Jiye
Feng‡
,
Danni
Shi‡
,
Fei
Wang
,
Yiming
Zou
,
Weicheng
Li
,
Wenbiao
Zhang
,
Huaijun
Lin
,
Yuying
Meng
and
Qingsheng
Gao
*
College of Chemistry and Materials Science, Jinan University, Guangzhou 510632, P. R. China. E-mail: tqsgao@jnu.edu.cn
First published on 2nd June 2025
The electrochemical reconstruction of metal–organic frameworks (MOFs) offers a promising approach for in situ fabrication of high-performance electrocatalysts. However, this innovation is often hindered by unpredictable structural transformations due to the complex thermodynamic and kinetic interplay of such multiple electrochemical and chemical processes. Herein, the reaction-atmosphere (Ar or CO2) guided reconstruction of Cu-based MOFs to Cu nanoparticles with mixed-valence surfaces/interfaces was investigated for the first time to unravel the kinetic contribution made by intermediate chemisorption. As shown, Cu-1,3,5-benzenetricarboxylate (HKUST-1) with frangible Cu–O4 nodes undergoes thermodynamically favored reduction quickly upon applying cathodic potentials, followed by varied surface changes kinetically governed by the intermediates of the hydrogen evolution reaction or CO2 reduction reaction (HER or CO2RR). Under an Ar atmosphere, the predominant HER increases the [OH−] in the microenvironment near the cathode and thereby boosts the re-oxidation of in situ formed Cu toward Cu/Cu2O interfaces. Conversely, the CO2RR facilitates the strong adsorption of *CO on Cu surfaces, effectively preserving Cu(0) species. Thanks to the rich Cu/Cu2O interfaces with a lowered energy barrier for *CO–*CO coupling during the subsequent CO2RR test, the electrocatalysts restructured under Ar afford the obviously improved CO2-to-C2H4 conversion as compared with their counterparts restructured under CO2. Such an atmosphere-controlled reconstruction strategy is further validated using CuBDC (BDC = 1,4-benzenedicarboxylate) with labile Cu–O4 nodes, while CuPz2 (Pz = pyrazole), with robust Cu–N4 coordination, remains stable, highlighting the framework-dependent nature. These findings establish atmosphere-controlled reconstruction of metastable MOFs as a powerful tool for rational electrocatalyst design.
Very recently, the electrochemical reconstruction of MOFs was introduced to produce active and selective catalysts on site, avoiding the time and energy-consuming preparation of catalysts and working electrodes.22–25 The in situ restructured electrocatalysts inherit the structural merits of MOFs, and more importantly generate accommodative surfaces with highly active and robust sites under controlled electrochemical conditions thanks to quasi-molecular-imprinting effects.21 Efforts have been devoted to boosting the electroreduction of CO2 to C2+ products by controlling the reconstruction of Cu-based MOFs.22,26–28 For instance, Wen et al. employed S doping to direct the in situ reconstruction of Cu-1,3,5-benzenetricarboxylate (HKUST-1) toward Cu/CuxSy interfaces, resulting in efficient ethylene production.22 Peng's group fabricated copper/polyamine hybrid composites by impregnating HKUST-1 with polyamines and then performing in situ electrochemical reconstruction, achieving favorable C2H4 production on polyaniline/HKUST-1 and CH4 production on polypyrrole/HKUST-1.27 However, these studies primarily focused on precursor modification, overlooking the thermodynamic and kinetic effects of reconstruction conditions (such as electrolytes and atmospheres), thereby limiting the general applicability of these strategies. It is noteworthy that HER invariably accompanies CO2RR in aqueous electrolytes, where their interplay critically influences the dynamic evolution of the Cu surface, although the underlying mechanisms remain elusive. Buonsanti and co-workers pointed out that the adsorption of either *H or *CO (* denotes an active-site) eventually alters the thermodynamic order of Cu facets, leading to potential-dependent nanoclustering of Cu electrocatalysts.29 By contrast, Lee et al. proved that continuous reconstruction from polycrystalline Cu to Cu(100) took place only in the presence of CO2, suggesting an aggravated reconstruction by the CO intermediate of CO2RR.30 Therefore, unravelling the directional functionalities of HER and CO2RR is necessary to consolidate the thermodynamic and kinetic fundamentals of electrochemical Cu-MOFs reconstruction,31,32 which is however a research gap to the best of our knowledge.
To decouple the above effects, the reaction-atmosphere (Ar or CO2) guided reconstruction of Cu-based MOFs was investigated herein for the first time. A series of characterization studies, including ex situ X-ray diffraction (XRD) and in situ Raman, showed that HKUST-1 consisting of frangible Cu–O4 nodes underwent thermodynamically favored reduction at cathodic potentials, followed by surface reconfigurations kinetically governed by HER or CO2RR intermediates (Scheme 1). Under an Ar flow, the predominant HER led to excessive OH− in the microenvironment of the working cathode, thereby facilitating the easy transformation to Cu–OH or Cu–Oad.33 Such enhanced re-oxidation of the in situ formed Cu surface ultimately produced rich Cu/Cu2O interfaces, which benefited the CO2-to-C2H4 conversion in the subsequent CO2RR thanks to the lowered energy barrier for *CO–*CO coupling. In sharp contrast, a CO2 atmosphere resulted in the strong adsorption of in situ formed CO from CO2RR on the metallic Cu surface, serving as protection to maintain Cu(0) species that however favors CO generation rather than coupling.34 Furthermore, similar reconstruction in a controlled manner and the correspondingly altered CO2RR were identified on another Cu-MOF with fragile Cu–O4 nodes, i.e., CuBDC (BDC = 1,4-benzenedicarboxylate), but negligible for CuPz2 (Pz = pyrazole) with much stronger Cu–N4. This revealed the coordination-dependent dynamic evolution of MOFs, which can be further modulated to design efficient catalysts simply by varying the atmospheric conditions.
E(vs. RHE) = E(vs. Ag/AgCl) + 0.059 × pH + 0.197 V |
The liquid products were analyzed using 1H nuclear magnetic resonance (NMR) spectroscopy. A 100 ppm DMSO solution was employed as the internal standard. For analysis, 800 μL of the electrolyte was mixed with 100 μL of D2O and 100 μL of the internal standard, and the product concentrations were calculated based on a standard calibration curve.
The binding energy (BE) of an adsorbate was calculated as:
BE(adsorbate) = E(slab+adsorbate) − E(slab) − E(adsorbate) |
The Gibbs free energy (G) of a species was calculated as:
G = E + ZPE − TS |
ΔG = ΔE + ΔZPE − TΔS |
The transition state search was performed based on the complete LST/QST method, with an RMS convergence criterion of 0.25 eV Å−1, utilizing the optimized geometric configurations of both reactants and products as initial structures. CASTEP employed algorithms including the nudged elastic band (NEB) method and the dimer approach to explore the potential energy surface and determine the transition state configuration. The transition state configuration was accurately resolved between the reactant and product geometries by implementing a 10-step QST energy minimization protocol.
![]() | ||
Fig. 1 (a) Experimental and simulated XRD patterns, and (b) SEM image of pristine HKUST-1. SEM images of HKUST-1 reconstructed under (c) Ar and (d) CO2 for 10, 30, 45 and 60 min. |
The reconfigured HKUST-CO2 and HKUST-Ar were characterized through a series of physical measurements to understand their structural and compositional differences. FT-IR analysis confirmed the collapse of the HKUST-1 framework following reconstruction, as evidenced by the disappearance of characteristic adsorption bands associated with the carboxylate ligand (Fig. S4, ESI†). As shown in the XRD patterns (Fig. 2a), the diffraction peaks of HKUST-1 disappeared, and new ones corresponding to Cu(111) and Cu(200) emerged at 2θ = 43.3° and 50.4°, respectively, confirming the complete transformation of the Cu–O4 framework into dominant metallic Cu. This could be ascribed to the thermodynamically favored reduction at the highly negative potential of −1.3 V (vs. RHE) compared to the Cu(OH)2/Cu redox potential of 0.61 V (vs. RHE). Different from HKUST-CO2, HKUST-Ar presented two additional peaks associated with Cu2O(111) and Cu2O(200), indicating the presence of Cu/Cu2O interfaces. XPS and AES were further utilized to investigate the chemical composition and elemental valence states. In the Cu 2p XPS profile (Fig. 2b), the absence of Cu(II) peaks and the presence of a peak related to Cu(I)/Cu(0) at 932.4 eV indicated the reduction of Cu(II) in HKUST-1 frameworks during electrolysis in Ar or CO2 atmospheres.22 In the AES profiles, HKUST-Ar displayed peaks at 914.3 and 917.3 eV, corresponding to Cu(I) and Cu(0),35 respectively, whereas HKUST-CO2 exhibited only the Cu(0) signal (Fig. 2c), consistent with the XRD results. High-resolution TEM (HR-TEM) images provided further insights into the structural configurations. HKUST-CO2 exclusively exhibited lattice fringes corresponding to Cu(111) (Fig. 2d), whereas HKUST-Ar distinctly revealed Cu/Cu2O interfaces, with clearly resolved lattice spacings of 0.204 nm and 0.245 nm assigned to Cu(111) and Cu2O(111), respectively (Fig. 2e). These comprehensive analyses confirm that under an Ar atmosphere, the Cu/Cu2O interface is restructured, while only metallic Cu is achieved in the presence of CO2.
![]() | ||
Fig. 2 (a) XRD patterns, (b) high-resolution Cu 2p3/2 XPS profiles, (c) Cu LMM AES profiles and (d and e) HR-TEM images of the restructured HKUST-CO2 and HKUST-Ar. |
Further attention was given to how the valence states of the restructured samples varied under different atmospheric conditions. At first, the changes in phase structures during reconstruction under CO2 or Ar atmospheres were analyzed by using ex situ XRD (Fig. 3a and b). Only the diffraction peaks of Cu appeared in the presence of CO2 over time. By contrast, the peaks of Cu2O gradually increased alongside a decrease in Cu peaks under Ar, suggesting the re-oxidation of the in situ formed metallic Cu. This evolution was then further examined by in situ Raman spectroscopy (Fig. 3c and d). At a working potential of −1.3 V vs. RHE, the absorption bands corresponding to Cu(II)–O and C–H disappeared rapidly, indicating the collapse of the metal–organic framework.36 Notably, Cu–CO peaks appeared under a CO2 atmosphere (Fig. 3c), attributed to the strong adsorption of *CO on the in situ formed Cu(0) surface.37 In sharp comparison, new bands at 520, 700, and 600 cm−1 emerged within only 5 minutes under Ar (Fig. 3d), corresponding to the characteristic vibrations of Cu2O, Cu–OH, and Cu–Oad, respectively,38,39 which gradually increased in intensity during electrolysis. The adsorbed Cu–Oad species are re-oxidized in the electrolyte, generating CuOx species,40 and thus the signals of Cu2O can be further observed at different reduction potentials by XRD (Fig. S5, ESI†) and Raman (Fig. S6, ESI†). This indicated that the re-oxidation process can prevent the reduction of Cu2O at highly negative potentials.33,41 Moreover, the restructured electrodes, HKUST-Ar and HKUST-CO2, were subjected for cyclic voltammetry (CV) analysis to identify the adsorption of *CO and *OH.42 After reconstruction under Ar for 1 h, the CV profile of HKUST-Ar exhibited visible adsorption/desorption peaks of OH− in 1.0 M KOH (Fig. 3e), which were subsequently weakened when the carrier gas was switched to CO, along with the emergence of a Cu–CO couple. As indicated, the restructured Cu preferentially adsorbs *CO due to their strong interactions. Concertedly, only the Cu–CO peaks could be observed on HKUST-CO2 in 1.0 M KOH initially (Fig. 3f), associated with the predominant *CO adsorption on Cu during HKUST-1 reconstruction. These peaks declined after switching the gas to Ar followed by 30 cycles of CV, and those of Cu–OH reappeared again.
There was still a doubt that whether variation in local [OH−] near electrodes would influence the valence states of the restructured surface. At a fixed potential of −1.3 V (vs. RHE) adopted for electrochemical reconstruction, the j under a CO2 atmosphere was approximately −200 mA cm−2 (Fig. S7, ESI†), lower than that observed under Ar (−300 mA cm−2). Given the synchronous proton consumption along with electron transfer in both the HER and CO2RR,20 we can infer that the local [OH−] formed in situ varies at different j values. To this end, the HKUST-1 pre-catalyst was restructured in a CO2 atmosphere at j = −200 and −300 mA cm−2, but the received HKUST-CO2 samples showed only the diffraction peaks of Cu in XRD (Fig. S8a, ESI†). Analogously, Cu2O was found in HKUST-Ar under an Ar atmosphere, regardless of the j value (−200 or −300 mA cm−2) used during reconstruction (Fig. S8b, ESI†). These results rule out the influence of local [OH−] variations caused by different rates of HER and CO2RR.
According to the above results, it's reasonable to propose that Cu(II) in HKUST-1 pre-catalysts undergoes rapid reduction toward metallic Cu(0), which is then re-oxidized to Cu–OH, Cu–Oad and even Cu2O under an Ar atmosphere because the predominant HER produces excess OH− at the local surface of the electrode and boosts the thermodynamic of re-oxidation according to the Nerst equation.33,43 By contrast, Cu(0) can be stabilized by the *CO intermediate of the CO2RR in a CO2 flow, preventing re-oxidation even in a microenvironment rich in OH−.
The varied valence states of the restructured electrocatalysts would lead to differences in CO2RR performance.44 In particular, mixed-valent Cu species, e.g., Cu(I)–Cu(0) ensembles, have demonstrated high efficiency for C2+ production.45,46 We afterward evaluated the CO2RR performance of HKUST-Ar and HKUST-CO2 in a flow cell with 1.0 M KOH as the electrolyte (Fig. S9, ESI†). The gaseous and liquid products were analyzed by GC (Fig. S10, ESI†) and 1H NMR (Fig. S11, ESI†), respectively. The FE of liquid products was less than 10% based on calculations derived from the standard curves (Fig. S12, ESI and Table S1, ESI†), so we only discussed gaseous products here, and focused on ethylene. Linear sweep voltammetry (LSV) curves revealed that the current density for HKUST-Ar was higher than that for HKUST-CO2 (Fig. 4a), indicating the enhanced activity of HKUST-Ar. Potentiostatic tests were conducted at sequentially decreasing potential from −1.1 to −1.5 V (vs. RHE), and the corresponding product FEs and current densities are depicted in Fig. 4b and c. As shown, the FE for C2H4 production through HKUST-Ar was substantially higher than that with HKUST-CO2. Notably, HKUST-Ar achieved a maximum FE of C2H4 of 47.8% at −1.2 V vs. RHE, compared to only 13.4% on HKUST-CO2 (Fig. S13, ESI†).
The electrochemically active surface areas of the catalysts were measured by deriving the double-layer capacitance (Cdl) through cyclic voltammetry (Fig. S14, ESI†). The results indicated that HKUST-Ar exhibited a Cdl value of 3.26 mF cm−2, surpassing the 2.04 mF cm−2 observed for HKUST-CO2, which suggested a higher number of active sites in HKUST-Ar. Furthermore, electrochemical impedance spectroscopy (EIS) was used to analyze electron transport dynamics during the CO2RR (Fig. S15, ESI†). The Nyquist plot revealed that HKUST-Ar has a smaller radius, indicative of lower electron transfer resistance (Rct). Meanwhile, HKUST-Ar exhibited a significantly lower Tafel slope compared to HKUST-CO2, indicating faster kinetics and better catalytic performance for CO2 electroreduction (Fig. S16, ESI†). This evidence suggests that in situ electrochemical reconstruction under Ar gas can effectively enhance electron transfer rates and accelerate reaction kinetics, thereby improving catalytic activity for CO2RR.
The long-term stability was evaluated by conducting a chronoamperometric test. The HKUST-Ar sample presented a significantly higher C2H4 FE of approximately 45% (Fig. 4d), which remained above 40% over 300 minutes of reaction. By contrast, the main products on HKUST-CO2 were H2 and CO with the FEs of ∼40% and ∼44%, respectively, and the FE of C2H4 was below 17% (Fig. 4e). Furthermore, we characterized the two catalysts after the long-term CO2RR test. There was no obvious change in the phase or valence on HKUST-CO2 (Fig. S17, ESI†). In contrast, the Cu2O(111) and (200) peaks of HKUST-Ar gradually weakened (Fig. S18a, ESI†), and the AES results confirmed a decrease in the Cu(I) signal (Fig. S18b, ESI†). Meanwhile, TEM images of both materials revealed only the presence of Cu(111), indicating the stable existence of Cu after prolonged reaction (Fig. S19, ESI†). By employing pulse electrolysis methods to stabilize the Cu(I) species, the long-term durability will be further improved.47
According to the abovementioned structural characterization, HKUST-CO2 predominantly consisted of metallic Cu, while HKUST-Ar featured multi-phase Cu/Cu2O interfaces. To gain more insights into the enhancement of C2H4 production, density functional theory (DFT) calculations were performed on the Cu and Cu/Cu2O structural models (Fig. S20, ESI†), which served as the representatives of HKUST-CO2 and HKUST-Ar, respectively. As a critical step toward C2+ products, the coupling of two *CO intermediates was brought into focus. The corresponding reaction models are illustrated in Fig. S21, ESI,† with computational data summarized in Table S2, ESI.† As shown in Fig. 4f, the kinetic barrier (0.36 eV) and free-energy change (−0.14 eV) for *CO coupling on Cu/Cu2O are lower than those on Cu (0.72 and 0.16 eV), suggesting that the Cu(0)–Cu(I) interfaces can effectively boost C2 production.48,49 This conclusion is consistent with previous studies.20,50 Additionally, the CO adsorption energies on both Cu and Cu/Cu2O were quite similar (Fig. S22, ESI†), further suggesting that the key contribution lies in reducing the *CO–*CO coupling energy barrier.
To demonstrate the universality of the proposed strategy, we substituted Ar gas with He for the reconstruction. The XRD analysis revealed the generation of Cu2O (Fig. S23, ESI†), and the SEM image showed a similar morphology to that of HKUST-Ar (Fig. S24, ESI†). The resulting CO2RR performance was essentially consistent (Fig. S25, ESI†). This finding suggests that the strategy is applicable across different inert gases. Subsequently, we expanded our study to include other Cu-based MOFs, first focusing on CuBDC (BDC = 1,4-benzenedicarboxylate), with a similar Cu–O4 coordination but a layered crystalline structure. The XRD pattern of the obtained CuBDC closely matched the simulated structure, and SEM showed that CuBDC exhibited a flake-like morphology (Fig. 5a). Then CuBDC was subjected to reconstruction under Ar or CO2 for 1 hour, followed by XRD analysis. The results mirrored those of HKUST-1, with Cu2O diffraction peaks observed under Ar gas (Fig. 5b). Furthermore, CuBDC-Ar significantly outperformed CuBDC-CO2 for the CO2RR to C2H4 production (Fig. 5d), indicative of the universality for Cu-MOFs with Cu–O4 coordination.
Next, we investigated CuPz2 (Pz = pyrazole), which possesses a Cu–N4 structure. The XRD results demonstrated that the MOF structure remained intact, and SEM observations corroborated this by showing an unchanged morphology (Fig. 5c). The applied potential was incrementally elevated to −1.5 V and −2.0 V (vs. RHE), with XRD analysis revealing no substantial alterations in the characteristic diffraction peaks corresponding to CuPz2 (Fig. S26, ESI†). This stability is likely due to the strong coordination of Cu–N4, making CuPz2 resistant to reconstruction.51 Subsequent performance testing revealed no significant differences compared to the measurement without reconstruction (Fig. 5e), further substantiating that our strategy is effective only in facilitating the reconstruction process for specific structural configurations and is not applicable to MOFs with inherently stable structures. The frangible Cu–O4 nodes are a prerequisite for reconstruction, while the strong Cu–N4 coordination requires harsher conditions with corrosive ions (e.g., halogen) to drive reconfiguration.51,52
Footnotes |
† Electronic supplementary information (ESI) available: Additional figures and data for CO2RR performance details, and additional Raman, XRD, IR, SEM, CV, and NMR spectra. See DOI: https://doi.org/10.1039/d5sc02601f |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |