X-ray spectroscopy for functional materials\(\def\hfill{\hskip 5em}\def\hfil{\hskip 3em}\def\eqno#1{\hfil {#1}}\)

Journal logoJOURNAL OF
SYNCHROTRON
RADIATION
ISSN: 1600-5775

Revealing the structure of the active sites for the electrocatalytic CO2 reduction to CO over Co single atom catalysts using operando XANES and machine learning

crossmark logo

aDepartment of Interface Science, Fritz Haber Institute of the Max Planck Society, 14195 Berlin, Germany
*Correspondence e-mail: martini@fhi-berlin.mpg.de, janis@fhi-berlin.mpg.de

Edited by K. Kvashnina, ESRF – The European Synchrotron, France (Received 8 March 2024; accepted 21 May 2024; online 25 June 2024)

Dedicated to the memory of Professor Carlo Lamberti.

Transition-metal nitro­gen-doped carbons (TM-N-C) are emerging as a highly promising catalyst class for several important electrocatalytic processes, including the electrocatalytic CO2 reduction reaction (CO2RR). The unique local environment around the singly dispersed metal site in TM-N-C catalysts is likely to be responsible for their catalytic properties, which differ significantly from those of bulk or nanostructured catalysts. However, the identification of the actual working structure of the main active units in TM-N-C remains a challenging task due to the fluctional, dynamic nature of these catalysts, and scarcity of experimental techniques that could probe the structure of these materials under realistic working conditions. This issue is addressed in this work and the local atomistic and electronic structure of the metal site in a Co–N–C catalyst for CO2RR is investigated by employing time-resolved operando X-ray absorption spectroscopy (XAS) combined with advanced data analysis techniques. This multi-step approach, based on principal component analysis, spectral decomposition and supervised machine learning methods, allows the contributions of several co-existing species in the working Co–N–C catalysts to be decoupled, and their XAS spectra deciphered, paving the way for understanding the CO2RR mechanisms in the Co–N–C catalysts, and further optimization of this class of electrocatalytic systems.

1. Introduction

Transition-metal nitro­gen-doped carbon (TM-N-C) catalysts are currently attracting great interest because of their promising performance in several electrocatalytic processes, such as oxygen reduction reactions (Zitolo et al., 2015[Zitolo, A., Goellner, V., Armel, V., Sougrati, M. T., Mineva, T., Stievano, L., Fonda, E. & Jaouen, F. (2015). Nat. Mater. 14, 937-942.], 2017[Zitolo, A., Ranjbar-Sahraie, N., Mineva, T., Li, J., Jia, Q., Stamatin, S., Harrington, G. F., Lyth, S. M., Krtil, P., Mukerjee, S., Fonda, E. & Jaouen, F. (2017). Nat. Commun. 8, 957.]), nitrate reduction (Wu et al., 2021[Wu, Z., Karamad, M., Yong, X., Huang, Q., Cullen, D. A., Zhu, P., Xia, C., Xiao, Q., Shakouri, M., Chen, F., Kim, J. Y., Xia, Y., Heck, K., Hu, Y., Wong, M. S., Li, Q., Gates, I., Siahrostami, S. & Wang, H. (2021). Nat. Commun. 12, 2870.]; Wang et al., 2022[Wang, S., Gao, H., Li, L., Hui, K. S., Dinh, D. A., Wu, S., Kumar, S., Chen, F., Shao, Z. & Hui, K. N. (2022). Nano Energy, 100, 107517.]) and electrocatalytic CO2 reduction (CO2RR) (Liang et al., 2021[Liang, S., Huang, L., Gao, Y., Wang, Q. & Liu, B. (2021). Adv. Sci. 8, 2102886.]; Genovese et al., 2018[Genovese, C., Schuster, M. E., Gibson, E. K., Gianolio, D., Posligua, V., Grau-Crespo, R., Cibin, G., Wells, P. P., Garai, D., Solokha, V., Krick Calderon, S., Velasco-Velez, J. J., Ampelli, C., Perathoner, S., Held, G., Centi, G. & Arrigo, R. (2018). Nat. Commun. 9, 935.]; Li et al., 2022[Li, K., Zhang, S. B., Zhang, X. L., Liu, S., Jiang, H. S., Jiang, T. L., Shen, C. Y., Yu, Y. & Chen, W. (2022). Nano Lett. 22, 1557-1565.]). The interest is based not only on the excellent catalytic properties of TM-N-C materials but also on their potential to decrease our dependence on precious metal resources and optimize the utilization of metal atoms (Hursán et al., 2024[Hursán, D., Timoshenko, J., Ortega, E., Jeon, H. S., Rüscher, M., Herzog, A., Rettenmaier, C., Chee, S. W., Martini, A., Koshy, D. & Roldán Cuenya, B. (2024). Adv. Mater. 36, 2307809.]). It is commonly believed that in the as-prepared TM-N-C catalysts the metal species are anchored to the carbon matrix by four N atoms, although some anchoring structures with lower coordination are reported in the literature as well (Fan et al., 2020[Fan, Q., Hou, P. F., Choi, C. H., Wu, T. S., Hong, S., Li, F., Soo, Y. L., Kang, P., Jung, Y. S. & Sun, Z. Y. (2020). Adv. Energy Mater. 10, 1903068.]; Yang et al., 2018[Yang, J., Qiu, Z. Y., Zhao, C. M., Wei, W. C., Chen, W. X., Li, Z. J., Qu, Y. T., Dong, J. C., Luo, J., Li, Z. Y. & Wu, Y. (2018). Angew. Chem. Int. Ed. 57, 14095-14100.]; Wang et al., 2018[Wang, X. Q., Chen, Z., Zhao, X. Y., Yao, T., Chen, W. X., You, R., Zhao, C. M., Wu, G., Wang, J., Huang, W. X., Yang, J. L., Hong, X., Wei, S. Q., Wu, Y. & Li, Y. D. (2018). Angew. Chem. Int. Ed. 57, 1944-1948.]; Gong et al., 2020[Gong, Y. N., Jiao, L., Qian, Y. Y., Pan, C. Y., Zheng, L. R., Cai, X. C., Liu, B., Yu, S. H. & Jiang, H. L. (2020). Angew. Chem. Int. Ed. 59, 2705-2709.]). In the context of CO2RR, Ni–N–C is the most popular choice for the CO2 reduction to CO, showing high selectivity [Faradaic efficiency FECO > 90% (Hursán et al., 2024[Hursán, D., Timoshenko, J., Ortega, E., Jeon, H. S., Rüscher, M., Herzog, A., Rettenmaier, C., Chee, S. W., Martini, A., Koshy, D. & Roldán Cuenya, B. (2024). Adv. Mater. 36, 2307809.])] and a large activity (current density) for this process (Liang et al., 2021[Liang, S., Huang, L., Gao, Y., Wang, Q. & Liu, B. (2021). Adv. Sci. 8, 2102886.]). Interestingly, while there is consensus in the literature regarding the origin of the catalytic performance of the Ni–N–C system, the same does not apply to the case of the Co–N–C catalyst, where the observed selectivity has been shown to vary strongly among different studies, with FECO values between 20% and 100% (Ju et al., 2017[Ju, W., Bagger, A., Hao, G. P., Varela, A. S., Sinev, I., Bon, V., Roldan Cuenya, B., Kaskel, S., Rossmeisl, J. & Strasser, P. (2017). Nat. Commun. 8, 944.]; Pan, Deng et al., 2018[Pan, F. P., Deng, W., Justiniano, C. & Li, Y. (2018). Appl. Catal. Environ. 226, 463-472.]; Wang et al., 2018[Wang, X. Q., Chen, Z., Zhao, X. Y., Yao, T., Chen, W. X., You, R., Zhao, C. M., Wu, G., Wang, J., Huang, W. X., Yang, J. L., Hong, X., Wei, S. Q., Wu, Y. & Li, Y. D. (2018). Angew. Chem. Int. Ed. 57, 1944-1948.]; Pan, Lin et al., 2018[Pan, Y., Lin, R., Chen, Y. J., Liu, S. J., Zhu, W., Cao, X., Chen, W. X., Wu, K. L., Cheong, W. C., Wang, Y., Zheng, L. R., Luo, J., Lin, Y., Liu, Y. Q., Liu, C. G., Li, J., Lu, Q., Chen, X., Wang, D. S., Peng, Q., Chen, C. & Li, Y. D. (2018). J. Am. Chem. Soc. 140, 4218-4221.]). On one hand, this suggests that the Co–N–C catalysts can be potentially optimized to reach the same high performance for CO2RR as Ni–N–C, or even outperform the latter. On the other hand, the large discrepancy between the results obtained in different studies suggests that the dynamic processes taking place in the working Co–N–C catalysts can be more complex than those in the Ni–N–C system, and need to be understood in detail before the Co–N–C can be used as a practical catalyst.

Operando X-ray absorption spectroscopy (XAS) measurements are a suitable tool for this purpose, providing valuable insights into the oxidation state and the local electronic and atomistic structure of the Co sites under applied potential. In particular, the analysis of the extended X-ray absorption fine-structure (EXAFS) region of the XAS spectrum has been proven to be an invaluable tool for understanding the structure of working electrocatalysts. (Timoshenko & Roldan Cuenya, 2021[Timoshenko, J. & Roldan Cuenya, B. (2021). Chem. Rev. 121, 882-961.]). However, for such disordered structures as TM-N-C catalysts, the information content in the EXAFS data is limited. Moreover, the EXAFS data quality of TM-N-C catalysts is often compromised by the low metal loading, by the signal attenuation due to the electrolyte and by the formation of gas bubbles.

An alternative approach consists of focusing instead on the X-ray absorption near-edge structure (XANES) part of the XAS spectrum. XANES usually has a higher signal-to-noise ratio and is highly sensitive to the electronic structure of the absorber, but also to the details of the local 3D geometry. However, the quantitative analysis of XANES features is challenging (Guda et al., 2021[Guda, A. A., Guda, S. A., Martini, A., Kravtsova, A. N., Algasov, A., Bugaev, A., Kubrin, S. P., Guda, L. V., Šot, P., van Bokhoven, J. A., Copéret, C. & Soldatov, A. V. (2021). NPJ Comput. Mater. 7, 203.]). In contrast to the EXAFS case, there is no analytical equation describing the XANES part (Guda et al., 2019[Guda, A. A., Guda, S. A., Lomachenko, K. A., Soldatov, M. A., Pankin, I. A., Soldatov, A. V., Braglia, L., Bugaev, A. L., Martini, A., Signorile, M., Groppo, E., Piovano, A., Borfecchia, E. & Lamberti, C. (2019). Catal. Today, 336, 3-21.]). Nonetheless, the development of ab initio codes for XANES simulations and of machine learning methods allows this issue to be addressed. For instance, Zitolo et al. (2015[Zitolo, A., Goellner, V., Armel, V., Sougrati, M. T., Mineva, T., Stievano, L., Fonda, E. & Jaouen, F. (2015). Nat. Mater. 14, 937-942.], 2017[Zitolo, A., Ranjbar-Sahraie, N., Mineva, T., Li, J., Jia, Q., Stamatin, S., Harrington, G. F., Lyth, S. M., Krtil, P., Mukerjee, S., Fonda, E. & Jaouen, F. (2017). Nat. Commun. 8, 957.]) utilized the MXAN code (Benfatto et al., 2021[Benfatto, M., Della Longa, S., Pace, E., Chillemi, G., Padrin, C., Natoli, C. R. & Sanna, N. (2021). Comput. Phys. Commun. 265, 107992.]) to fit the XANES spectra of Co–N–C and Fe–N–C compounds used for the oxygen reduction reaction. Similarly, Xiang et al. (2022[Xiang, S. T., Huang, P. P., Li, J. Y., Liu, Y., Marcella, N., Routh, P. K., Li, G. H. & Frenkel, A. I. (2022). Phys. Chem. Chem. Phys. 24, 5116-5124.]) employed a neural network approach to decipher the XANES spectrum of a Co–N–C catalyst for the photoreduction reaction of CO2. Crucially, these approaches work only for phase-pure catalysts, and are not applicable for tracking the evolution of working catalysts, where the co­existence of different species or states of the metal site is expected. In our recent work on the characterization of Ni–N–C catalysts (Martini et al., 2023[Martini, A., Hursán, D., Timoshenko, J., Rüscher, M., Haase, F., Rettenmaier, C., Ortega, E., Etxebarria, A. & Roldan Cuenya, B. (2023). J. Am. Chem. Soc. 145, 17351-17366.]) we addressed this problem through a multi-step approach. We first identified the number of different coexisting metal species, their corresponding kinetic profiles and XANES spectra using unsupervised machine learning methodologies (Martini & Borfecchia, 2020[Martini, A. & Borfecchia, E. (2020). Crystals, 10, 664.]; Timoshenko & Frenkel, 2019[Timoshenko, J. & Frenkel, A. I. (2019). ACS Catal. 9, 10192-10211.]). Afterwards, we deduced the atomistic structures for each of the identified species through a XANES fitting procedure facilitated by a supervised machine learning approach (Martini, Guda et al., 2020[Martini, A., Guda, S. A., Guda, A. A., Smolentsev, G., Algasov, A., Usoltsev, O., Soldatov, M. A., Bugaev, A., Rusalev, Y., Lamberti, C. & Soldatov, A. V. (2020). Comput. Phys. Commun. 250, 107064.]; Guda et al., 2019[Guda, A. A., Guda, S. A., Lomachenko, K. A., Soldatov, M. A., Pankin, I. A., Soldatov, A. V., Braglia, L., Bugaev, A. L., Martini, A., Signorile, M., Groppo, E., Piovano, A., Borfecchia, E. & Lamberti, C. (2019). Catal. Today, 336, 3-21.]). In this study, we will extend the methodology developed for the Ni–N–C catalyst to explore the structures and behaviour of Co–N–C catalysts. Specifically, we aim to elucidate the evolution of cobalt centres during the CO2RR and understand their electronic and structural characteristics. By combining principal component analysis with spectral decomposition techniques we identified the distinct contributions from coexisting cobalt species to the averaged XANES spectra. The obtained spectra for pure species are interpreted using XANES simulations, including those carried out within the time-dependent density functional theory (DFT) framework, XANES data fitting and EXAFS analysis. Taken together, these techniques ultimately confirm the interactions between singly dispersed Co sites and the CO adsorbates, demonstrating the key role of these species for the electrocatalytic conversion of CO2 to CO. Nonetheless, our analysis suggests that these singly dispersed Co sites coexist in the working catalysts with partially reduced agglomerates of Co atoms, which limit the efficiency of Co–N–C catalysts for CO2RR.

2. Experimental and methods

2.1. Sample preparation

The Co–N–C catalysts were synthesized using an impregnation-calcination method (Hursán et al., 2024[Hursán, D., Timoshenko, J., Ortega, E., Jeon, H. S., Rüscher, M., Herzog, A., Rettenmaier, C., Chee, S. W., Martini, A., Koshy, D. & Roldán Cuenya, B. (2024). Adv. Mater. 36, 2307809.]) starting from a zeolitic imidazolate framework (ZIF-8) precursor. ZIF-8 crystals underwent carbonization in an Ar atmosphere at 1000°C for 1 h. At this temperature the majority of metallic Zn evaporated, resulting in a porous (Zn)-N-doped carbon structure, referred to as N–C. To eliminate all crystalline Zn species from the N–C support, a room temperature acid wash was performed with 20 wt% nitric acid (HNO3, ≥65%, Carl Roth) over 24 h. The treated sample underwent thorough washing and vacuum filtration with ultrapure water (at least 3 × 600 ml) until the pH of the supernatant solution exceeded 5.

For the synthesis of Co–N–C, 200 mg of N–C was dispersed in a 20 ml solution containing 6 mM of Co-nitrate [Co(NO3)2·6H2O, Sigma-Aldrich, ≥98%). The suspension underwent sonication for 2 h in an ultrasonic bath at ∼30–40°C and stirring for an additional 2 h at room temperature. The Co-impregnated N–C was collected through centrifugation, dried at 60°C in air, and subsequently heat-treated in Ar atmosphere at 700°C for 1 h. The final Co–N–C catalysts were obtained after an additional washing with ultrapure water and drying.

The corresponding potential-dependent CO2RR selectivity trends retrieved for this sample are reported in our previous work (Hursán et al., 2024[Hursán, D., Timoshenko, J., Ortega, E., Jeon, H. S., Rüscher, M., Herzog, A., Rettenmaier, C., Chee, S. W., Martini, A., Koshy, D. & Roldán Cuenya, B. (2024). Adv. Mater. 36, 2307809.]).

2.2. Experimental setup and X-ray absorption measurements

Operando Co K-edge XAS measurements were performed in our home-built single-compartment electrochemical cell (Timoshenko & Roldan Cuenya, 2021[Timoshenko, J. & Roldan Cuenya, B. (2021). Chem. Rev. 121, 882-961.]; Martini et al., 2023[Martini, A., Hursán, D., Timoshenko, J., Rüscher, M., Haase, F., Rettenmaier, C., Ortega, E., Etxebarria, A. & Roldan Cuenya, B. (2023). J. Am. Chem. Soc. 145, 17351-17366.]) in fluorescence mode. The Co–N–C sample was spray-coated over a carbon paper acting as both a working electrode and a window for the X-ray fluorescence and the incoming X-rays. A Pt mesh and a leak-free Ag/AgCl electrode were used as a counter electrode and as a potential reference, respectively. The continuous flow of the electrolyte – CO2-saturated 0.1 M KHCO3 – was ensured by a peristaltic pump. The applied potential was controlled by a BioLogic potentiostat and was set to −1.2 VRHE. XAS measurements were carried out at the SAMBA beamline of the SOLEIL synchrotron radiation facility (Saint-Aubin, France). A Si(220) monochromator was used for the energy selection, while the XAS data were collected in fluorescence mode using an energy-selective 32-channel Ge detector. Alignment, background subtraction and normalization of the collected XANES data were performed using the Athena software (Ravel & Newville, 2005[Ravel, B. & Newville, M. (2005). J. Synchrotron Rad. 12, 537-541.]).

2.3. Quantitative XANES analysis

2.3.1. PCA and spectral decomposition of XANES dataset

The principal component analysis (PCA) and the spectral decomposition of the XANES dataset were performed using the PyFitIt code (Martini, Guda et al., 2020[Martini, A., Guda, S. A., Guda, A. A., Smolentsev, G., Algasov, A., Usoltsev, O., Soldatov, M. A., Bugaev, A., Rusalev, Y., Lamberti, C. & Soldatov, A. V. (2020). Comput. Phys. Commun. 250, 107064.]). The identification of the number of pure species contributing to the dataset was performed by examining the extracted principal components (PCs) and keeping only those k PCs with distinct spectral features. Additionally, four quantitative tests [scree plot, Malinowsky indicator factor (IND), imbedded error (IE) and Fisher test (Malinowski, 2002[Malinowski, E. R. (2002). Factor Analysis in Chemistry. Wiley.]; Manceau et al., 2014[Manceau, A., Marcus, M. & Lenoir, T. (2014). J. Synchrotron Rad. 21, 1140-1147.])] were used to further validate the number of pure species. The details of these statistical tests are discussed by Martini et al. (2023[Martini, A., Hursán, D., Timoshenko, J., Rüscher, M., Haase, F., Rettenmaier, C., Ortega, E., Etxebarria, A. & Roldan Cuenya, B. (2023). J. Am. Chem. Soc. 145, 17351-17366.]).

To convert the obtained PCs into the spectra of pure species we relied on a transformation matrix (TM) approach (Smolentsev et al., 2009[Smolentsev, G., Guilera, G., Tromp, M., Pascarelli, S. & Soldatov, A. V. (2009). J. Chem. Phys. 130, 174508.]; Martini, Guda et al., 2020[Martini, A., Guda, S. A., Guda, A. A., Smolentsev, G., Algasov, A., Usoltsev, O., Soldatov, M. A., Bugaev, A., Rusalev, Y., Lamberti, C. & Soldatov, A. V. (2020). Comput. Phys. Commun. 250, 107064.]; Martini et al., 2021[Martini, A., Guda, A. A., Guda, S. A., Dulina, A., Tavani, F., D'Angelo, P., Borfecchia, E. & Soldatov, A. V. (2021). Synchrotron Radiation Science and Applications, edited by A. Di Cicco, G. Giuli & A. Trapananti, pp. 65-84. Cham: Springer International Publishing.]; Martini & Borfecchia, 2020[Martini, A. & Borfecchia, E. (2020). Crystals, 10, 664.]). Here the dataset containing the original XANES spectra X was decomposed as X = [PTT^{-1}{{\Pi}}+{\cal{E}}]. Matrix P contains the k significant PCs identified as discussed above, while the rows of Π, referred to here as PC weights, are the projections of X onto each of the PCs. T is a k × k matrix whose elements can be varied until a set of chemical/physical interpretable spectra S = PT and concentration profiles C = T−1Π, satisfying the imposed constraints (i.e., non-negativity of the XANES and concentration profiles and mass balance condition), are identified. Finally, the error matrix [{\cal E}] contains the residuals between the original XANES spectra in X and their approximations using k PCs.

2.3.2. XANES fitting procedure

Identification of 3D structural motifs corresponding to the XANES spectra for the pure species was carried out based on the EXAFS analysis and XANES data fitting. The latter was facilitated by the indirect supervised machine learning (ML) approach, as implemented in the PyFitIt code (Martini, Guda et al., 2020[Martini, A., Guda, S. A., Guda, A. A., Smolentsev, G., Algasov, A., Usoltsev, O., Soldatov, M. A., Bugaev, A., Rusalev, Y., Lamberti, C. & Soldatov, A. V. (2020). Comput. Phys. Commun. 250, 107064.]). Herein, the FMDNES code (Joly, 2001[Joly, Y. (2001). Phys. Rev. B, 63, 125120.]; Guda et al., 2015[Guda, S. A., Guda, A. A., Soldatov, M. A., Lomachenko, K. A., Bugaev, A. L., Lamberti, C., Gawelda, W., Bressler, C., Smolentsev, G., Soldatov, A. V. & Joly, Y. (2015). J. Chem. Theory Comput. 11, 4512-4521.]) is used to calculate explicitly the XANES spectra for a few hundreds (or in some cases thousands) of relevant structure models, with different values for those structure parameters that characterize the environment around the single metal site. These structures for the ML training are chosen based on an adaptive sampling approach (active learning) (Tereshchenko et al., 2022[Tereshchenko, A., Pashkov, D., Guda, A., Guda, S., Rusalev, Y. & Soldatov, A. (2022). Molecules, 27, 357.]). Note that the FDMNES code has been demonstrated to be very accurate in the reproduction of the Co K-edge (Shapovalova et al., 2021[Shapovalova, S. O., Guda, A. A., Bubnov, M. P., Smolentsev, G., Rusalev, Y. V., Shapovalov, V. V., Zolotukhin, A. A., Cherkasov, V. K., Starikov, A. G., Vlasenko, V. G. & Soldatov, A. V. (2021). Chem. Lett. 50, 1933-1937.]). FDMNES simulations were performed by employing the finite difference method. Real Hedin–Lundqvist and von Barth local exchange correlation potentials (Hedin & Lundqvist, 1970[Hedin, L. & Lundqvist, S. (1970). Solid State Physics, Vol. 23, edited by F. Seitz, D. Turnbull & H. Ehrenreich, pp. 1-181. Academic Press.]; von Barth & Hedin, 1972[Barth, U. & Hedin, L. (1972). J. Phys. C.: Solid State Phys. 5, 1629-1642.]) were used. The cluster size was set to 5.5 Å. The increase of cluster size did not result in significant changes in the simulated spectra. All theoretical XANES data were aligned by correcting each energy grid by their related EPSII parameter (Joly, 2021[Joly, Y. (2021). The FDMNES project, https://fdmnes.neel.cnrs.fr/.]). At the same time, a common shift of 142 eV was applied to all of them as described by Martini et al. (2023[Martini, A., Hursán, D., Timoshenko, J., Rüscher, M., Haase, F., Rettenmaier, C., Ortega, E., Etxebarria, A. & Roldan Cuenya, B. (2023). J. Am. Chem. Soc. 145, 17351-17366.]).

The obtained theoretical spectra are then used to train an ML model, establishing a link between the values of the structure parameters and spectral features. The trained ML is then able to generate the corresponding XANES spectra for the intermediate values of the structure parameters that were not explicitly included in the training dataset, thus allowing us to obtain theoretical XANES much faster than in the FDMNES simulations, providing a significant speed-up for XANES data fitting. Our ML model, describing a XANES spectrum [\hat\mu\left(E{\semi}p\right)] as a function of the energy and of the structural parameters p, consists of a set of radial basis functions (RBF),

[\hat\mu\left(E{\semi}\,p\right) = \sum\limits_{i\,=\,1}^{N} w_i(E)\,K(\|p-p_i\|) + P(E{\semi}\,p). \eqno(1)]

Here the [\|\ldots\|] operator is the L2 norm, K(r) is a linear radial basis function while P(E;p) is a second-order polynomial with energy-dependent coefficients (Guda et al., 2021[Guda, A. A., Guda, S. A., Martini, A., Kravtsova, A. N., Algasov, A., Bugaev, A., Kubrin, S. P., Guda, L. V., Šot, P., van Bokhoven, J. A., Copéret, C. & Soldatov, A. V. (2021). NPJ Comput. Mater. 7, 203.]). The unknown factors wi and the polynomial coefficients are obtained during the ML training using the ridge regression. The accuracy of the trained ML model is tested using a ten-fold cross validation technique. See Martini et al. (2023[Martini, A., Hursán, D., Timoshenko, J., Rüscher, M., Haase, F., Rettenmaier, C., Ortega, E., Etxebarria, A. & Roldan Cuenya, B. (2023). J. Am. Chem. Soc. 145, 17351-17366.]), Martini, Guda et al. (2020[Martini, A., Guda, S. A., Guda, A. A., Smolentsev, G., Algasov, A., Usoltsev, O., Soldatov, M. A., Bugaev, A., Rusalev, Y., Lamberti, C. & Soldatov, A. V. (2020). Comput. Phys. Commun. 250, 107064.]) and Guda et al. (2021[Guda, A. A., Guda, S. A., Martini, A., Kravtsova, A. N., Algasov, A., Bugaev, A., Kubrin, S. P., Guda, L. V., Šot, P., van Bokhoven, J. A., Copéret, C. & Soldatov, A. V. (2021). NPJ Comput. Mater. 7, 203.]) for additional details on the ML training process.

The fit of the XANES spectra derived from the experimental measurements was performed using the coordinate descendent algorithm (Wright, 2015[Wright, S. J. (2015). Math. Program. 151, 3-34.]), minimizing the Rfactor between the experimental and the ML-derived XANES spectra, defined as

[R_{\rm{factor}} = {{ \|\mu^{\exp}(E)-\hat\mu(E\semi\,p)\|^2 }\over{ \|\mu^{\exp}(E)\|^2 }}. \eqno(2)]

The uncertainties were estimated by calculating the intervals of the structure parameter values which would result in an increase of the residual by less than 10%.

2.3.3. Pre-edge simulation

For the Co K-edge XANES pre-edge simulations, time-dependent DFT (TDDFT) calculations, implemented in the ORCA 5.0.4 code (Neese, 2012[Neese, F. (2012). WIREs Comput. Mol. Sci. 2, 73-78.], 2022[Neese, F. (2022). WIREs Comput. Mol. Sci. 12, e1606.]; Neese et al., 2020[Neese, F., Wennmohs, F., Becker, U. & Riplinger, C. (2020). J. Chem. Phys. 152, 224108.]), were employed. We used the CAM-B3LYP functional (Yanai et al., 2004[Yanai, T., Tew, D. P. & Handy, N. C. (2004). Chem. Phys. Lett. 393, 51-57.]) with the D3BJ dispersion correction (Grimme et al., 2011[Grimme, S., Ehrlich, S. & Goerigk, L. (2011). J. Comput. Chem. 32, 1456-1465.]) using the ZORA-def2-TZVP (vanLenthe et al., 1996[Lenthe, E. van, van Leeuwen, R., Baerends, E. J. & Snijders, J. G. (1996). Int. J. Quantum Chem. 57, 281-293.]; Schäfer et al., 1992[Schäfer, A., Horn, H. & Ahlrichs, R. (1992). J. Chem. Phys. 97, 2571-2577.]; Weigend & Ahlrichs, 2005[Weigend, F. & Ahlrichs, R. (2005). Phys. Chem. Chem. Phys. 7, 3297-3305.]) with the SARC/J auxiliary basis set (Rolfes et al., 2020[Rolfes, J. D., Neese, F. & Pantazis, D. A. (2020). J. Comput. Chem. 41, 1842-1849.]). The calculated XANES pre-edge transitions for a Co +2 state (doublet and quadruplet) were shifted by 17.1 eV until they matched the experimental pre-edge features and then broadened with 2 eV FWHM Lorentzian functions.

3. Results and discussion

Fig. 1[link] shows the Co K-edge XANES data for the as-prepared catalyst, and their evolution during the CO2RR. The obtained spectra for the as-prepared state match well with those in our prior work (Hursán et al., 2024[Hursán, D., Timoshenko, J., Ortega, E., Jeon, H. S., Rüscher, M., Herzog, A., Rettenmaier, C., Chee, S. W., Martini, A., Koshy, D. & Roldán Cuenya, B. (2024). Adv. Mater. 36, 2307809.]), but the new dataset allows us to track the changes in the catalyst structure as a function of the reaction time, thus providing valuable information about the possible reaction intermediates.

[Figure 1]
Figure 1
Co K-edge XANES dataset (27 spectra) collected under operando conditions during the CO2RR process. The inset shows a magnification of the pre-edge region. The arrows highlight the evolution of the XANES features with time under reaction conditions. Measurements performed at −1.2 VRHE in 0.1 M KHCO3 as electrolyte.

As discussed in our previous study (Hursán et al., 2024[Hursán, D., Timoshenko, J., Ortega, E., Jeon, H. S., Rüscher, M., Herzog, A., Rettenmaier, C., Chee, S. W., Martini, A., Koshy, D. & Roldán Cuenya, B. (2024). Adv. Mater. 36, 2307809.]), the spectrum collected at open circuit potential conditions before the reaction started, labelled as t = 0 min, can be associated with octahedrally coordinated Co sites, with the local geometry resembling that of an Ni–N–C catalyst (Martini et al., 2023[Martini, A., Hursán, D., Timoshenko, J., Rüscher, M., Haase, F., Rettenmaier, C., Ortega, E., Etxebarria, A. & Roldan Cuenya, B. (2023). J. Am. Chem. Soc. 145, 17351-17366.]). As the reaction proceeds, the XANES white-line intensity decreases, while the pre-edge increases (see the inset of Fig. 1[link]). Changes in the post-edge region of the XANES spectra indicate strong transformations in the catalyst atomistic structure. The observed changes can be associated with a reduction of the Co coordination number and/or the increased distortion of the local geometry.

To understand the number of different species contributing to the collected dataset, and to reveal their nature, we relied first on the PCA approach. The visual analysis of each PC, see Fig. 2[link](a), indicates that only the first three PCs exhibit meaningful spectral features. This suggests that only three spectroscopically distinct species are contributing to the collected XANES dataset. This conclusion is confirmed also by the analysis of the scree plot, IND and IE indicators, and also by the Fisher test, see Figs. 2(b)–2(e).

[Figure 2]
Figure 2
(a) First five principal components (PCs) extracted from the Co K-edge XANES dataset, weighted by their corresponding singular values. The numbers in brackets indicate the multiplicative factors used to enhance the PCs features for a clearer visualization. (b) Scree plot. (c) Imbedded error plot. (d) Malinowski indicator factor plot. (e) Significance level plot derived from the Fisher test. The arrows point to the number of pure species contributing to the XANES dataset, as obtained from each test. For all of them, this number was found to correspond to 3.

We note here that the PCs themselves do not correspond to the spectra of the pure species. Nonetheless, the latter can be obtained as some linear combination of the PCs. The spectra for the pure species and the corresponding concentration profiles can be retrieved by using the TM approach. Considering that the PCA revealed the presence of three spectroscopically distinct species, the TM matrix contains 3 × 3 = 9 unknown coefficients (projections of each of the pure species on each of the significant PCs). This number of unknowns can be further reduced by applying a set of constraints. First, we note that for the normalized XANES spectra, the weight of the first PC should be the same for all spectra in the dataset, and should be equal also to the weight of this PC to the spectra of the pure species (Martini, Guda et al., 2020[Martini, A., Guda, S. A., Guda, A. A., Smolentsev, G., Algasov, A., Usoltsev, O., Soldatov, M. A., Bugaev, A., Rusalev, Y., Lamberti, C. & Soldatov, A. V. (2020). Comput. Phys. Commun. 250, 107064.]). The latter enforces also the mass balance condition, i.e., that for each experimental XANES in the dataset the sum of the concentrations of all pure species has to be equal to 1 (Martini & Borfecchia, 2020[Martini, A. & Borfecchia, E. (2020). Crystals, 10, 664.]). This useful property is satisfied if the normalization of the experimental XANES spectra [\mu_i^{\exp}(E)] is carried out so that the quantity

[\xi_i = {{1}\over{E_2-E_1}}\int_{E_1}^{\,E_2} \mu_i^{\exp}(E)^2 \,{\rm{d}}E \eqno(3)]

is the same for all spectra. Here E1 and E2 define the limits of the energy range used for the analysis (7700 eV and 7825 eV in our case). Setting the first row of T to be equal to ξi thus ensures that both the normalization of the XANES spectra and mass balance condition are satisfied, and reduces the number of unknown elements in T from 9 to 6. Finally, a further constraint was imposed: we assumed that the first spectrum of our dataset (t = 0 min) corresponds to one of the pure species. This assumption is plausible since this spectrum corresponds to a stable state of the as-prepared sample. The projections of the first spectrum in our dataset on the PCs thus provides us with the values of the coefficients in the first column of T, and reduces the number of unknown coefficients to 4. The ranges of possible values for these unknown coefficients can be further reduced to ensure that the spectra for the pure species and the concentration profiles are non-negative. The obtained spectra for the pure species and the corresponding concentration profiles are shown in Fig. 3[link], and their validation will be provided in the following paragraphs.

[Figure 3]
Figure 3
Pure XANES spectra (a) and concentration profiles (b) extracted by the TM approach. The inset in (a) shows a magnification of the pre-edge region for the three XANES components.

The obtained XANES spectra for the first and third pure species in the working Co–N–C catalyst resemble strongly the pure spectra isolated using the same technique for the Ni–N–C under similar reaction conditions (Martini et al., 2023[Martini, A., Hursán, D., Timoshenko, J., Rüscher, M., Haase, F., Rettenmaier, C., Ortega, E., Etxebarria, A. & Roldan Cuenya, B. (2023). J. Am. Chem. Soc. 145, 17351-17366.]). Based on our prior works (Martini et al., 2023[Martini, A., Hursán, D., Timoshenko, J., Rüscher, M., Haase, F., Rettenmaier, C., Ortega, E., Etxebarria, A. & Roldan Cuenya, B. (2023). J. Am. Chem. Soc. 145, 17351-17366.]; Hursán et al., 2024[Hursán, D., Timoshenko, J., Ortega, E., Jeon, H. S., Rüscher, M., Herzog, A., Rettenmaier, C., Chee, S. W., Martini, A., Koshy, D. & Roldán Cuenya, B. (2024). Adv. Mater. 36, 2307809.]) we thus conclude that the first species (as-prepared state of the catalyst) can be associated with a Co site surrounded by four N atoms within the carbon ring plane and two axial O atoms or OH groups. We validate this assumption using Co K-edge XANES data fitting, which is discussed in detail in Section S6.1 of the supporting information. The third XANES component can be tentatively attributed, instead, to a more distorted environment for single cationic Co sites where the axial O atoms of the first component have been substituted by two CO. By analogy with our observations for the Ni–N–C system, we associate this species with the main active site for CO2RR. We confirm and investigate its structure in more detail below. However, we first need to comment on the second pure species, identified by the TM approach, whose XANES spectrum does not resemble the spectra of the pure species obtained during CO2RR for the Ni–N–C catalyst (Martini et al., 2023[Martini, A., Hursán, D., Timoshenko, J., Rüscher, M., Haase, F., Rettenmaier, C., Ortega, E., Etxebarria, A. & Roldan Cuenya, B. (2023). J. Am. Chem. Soc. 145, 17351-17366.]).

The XANES spectrum for this species exhibits the first pre-edge peak at ∼7709 eV, i.e., at lower energies than for the first and third component, which could suggest a more reduced chemical state of Co. Moreover, the second XANES feature at ∼7716 eV is more intense compared with that in the XANES spectra for the other pure species, see the inset of Fig. 3[link](a) and Fig. S2 of the supporting information. The intense second pre-edge peak could be associated with the Co 1s → 4pz shakedown transition and can suggest the presence of a four- or three-coordinated distorted geometry (Colpas et al., 1991[Colpas, G. J., Maroney, M. J., Bagyinka, C., Kumar, M., Willis, W. S., Suib, S. L., Mascharak, P. K. & Baidya, N. (1991). Inorg. Chem. 30, 920-928.]; Baker et al., 2017[Baker, M. L., Mara, M. W., Yan, J. J., Hodgson, K. O., Hedman, B. & Solomon, E. I. (2017). Coord. Chem. Rev. 345, 182-208.]). Further clues about the nature of this site can be extracted from the EXAFS analysis.

Since the concentration profiles for all species are known from the XANES analysis, the EXAFS spectra can be extracted from the original dataset as SXAS = XXASC(CCt)−1, where t denotes the matrix transpose and XXAS contains absorption spectra extended to 8290 eV and normalized in the same way as used for XANES data processing. The obtained EXAFS spectra for all three species are shown in Fig. S3.

For the qualitative interpretation of the EXAFS data we relied on wavelet transform (WT) analysis (Timoshenko & Kuzmin, 2009[Timoshenko, J. & Kuzmin, A. (2009). Comput. Phys. Commun. 180, 920-925.]; Funke et al., 2005[Funke, H., Scheinost, A. C. & Chukalina, M. (2005). Phys. Rev. B, 71, 094110.]; Martini, Signorile et al., 2020[Martini, A., Signorile, M., Negri, C., Kvande, K., Lomachenko, K. A., Svelle, S., Beato, P., Berlier, G., Borfecchia, E. & Bordiga, S. (2020). Phys. Chem. Chem. Phys. 22, 18950-18963.]) (see Section S5.1 for more technical details). While for the first and third pure species the WT-EXAFS spectra are consistent with those expected for the singly dispersed cationic site (Martini et al., 2023[Martini, A., Hursán, D., Timoshenko, J., Rüscher, M., Haase, F., Rettenmaier, C., Ortega, E., Etxebarria, A. & Roldan Cuenya, B. (2023). J. Am. Chem. Soc. 145, 17351-17366.]; Hursán et al., 2024[Hursán, D., Timoshenko, J., Ortega, E., Jeon, H. S., Rüscher, M., Herzog, A., Rettenmaier, C., Chee, S. W., Martini, A., Koshy, D. & Roldán Cuenya, B. (2024). Adv. Mater. 36, 2307809.]), for the second component [Fig. 4[link](a)] we observe an intense contribution of the second coordination shell in the R range between 2 and 2.7 Å, which appears to be split into two regions in the k-space, namely 1–3 Å−1 and 6–8 Å−1. The contribution at lower k values can be attributed to Co—C (or Co—O) and Co—N bonds. On the other hand, the position of the second contribution both in R- and k-spaces resembles that of Co—Co bonds in metallic Co, see Fig. 4[link](b), suggesting the existence of Co—Co moieties.

[Figure 4]
Figure 4
Moduli of the Morlet wavelet transform calculated for the EXAFS signal of the second pure species (a) and for the metallic Co foil (b). The wavelet resolution parameters σ and η were set to 1 and 6, respectively. The wavelet transforms are not corrected for the phase shift.

In order to confirm this conclusion, we performed EXAFS data fitting. Details on the EXAFS fitting and the complete list of the refined parameters can be found in Section S5.2 and in Table S2. The fit indicates the existence of a Co site with a highly distorted environment coordinating two N atoms (2 ± 1) and one additional Co atom (1.0 ± 0.9) located at 2.54 ± 0.08 Å from the absorbing atom. The large uncertainties stem from a large number of fitting parameters that need to be considered to describe the environment of the Co site in this case. However, the obtained results confirm the presence of Co—Co bonds that could be associated with the presence of Co moieties (e.g., Co dimers) or ultrasmall metallic clusters.

From the concentration profiles in Fig. 3[link](b) it is evident that the second component appears very quickly after the onset of CO2RR (within 5 min). After that, its concentration reaches ∼40% and it remains almost constant. On the other hand, the transformation of the first component into the third one is more gradual. After ∼35 min, the third component constitutes the majority of the chemical species, with a concentration of ∼60%. The obtained results agree well with our previous work (Hursán et al., 2024[Hursán, D., Timoshenko, J., Ortega, E., Jeon, H. S., Rüscher, M., Herzog, A., Rettenmaier, C., Chee, S. W., Martini, A., Koshy, D. & Roldán Cuenya, B. (2024). Adv. Mater. 36, 2307809.]), where we demonstrated, based on EXAFS analysis, that after 1 h of CO2RR ∼50% of the cobalt sites in the Co–N–C catalysts seem to aggregate and form metallic-like species. However, due to the limited information amount in EXAFS data for the TM-N-C systems and the lack of sufficient time resolution, our prior work could not provide definitive answers about the kinetics of the catalyst evolution, and, crucially, also about the nature of the remaining cationic Co species, which we believe are the main active sites for CO2RR. By focusing here on the analysis of time-resolved XANES data, we address this issue. Indeed, as demonstrated in our prior work (Hursán et al., 2024[Hursán, D., Timoshenko, J., Ortega, E., Jeon, H. S., Rüscher, M., Herzog, A., Rettenmaier, C., Chee, S. W., Martini, A., Koshy, D. & Roldán Cuenya, B. (2024). Adv. Mater. 36, 2307809.]), the concentration of metallic species depends also on the applied potential, and its increased content at lower cathodic potentials appears to correlate with the enhanced contribution of the parasitic hydrogen evolution reaction that is competing with CO2RR. This allows us to suggest that these species are not active for CO2RR, but the main active sites for the CO2 to CO conversion are instead associated with the third pure species, which preserves their singly dispersed cationic state, but, nonetheless, has a different Co environment than the single sites in the as-prepared catalyst. To clarify the local structure of the Co in these working sites, a XANES fitting was performed using the indirect machine learning approach.

The initial structure model was constructed based on the structure of the Co phthalocyanine complex (Crystallography Open Database, 2024[Crystallography Open Database (2024). https://www.crystallography.net/cod/2100746.html.]) with the addition of two axial CO ligands with initial Co—CO distance of 2 Å and initial C—O distance of 1.1 Å (National Institute of Standards and Technology Database, 2024[National Institute of Standards and Technology Database (2024). https://cccbdb.nist.gov/exp2x.asp?casno=630080&charge=0.]). To construct the training dataset, we distorted this initial structural model as shown in Fig. 5[link] and described in Table 1[link]. We emphasize here that, since XAS is a local technique, we are not sensitive to the arrangement of atoms that are further away from the central Co atom (e.g., the exact arrangement of C atoms in the carbon rings). The employed distortions are matching those used by us for the Ni–N–C case (Martini et al., 2023[Martini, A., Hursán, D., Timoshenko, J., Rüscher, M., Haase, F., Rettenmaier, C., Ortega, E., Etxebarria, A. & Roldan Cuenya, B. (2023). J. Am. Chem. Soc. 145, 17351-17366.]). In total, ∼5000 distorted structures were generated and used to simulate the ab initio XANES spectra. Afterwards, these spectra were normalized by a common factor α = 0.00248, and used as the training set for the RBF regressor. The trained ML model has a prediction accuracy value higher than 0.94, indicating that the model yields the XANES spectra closely matching those obtained using the explicit FDMNES simulations.

Table 1
List of structural parameters for the model shown in Fig. 5[link], employed in the fit of the third XANES component shown in Fig. 3[link](a)

Model used to describe the third Co K-edge XANES component
p1 Contraction/expansion of the N4 square [−0.2: +0.2] Å
p2 Shift of the Co atom and of the CO groups towards the edge of the N4 square [0: +0.3] Å
p3 Contraction/expansion of the axial Co—C bonds [−0.1: +0.2] Å
p4 Contraction/expansion of the C—O bonds. [−0.2: +0.2] Å
p5 Co—C—O bond angle. [135: 180]°
[Figure 5]
Figure 5
Set of structural deformations employed for the XANES fit of the third pure XANES component shown in Fig. 3[link](a).

Using the trained ML model, the XANES fit was performed in the energy range spanning from 7719 to 7795 eV, with the exclusion of the pre-edge region. The latter was necessary due to the lower accuracy of the FDMNES code in the simulations of this part of the XANES spectrum (Joly, 2021[Joly, Y. (2021). The FDMNES project, https://fdmnes.neel.cnrs.fr/.]; Guda et al., 2019[Guda, A. A., Guda, S. A., Lomachenko, K. A., Soldatov, M. A., Pankin, I. A., Soldatov, A. V., Braglia, L., Bugaev, A. L., Martini, A., Signorile, M., Groppo, E., Piovano, A., Borfecchia, E. & Lamberti, C. (2019). Catal. Today, 336, 3-21.]).

The XANES fitting results, together with the optimized structure, are reported in Fig. 6[link](a). The full list of the refined structural parameters can be found in Table S3. Table 2[link] reports the derived interatomic distances in the first coordination shell around Co, as well as the C—O interatomic distances and the ∠Co—C—O bonding angle.

Table 2
Interatomic distances and Co—CO angle, obtained from the XANES fit

The uncertainties are derived from the ones reported in Table S3 and are indicated in parenthesis.

Distance (average)/angle Co K-edge XANES best-fit values
Misfit (Rfactor): 0.96%
Co—C (C of the CO groups) 1.71 (2) Å
Co—N (two N atoms that are closer to Co) 1.77 (3) Å
Co—N (two N atoms that are further away from Co) 2.12 (3) Å
C—O (C and O of the CO groups) 1.25 (2) Å
∠Co—C—O bond angle 178 (6)°
[Figure 6]
Figure 6
(a) Best-fit of the XANES third component showed in Fig. 3[link](a) using the machine learning indirect approach. (b) Reproduction of the pre-edge region of the third component using TDDFT calculations (quartet state). The grey bars represent the transition strengths involved in the total spectrum, while the numbers 1–3 indicate the principal (acceptor) natural transition orbitals (Martin, 2003[Martin, R. L. (2003). J. Chem. Phys. 118, 4775-4777.]). The calculated spectrum and the single transitions are shifted by 17.1 eV in order to have the best match with the experimental features of the pre-edge. For the three main orbital representations the iso-value is set to 0.06.

One can see that the agreement between the XANES data extracted from the experimental dataset and the fitting results is very good, with an Rfactor value of lower than 1%. The obtained interatomic distances, in turn, are comparable with the ones obtained for the Ni–N–C case (Martini et al., 2023[Martini, A., Hursán, D., Timoshenko, J., Rüscher, M., Haase, F., Rettenmaier, C., Ortega, E., Etxebarria, A. & Roldan Cuenya, B. (2023). J. Am. Chem. Soc. 145, 17351-17366.]). The XANES fit shows that the Co atom is slightly displaced from the centre of the N4 square, resulting in two shorter and two longer Co—N4 bonds (with distances of 1.77 ± 0.03 Å and 2.12 ± 0.03 Å, respectively). Such symmetry breaking can be an indication that these singly dispersed Co sites are not distributed uniformly in the carbon support but are localized in the vicinity of some defects, as has been shown for similar systems elsewhere (see Fan et al., 2020[Fan, Q., Hou, P. F., Choi, C. H., Wu, T. S., Hong, S., Li, F., Soo, Y. L., Kang, P., Jung, Y. S. & Sun, Z. Y. (2020). Adv. Energy Mater. 10, 1903068.]; Kramm et al., 2012[Kramm, U. I., Herranz, J., Larouche, N., Arruda, T. M., Lefèvre, M., Jaouen, F., Bogdanoff, P., Fiechter, S., ABSWurmbach, I., Mukerjee, S. & Dodelet, J. P. (2012). Phys. Chem. Chem. Phys. 14, 11673-11688.]; Mou et al., 2019[Mou, K. W., Chen, Z. P., Zhang, X. X., Jiao, M. Y., Zhang, X. P., Ge, X., Zhang, W. & Liu, L. C. (2019). Small, 15, 1903668.]; Wang et al., 2018[Wang, X. Q., Chen, Z., Zhao, X. Y., Yao, T., Chen, W. X., You, R., Zhao, C. M., Wu, G., Wang, J., Huang, W. X., Yang, J. L., Hong, X., Wei, S. Q., Wu, Y. & Li, Y. D. (2018). Angew. Chem. Int. Ed. 57, 1944-1948.]; Zheng et al., 2018[Zheng, T., Jiang, K. & Wang, H. (2018). Adv. Mater. 30, 1802066.]; Yan et al., 2018[Yan, C. C., Li, H. B., Ye, Y. F., Wu, H. H., Cai, F., Si, R., Xiao, J. P., Miao, S., Xie, S. H., Yang, F., Li, Y. S., Wang, G. X. & Bao, X. H. (2018). Energy Environ. Sci. 11, 1204-1210.]; Rong et al., 2020[Rong, X., Wang, H. J., Lu, X. L., Si, R. & Lu, T. B. (2020). Angew. Chem. Int. Ed. 59, 1961-1965.]; Zhao et al., 2017[Zhao, C. M., Dai, X. Y., Yao, T., Chen, W. X., Wang, X. Q., Wang, J., Yang, J., Wei, S. Q., Wu, Y. E. & Li, Y. D. (2017). J. Am. Chem. Soc. 139, 8078-8081.]; Zhou et al., 2023[Zhou, Y., Zhou, Q., Liu, H., Xu, W., Wang, Z., Qiao, S., Ding, H., Chen, D., Zhu, J., Qi, Z., Wu, X., He, Q. & Song, L. (2023). Nat. Commun. 14, 3776.]). Interestingly, while the average Co—N bond length in Co–N–C is slightly lower than the Ni—N bond length in the Ni–N–C (where the obtained interatomic distances for the shorter and longer Ni—N bonds were 1.82 and 2.17 Å, respectively), the difference between the longer and shorter bonds in both cases is ∼0.35 Å, suggesting a similar degree of the off-centre displacement. The axial CO bonds, in the case of Co–C–N, form a ∠Co—C—O angle of 178 ± 6°, with a Co—C distance of 1.71 ± 0.02 Å and a C—O distance of 1.25 ± 0.02 Å. Within the uncertainty, the C—O distance is the same for Co–N–C obtained for Ni–N–C, while the Co—C distance is slightly shorter than the Ni—C distance in Ni–N–C (1.78 Å), which can be attributed to the smaller size of the Co cation. As we mentioned in Section 2.3.2[link], the pre-edge region was not included in the fit. Nonetheless, its overall shape is reasonably reproduced by our fitting results, including the relative positions and the shape of the pre-edge peaks, see Fig. 6[link](a) and Fig. S8. This is particularly important, considering that this XANES part can be rich in spectroscopic details about the molecular structure of the Co site because it mainly involves the transitions of the Co 1s electrons to empty 3d states and to the unoccupied molecular orbitals of the possible adsorbates (Gallo et al., 2011[Gallo, E., Lamberti, C. & Glatzel, P. (2011). Phys. Chem. Chem. Phys. 13, 19409-19419.]).

To exploit this sensitivity of the pre-edge, and to validate the obtained structural model, including the presence of CO adsorbates, the simulations of the pre-edge were performed using a TDDFT approach. The ORCA 5.0.4 code was employed and the refined Co–N–C structure model, obtained in PyFitIt/FDMNES simulations, was used as the input for the calculation. All the C terminal atoms of the carbon rings were saturated with H to ensure charge neutrality of the model. Both doublet (low spin) and quadruplet (high spin) spin states were considered. However, the differences between these obtained results for these two cases were found minimal, see Fig. S12, thus they could not be reliably discriminated within the energy resolution of our experiment.

Alternative spectroscopic techniques could be employed here, such as Kα X-ray emission spectroscopy (Saveleva et al., 2021[Saveleva, V. A., Ebner, K., Ni, L. M., Smolentsev, G., Klose, D., Zitolo, A., Marelli, E., Li, J. K., Medarde, M., Safonova, O. V., Nachtegaal, M., Jaouen, F., Kramm, U. I., Schmidt, T. J. & Herranz, J. (2021). Angew. Chem. Int. Ed. 60, 11707-11712.]; Saveleva, Kumar et al., 2023[Saveleva, V. A., Kumar, K., Theis, P., Salas, N. S., Kramm, U. I., Jaouen, F., Maillard, F. & Glatzel, P. (2023). ACS Appl. Energy Mater. 6, 611-616.]). Overall, the quadruplet simulated spectrum shows a slightly better agreement with the experiment, and is thus reported in Fig. 6[link](b) together with the schematics of the acceptor orbitals responsible for the most intense spectroscopic transitions. We conclude that the first pre-edge peak at ∼7711 eV stems from two main transitions involving the Co core 1s orbital and two molecular orbitals (points 1 and 2 in Fig. 6[link]) dominated mainly by the Co eg orbitals characters: [3d_{{x^2}-{y^2}}] and 3dz2, respectively. At the same time, the peak at ∼7716 eV is attributed to a metal-to-ligand charge transfer. The simulation shows, in fact, that it derives from the non-occupied states of the 2p(C)–2p(O) π* anti-bonding molecular orbital (point 3), thus confirming the presence of CO adsorbates.

4. Conclusions

In this work we demonstrate how the operando time-resolved XAS in combination with advanced data analysis approaches can be employed to probe the heterogeneous, dynamic structure of the TM-N-C catalysts. The combination of un­supervised ML approaches as PCA and TM methods allowed us to identify the main species available and co-existing in the working Co–N–C catalysts for the electrocatalytic CO2 reduction, and for the first time to track their kinetics. XANES data fitting, facilitated by supervised ML and XANES simulations within the TDDFT approach, in turn clarified the structure of the main active species for CO2 conversion to CO. We observed that the speciation of Co–N–C catalysts is more complex than that for the Ni–N–C catalyst, and, in particular, that the formation of reduced Co clusters plays an important role. The amount of reduced Co clusters, and the sizes of the metallic Co particles are likely to depend strongly on the details of the electrochemical system, including the sample preparation, metal loading, applied potential and duration of the experiment. The presence of these species, which we believe are not active for CO2RR but facilitate hydrogen evolution reaction, could explain the large discrepancies between the CO2RR activities for Co–N–C materials reported in the literature. Indeed, our analysis of the local structure of originally singly dispersed cationic Co species that coexist with multi-Co moieties under CO2RR conditions suggest that their structure is highly similar to that of active Ni–N–C catalysts. In particular, the interaction between Co and CO adsorbates can be directly observed in our XANES data. The obtained results thus show promise that Co–N–C catalysts can potentially reach similar CO2RR activities as their Ni-based counterparts, provided that the highly dispersed Co–N–C species can be maximized during CO2RR, for instance by their careful initial placement at/close defects on the N-doped C support matrix.

Intriguingly, our results suggest that the kinetics of the formation of metallic clusters and active single Co sites under CO2RR conditions are vastly different. The provided guidance from operando measurements with increased temporal resolution, involving sub-second quick-XAS measurements (Timoshenko et al., 2022[Timoshenko, J., Bergmann, A., Rettenmaier, C., Herzog, A., Arán-Ais, R. M., Jeon, H. S., Haase, F. T., Hejral, U., Grosse, P., Kühl, S., Davis, E. M., Tian, J., Magnussen, O. & Roldan Cuenya, B. (2022). Nat. Catal. 5, 259-267.]), thus would be highly instrumental for the rational design of these future experiments. On the other hand, the measurements of high-energy-resolution fluorescence-detected XANES (Saveleva, Retegan et al., 2023[Saveleva, V. A., Retegan, M., Kumar, K., Maillard, F. & Glatzel, P. (2023). J. Mater. Chem. A, 11, 18862-18871.]), more sensitive to the geometry of the site because of the lower order of the core–hole lifetime broadening, would help to provide further details about the interactions between the metal site and adsorbates, yielding new insight into the working mechanism of TM-N-C catalysts.

Supporting information


Footnotes

Current address: eChemicles Zrt, Alsó Kikötő sor 11, Szeged H-6726, Hungary.

Acknowledgements

This article is dedicated to Professor Carlo Lamberti, a pioneer in the field of study discussed here who has made seminal contributions to the fundamental understanding of functional materials using experimental methods such as XAS. We, as his students and colleagues, will now do our best to carry the torch bringing the field forward into future applications. We acknowledge SOLEIL for provision of synchrotron radiation facilities and we would like to thank Dr Andrea Zitolo and Dr Emiliano Fonda for assistance in using the beamline SAMBA. We also thank Dr Gabriele Deplano (University of Turin) for fruitful discussions and advice concerning the TDDFT calculations performed with the ORCA software. Open access funding enabled and organized by Projekt DEAL.

Funding information

The following funding is acknowledged: Alexander von Humboldt Foundation (to DH).

References

First citationBaker, M. L., Mara, M. W., Yan, J. J., Hodgson, K. O., Hedman, B. & Solomon, E. I. (2017). Coord. Chem. Rev. 345, 182–208.  Web of Science CrossRef CAS PubMed Google Scholar
First citationBarth, U. & Hedin, L. (1972). J. Phys. C.: Solid State Phys. 5, 1629–1642.  CrossRef Web of Science Google Scholar
First citationBenfatto, M., Della Longa, S., Pace, E., Chillemi, G., Padrin, C., Natoli, C. R. & Sanna, N. (2021). Comput. Phys. Commun. 265, 107992.  CrossRef Google Scholar
First citationColpas, G. J., Maroney, M. J., Bagyinka, C., Kumar, M., Willis, W. S., Suib, S. L., Mascharak, P. K. & Baidya, N. (1991). Inorg. Chem. 30, 920–928.  CrossRef CAS Web of Science Google Scholar
First citationCrystallography Open Database (2024). https://www.crystallography.net/cod/2100746.htmlGoogle Scholar
First citationFan, Q., Hou, P. F., Choi, C. H., Wu, T. S., Hong, S., Li, F., Soo, Y. L., Kang, P., Jung, Y. S. & Sun, Z. Y. (2020). Adv. Energy Mater. 10, 1903068.  CrossRef Google Scholar
First citationFunke, H., Scheinost, A. C. & Chukalina, M. (2005). Phys. Rev. B, 71, 094110.  Web of Science CrossRef Google Scholar
First citationGallo, E., Lamberti, C. & Glatzel, P. (2011). Phys. Chem. Chem. Phys. 13, 19409–19419.  CrossRef CAS PubMed Google Scholar
First citationGenovese, C., Schuster, M. E., Gibson, E. K., Gianolio, D., Posligua, V., Grau-Crespo, R., Cibin, G., Wells, P. P., Garai, D., Solokha, V., Krick Calderon, S., Velasco-Velez, J. J., Ampelli, C., Perathoner, S., Held, G., Centi, G. & Arrigo, R. (2018). Nat. Commun. 9, 935.  CrossRef PubMed Google Scholar
First citationGong, Y. N., Jiao, L., Qian, Y. Y., Pan, C. Y., Zheng, L. R., Cai, X. C., Liu, B., Yu, S. H. & Jiang, H. L. (2020). Angew. Chem. Int. Ed. 59, 2705–2709.  CrossRef CAS Google Scholar
First citationGrimme, S., Ehrlich, S. & Goerigk, L. (2011). J. Comput. Chem. 32, 1456–1465.  Web of Science CrossRef CAS PubMed Google Scholar
First citationGuda, A. A., Guda, S. A., Lomachenko, K. A., Soldatov, M. A., Pankin, I. A., Soldatov, A. V., Braglia, L., Bugaev, A. L., Martini, A., Signorile, M., Groppo, E., Piovano, A., Borfecchia, E. & Lamberti, C. (2019). Catal. Today, 336, 3–21.  Web of Science CrossRef CAS Google Scholar
First citationGuda, A. A., Guda, S. A., Martini, A., Kravtsova, A. N., Algasov, A., Bugaev, A., Kubrin, S. P., Guda, L. V., Šot, P., van Bokhoven, J. A., Copéret, C. & Soldatov, A. V. (2021). NPJ Comput. Mater. 7, 203.  Google Scholar
First citationGuda, S. A., Guda, A. A., Soldatov, M. A., Lomachenko, K. A., Bugaev, A. L., Lamberti, C., Gawelda, W., Bressler, C., Smolentsev, G., Soldatov, A. V. & Joly, Y. (2015). J. Chem. Theory Comput. 11, 4512–4521.  Web of Science CrossRef CAS PubMed Google Scholar
First citationHedin, L. & Lundqvist, S. (1970). Solid State Physics, Vol. 23, edited by F. Seitz, D. Turnbull & H. Ehrenreich, pp. 1–181. Academic Press.  Google Scholar
First citationHursán, D., Timoshenko, J., Ortega, E., Jeon, H. S., Rüscher, M., Herzog, A., Rettenmaier, C., Chee, S. W., Martini, A., Koshy, D. & Roldán Cuenya, B. (2024). Adv. Mater. 36, 2307809.  Google Scholar
First citationJoly, Y. (2001). Phys. Rev. B, 63, 125120.  Web of Science CrossRef Google Scholar
First citationJoly, Y. (2021). The FDMNES project, https://fdmnes.neel.cnrs.fr/Google Scholar
First citationJu, W., Bagger, A., Hao, G. P., Varela, A. S., Sinev, I., Bon, V., Roldan Cuenya, B., Kaskel, S., Rossmeisl, J. & Strasser, P. (2017). Nat. Commun. 8, 944.  CrossRef PubMed Google Scholar
First citationKramm, U. I., Herranz, J., Larouche, N., Arruda, T. M., Lefèvre, M., Jaouen, F., Bogdanoff, P., Fiechter, S., ABSWurmbach, I., Mukerjee, S. & Dodelet, J. P. (2012). Phys. Chem. Chem. Phys. 14, 11673–11688.  CrossRef CAS PubMed Google Scholar
First citationLenthe, E. van, van Leeuwen, R., Baerends, E. J. & Snijders, J. G. (1996). Int. J. Quantum Chem. 57, 281–293.  Google Scholar
First citationLi, K., Zhang, S. B., Zhang, X. L., Liu, S., Jiang, H. S., Jiang, T. L., Shen, C. Y., Yu, Y. & Chen, W. (2022). Nano Lett. 22, 1557–1565.  CrossRef CAS PubMed Google Scholar
First citationLiang, S., Huang, L., Gao, Y., Wang, Q. & Liu, B. (2021). Adv. Sci. 8, 2102886.  CrossRef Google Scholar
First citationMalinowski, E. R. (2002). Factor Analysis in Chemistry. Wiley.  Google Scholar
First citationManceau, A., Marcus, M. & Lenoir, T. (2014). J. Synchrotron Rad. 21, 1140–1147.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationMartin, R. L. (2003). J. Chem. Phys. 118, 4775–4777.  CrossRef CAS Google Scholar
First citationMartini, A. & Borfecchia, E. (2020). Crystals, 10, 664.  CrossRef Google Scholar
First citationMartini, A., Guda, A. A., Guda, S. A., Dulina, A., Tavani, F., D'Angelo, P., Borfecchia, E. & Soldatov, A. V. (2021). Synchrotron Radiation Science and Applications, edited by A. Di Cicco, G. Giuli & A. Trapananti, pp. 65–84. Cham: Springer International Publishing.  Google Scholar
First citationMartini, A., Guda, S. A., Guda, A. A., Smolentsev, G., Algasov, A., Usoltsev, O., Soldatov, M. A., Bugaev, A., Rusalev, Y., Lamberti, C. & Soldatov, A. V. (2020). Comput. Phys. Commun. 250, 107064.  Web of Science CrossRef Google Scholar
First citationMartini, A., Hursán, D., Timoshenko, J., Rüscher, M., Haase, F., Rettenmaier, C., Ortega, E., Etxebarria, A. & Roldan Cuenya, B. (2023). J. Am. Chem. Soc. 145, 17351–17366.  CrossRef CAS PubMed Google Scholar
First citationMartini, A., Signorile, M., Negri, C., Kvande, K., Lomachenko, K. A., Svelle, S., Beato, P., Berlier, G., Borfecchia, E. & Bordiga, S. (2020). Phys. Chem. Chem. Phys. 22, 18950–18963.  CrossRef CAS PubMed Google Scholar
First citationMou, K. W., Chen, Z. P., Zhang, X. X., Jiao, M. Y., Zhang, X. P., Ge, X., Zhang, W. & Liu, L. C. (2019). Small, 15, 1903668.  CrossRef Google Scholar
First citationNational Institute of Standards and Technology Database (2024). https://cccbdb.nist.gov/exp2x.asp?casno=630080&charge=0Google Scholar
First citationNeese, F. (2012). WIREs Comput. Mol. Sci. 2, 73–78.  Web of Science CrossRef CAS Google Scholar
First citationNeese, F. (2022). WIREs Comput. Mol. Sci. 12, e1606.  Google Scholar
First citationNeese, F., Wennmohs, F., Becker, U. & Riplinger, C. (2020). J. Chem. Phys. 152, 224108.  Web of Science CrossRef PubMed Google Scholar
First citationPan, F. P., Deng, W., Justiniano, C. & Li, Y. (2018). Appl. Catal. Environ. 226, 463–472.  CrossRef CAS Google Scholar
First citationPan, Y., Lin, R., Chen, Y. J., Liu, S. J., Zhu, W., Cao, X., Chen, W. X., Wu, K. L., Cheong, W. C., Wang, Y., Zheng, L. R., Luo, J., Lin, Y., Liu, Y. Q., Liu, C. G., Li, J., Lu, Q., Chen, X., Wang, D. S., Peng, Q., Chen, C. & Li, Y. D. (2018). J. Am. Chem. Soc. 140, 4218–4221.  CrossRef CAS PubMed Google Scholar
First citationRavel, B. & Newville, M. (2005). J. Synchrotron Rad. 12, 537–541.  Web of Science CrossRef CAS IUCr Journals Google Scholar
First citationRolfes, J. D., Neese, F. & Pantazis, D. A. (2020). J. Comput. Chem. 41, 1842–1849.  CrossRef CAS PubMed Google Scholar
First citationRong, X., Wang, H. J., Lu, X. L., Si, R. & Lu, T. B. (2020). Angew. Chem. Int. Ed. 59, 1961–1965.  CrossRef CAS Google Scholar
First citationSaveleva, V. A., Ebner, K., Ni, L. M., Smolentsev, G., Klose, D., Zitolo, A., Marelli, E., Li, J. K., Medarde, M., Safonova, O. V., Nachtegaal, M., Jaouen, F., Kramm, U. I., Schmidt, T. J. & Herranz, J. (2021). Angew. Chem. Int. Ed. 60, 11707–11712.  CrossRef CAS Google Scholar
First citationSaveleva, V. A., Kumar, K., Theis, P., Salas, N. S., Kramm, U. I., Jaouen, F., Maillard, F. & Glatzel, P. (2023). ACS Appl. Energy Mater. 6, 611–616.  CrossRef CAS Google Scholar
First citationSaveleva, V. A., Retegan, M., Kumar, K., Maillard, F. & Glatzel, P. (2023). J. Mater. Chem. A, 11, 18862–18871.  CrossRef CAS Google Scholar
First citationSchäfer, A., Horn, H. & Ahlrichs, R. (1992). J. Chem. Phys. 97, 2571–2577.  CrossRef Web of Science Google Scholar
First citationShapovalova, S. O., Guda, A. A., Bubnov, M. P., Smolentsev, G., Rusalev, Y. V., Shapovalov, V. V., Zolotukhin, A. A., Cherkasov, V. K., Starikov, A. G., Vlasenko, V. G. & Soldatov, A. V. (2021). Chem. Lett. 50, 1933–1937.  CrossRef CAS Google Scholar
First citationSmolentsev, G., Guilera, G., Tromp, M., Pascarelli, S. & Soldatov, A. V. (2009). J. Chem. Phys. 130, 174508.  Web of Science CrossRef PubMed Google Scholar
First citationTereshchenko, A., Pashkov, D., Guda, A., Guda, S., Rusalev, Y. & Soldatov, A. (2022). Molecules, 27, 357.  CrossRef PubMed Google Scholar
First citationTimoshenko, J., Bergmann, A., Rettenmaier, C., Herzog, A., Arán-Ais, R. M., Jeon, H. S., Haase, F. T., Hejral, U., Grosse, P., Kühl, S., Davis, E. M., Tian, J., Magnussen, O. & Roldan Cuenya, B. (2022). Nat. Catal. 5, 259–267.  CrossRef CAS Google Scholar
First citationTimoshenko, J. & Frenkel, A. I. (2019). ACS Catal. 9, 10192–10211.  Web of Science CrossRef CAS Google Scholar
First citationTimoshenko, J. & Kuzmin, A. (2009). Comput. Phys. Commun. 180, 920–925.  Web of Science CrossRef CAS Google Scholar
First citationTimoshenko, J. & Roldan Cuenya, B. (2021). Chem. Rev. 121, 882–961.  Web of Science CrossRef CAS PubMed Google Scholar
First citationWang, S., Gao, H., Li, L., Hui, K. S., Dinh, D. A., Wu, S., Kumar, S., Chen, F., Shao, Z. & Hui, K. N. (2022). Nano Energy, 100, 107517.  CrossRef Google Scholar
First citationWang, X. Q., Chen, Z., Zhao, X. Y., Yao, T., Chen, W. X., You, R., Zhao, C. M., Wu, G., Wang, J., Huang, W. X., Yang, J. L., Hong, X., Wei, S. Q., Wu, Y. & Li, Y. D. (2018). Angew. Chem. Int. Ed. 57, 1944–1948.  CrossRef CAS Google Scholar
First citationWeigend, F. & Ahlrichs, R. (2005). Phys. Chem. Chem. Phys. 7, 3297–3305.  Web of Science CrossRef PubMed CAS Google Scholar
First citationWright, S. J. (2015). Math. Program. 151, 3–34.  CrossRef Google Scholar
First citationWu, Z., Karamad, M., Yong, X., Huang, Q., Cullen, D. A., Zhu, P., Xia, C., Xiao, Q., Shakouri, M., Chen, F., Kim, J. Y., Xia, Y., Heck, K., Hu, Y., Wong, M. S., Li, Q., Gates, I., Siahrostami, S. & Wang, H. (2021). Nat. Commun. 12, 2870.  CrossRef PubMed Google Scholar
First citationXiang, S. T., Huang, P. P., Li, J. Y., Liu, Y., Marcella, N., Routh, P. K., Li, G. H. & Frenkel, A. I. (2022). Phys. Chem. Chem. Phys. 24, 5116–5124.  CrossRef CAS PubMed Google Scholar
First citationYan, C. C., Li, H. B., Ye, Y. F., Wu, H. H., Cai, F., Si, R., Xiao, J. P., Miao, S., Xie, S. H., Yang, F., Li, Y. S., Wang, G. X. & Bao, X. H. (2018). Energy Environ. Sci. 11, 1204–1210.  CrossRef CAS Google Scholar
First citationYanai, T., Tew, D. P. & Handy, N. C. (2004). Chem. Phys. Lett. 393, 51–57.  Web of Science CrossRef CAS Google Scholar
First citationYang, J., Qiu, Z. Y., Zhao, C. M., Wei, W. C., Chen, W. X., Li, Z. J., Qu, Y. T., Dong, J. C., Luo, J., Li, Z. Y. & Wu, Y. (2018). Angew. Chem. Int. Ed. 57, 14095–14100.  CrossRef CAS Google Scholar
First citationZhao, C. M., Dai, X. Y., Yao, T., Chen, W. X., Wang, X. Q., Wang, J., Yang, J., Wei, S. Q., Wu, Y. E. & Li, Y. D. (2017). J. Am. Chem. Soc. 139, 8078–8081.  CrossRef CAS PubMed Google Scholar
First citationZheng, T., Jiang, K. & Wang, H. (2018). Adv. Mater. 30, 1802066.  CrossRef Google Scholar
First citationZhou, Y., Zhou, Q., Liu, H., Xu, W., Wang, Z., Qiao, S., Ding, H., Chen, D., Zhu, J., Qi, Z., Wu, X., He, Q. & Song, L. (2023). Nat. Commun. 14, 3776.  CrossRef PubMed Google Scholar
First citationZitolo, A., Goellner, V., Armel, V., Sougrati, M. T., Mineva, T., Stievano, L., Fonda, E. & Jaouen, F. (2015). Nat. Mater. 14, 937–942.  CrossRef CAS PubMed Google Scholar
First citationZitolo, A., Ranjbar-Sahraie, N., Mineva, T., Li, J., Jia, Q., Stamatin, S., Harrington, G. F., Lyth, S. M., Krtil, P., Mukerjee, S., Fonda, E. & Jaouen, F. (2017). Nat. Commun. 8, 957.  CrossRef PubMed Google Scholar

This is an open-access article distributed under the terms of the Creative Commons Attribution (CC-BY) Licence, which permits unrestricted use, distribution, and reproduction in any medium, provided the original authors and source are cited.

Journal logoJOURNAL OF
SYNCHROTRON
RADIATION
ISSN: 1600-5775
Follow J. Synchrotron Rad.
Sign up for e-alerts
Follow J. Synchrotron Rad. on Twitter
Follow us on facebook
Sign up for RSS feeds