 (2023) e202200244. doi: 10.1002/tcr.202200244

    98. [98]

      S. Zhou, H. Jang, Q. Qin, et al., Angew. Chem. Int. Ed. (2022) e202212196.

    99. [99]

      Y. Liu, Y. Chen, Y. Tian, et al., Adv. Mater. 34 (2022) e2203615. doi: 10.1002/adma.202203615

    100. [100]

      J. Li, Y. Tan, M. Zhang, W. Gou, et al., ACS Energy Lett. 7 (2022) 1330–1337. doi: 10.1021/acsenergylett.1c02769

    101. [101]

      Y. Tan, Y. Zhu, X. Cao, et al., ACS Catal. 12 (2022) 11821–11829. doi: 10.1021/acscatal.2c02594

    102. [102]

      H. Xu, J. Li, X. Chu, Nanoscale Horiz. 8 (2023) 441–452. doi: 10.1039/d2nh00549b

    103. [103]

      X. Chu, Y. Liao, L. Wang, et al., Chin. Chem. Lett. 34 (2023) 108285. doi: 10.1016/j.cclet.2023.108285

    104. [104]

      M. Yuan, C. Wang, Y. Wang, et al., Nanoscale 13 (2021) 13042–13047. doi: 10.1039/d1nr02752b

    105. [105]

      J. Chen, C. Chen, M. Qin, et al., Nat. Commun. 13 (2022) 5382. doi: 10.1038/s41467-022-33007-3

    106. [106]

      M.J. Hulsey, V. Fung, et al., Angew. Chem. Int. Ed. 61 (2022) e202208237. doi: 10.1002/anie.202208237

    107. [107]

      X. Kong, J. Xiao, A. Chen, et al., J. Colloid Interface Sci. 608 (2022) 2973–2984. doi: 10.1016/j.jcis.2021.11.018

    108. [108]

      J. Li, J. Hu, M. Zhang, et al., Nat. Commun. 12 (2021) 3502. doi: 10.1039/d1qo00111f

    109. [109]

      Z.W. Wei, H.J. Wang, C. Zhang, et al., Angew. Chem. Int. Ed. 60 (2021) 16622–16627. doi: 10.1002/anie.202104856

    110. [110]

      Y. Wang, C. Li, Z. Fan, et al., Nano Lett. 20 (2020) 8074–8080. doi: 10.1021/acs.nanolett.0c03073

    111. [111]

      Y. Chen, Z. Fan, J. Wang, et al., J. Am. Chem. Soc. 142 (2020) 12760–12766. doi: 10.1021/jacs.0c04981

    112. [112]

      N. Zhang, F. Zheng, B. Huang, et al., Adv. Mater. 32 (2020) e1906477. doi: 10.1002/adma.201906477

    113. [113]

      H. Tao, X. Sun, S. Back, et al., Chem. Sci. 9 (2018) 483–487. doi: 10.1039/C7SC03018E

    114. [114]

      S. Yu, S. Louisia, P. Yang, JACS Au 2 (2022) 562–572. doi: 10.1021/jacsau.1c00562

    115. [115]

      Y. Yang, S. Louisia, S. Yu, et al., Nature 614 (2023) 262–269. doi: 10.1038/s41586-022-05540-0

  • Scheme 1  Scheme of the in situ technologies and their applications in energy conversion.

    Figure 1  (a-c) In situ TEM observation of Pt nanowire growth at different stages. Reproduced with permission [34]. Copyright 2017, Wiley-VCH.

    Figure 2  (a) Illustration of the fabrication of Bi(OH)3 and the reconstruction to pits-Bi nanosheets. (b) TEM image, (c, d) STEM image and (e) AC-STEM image of the pits-Bi nanosheets. (f) Ex situ XRD and (g) Raman spectra for the in situ reconstruction process. Reproduced with permission [39]. Copyright 2022, Wiley-VCH.

    Figure 3  In situ Raman spectroscopy tests for studying the transformation of (a) the pristine NiFe-LDH and (b) d-NiFe-LDH during OER. (c, d) The evolution of I528/I457 of NiFe-LDH and d-NiFe-LDH. Reproduced with permission [44]. Copyright 2021, Wiley-VCH.

    Figure 4  In situ ATR-FTIR spectra of C3H6 oxidation on (a) PdO/C and (b) Pd/C at different potentials in a C3H6-saturated 0.1 mol/L HClO4 solution. Different propene adsorbed configurations (c) on PdO below 0.80 V, (d) above 0.8 V, (e) on Pd below 0.80 V, and (f) above 0.80 V. Reproduced with permission [50]. Copyright 2021, American Chemical Society.

    Figure 5  (a) Operando XANES of Cu K-edge of the I-AgCu as a function of the reaction time at −1.0 V vs. RHE in CO2-saturated 0.1 mol/L KHCO3. (b) XANES pre-edges of I-AgCu magnified from the dashed box region in (a) and the comparison with Cu and Cu2O references (dashed lines). (c, d) Operando XANES and EXAFS spectra of I-AgCu for CO2RR. (e) The in situ evolution of intermixed and phase-separated AgCu catalysts in electrochemical CO2RR. Reproduced with permission [58]. Copyright 2023, American Chemical Society.

    Figure 6  In situ XAS tests. (a) Scheme of the electrochemical setup for the in situ XAS experiment. Pt L3-edge XANES spectra of (b) Ru-Pt3Co/C and (c) Pt3Co/C during the ORR. (d) Peak positions of Pt 4f7/2 XPS spectra for Ru-Pt3Co/C, Pt3Co/C, and commercial Pt/C. (e) The corresponding ORR mechanism of Ru-Pt3Co/C. Schematic illustration of the (f) SHINERS study of the ORR at Pt (311) surface and the (g) ORR mechanism. (a-e) Reproduced with permission [72]. Copyright 2021, American Chemical Society. (f, g) Reproduced with permission [74]. Copyright 2020, American Chemical Society.

    Figure 7  (a, b) HAADF-STEM images and the (c) elemental mapping images of the Pd8Sb3 HPs. (d) CV curves of the Pd8Sb3 HPs/C, Pd/C, and Pt/C for EOR. In situ ATR-SEIRAS spectra of (e) Pd8Sb3 HPs/C and (f) Pd/C during CV cycling. (g) Atomic-resolution HAADF-STEM image of the 22% YOx/MoOx-Pt nanowires. (h) Pt mass-normalized MOR, (i) comparison on the specific activities and mass activities of 22% YOx/MoOx-Pt nanowires and other references. (j) In situ FTIR spectra recorded from 0.15 V to 1.20 V and (k) magnified in situ FTIR spectra recorded from 0.15 V to 0.60 V versus RHE during MOR on 22% YOx/MoOx-Pt. (l, m) In situ FTIR spectra during MOR. (a-f) Reproduced with permission [83]. Copyright 2022, Wiley-VCH. (g-m) Reproduced with permission [84]. Copyright 2021, Wiley-VCH.

    Figure 8  (a) In situ UV–vis and (b) Raman devices. (c) In situ UV–vis spectroelectrochemical study between 0.925 V and 1.705 V. (d) In situ Raman spectroelectrochemical study between 0.925 V and 1.585 V. (e) Scheme for the transformation of ZIF-67 to α-Co(OH)2 and β-Co(OH)2. Reproduced with permission [96]. Copyright 2019, American Chemical Society.

    Figure 9  (a) Illustration of the synthesis procedures of the VO-rich Pt/TiO2 and VO-deficient Pt/TiO2 catalysts. (b) ESR spectra and (c) XPS profiles. (d) HER polarization curves and (e) corresponding mass activity at the overpotential of 100 mV. (f) DFT calculation of the free energy. (g) In situ Raman spectra of VO-rich Pt/TiO2 at −0.1 V vs. RHE with different electrolysis durations. (h) Magnified view of the Eg(1) mode. (i–k) In situ Raman spectra of VO-rich Pt/TiO2 (i), VO-deficient Pt/TiO2 (j), and TiO2 (k) at a potential of −0.2 V vs. RHE. Reproduced with permission [109]. Copyright 2021, Wiley-VCH.

    Figure 10  (a) Operando/in situ methods used to reveal the dynamic changes during electrocatalytic CO2RR. (b) Scheme of operando EC-STEM and 4D-STEM under CO2RR-relevant conditions. (c-m) EC-STEM images of the initial growth of Cu nanocatalysts after a single negative-direction LSV scan, from 0.4 V to 0 V. (n) Quantitative analysis of relative fraction of metallic Cu. Operando XANES (o) and corresponding quantitative analysis (p) of 18 nm nanoparticles at −0.8 V. Reproduced with permission [115]. Copyright 2023, Nature Publishing Group.

  • 加载中
计量
  • PDF下载量:  3
  • 文章访问数:  929
  • HTML全文浏览量:  30
文章相关
  • 发布日期:  2024-09-15
  • 收稿日期:  2023-05-31
  • 接受日期:  2023-10-23
  • 修回日期:  2023-10-18
  • 网络出版日期:  2023-10-24
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