Open AccessCCS ChemistryMINI REVIEW3 May 2024

Heterogeneous Catalysis with Metal-Containing Crystalline Porous Materials: Element- and Site-Specific Insights

    Metal-containing crystalline porous materials (CPMs) are gaining popularity in heterogeneous catalysis because of their highly crystalline and porous systems, and their excellent chemical tunability. Modification of the metal species and framework structure permits them to have greater activity, selectivity, and stability over other materials. An in-depth understanding of the complex nature of metal active sites in CPMs is essential for revealing the structure-performance relationships and directing the rational design of such catalysts. Compared to conventional characterization techniques, the rapid development of X-ray absorption spectroscopy (XAS) has provided element- and site-specific deep insights into the electronic and structural information of metal species in CPMs. As such, this review begins by summarizing novel XAS techniques and analysis methods in accurately obtaining such data. Next, the combination of XAS with other high-level characterization methods into disclosing the configuration of active sites in metal-containing CPMs is presented. Then, the utilization of theory-assisted XAS data analysis in examining complex metal-containing CPM catalysts is discussed. Afterwards, advanced in-situ/operando XAS studies into revealing the working sites in metal-containing CPMs under catalytic conditions are highlighted. We conclude by outlining the future challenges and prospects of XAS measurements, data analyses, and in-situ/operando setups in advancing the study of metal-in-CPM catalysts.

    Introduction

    Crystalline porous materials (CPMs), such as zeolites, metal–organic frameworks (MOFs), covalent organic frameworks (COFs), and zeolitic imidazolate frameworks (ZIFs), are of great interest due to their high crystallinity, superior porosity, and excellent chemical tunability, making them ideal materials in the design of heterogeneous catalysts.13 Implanting active metal species into CPMs can endow them with tremendous catalytic potential in heterogeneous catalysis, including thermocatalysis, electrocatalysis, and photocatalysis.46 Owing to the different forms of metal species (nanoparticles, nanoclusters, single-atoms, cations, etc.) and the diversity in framework structure of CPMs, such catalysts can exhibit activity, selectivity, and stability in heterogeneous catalysis.712 Despite the promising aspects of metal-containing CPMs, understanding the complex configurations of metal active sites in CPMs remains challenging.

    The interaction between the metal species and the long-range ordered frameworks of CPMs modulates the unique geometric and electronic structures of the metal species.5 Therefore, the determination of the precise configuration of metal species in CPMs plays a crucial role in designing and optimizing such catalysts. However, the routine use of conventional characterization methods such as X-ray diffraction, infrared spectroscopy, ultraviolet–visible spectroscopy, X-ray photoelectron spectroscopy (XPS), and electron microscopy, can provide fundamental identification of particle sizes, species, and valence states for metal active sites in CPMs.1315 To dig out atomic level structural and dynamical information from metal-in-CPM catalysts, more powerful characterizations techniques are still required.

    X-ray absorption spectroscopy (XAS) is a characterization technique that element-selectively examines the bulk geometries, electronic properties, and atomic structures of metal active sites in multielement CPMs.16 Typically, an XAS spectrum can be divided into two regions (Figure 1): X-ray absorption near edge structure (XANES) and extended X-ray absorption fine structure (EXAFS). The XANES region, occurring approximately 50 eV before and after the absorption edge, exhibits sensitivity towards the valence states and electronic structures of the detected elements. This is because the core electron energy is influenced by the electron distribution in the valence state.17 Notably, for 3d transition metals, the main absorption edge pertains to transitions from the 1s to 4p states while the corresponding pre-edge, on the other hand, covers transitions from the 1s to 3d states, which is useful in probing the geometries of metal clusters.18 The EXAFS region starts from about 50 eV and extends to roughly 1000 eV beyond the absorption edge, which is dominated by photoelectron absorption. The EXAFS region provides limited electronic structure details due to the high density of available electronic states and the high energy of the photoelectrons; yet, it is capable of providing information on coordination numbers (CNs) and bond distances around the absorber, as it is a sensitive probe of spatial arrangements of nearest neighbors around the absorbing atom(s).19 The rapid development of XAS, both as an X-ray technique [e.g., high energy resolution fluorescence detection (HERFD)] and as an analysis method [e.g., wavelet transformation (WT) for EXAFS, linear combination fitting (LCF) for XANES] has drawn significant attention since more detailed information on the atomic structures, electronic properties, and species determination of the metal active sites in CPMs can be obtained.2022

    Figure 1

    Figure 1 | Illustration of an XAS spectrum and the advantages of XAS in metal-in-CPMs’ characterization.

    Additionally, when XAS data is coupled with experimental results obtained using complementary advanced techniques such as scanning transmission electron microscope (STEM), CO-diffuse reflectance infrared Fourier transform spectroscopy (CO-DRIFTS), electron paramagnetic resonance (EPR) spectroscopy, Mössbauer spectroscopy, and XPS, more integrated characterization for metal-in-CPM catalysts can be achieved.2325 Moreover, the successful adoption of computational theory [e.g., XANES/EXAFS simulation, multivariate curve resolution-alternating least squares analysis (MCR-ALS), new XAS analysis program, etc.] is helping to unravel the complex nature of metal active sites in CPMs.2628

    In recent years, in-situ/operando XAS is becoming more commonly used in a broad range of metal-in-CPM materials (ordered, disordered, nanostructured, etc.) under different working conditions, which have facilitated the capture of oxidation states and local geometric coordination changes, thus promoting the elucidation of catalysis mechanisms.2931 Furthermore, thanks to the development of in-situ/operando XAS, particularly in time-resolved measurements, simultaneous multispectra measurements, and machine-learning (ML) assisted analyses, such advanced characterization techniques allow deeper insight of the evolution of coordination environments of active metal sites in CPMs under catalytic conditions.3234

    XAS shows many advantages in the characterization of metal-in-CPMs (Figure 1) and is becoming increasingly important in the design and optimization of metal-in-CPM catalysts. Therefore, in this minireview, we will summarize the progress of some representative works about advanced XAS analyses for metal-containing CPMs in heterogeneous catalysis over the last 5 years. We will first introduce the novel X-ray techniques and analysis methods which have been used to accurately identify the chemical states and atomic structures of metal sites in CPMs. Then, we will outline the combined use of XAS and other advanced characterization techniques to disclose the configuration of metal active centers in CPMs. Next, we will discuss theory-assisted XAS analyses which greatly deepen the understanding on complex metal-in-CPM systems. In-situ/operando XAS, which is divided into thermocatalysis, electrocatalysis/photocatalysis, and advanced in-situ/operando XAS techniques, will also be introduced, thus shedding light on the structure-performance relationships of metal-containing CPM catalysts. Finally, we will highlight the current challenges and prospects of XAS measurements, data analyses, and in-situ/operando setups for metal-in-CPM catalysts.

    Advanced XAS Analysis for Metal-Containing CPM Catalysts

    The cutting-edge XAS techniques

    Conventional XAS has been beneficial in understanding the oxidation states and atomic structures for metal active sites in CPMs, however its resolution (especially in XANES) is limited by core-hole lifetime broadening.35,36 The development of X-ray techniques such as HERFD mode, which reduces the effective core-hole lifetime, opens the door to disclosing the complex nature of metal species in CPMs.20,37 For instance, HERFD-XANES is a technique that can be used to investigate the pre-edge features of 3d transition metals, which correspond to the 1s to 3d electronic transitions and reflect the oxidation states and coordination geometries of these metals, which was done by Zhang et al.37 using Fe K-edge HERFD-XANES to identify the iron species in different Fe-containing zeolites (Figure 2a,b). By comparing the pre-edge peak information of [Fe]zeolite catalysts and reference samples in their HERFD-XANES spectra (Figure 2a), the iron species were shown to be isomorphously substituted into the zeolitic framework (FeCHA100 and FeMFI100), which were directly hydrothermally synthesized in the form of tetrahedral coordination (Fe3+ Td). In contrast, the octahedral coordinated (Fe3+ Oh) iron species, like Fe2O3 nanoparticles and Fe-oxo clusters, were mainly present in impregnated [Fe]zeolite catalysts (FeCHA100-imp and FeMFI100-imp). The HERFD-XANES spectroscopy results were combined with those from X-ray emission spectroscopy, which further confirmed the iron species in zeolite. Therefore, compared to the impregnation method, the hydrothermal synthesis was proven to be beneficial for the dispersion of Fe atoms in the zeolitic framework. Consequently, isomorphously substituted [Fe] zeolite catalysts can exhibit enhanced catalytic performance in the nonoxidative coupling of methane.

    Figure 2

    Figure 2 | (a) Pre-edge region of Fe K-edge HERFD-XANES, and (b) the related pre-edge peak analysis of various [Fe]zeolite catalysts. Reprinted with permission from ref 37. Copyright 2023 Wiley-VCH. (c) Dynamic changes in the average oxidation states of cobalt according to the edge position of Co K-edge XANES spectra of the reference Co@Y sample and Co@Y after the successive feeding of O2 and C3H6 at 500 °C. Reprinted with permission from ref 38. Copyright 2022 American Chemical Society. (d) LCF of the Zn K-edge spectra of Zn1AlOx/AlPO-18. Reprinted with permission from ref 39. Copyright 2023 American Chemical Society. 2D plots of WT-EXAFS for (e) CuO and (f) Cu-exchanged FAU zeolite. Reprinted with permission from ref 40. Copyright 2020 Royal Society of Chemistry.

    With the development of new analysis methods for XANES spectra, quantitative valence state analysis for metal centers in CPMs (cast into a single energy value) can be achieved by establishing a linear correlation between the edge energy and the valence state of reference samples.41 For example, Li et al.38 revealed that the average oxidation state was +2.11 for cobalt in Co@Y according to the Co K-edge absorption edge positions (Figure 2c). In addition, the valence state of Co was around +2 (+2.11, +2.21, +2.00) during redox cycles of propylene aerobic epoxidation, indicating its stability at high temperature (500 °C). Despite the valence state determination for metal-containing zeolite, such analysis methods have also been used to resolve the oxidation states of some metal active centers (e.g., Ni, Ru) in ZIF and MOF catalysts.42,43

    In addition to valence state analysis, LCF of XANES can be used to quantitatively distinguish metal species from a mixture by comparing the obtained spectra with well-known reference materials.21,39,44 An example of the use of LCF for XANES was shown by Su et al.,39 who determined the active zinc species in the AlPO-18 zeolite bifunctional catalyst during syngas conversion. Due to the existence of different Zn2+ species, a direct comparison of XANES spectra was unable to characterize the catalytic sites in the bifunctional catalyst. Upon comparison with reference spectra, it was found that the XANES curve of Zn1AlOx in the bifunctional catalyst exhibited characteristics of both ZnO and ZnAl2O4 phases. Therefore, LCF of the XANES was conducted to unveil the concentrations of ZnO (56 mol %) and the ZnAl2O4 spinel (44 mol %) in the bifunctional catalyst (Figure 2d), which were close to the theoretical value (50:50 in mol ratio) of the actual feeding. Based on this result, the ZnAl2O4 spinel was revealed as the active species in syngas conversion from the mixture.

    Other than XANES analysis, WT-EXAFS analysis can aid in the chemical and structural identification for metal active sites in CPMs.40,45,46 As WT has both radial distance and k-space resolution, which can detect backscattering in low k-regions by lighter elements (unlike traditional Fourier transform (FT)-EXAFS analysis), it can provide higher resolution for EXAFS (especially for directly distinguishing different types of neighbors).46 As an example, by comparing the Pt L3- and Ge K-edge 3D contour maps with the 2D projection of the Pt@Ge-UTL zeolite catalyst, Ma et al.45 confirmed the bonding environments of Ge-zeolite confined sub-nanometric Pt clusters. Moreover, by carefully studying the WT-EXAFS of a catalyst structure with known reference structures, better understanding of test samples can be obtained as done Sushkevich et al.40 for Cu-exchanged faujasite (FAU) zeolite. The WT-EXAFS spectrum of CuO (Figure 2e) helped distinguish the different chemical bonds in CuO, including Cu-Oshort, Cu-Olong, and Cu-Cuoxide. Based on these results, the WT-EXAFS spectrum of Cu-exchanged FAU zeolite (Figure 2f) was employed to reveal the bonding environment of the Cu-exchanged FAU zeolite, including Cu-O, Cu-Al-Cu (superposition), and Cu-Cuoxide.

    XAS-guided multitechniques

    Since particle sizes can be estimated from a first shell analysis (Figure 3a,b), XAS can be used for the determination of the sizes, shapes, and distributions of metal nanoparticles/nanoclusters, but the information obtained from XAS is averaged.47 More accurate structural results can be achieved by combining XAS with other advanced characterization techniques. Notably, STEM is regarded as a direct visualization tool, but only offers local information for the size/spatial distribution and shape determination of metal species in CPMs. Therefore, by combining the XAS and STEM, the size distribution of metal nanoparticles/nanoclusters in CPMs can be accurately determined. For example, Sun et al.48 proved that Pt@S-1-C (calcination in air then reduction under H2) with a CNPt-Pt near 11 contained large Pt nanoparticles, whereas Pt@S-1-H (ligand-protected direct reduction in H2) with a CNPt-Pt below 2 exhibited subnanometer Pt3 clusters, revealing the advantage of pure H2 reduction method in reducing the size of Pt species. Moreover, Cs-corrected STEM-high-angle annular dark-field (HAADF) images (Figure 3c,d) confirmed the Pt particle sizes of Pt@S-1-C (3.1 nm) and Pt@S-1-H (<1 nm) and that the Pt nanoparticles/nanoclusters were confined into zeolite channels. Similar studies into Mo3S4@ZSM-5, MOF-confined PdZn intermetallic nanoparticles, and zeolite-confined PtSn clusters have also been reported.5254

    Figure 3

    Figure 3 | (a) Comparison of first-shell CNs based on the results of conventional EXAFS analysis and theory-assisted XAS analysis for Pt nanoparticles on γ-Al2O3; (b) corresponding possible 3D models of Pt nanoparticles. Reprinted with permission from ref 47. Copyright 2017 American Chemical Society. Cs-corrected STEM-HAADF images of (c) Pt@S-1-C and (d) Pt@S-1-H. Reprinted with permission from ref 48. Copyright 2020 Wiley-VCH. (e) Fourier transform of k2-weighted EXAFS spectra of Rh foil and various zeolite-encaged Rh zeolite catalysts at Rh K-edge; (f) In-situ CO-DRIFTS of the Rh zeolite catalysts. Reprinted with permission from ref 49. Copyright 2019 Wiley-VCH. (g) EPR spectra and (h) XANES spectra of auto-reduced (black), CH4-reacted (green), and N2O-activated (pink) Cu-MOR. Reprinted with permission from ref 50. Copyright 2022 American Chemical Society. (i) Fe K-edge XANES spectra of Fe zeolites and reference samples; 57Fe Mössbauer spectra for (j) Fe-Z5-C, (k) Fe-Z5-TF, (l) Fe-HZ5-C, and (m) Fe-HZ5-TF. ISO FeFW, ESSR FeFW, and FeEFW refer to isolated framework Fe, framework Fe with the ESSR effect, and extra-framework Fe species, respectively. Reprinted with permission from ref 51. Copyright 2023 American Chemical Society.

    The interaction between metal species and CPMs permit single-atom/single-site metals to be confined or immobilized in CPMs, leading to significantly enhanced catalytic properties in a variety of reactions. The combined use of XAS and CO-DRIFTS (using CO as probe molecules) can aid the identification of isolated single atoms in CPMs.49,5557 For example, Sun et al.49 reported the structural characterization of Rh species in S-1 zeolite by using XAS and in-situ CO-DRIFTS. As shown in FT-EXAFS spectra (Figure 3e), it was found that there was a sole Rh-O bond in Rh@S-1-H with no peaks due to Rh-Rh or metal oxides, indicating the oxygen-stabilized single Rh atoms are formed in the zeolite framework. Moreover, as depicted in Figure 3f, two peaks at 2092 and 2023 cm−1 in the CO-DRIFTS of Rh@S-1-H were ascribed to the symmetrical and asymmetrical stretching of CO from the isolated mononuclear RhI(CO)2 species. In addition, the peak at 2048 cm−1 due to linear-adsorbed CO on Rh NPs is absent in Rh@S-1-H, further revealing the atomic dispersion of Rh species in Rh@S-1-H. Apart from zeolite-encaged metal single-atom catalysts, XAS and CO-DRIFTS coupling has been used to resolve the atomic structures for MOF-immobilized single-atom/site catalysts in multimetal systems.55,56

    Since XAS is element-specific characterization technique, the combined use of multielements’ XAS for one metal-in-CPM catalyst can provide more detailed structure information from different perspectives. For instance, Ma et al.57 revealed the unique Ni-S coordination of Ni1-S/MOF by jointly using conventional K-edge XAS (for Ni active sites) and soft K-edge XAS (for coordinated S atoms). In addition, Sun et al.48 extracted structure information for subnanometer PtZn alloy from Pt L3-edge and Zn K-edge XAS, confirming the configuration of bimetallic PtZn4 clusters in S-1 zeolite.

    In addition to atomic structure investigation, the joint use of XANES and EPR can accurately and quantitatively (in some cases) unveil the oxidation states for transition metal species in CPMs.50,58,59 For instance, Heyer et al.50 reported the use of XANES and EPR spectra for quantitatively analyzing the Cu species in mordenite (MOR) zeolite (Figure 3g,h). By conducting the spin quantification of EPR spectra and LCF of XANES spectra, changes in the Cu(I)/Cu(II) ratio during auto-reduced, CH4-reaction and N2O activation conditions were resolved, which allows for better understanding of the Cu speciation in MOR zeolite in the conversion of methane to methanol. Likewise, by combining XPS and XANES, valence state analysis for metal species in CPMs, such as investigating electron transferring phenomenon in multimetal-containing CPM systems, can be achieved.43,60,61

    Coupling XANES spectra with Mössbauer spectroscopy can provide more comprehensive and quantitative information of the chemical states, spin states, and coordination symmetries of specific elements such as Fe and Sn.51,62,63 One of this coupling was exhibited by Cheng et al.51 who studied the Fe species in zeolite socony mobil-5 (ZSM-5) zeolite. According to Figure 3i, the XANES spectra for Fe-Z5-TF and Fe-Z5-C revealed the appearance of a pre-edge peak at 7114 eV. Notably, this pre-edge peak was rarely observed in Fe3+ with centrosymmetric octahedral coordination (as in the Fe2O3 reference). That is to say, such a peak suggested the formation of Fe3+ within the tetrahedral zeolite framework and indicated a significant interaction between the 3d and 4p metal orbitals caused by electric-quadrupole coupling. Moreover, the intensity of the pre-edge peak detected in Fe-HZ5-C, which was assigned to the partial conversion of framework Fe to octa-coordinated extraframework Fe, was comparatively lower than that of Fe-Z5-C. In the XANES spectrum of Fe-HZ5-TF, the remarkably weak pre-edge peak indicated complete migration of framework Fe. Furthermore, more accurate analysis for Fe species in different states was investigated by 57Fe Mössbauer spectroscopy (Figure 3jm). Such a quantitative study disclosed that most of the Fe (∼93%) in Fe-Z5-C was enhanced spin–spin relaxation framework Fe (ESSR FeFW) whereas most of the Fe (∼65%) in Fe-Z5-TF was isolated framework Fe (ISO FeFW) (Figure 3j,k). In contrast, a doublet component was observed in the analysis of Fe-HZ5-C and Fe-HZ5-TF (Figure 3l,m), which was assigned to extraframework Fe3+ species (denoted as FeEFW). The results indicated that during the conversion from Fe-Z5-C to Fe-HZ5-C, approximately 48% of the framework Fe underwent transformation into extraframework Fe. Conversely, when transforming Fe-Z5-TF to Fe-HZ5-TF, nearly 84% of the framework Fe was converted. As a whole, the remarkable differences between the Fe-ZSM-5 samples synthesized with and without the organic template were identified.

    Theory-assisted XAS analysis

    Theory-guided XAS spectra simulation

    The well-known structure of CPMs make it suitable to employ computational methods in the analysis of XAS data, which can greatly extend the power for investigating the electronic properties and atomic structures of metal active sites in CPMs.6468 As an example, Signorile et al.67 constructed a series of Ti cluster structural models for titanium silicalite-1 (TS-1) zeolite using a computational approach (B3LYP-D2). Then, the simulated spectra were obtained via finite difference method (for XANES) and the density functional theory (DFT)-based method (for the pre-edge region). The structural differences between two different TS-1 samples were distinguished by comparing their measured spectra with the simulated XANES spectra of different cluster models in TS-1 zeolites while the configuration of Ti in perfect tetracoordinated TS-1 and defective hexacoordinated TS-1 was well-resolved. This study more accurately depicted the catalytic role of defects in TS-1 in partial oxidation reactions.

    Another work employing this approach was done by Lee et al.68 who used both XANES and EXAFS simulations to determine the structure of Cu species in MOR zeolite with assistance from DFT-based calculations. Figure 4a shows snapshots of the simulated structures. Figure 4b indicates that the calculated k2χ(k) EXAFS spectrum of the Cu2AlO3 species agreed with the experimental spectrum. Based on the Cu K-edge (Figure 4c) and Cu L3-edge (Figure 4d) XANES simulated spectra, the presence of a [Cu2AlO3]2+ cluster exchanged on MOR Al pairs, which showed the ability to oxidize up to two CH4 molecules per cluster at ambient pressure, was confirmed.

    Figure 4

    Figure 4 | (a) Annealed structure of Cu clusters; (b) compared k2-weighted EXAFS (k2χ(k)) from calculation (blue) and measurement (red) for Cu clusters; (c) Cu K-edge XANES and (d) Cu L3-edge XANES spectra of the Cu trimer and Cu2AlO3. (c) and (d) Compare theoretical spectra (blue) with experimental ones (red) along with red vertical guiding lines and energies, respectively. Reprinted with permission from ref 68. Copyright 2021 American Chemical Society.

    Theory-led extraction of structural parameters from EXAFS

    To gain an atomistic picture of metal-containing CPMs, an interpretation of the EXAFS data is essential. For the conventional analysis, the EXAFS data are typically compared to a model, and the suitability of the fit is evaluated using a goodness of fit parameter known as the R factor (a lower R factor indicates a better fit). The fitting process of the EXAFS data can often be challenging and demanding. Recently, applying DFT-computed structures as a beginning for EXAFS explanation has been proven to be superior to traditional analyses, particularly in a complicated multishell fitting process.6972 As an example, Zhang and coworkers70 used the DFT-optimized MOF structure as the fitting model to do multishell fitting for the MOF-808-Zn catalyst during the hydrogenation of CO2 to methanol (Figure 5ad). Fitting of the Zn K-edge EXAFS data of MOF-808-Zn-BR (as-synthesized, BR stands for before catalytic reaction) and MOF-808-Zn-AR (AR stands for after catalytic reaction) showed single-site Zn2+ centers instead of ZnO nanoparticles or clusters. Moreover, the multishell EXAFS fittings of MOF-808-BR and MOF-808-AR disclosed the existence of Zn-Zr bonds (3.41 and 3.01 Å, respectively).

    Figure 5

    Figure 5 | Local structures of (a) MOF-808-Zn-BR and (b) MOF-808-Zn-AR. R-space EXAFS fitting plot of (c) MOF-808-Zn-BR and (d) MOF-808-Zn-AR. The sample before catalysis was not exposed to air for the EXAFS analysis, so there is still an ethyl group on the Zn center. Reprinted with permission from ref 70. Copyright 2021 American Chemical Society. Local environment of platinum in Pt/ZSM-5: interpretation of EXAFS spectra. (e) R-factors representative of EXAFS fits of each of 301 DFT-optimized structures of 2Al-Pt and 1Al-PtOH in ZSM-5. Vertical white gridlines mark increments of 20 structures. (f) Dependence of relative energy on the zeolite Al-Al distance of 2Al-Pt structures with relative energies ≤2.6 eV. (g) Magnitudes (fit, blue; experiment, black) and imaginary parts (fit, green; experiment, gray) of Fourier-transformed EXAFS of Pt/ZSM-5 after exposure to O2 at 700 °C with fitting based on a 2Al-Pt structure. (h) Visual representation of the 2Al-Pt structure corresponding to fit shown in (g). Red, O; yellow, Si; pink, Al; gray, Pt; white, H. (i) Magnitudes (fit, blue; experiment, black) and imaginary parts (fit, green; experiment, gray) of Fourier-transformed EXAFS of Pt/ZSM-5 after exposure to O2 at 700 °C with fitting based on the lowest R-factor 1Al-PtOH structure. (j) Visual representation of the 1Al-PtOH structure from (i). Reprinted with permission from ref 28. Copyright 2022 American Chemical Society. (k) Illustration of the MFI intersection region and the confined PtxSnyOz clusters. (l) The minimum distance between metal atom and zeolite oxygen (Ozeo) for different PtxSnyOz clusters. (m) The probability of Pt facing the channels (P) and the Pt concentration (cPt) in the most stable PtSn alloy clusters. Reprinted with permission from ref 73. Copyright 2022 Springer Nature.

    Another example of DFT-guided multishell fitting for EXAFS spectra was done by Felvey et al.28 who used a newly developed analysis method called theory-guided QuantEXAFS to analyze the platinum atomic structure in ZSM-5 zeolite. According to the DFT calculation, a total of 301 unique geometrically arranged aluminum structures were optimized for 2Al-Pt and 1Al-PtOH frameworks. As shown in Figure 5e,f, such calculated structures were used to fit the experimental EXAFS data and the R factor (Figure 5e) was employed to quantify the accuracy of each fit. The 2Al-Pt structures were resolved from the 118 EXAFS fits with the lowest R factors. In Figure 5f, the zeolite framework Al-Al distance was plotted against the relative energies determined by DFT. The relative energies of all 2Al-Pt structures were ≤2.6 eV, with the most stable structures having Al-Al distances of approximately 5–6 Å. Based on these findings, it was determined that the most favorable structural model involves platinum being coordinated to four oxygen atoms of a six-membered ring (6-MR) which includes two next-next-nearest neighbor aluminum atoms (Figure 5h). Moreover, the theoretical FT-EXAFS spectrum of this structure closely matched the experimental spectrum characterizing calcined Pt/ZSM-5 (Figure 5g). Notably, this structure (Figure 5j) had the second lowest energy of all the 2Al-Pt structures found and included platinum at a 6-MR paired aluminum site. However, this structure was deemed improbable and subsequently rejected due to limited access to the site, which is obstructed by 5-MR openings that are not typically penetrated by platinum. Nevertheless, the area of the FT-EXAFS spectrum, which is indicative of Pt-O bonding, in the range of approximately 1–2 Å was well-matched with the experimental results (Figure 5i). In contrast, the region beyond 2 Å, which signifies the existence of Si, Al, and O atoms at higher distances from platinum, did not fit the experimental results.

    Furthermore, owing to the strength of ML-based atomic simulations, Ma and Liu73 scanned the thermodynamically stable MFI zeolite-confined PtSn alloy catalysts from millions of possible structures (Figure 5k) by carefully comparing the Pt-Sn CN and bond distance (Figure 5l,m) between the experimental EXAFS fitting results and the computational structures. They determined that the small PtSn alloy and Sn4O4 clusters were major components in MFI channels, which are beneficial for the pyruvate dehydrogenation reaction. Thus, stabilizing and increasing the concentration of these clusters should be a key goal.

    In-situ/operando XAS analysis

    Using in-situ/operando XAS techniques to determine the relationships between the dynamic electronic and atomic structure of the catalysts and their catalytic performance under actual reaction conditions is essential to gain a deeper understanding of reaction mechanisms and reveal the configuration of working catalytic sites in CPMs.14,7476 Recently, in-situ/operando XAS techniques have been broadly adopted in a variety of catalytic systems, such as thermocatalysis, photocatalysis, and electrocatalysis (Table 1).

    Table 1 | Selected In-Situ/Operando XAS Study for Metal-Containing CPMs Catalysts

    Catalysts Targeted Edge Application In-Situ/Operando Condition Main Findings Ref.
    Cu-zeolite (CHA) Cu K-edge SCR of NOx Temperature up to 400 °C;Atmosphere: O2/N2, H2/He Determine the correlations between the SCR performance and Cu-speciation behavior of the SO2-poisoned catalyst 77
    Cu-zeolite (CHA) Cu K-edge SCR of NOx Temperature at 200 °C;Atmosphere: O2/NO/NH3/CO2/H2O/N2 Verify the faster rates of CuI oxidation relative to CuII reduction at higher framework Al densities (>1 Al center per d6r) 29
    Ta-zeolite (MFI) Ta L3-edge Methanol-to-olefin Temperature up to 370 °C;Atmosphere: CH3OH/N2;Quick XAS: 4s/spectrum Reveal the high stability of the framework Ta(V) sites under reaction condition 32
    Ni-zeolite (BEA) Ni K-edge Alkane dehydrogenation Temperature at 300 °C;Atmosphere: O2/He, iC4H10/He, H2 Confirm the dehydrogenation active sites are isolated, four-coordinate Ni(2+) in beta zeolite 78
    Fe-MIL-100 Fe K-edge Intermediate study Temperature at 250 °C; Atmosphere: N2O/CO, N2O/O2 Prove the active oxygen intermediate as an iron-oxyl (Fe3+–O) during NO2-mediated CO oxidation reaction 79
    Cu-zeolite (MOR) Cu K-edge Methane oxidation to methanol Temperature at 210 °C;Atmosphere: O2, CH4;Quick XAS: 1s/spectrum Mono-μ-oxo dicopper and Cu(OH)+ monomers with extraframework oxygen in Cu-MOR zeolite are active sites in methane oxidation reaction 80
    Pt-zeolite (MFI) Pt L3-edge Intermediate study Temperature up to 700 °C;Atmosphere: O2/He, H2/He, CO/He Upon exposure of platinum clusters to O2 at 700 °C, oxidative fragmentation occurs, and Pt2+ ions are stabilized at 6-MRs in the zeolite that contain paired aluminum sites. When exposed to CO under mild conditions, these Pt2+ ions form highly uniform platinum gem-dicarbonyls, which can be converted in H2 to Ptδ+ monocarbonyls 28
    TS-1 (MFI) Ti K-edge Intermediate study The addition of H2O2 aqueous Six-coordinate Ti species is the active site for the vinyl-to-carbonyl reaction pathway 81
    Ru-zeolite (BEA) Ru K-edge Amination of fatty alcohol Temperature up to 450 °C;Atmosphere: O2/He, H2/He, NH3/He, 1-hexanol vapor The N-containing species are present at the surface of Ru nanoparticles in beta zeolite 26
    Pt-NU-1000 Pt L3-edge Ethylene hydrogenation Temperature at 150 °C;Atmosphere: H2, H2/He, H2/C2H6 The Pt nanoclusters in NU-1000 are sinter-resistance under the strongly exothermic hydrogenation reaction of ethylene 82
    Cu-UiO-66 Cu K-edge CO oxidation Temperature up to 250 °C;Atmosphere: CO/N2, H2/N2, CO/O2/N2 The mean Cu oxidation state (Cu2+–Cu+) can vary significantly with the reaction temperature (120–250 °C) 23
    CoNi-MOF-74 Co K-edgeNi K-edge Electrocatalytic oxygen evolution XAS signal is recorded at different potentials (0–1.5 V) in 1 M KOH Ni0.5Co0.5OOH0.75 with abundant oxygen vacancies and high oxidation states forms in NiCo-MOF-74 during OER 30
    LaNi-COF-5 Ni K-edgeLa L3-edge Photocatalytic CO2 reduction XAS signal is recorded at Ar or CO2-saturated aqueous solution with xenon lamp Ni atoms are the active centers for the CO2 reduction, while La atoms are not only the optically active center but also the catalytically active center for CO2 adsorption and activation 31

    In-situ/operando XAS analysis in thermocatalysis

    One of the most reported applications for in-situ/operando XAS (Figure 6a: the in-situ cell) for metal-in-CPM catalysts is studying under thermocatalysis conditions.83,84 As an example, Qi et al.84 used in-situ XANES spectra to monitor the evolution of Pt species in dealuminated beta zeolite (DeAlBEA) under thermal conditions. As shown in Figure 6b, the Pt L3-edge XANES spectra of the 0.04Pt0.36Zn-DeAlBEA catalyst were measured during propane dehydrogenation pretreatment process, with the temperature gradually increasing from 25 °C to 550 °C in a flowing He environment. In the spectra, the presence of three isosbestic points was noted, signifying that there was no accumulation of intermediates throughout the treatment process. This phenomenon would not be anticipated if the initially isolated Pt atoms were converted into Pt and/or PtZn alloy nanoparticles, which could have a variety of compositions and sizes. This further implies that the Pt atoms were in an isolated state both prior to and following the pretreatment.

    Figure 6

    Figure 6 | (a) Schematic view of the heatable/coolable in situ fluorescence/transmission EXAFS cell. Reprinted with permission from ref 83. Copyright 2007 International Union of Crystallography. (b) Isosbestic points of Pt L3-edge XANES spectra of 0.04Pt0.36Zn-DeAlBEA recorded during treatment in flowing He as temperature was ramped from 25 °C to 550 °C. Reprinted with permission from ref 84. Copyright 2021 American Chemical Society. (c) In-situ Pd K-edge XANES spectra and (d) corresponding Pd K-edge ΔXANES spectra during the reaction of Pd+CHA under 4% NO/He flow at 600 °C and O2-treated Pdimp/CHA(H2_NO) under He flow at 100 °C. The data for the ΔXANES spectra were obtained by subtraction with the spectrum taken at 0 min. Reprinted with permission from ref 85. Copyright 2021 American Chemical Society. (e) Theoretical “pure” Cu K-edge HERFD-XANES spectra of PC1, PC2, PC3, PC4, and PC5 from MCR analysis. (f) Corresponding temperature-dependent concentration profiles of 0.18Cu-HMOR(7)-He-act., 0.18Cu-HMOR(7)-O2-act., 0.36Cu-HMOR(11)-He-act., and 0.36Cu-HMOR(11)-O2-act. Reprinted with permission from ref 86. Copyright 2018 American Chemical Society. Time evolution of Cu-O (g) CN and (h) bond distance based on the EXAFS data recorded during CO oxidation (1% CO, 1% O2, N2 balance-30 NmL min−1) at 250 °C after the H250 pretreatment. Reprinted with permission from ref 87. Copyright 2019 American Chemical Society.

    Another technique used to clearly investigate changes of in-situ XANES spectra is the ΔXANES analysis.26,67,85 For example, Yasumura et al.85 studied Pd species evolution in chabazite (CHA) zeolite (Pd+CHA) by the in-situ ΔXANES analysis. In order to shed light on the structural changes of Pd species during atomic dispersion, in-situ XAS measurements were conducted on Pd+CHA during the nitrogen oxide (NO) treatment process in 4% NO/He atmosphere at 600 °C. As evidenced by the Pd XANES spectra acquired at t = 0, 2, 20, 60, 120, and 145 min (Figure 6c), the changes were distinctly noticeable in the Pd K-edge ΔXANES spectra (Figure 6d). According to this spectra difference analysis, the intensity of the peak originating from Pd0 species sharply declined, concurrently with the initial emergence of a Pd2+ peak (t = 0–20 min) that consistently diminished until t = 120 min. These findings corroborated the atomic dispersion of Pd black through the oxidation of Pd0 species to Pd2+.

    As previously discussed, it is feasible to separate different contributions using XANES LCF analysis. However, the XANES study of some Cu-containing CPMs presents a unique and often hard-to-replicate coordination environment (e.g. cluster geometry and oxidation state). This complexity makes the LCF analysis of synthetic model compounds challenging. To overcome this problem, the MCR-ALS algorithm is frequently employed. This approach involves the iterative refinement of concentration profiles and pure spectra through an alternating least squares routine, all the while maintaining specific constraints.88 By using MCR-ALS analysis, Pappas et al.86 compared HERFD-XANES spectra of two most different Cu-MOR materials (Cu/Al = 0.18 and Si/Al = 7; Cu/Al = 0.36 and Si/Al = 11) to quantitatively resolve the Cu speciation (active/inactive species) during the conversion of methane to methanol in the Cu K-edge XANES region (Figure 6e,f). Specifically, the activation in helium notably encouraged the creation of CuI via auto-reduction. This could potentially be due to the carbonaceous residues left within the materials after synthesis. During aerobic activation, the auto-reduction process that produced CuI was practically nonexistent at higher temperatures and is only briefly noticeable at lower temperatures. Moreover, varying activation conditions shifted the equilibrium between two types of CuII species. These species originated from a four-coordinate CuII intermediate species (PC4), and they exhibited markedly different tendencies to resist auto-reduction (as indicated in Figure 6e,f as PC3 and PC5). When the accumulated evidence was linked with reactivity studies, it suggested that PC5 was the CuII species that showed the highest resistance to auto-reduction during anaerobic activation. Therefore, the PC5 Cu species drove the activity of Cu/MOR in the selective oxidation of methane to methanol.

    Recently, Krishna et al.29 conducted another operando XAS study on Cu-containing zeolite catalysts (Table 1). They presented the steady-state selective catalytic reduction (SCR) mechanistic explanations for Cu-CHA zeolites with different compositions at 1, 10, and 60 kPa O2. The XAS data offered conclusive proof that the oxidation rates of Cu(I) were quicker compared to the reduction rates of Cu(II) at elevated framework Al densities (>1 Al center per d6r).

    Other than in-situ XANES analysis, in-situ EXAFS analysis can be conducted as shown by Abdel-Mageed et al. for the Cu/UiO-66 catalyst.23,87 The EXAFS data, measured over a period of 375 min, revealed minimal alterations in both the Cu-O CN (Figure 6g) and the bond distance (Figure 6h) of the Cu species during CO oxidation, after just a few minutes into the reaction. Also, these results suggested high degree of stability in coordination environment surrounding the Cu atoms during CO oxidation. Altogether, these findings emphasized the existence of individual Cu atoms, each coordinated to two oxygen atoms, in an atomically dispersed state within the defect site of UiO-66. This arrangement was distinct from those found in Cu2O or CuO clusters/nanoparticles.

    Operando EXAFS can also be analyzed in WT; as an example, Negri et al.89 utilized in-situ Cu K-edge WT-EXAFS for the analysis of Cu species in NH3-mediated SCR (NH3-SCR) of NOx. The operando WT-EXAFS analysis provided a more precise and in-depth differentiation of the scattering contributions surrounding the absorber atom when compared to traditional FT-EXAFS analysis. Through a comprehensive investigation of operando WT-EXAFS spectra, the species evolution of solvated Cu-pairs in Cu-CHA catalysts during the NH3-SCR reaction cycle can be obtained.

    In-situ/operando XAS analysis in electro and photo-catalysis

    Recently, with the rapid development of metal-containing MOF electrocatalysts, in-situ/operando XAS characterization (the in-situ cell showed in Figure 7a) in electrocatalysis has become increasingly important.29,32,90 An example is done by Zhao et al.30 who used operando cell for XAS measurement under electrocatalysis condition (Table 1). The variations in the electronic and atomic structures of the metal sites inside NiCo-MOF-74 throughout the entire electrocatalytic cycle were monitored. As a result, the high oxygen evolution reaction (OER) activity observed can be attributed to the in-situ formation of Ni0.5Co0.5OOH0.75 with abundant oxygen vacancies and elevated oxidation states. Likewise, another in-situ XAS study for metal-containing MOFs in electrocatalysis was reported by Xu et al.91 During the electrocatalysis process (potential drop from 0.3 to −0.9 V), the catalytic center for the nitrate-to-ammonia reduction reaction is formed by the in situ clustering of single-atom Cu, resulting in MOF-confined ultrasmall Cu nanoclusters, as depicted in in-situ XANES and FT-EXAFS spectra (Figure 7b,c).

    Figure 7

    Figure 7 | (a) In situ XAS setup for electrocatalysis, where CE, RE, and WE stand for counter, reference, and working electrodes, respectively. Reprinted with permission from ref 90. Copyright 2013 American Chemical Society. In-situ (b) XANES and (c) FT-EXAFS at the Cu K-edge for Cu-MOF during the electrocatalytic nitrate-to-ammonia reduction reaction. Reprinted with permission from ref 91. Copyright 2022 American Chemical Society. (d) Schematic of the photocatalytic cell for in-situ XAS study. Reprinted with permission from ref 92. Copyright 2023 Multidisciplinary Digital Publishing Institute. (e) Ni K-edge XANES spectra of LaNi-Phen/COF-5 in CO2 photoreduction reduction reaction at room temperature and 1 atm of Ar or CO2-saturated aqueous solution (inset: the enlarged Ni K-edge XANES spectra). Reprinted with permission from ref 31. Copyright 2023 Springer Nature. (f) FT-EXAFS spectra of Co-Ru-UiO-67(bpy) before illumination and the intermediate species formed after induction period ends. Reprinted with permission from ref 93. Copyright 2018 American Chemical Society.

    As with electrocatalysis conditions, in-situ/operando XAS measurements can also be conducted for the characterization of metal-in-MOFs under photocatalysis conditions (the in-situ cell showed in Figure 7d).92 For instance, Zhou et al.31 reported the use of Ni K-edge and La L3-edge operando XAS measurements under photocatalytic CO2 reduction conditions to dynamically monitor the oxidation state and gain insights into the atomic-scale functionality of active sites for LaNi-Phen/COF-5 (Figure 7e). According to the Ni K-edge XANES spectra of LaNi-Phen/COF-5, the white-line intensity was slightly higher in a CO2-saturated CH3CN solution than that in an Ar-saturated CH3CN solution, implying the increase of valence state of Ni, which can be attributed to the transfer of electrons from the active Ni center to CO2. In contrast, based on the change of XANES spectra under xenon lamp irradiation, the active Ni center’s oxidation state was gradual recovery during CO2 reduction. Additionally, the alteration in oxidation state was also noticed in the La L3-edge XANES spectra. To summarize, the in-situ XAFS analysis revealed that Ni atoms played a crucial role as active centers in the CO2 photoreduction, while La atoms served as both the optically active center and the catalytically active center in the process of CO2 adsorption and activation. In addition to in-situ XANES analysis for oxidation state, the in-situ EXAFS study under photocatalysis condition for Co-Ru-UiO-67(bpy) was done by Yang et al.93 As shown in Figure 7f, by quantitatively analyzing the coordination environment of a long-lived CoI intermediate, they identified the rate-limiting step in the light-driven H2 production, shedding light on the catalysis mechanism of these single-site Co-Ru-MOF photocatalysts.

    Other advanced in-situ/operando XAS techniques

    Typically, it takes several minutes or more to acquire a complete scan for in-situ XAS spectrum in conventional beamline, which poses challenges for time-resolve XAS studies on metal-in-CPM catalysts. To overcome such difficulty, by continuously moving the monochromator (which is conventionally moved stepwise) for swift XAS measurements, high quality XAS data can be collected in several tenths of a second in some advanced beamlines.25,32,37,94

    An example employing these measurements was shown by Petrov et al.94 in assessing the redox properties of palladium in MOR zeolites (Pd/Na-MOR and Pd/H-MOR) of in a quick XAS experiment at the time resolution of 1 s (Figure 8ad, respectively, 90 min at 400 °C). O2 was removed from the feed to reduce the catalyst using CH4 and then reintroduced to allow palladium to be oxidized again under reaction conditions. The XANES LCF analysis confirmed that both Pd/H-MOR and Pd/Na-MOR exhibited similar evolution behavior of Pd species during reduction and reoxidation. However, the fraction of Pd2+ in the steady state under reaction conditions was considerably higher in Pd/Na-MOR. The observed phenomenon can be attributed to the accelerated rate of reoxidation from Pd0 to Pd2+ in the presence of sodium, which was four times faster (as evidenced by the initial rates in Figure 8a,b). This can be explained by the smaller size of the metal nanoparticles in Pd/Na-MOR (as indicated by the smaller CNPd–Pd in Figure 8c,d) and their larger specific surface areas, which generally enhanced reactivity towards oxygen.

    Figure 8

    Figure 8 | Transient operando EXAFS experiment with oxygen cut-off. (a) and (b) Fraction of oxidized palladium in Pd/Na-MOR and Pd/H-MOR and corresponding catalytic activity in 1 vol % CH4, 4 vol % O2, 5 vol % H2O, bal. N2, gas hourly space velocity (GHSV) = 350,000 h−1 at 350 °C. Oxygen was removed from the feed to perform reduction of palladium by methane and then added for subsequent reoxidation in reaction conditions. (c) and (d) Fourier transforms of in situ Pd K-edge EXAFS spectra (nonphase shift corrected) of Pd/Na-MOR and Pd/H-MOR upon averaging 15 spectra at 1, 90, 210, 280, and 550 s of the experiment. Initial rates were obtained from fitting the linear function over the first 10 points during reoxidation. Reprinted with permission from ref 94. Copyright 2018 Springer Nature. (e) Design of the in-situ XAS experiment; (f) The combined in-situ XANES-MS experimental setup at beamline P65, Petra III. Reprinted with permission from ref 33. Copyright 2023 American Chemical Society. (g) Molecular models for the structural components included in the ML-assisted EXAFS fitting model: 1 [CuI(NH3)2]+, 2 “planar”, and 3 “bent” motifs for μ-η2,η2-peroxo diamino dicopper(II); when relevant, characteristic EXAFS-derived ranges for Cu-Cu interatomic distances are reported, in Å. (h–j) Comparison between magnitudes of experimental (colored circles) and best fit (thick lines) in-situ FT-EXAFS spectra at the oxidation condition for (h) 0.1_5, (i) 0.5_15, and (j) 0.6_29 Cu-CHA zeolites. The scaled components for Cu-species 1, 2/2′, 3 are also reported, vertically translated for the sake of clarity, together with percentages of each component over total Cu refined by ML-EXAFS fitting. Reprinted with permission from ref 27. Copyright 2022 American Chemical Society.

    The combined use of XAS and other characterization techniques has also been conducted under in-situ conditions. For example, by a combination of Sr K-edge XANES-mass spectroscopy (MS), Liutkova et al.33 investigated the evolution of Sr species in ZSM-5 zeolite under the methanol to hydrocarbon (MTH) reaction conditions (Figure 8e,f). Firstly, they confirmed the high conversion of methanol (indicated by a low m/z = 31 signal) and the formation of propylene (highlighted by an intense m/z = 41 signal), which demonstrated the catalyst’s activity in the MTH reaction and also indicated the development of a functional hydrocarbon pool within the pores of the zeolite. In addition, they carefully analyzed the simultaneous recorded operando XAS data. The operando ΔXANES analysis revealed the interaction between Sr cations in the ZSM-5 catalyst and adsorbates. Despite a decrease in coverage at higher temperatures, there were still residual adsorbates present at 450 °C. Moreover, the operando EXAFS analysis revealed that the Sr CN elevated in the presence of adsorbates and stayed high when the feed was switched to He, suggesting that the Sr moieties can possess a strong ability to retain adsorbates and thereby influenced the local environment under MTH reaction conditions.

    With the advancement of artificial intelligence technology, ML was also applied to assist the in-situ XAS analysis, which in theory enables the refinement of any structure beginning from an initially estimated molecular complex, even in cases where there are notable disparities in interatomic distances and angles. An example of this approach was performed by Martini et al.27 to understand the emerging structural Cu complex in CHA zeolite (Figure 8gj) under SCR conditions. First, by comparing the in-situ XANES features, they identified the formation of [Cu2(NH3)4O2]2+ motif in the Al rich 0.1_5 catalyst, which differs from those observed for the 0.5_15 and 0.6_29 catalysts. Additionally, in the high-k range characteristic of Cu–Cu scattering (WT maps), two local maxima were observed along the R direction at approximately 2.5 Å (low-R) and 3.2 Å (high-R), pointing to an EXAFS resolvable bimodal distribution of Cu–Cu interatomic distances. However, the conventional refinement method (solely based on one DFT-optimized model) is insufficient to extract the real structure information from these in-situ EXAFS data. Because ML-assisted fitting can start from models with significantly different interatomic distances and angles, such an approach was employed to resolve in-situ EXAFS data and accurately analyze the subtle differences observed in the XANES spectral shape and the high-R portion of WT-EXAFS. As a result, the low-R/high-R features were well-matched with the optimized Cu–Cu distances corresponding to the planar and bent μ-η2,η2-peroxo diamino dicopper(II) structures. Hence, the potential of in-situ XAS combined with integrated ML-assisted analysis should be highlighted.

    A newly discovered XAS-related characterization method, X-ray adsorbate quantification (XAQ), have been used to generate thermogravimetric analysis-, temperature-programmed desorption-, and temperature-programmed reduction-like information from raw XAS data under in-situ conditions. The XAQ data can be derived from transmission XAS spectra collected simultaneously, without requiring any adjustments of the experimental procedure or additional equipment or software. Moreover, combining it with XANES/EXAFS does not result in an increase in the measurement time because the data in the XAQ region is usually collected as a necessary step for normalizing XANES and EXAFS spectra. Such an example was reported by Lomachenko et al. for quantitatively investigating the absorption behavior of Cu-CHA zeolite under in-situ conditions.95,96

    Summary and Outlook

    This minireview summarizes the most recent progresses in advanced XAS studies for metal-containing CPMs in heterogenous catalysis. Depending on the high sensitivity in electronic behavior and atomic resolution in local structure required, different catalysis model systems can be investigated by such element-selective characterization techniques. Pre-edge fingerprints and XANES region have been used for the qualitative and even quantitative identification of different metal species in CPMs. EXAFS analysis has shown strength in resolving the atomic structure for metal active sites in CPMs. Furthermore, with the help of other advanced characterization techniques and theories, the power of XAS has been significantly enhanced. In particular, in-situ/operando XAS has been highlighted for the structure-performance relationship studies of metal-in-CPM due to the unparalleled advantage in determining working sites in CPMs under real catalytic conditions.

    Although there have been many breakthroughs in the investigation of metal-containing CPMs in heterogeneous catalysis, ongoing challenges with advanced X-ray spectroscopic approaches remain. The quality of XAS data is greatly influenced by sample preparation and the synchrotron light source’s beamline, which can introduce data artifacts like self-absorption and pinhole effects, and so on. Therefore, X-ray sources with higher flux, stronger coherence, and better time resolution need to be developed. Also, combining XAS with other advanced characterization techniques adopts time-consuming alternating signal collection, which make them inappropriate for time-resolved studies. Thus, the development of simultaneously measured XAS-guided multitechniques is needed to more precisely unveil the catalytic sites in CPMs. Moreover, researchers who possess moderate XAS training can only achieve fundamental details regarding catalyst oxidation states and its overall structure from the absorption edge and the main peaks in the FT-EXAFS spectra. However, such qualitative information is sometimes insufficient for obtaining crucial insights into the electronic and atomic structures of metal active sites in CPMs. Fitting of the FT-EXAFS data is usually complicated and time-consuming, and it is easy to get erroneous results when the fitting method or chosen model is not reliable. Consequently, the availability of ab initio codes and user-friendly data analysis packages must be developed to easily achieve fully resolved and accurate results for the in-depth understanding of metal-sites in CPMs. Additionally, in-situ/operando XAS studies have been demonstrated as indispensable tools to analyze the dynamic electronic behavior and local structural changes of metal active sites in CPMs, which have revealed the underlying reaction mechanisms and the structure-performance relationships of these catalysts. To guarantee signal collection efficiency and high signal-to-noise ratio in the collected data, the in-situ cells need to be rationally designed to closely mimic the true reaction environment. Also, advanced in-situ/operando XAS studies with higher spatial, energy, and time resolution can produce more useful information than currently available. As such, it is important to give special attention to the state-of-the-art analysis technologies such as artificial intelligence, machine learning, and so on.

    Overall, the vast quantity of information extracted from XAS data concerning the electronic/atomic structure, composition, and dynamics under reaction conditions makes it well-suited for in-depth analysis of the metal active sites in CPMs. In the near future, we foresee the following advances in the use of XAS in heterogeneous catalysis with metal-containing CPMs: (1) With HERFD-XANES and EXAFS becoming more readily available, deeper insights on oxidation state and coordination geometry of metal active sites will be discovered; (2) Thanks to the newly established X-ray free electron laser source with ultra high brightness and extreme short duration, XAS studies will be greatly improved, especially in time-resolves experiments; (3) In-situ/operando XAS studies will rapidly increase. In the meantime, in-situ/operando equipment will become more versatile and efficient; (4) With the development of theoretical codes, the strength of XANSES simulation will become increasingly important. It can be used in conjunction with EXAFS refinement to resolve the real atomic structure of metal active sites in CPMs. While many opportunities and challenges arising from recent advanced XAS studies in metal-containing CPMs remain, XAS has been demonstrated to be a major tool for the characterization of metal-in-CPM catalysts and should be expected to play an increasingly crucial role in advancing characterization techniques in this field.

    Conflict of Interest

    The authors declare no competing financial interests.

    Acknowledgments

    T.Z. acknowledges the funding from the National Natural Science Foundation of China (grant no. 22301057), the financial support by the Natural Science Foundation of Hebei Province (grant no. B2023201065), and Hebei University High-level Talent Research Program (grant no. 521100223025). Y.L. thanks the funding from the National Natural Science Foundation of China (grant no. 22305060), and Hebei University High-level Talent Research Program (grant no. 521100222060). P.Z. acknowledges the financial support from an Natural Sciences and Engineering Research Council of Canada (NSERC) Discovery Grant. A.G.W. thanks financial support from an NSERC Canada Graduate Scholarships - Doctoral Program (CGS-D) scholarship.

    References

    • 1. Li Y.; Yu J.Emerging Applications of Zeolites in Catalysis, Separation and Host–Guest Assembly.Nat. Rev. Mater.2021, 6, 1156–1174. CrossrefGoogle Scholar
    • 2. Wei Y.; Zhang M.; Zou R.; Xu Q.Metal–Organic Framework-Based Catalysts with Single Metal Sites.Chem. Rev.2020, 120, 12089–12174. Google Scholar
    • 3. Huang S.; Chen K.; Li T.Porphyrin and Phthalocyanine Based Covalent Organic Frameworks for Electrocatalysis.Coord. Chem. Rev.2022, 464, 214563. CrossrefGoogle Scholar
    • 4. Zhang Q.; Gao S.; Yu J.Metal Sites in Zeolites: Synthesis, Characterization, and Catalysis.Chem. Rev.2023, 123, 6039–6106. CrossrefGoogle Scholar
    • 5. Liu L.; Corma A.Confining Isolated Atoms and Clusters in Crystalline Porous Materials for Catalysis.Nat. Rev. Mater.2021, 6, 244–263. Google Scholar
    • 6. Zhang T.; Chen Z.; Walsh A. G.; Li Y.; Zhang P.Single-Atom Catalysts Supported by Crystalline Porous Materials: Views from the Inside.Adv. Mater.2020, 32, 2002910. Google Scholar
    • 7. Wen L.; Sun K.; Liu X.; Yang W.; Li L.; Jiang H.Electronic State and Microenvironment Modulation of Metal Nanoparticles Stabilized by MOFs for Boosting Electrocatalytic Nitrogen Reduction.Adv. Mater.2023, 35, 2210669. Google Scholar
    • 8. Dai S.; Kajiwara T.; Ikeda M.; Romero-Muñiz I.; Patriarche G.; Platero-Prats A. E.; Vimont A.; Daturi M.; Tissot A.; Xu Q.; Serre C.Ultrasmall Copper Nanoclusters in Zirconium Metal-Organic Frameworks for the Photoreduction of CO2.Angew. Chem. Int. Ed.2022, 61, 202211848. Google Scholar
    • 9. Liu Y.; Li Z.; Yu Q.; Chen Y.; Chai Z.; Zhao G.; Liu S.; Weng-Chon C.; Pan Y.; Zhang Q.; Gu L.; Zheng L.; Wang Y.; Lu Y.; Wang D.; Chen C.; Peng Q.; Liu Y.; Liu L.; Chen J.; Li Y.A General Strategy for Fabricating Isolated Single Metal Atomic Site Catalysts in Y Zeolite.J. Am. Chem. Soc.2019, 141, 9305–9311. Google Scholar
    • 10. Deng X.; Bai R.; Chai Y.; Hu Z.; Guan N.; Li L.Homogeneous-Like Alkyne Selective Hydrogenation Catalyzed by Cationic Nickel Confined in Zeolite.CCS Chem.2022, 4, 949–962. LinkGoogle Scholar
    • 11. Xu Y.; Hu W.; Li Y.; Su H.; Liang W.; Liu B.; Gong J.; Liu Z.; Liu X.Manipulating the Cobalt Species States to Break the Conversion–Selectivity Trade-Off Relationship for Stable Ethane Dehydrogenation over Ligand-Free-Synthesized Co@MFI Catalysts.ACS Catal.2023, 13, 1830–1847. Google Scholar
    • 12. Jiao F.; Bai B.; Li G.; Pan X.; Ye Y.; Qu S.; Xu C.; Xiao J.; Jia Z.; Liu W.; Peng T.; Ding Y.; Liu C.; Li J.; Bao X.Disentangling the Activity-Selectivity Trade-Off in Catalytic Conversion of Syngas to Light Olefins.Science2023, 380, 727–730. Google Scholar
    • 13. Juneau M.; Liu R.; Peng Y.; Malge A.; Ma Z.; Porosoff M. D.Characterization of Metal-Zeolite Composite Catalysts: Determining the Environment of the Active Phase.ChemCatChem2020, 12, 1826–1852. Google Scholar
    • 14. Sun K.; Qian Y.; Jiang H.Metal-Organic Frameworks for Photocatalytic Water Splitting and CO2 Reduction.Angew. Chem. Int. Ed.2023, 62, 202217565. Google Scholar
    • 15. Li W.; Chai Y.; Wu G.; Li L.Stable and Uniform Extraframework Cations in Faujasite Zeolites.J. Phys. Chem. Lett.2022, 13, 11419–11429. CrossrefGoogle Scholar
    • 16. Rhoda H. M.; Heyer A. J.; Snyder B. E. R.; Plessers D.; Bols M. L.; Schoonheydt R. A.; Sels B. F.; Solomon E. I.Second-Sphere Lattice Effects in Copper and Iron Zeolite Catalysis.Chem. Rev.2022, 122, 12207–12243. Google Scholar
    • 17. Timoshenko J.; Roldan Cuenya B.In Situ/Operando Electrocatalyst Characterization by X-Ray Absorption Spectroscopy.Chem. Rev.2021, 121, 882–961. CrossrefGoogle Scholar
    • 18. van Oversteeg C. H. M.; Doan H. Q.; de Groot F. M. F.; Cuk T.In Situ X-Ray Absorption Spectroscopy of Transition Metal Based Water Oxidation Catalysts.Chem. Soc. Rev.2017, 46, 102. Google Scholar
    • 19. Zhang Z.; Zhang P.; Andrew G.Dynamic Structure of Metal Nanoclusters from Synchrotron X-Ray Spectroscopy.J. Phys. Chem. C2021, 125, 5982–5994. Google Scholar
    • 20. Li L.; Chalmers J. A.; Bare S. R.; Scott S. L.; Vila F. D.Rigorous Oxidation State Assignments for Supported Ga-Containing Catalysts Using Theory-Informed X-Ray Absorption Spectroscopy Signatures from Well-Defined Ga(I) and Ga(III) Compounds.ACS Catal.2023, 13, 6549–6561. Google Scholar
    • 21. Deplano G.; Martini A.; Signorile M.; Borfecchia E.; Crocellà V.; Svelle S.; Bordiga S.Copper Pairing in the Mordenite Framework as a Function of the CuI/CuII Speciation.Angew. Chem. Int. Ed.2021, 60, 25891–25896. Google Scholar
    • 22. Mesilov V.; Dahlin S.; Bergman S. L.; Xi S.; Han J.; Olsson L.; Pettersson L. J.; Bernasek S. L.Regeneration of Sulfur-Poisoned Cu-SSZ-13 Catalysts: Copper Speciation and Catalytic Performance Evaluation.Appl. Catal. B-Environ.2021, 299, 120626. Google Scholar
    • 23. Abdel-Mageed A. M.; Rungtaweevoranit B.; Impeng S.; Bansmann J.; Rabeah J.; Chen S.; Häring T.; Namuangrak S.; Faungnawakij K.; Brückner A.; Behm R. J.Unveiling the CO Oxidation Mechanism over a Molecularly Defined Copper Single-Atom Catalyst Supported on a Metal-Organic Framework.Angew. Chem. Int. Ed.2023, 62, 202301920. Google Scholar
    • 24. Yan P.; Xi S.; Peng H.; Mitchell D. R. G.; Harvey L.; Drewery M.; Kennedy E. M.; Zhu Z.; Sankar G.; Stockenhuber M.Facile and Eco-Friendly Approach to Produce Confined Metal Cluster Catalysts.J. Am. Chem. Soc.2023, 145, 9718–9728. Google Scholar
    • 25. Kornas A.; Tabor E.; Wierzbicki D. K.; Olszowka J. E.; Pilar R.; Dedecek J.; Sliwa M.; Jirglova H.; Sklenak S.; Rutkowska-Zbik D.; Mlekodaj K.Activation of Molecular Oxygen over Binuclear Iron Centers in Al-Rich BEA Zeolite.Appl. Catal. B-Environ.2023, 336, 122915. Google Scholar
    • 26. Fang L.; Yan Z.; Wu J.; Bugaev A.; Lamberti C.; Pera-Titus M.Highly Selective Ru/HBEA Catalyst for the Direct Amination of Fatty Alcohols with Ammonia.Appl. Catal. B-Environ.2021, 286, 119942. Google Scholar
    • 27. Martini A.; Negri C.; Bugarin L.; Deplano G.; Abasabadi R. K.; Lomachenko K. A.; Janssens T. V. W.; Bordiga S.; Berlier G.; Borfecchia E.Assessing the Influence of Zeolite Composition on Oxygen-Bridged Diamino Dicopper(II) Complexes in Cu-CHA DeNOx Catalysts by Machine Learning-Assisted X-ray Absorption Spectroscopy.J. Phys. Chem. Lett.2022, 13, 6164–6170. Google Scholar
    • 28. Felvey N.; Guo J.; Rana R.; Xu L.; Bare S. R.; Gates B. C.; Katz A.; Kulkarni A. R.; Runnebaum R. C.; Kronawitter C. X.Interconversion of Atomically Dispersed Platinum Cations and Platinum Clusters in Zeolite ZSM-5 and Formation of Platinum Gem-Dicarbonyls.J. Am. Chem. Soc.2022, 144, 13874–13887. Google Scholar
    • 29. Krishna S. H.; Goswami A.; Wang Y.; Jones C. B.; Dean D. P.; Miller J. T.; Schneider W. F.; Gounder R.Influence of Framework Al Density in Chabazite Zeolites on Copper Ion Mobility and Reactivity During NOx Selective Catalytic Reduction with NH3.Nat. Catal.2023, 6, 276–285. Google Scholar
    • 30. Zhao S.; Tan C.; He C.; An P.; Xie F.; Jiang S.; Zhu Y.; Wu K.; Zhang B.; Li H.; Zhang J.; Chen Y.; Liu S.; Dong J.; Tang Z.Structural Transformation of Highly Active Metal–Organic Framework Electrocatalysts during the Oxygen Evolution Reaction.Nat. Energy2020, 5, 881–890. Google Scholar
    • 31. Zhou M.; Wang Z.; Mei A.; Yang Z.; Chen W.; Ou S.; Wang S.; Chen K.; Reiss P.; Qi K.; Ma J.; Liu Y.Photocatalytic CO2 Reduction Using La-Ni Bimetallic Sites within a Covalent Organic Framework.Nat. Commun.2023, 14, 2473. Google Scholar
    • 32. Lin L.; Fan M.; Sheveleva A. M.; Han X.; Tang Z.; Carter J. H.; da Silva I.; Parlett C. M. A.; Tuna F.; McInnes E. J. L.; Sastre G.; Rudić S.; Cavaye H.; Parker S. F.; Cheng Y.; Daemen L. L.; Ramirez-Cuesta A. J.; Attfield M. P.; Liu Y.; Tang C. C.; Han B.; Yang S.Control of Zeolite Microenvironment for Propene Synthesis from Methanol.Nat. Commun.2021, 12, 822. Google Scholar
    • 33. Liutkova A.; Drozhzhin V.; Heinrichs J. M. J. J.; Jestl V.; Evtushkova A.; Mezari B.; Mayoral Á.; Kosinov N.; Hensen E. J. M.Role of Strontium Cations in ZSM-5 Zeolite in the Methanol-to-Hydrocarbons Reaction.J. Phys. Chem. Lett.2023, 14, 6506–6512. Google Scholar
    • 34. Mesilov V. V.; Bergman S. L.; Dahlin S.; Xiao Y.; Xi S.; Ma Z.; Xu L.; Chen W.; Pettersson L. J.; Bernasek S. L.Differences in Oxidation-Reduction Kinetics and Mobility of Cu Species in Fresh and SO2-Poisoned Cu-SSZ-13 Catalysts.Appl. Catal. B-Environ.2021, 284, 119756. Google Scholar
    • 35. Bauer M.HERFD-XAS and Valence-to-Core-XES: New Tools to Push the Limits in Research with Hard X-Rays?Phys. Chem. Chem. Phys.2014, 16, 13827–13837. Google Scholar
    • 36. Cao D.; Xu W.; Chen S.; Liu C.; Sheng B.; Song P.; Moses O.; Song L.; Wei S.Visualizing Catalytic Dynamics Processes via Synchrotron Radiation Multitechniques.Adv. Mater.2023, 35, 2205346. Google Scholar
    • 37. Zhang H.; Bolshakov A.; Meena R.; Garcia G. A.; Dugulan A. I.; Parastaev A.; Li G.; Hensen E. J. M.; Kosinov N.Revealing Active Sites and Reaction Pathways in Methane NonOxidative Coupling over Iron-Containing Zeolites.Angew. Chem. Int. Ed.2023, 62, 202306196. Google Scholar
    • 38. Li W.; Wu J.; Hu W.; Dang J.; Wang C.; Weng X.; da Silva I.; Manuel P.; Yang S.; Guan N.; Li L.Direct Propylene Epoxidation with Molecular Oxygen over Cobalt Containing Zeolites.J. Am. Chem. Soc.2022, 144, 4260–4268. CrossrefGoogle Scholar
    • 39. Su J.; Zhang L.; Zhou H.; Ye Y.; Zheng X.; Liu C.; Liu S.; Jiao W.; Liu X.; Wang C.; Wang Y.; Xie Z.Unveiling the Anti-Trap Effect for Bridging Intermediates on ZnAlOx/AlPO-18 Bifunctional Catalysts to Boost Syngas to Olefin Conversion.ACS Catal.2023, 13, 2472–2481. Google Scholar
    • 40. Sushkevich V. L.; Safonova O. V.; Palagin D.; Newton M. A.; van Bokhoven J. A.Structure of Copper Sites in Zeolites Examined by Fourier and Wavelet Transform Analysis of EXAFS.Chem. Sci.2020, 11, 5299. Google Scholar
    • 41. Timoshenko J.; Frenkel A.“Inverting” X-Ray Absorption Spectra of Catalysts by Machine Learning in Search for Activity Descriptors.ACS Catal.2019, 9, 10192–10211. Google Scholar
    • 42. Liu Z.; Chen Z.; Li M.; Li J.; Zhuang W.; Yang X.; Wu S.; Zhang J.Construction of Single Ni Atom-Immobilized ZIF-8 with Ordered Hierarchical Pore Structures for Selective CO2 Photoreduction.ACS Catal.2023, 13, 6630–6640. Google Scholar
    • 43. Fang G.; Wei F.; Lin J.; Zhou Y.; Sun L.; Shang X.; Lin S.; Wang X.Retrofitting Zr-Oxo Nodes of UiO-66 by Ru Single Atoms to Boost Methane Hydroxylation with Nearly Total Selectivity.J. Am. Chem. Soc.2023, 145, 13169–13180. Google Scholar
    • 44. He J.; Wu Z.; Gu Q.; Liu Y.; Chu S.; Chen S.; Zhang Y.; Yang B.; Chen T.; Wang A.; Weckhuysen B.; Zhang T.; Luo W.Zeolite-Tailored Active Site Proximity for the Efficient Production of Pentanoic Biofuels.Angew. Chem. Int. Ed.2021, 60, 23713–23721. Google Scholar
    • 45. Ma Y.; Song S.; Liu C.; Liu L.; Zhang L.; Zhao Y.; Wang X.; Xu H.; Guan Y.; Jiang J.; Song W.; Han Y.; Zhang J.; Wu P.Germanium-Enriched Double-Four-Membered-Ring Units Inducing Zeolite-Confined Subnanometric Pt Clusters for Efficient Propane Dehydrogenation.Nat. Catal.2023, 6, 506–518. Google Scholar
    • 46. Zhao Y.; Lu X.; Wu Z.; Pei Z.; Luan D.; Lou X.Supporting Trimetallic Metal-Organic Frameworks on S/N-Doped Carbon Macroporous Fibers for Highly Efficient Electrocatalytic Oxygen Evolution.Adv. Mater.2023, 35, 2207888. Google Scholar
    • 47. Janis T.; Lu D.; Lin Y.; Frenkel A. I.Supervised Machine-Learning-Based Determination of Three Dimensional Structure of Metallic Nanoparticles.J. Phys. Chem. Lett.2017, 8, 5091–5098. Google Scholar
    • 48. Sun Q.; Wang Q.; Fan Q.; Zeng L.; Alvaro M.; Miao S.; Yang R.; Jiang Z.; Zhou W.; Zhang J.; Zhang T.; Xu J.; Zhang P.; Cheng J.; Yang D.; Jia R.; Li L.; Zhang Q.; Wang Y.; Terasaki O.; Yu Y.Subnanometer Bimetallic Platinum-Zinc Clusters in Zeolites for Propane Dehydrogenation.Angew. Chem. Int. Ed.2020, 59, 19450–19459. Google Scholar
    • 49. Sun Q.; Wang N.; Zhang T.; Bai R.; Mayoral A.; Zhang P.; Zhang Q.; Terasaki O.; Yu J.Zeolite-Encaged Single-Atom Rhodium Catalysts: Highly-Efficient Hydrogen Generation and Shape-Selective Tandem Hydrogenation of Nitroarenes.Angew. Chem. Int. Ed.2019, 58, 2–9. Google Scholar
    • 50. Heyer A. J.; Plessers D.; Braun A.; Rhoda H. M.; Bols M. L.; Hedman B.; Hodgson K. O.; Schoonheydt R. A.; Sels B. F.; Solomon E. I.Methane Activation by a Mononuclear Copper Active Site in the Zeolite Mordenite: Effect of Metal Nuclearity on Reactivity.J. Am. Chem. Soc.2022, 144, 19305–19316. Google Scholar
    • 51. Cheng Q.; Li G.; Yao X.; Zheng L.; Wang J.; Emwas A. H.; Castaño P.; Ruiz-Martínez J.; Han Y.Maximizing Active Fe Species in ZSM-5 Zeolite Using Organic-Template-Free Synthesis for Efficient Selective Methane Oxidation.J. Am. Chem. Soc.2023, 145, 5888–5898. Google Scholar
    • 52. Liu G.; Liu P.; Meng D.; Zhao T.; Qian X.; He Q.; Guo X.; Qi J.; Peng L.; Xue N.; Zhu Y.; Ma J.; Wang Q.; Liu X.; Chen L.; Ding W.COx Hydrogenation to Methanol and Other Hydrocarbons under Mild Conditions with Mo3S4@ZSM-5.Nat. Commun.2023, 14, 513. Google Scholar
    • 53. Hu M.; Zhao S.; Liu S.; Chen C.; Chen W.; Zhu W.; Liang C.; Cheong W.; Wang Y.; Yu Y.; Peng Q.; Zhou K.; Li J.; Li Y.MOF-Confined Sub-2 nm Atomically Ordered Intermetallic PdZn Nanoparticles as High-Performance Catalysts for Selective Hydrogenation of Acetylene.Adv. Mater.2018, 30, 1801878. Google Scholar
    • 54. Liu L.; Lopez-Haro M.; Lopes C. W.; Rojas-Buzo S.; Concepcion P.; Manzorro R.; Simonelli L.; Sattler A.; Serna P.; Calvino J. J.; Corma A.Structural Modulation and Direct Measurement of Subnanometric Bimetallic PtSn Clusters Confined in Zeolites.Nat. Catal.2020, 3, 628–638. Google Scholar
    • 55. Sui J.; Liu H.; Hu S.; Sun K.; Wan G.; Zhou H.; Zheng X.; Jiang H.A General Strategy to Immobilize Single-Atom Catalysts in Metal-Organic Frameworks for Enhanced Photocatalysis.Adv. Mater.2021, 34, 2109203. Google Scholar
    • 56. Zhao L.; Bian J.; Zhang X.; Bai L.; Xu L.; Qu Y.; Li Z.; Li Y.; Jing L.Construction of Ultrathin S-Scheme Heterojunctions of Single Ni Atom Immobilized Ti-MOF and BiVO4 for CO2 Photoconversion of Nearly 100% to CO by Pure Water.Adv. Mater.2022, 34, 2205303. Google Scholar
    • 57. Ma X.; Liu H.; Yang W.; Mao G.; Zheng L.; Jiang H.Modulating Coordination Environment of Single-Atom Catalysts and Their Proximity to Photosensitive Units for Boosting MOF Photocatalysis.J. Am. Chem. Soc.2021, 143, 12220–12229. Google Scholar
    • 58. Tang X.; Ye J.; Guo L.; Pu T.; Cheng L.; Cao X.; Guo Y.; Wang L.; Guo Y.; Zhan W.Atomic Insights into the Cu Species Supported on Zeolite for Direct Oxidation of Methane to Methanol via Low-Damage HAADF-STEM.Adv. Mater.2023, 35, 2208504. Google Scholar
    • 59. Luo D.; Li C.; Zhao Y.; Ma Q.; Ma C.; Nie Y.; Li M.; Weng X.; Huang R.; Zhao Y.; Shui L.; Wang X.; Chen Z.Design of Quasi-MOF Nanospheres as a Dynamic Electrocatalyst Toward Accelerated Sulfur Reduction Reaction for High-Performance Lithium–Sulfur Batteries.Adv. Mater.2022, 34, 2105541. Google Scholar
    • 60. Antil N.; Chauhan M.; Akhtar N.; Kalita R.; Manna K.Selective Methane Oxidation to Acetic Acid Using Molecular Oxygen over a Mono-Copper Hydroxyl Catalyst.J. Am. Chem. Soc.2023, 145, 6156–6165. Google Scholar
    • 61. Han S.; Li H.; Li T.; Chen F.; Yang R.; Yu Y.; Zhang B.Ultralow Overpotential Nitrate Reduction to Ammonia via a Three-Step Relay Mechanism.Nat. Catal.2023, 6, 402–414. Google Scholar
    • 62. Zeng Y.; Li X.; Wang J.; Sougrati M. T.; Huang Y.; Zhang T.; Liu B.In Situ/Operando Mössbauer Spectroscopy for Probing Heterogeneous Catalysis.Chem Catal.2021, 1, 1215–1233. Google Scholar
    • 63. Simons M. C.; Vitillo J. G.; Babucci M.; Hoffman A. S.; Boubnov A.; Beauvais M. L.; Chen Z.; Cramer C. J.; Chapman K. W.; Bare S. R.; Gates B. C.; Lu C. C.; Gagliardi L.; Bhan A.Structure, Dynamics, and Reactivity for Light Alkane Oxidation of Fe(II) Sites Situated in the Nodes of a Metal–Organic Framework.J. Am. Chem. Soc.2019, 141, 18142–18151. Google Scholar
    • 64. Rana R.; Vila F. D.; Kulkarni A. R.; Bare S. R.Bridging the Gap Between the X-Ray Absorption Spectroscopy and the Computational Catalysis Communities in Heterogeneous Catalysis: A Perspective on the Current and Future Research Directions.ACS. Catal.2022, 12, 13813–13830. Google Scholar
    • 65. Shan J.; Ye C.; Jiang Y.; Mietek J.; Zheng Y.; Qiao S.Metal-Metal Interactions in Correlated Single-Atom Catalysts.Sci. Adv.2022, 8, eabo0762. Google Scholar
    • 66. Liu Y.; Wang B.; Fu Q.; Liu W.; Wang Y.; Gu L.; Wang D.; Li Y.Polyoxometalate-Based Metal–Organic Framework as Molecular Sieve for Highly Selective Semi-Hydrogenation of Acetylene on Isolated Single Pd Atom Sites.Angew. Chem. Int. Ed.2021, 60, 22522–22528. Google Scholar
    • 67. Signorile M.; Braglia L.; Crocella V.; Torelli P.; Groppo E.; Ricchiardi G.; Bordiga S.; Bonino F.Titanium Defective Sites in TS-1: Structural Insights by Combining Spectroscopy and Simulation.Angew. Chem. Int. Ed.2020, 59, 18145–18150. Google Scholar
    • 68. Lee I.; Lee M. S.; Tao L.; Ikuno T.; Khare R.; Jentys A.; Huthwelker T.; Borca C.N.; Kalinko A.; Gutiérrez O.Y.; Govind N.; Fulton J. L.; Hu J.; Glezakou V. A.; Rousseau R.; Sanchez-Sanchez M.; Lercher J. A.Activity of Cu–Al–Oxo Extra-Framework Clusters for Selective Methane Oxidation on Cu-Exchanged Zeolites.JACS Au2021, 1, 1412–1421. Google Scholar
    • 69. Zhang J.; An B.; Li Z.; Cao Y.; Dai Y.; Wang W.; Zeng L.; Lin W.; Wang C.Neighboring Zn–Zr Sites in a Metal–Organic Framework for CO2 Hydrogenation.J. Am. Chem. Soc.2021, 143, 8829–8837. Google Scholar
    • 70. An B.; Li Z.; Song Y.; Zhang J.; Zeng L.; Wang C.; Lin W.Cooperative Copper Centres in a Metal–Organic Framework for Selective Conversion of CO2 to Ethanol.Nat. Catal.2019, 2, 709–717. Google Scholar
    • 71. Zhu Y.; Zheng J.; Ye J.; Cui Y.; Koh K.; Kovarik L.; Camaioni D. M.; Fulton J. L.; Truhlar D. G.; Neurock M.; Cramer C. J.; Gutiérrez O. Y.; Lercher J. A.Copper-Zirconia Interfaces in UiO-66 Enable Selective Catalytic Hydrogenation of CO2 to Methanol.Nat. Commun.2020, 11, 5849. Google Scholar
    • 72. Feng X.; Song Y.; Chen J. S.; Xu Z.; Dunn S. J.; Lin W.Rational Construction of an Artificial Binuclear Copper Monooxygenase in a Metal–Organic Framework.J. Am. Chem. Soc.2021, 143, 1107–1118. Google Scholar
    • 73. Ma S.; Liu Z.Zeolite-Confined Subnanometric PtSn Mimicking Mortise-and-Tenon Joinery for Catalytic Propane Dehydrogenation.Nat. Commun.2022, 13, 2716. Google Scholar
    • 74. Kiani D.; Sourav S.; Tang Y.; Baltrusaitis J.; Wachs I. E.Methane Activation by ZSM-5-Supported Transition Metal Centers.Chem. Soc. Rev.2021, 50, 1251–1268. Google Scholar
    • 75. Signorile M.; Borfecchia E.; Bordiga S.; Berlier G.Influence of Ion Mobility on the Redox and Catalytic Properties of Cu Ions in Zeolites.Chem. Sci.2022, 13, 10238–10250. Google Scholar
    • 76. Zhu Y.; Kuo T.; Li Y.; Qi M.; Chen G.; Wang J.; Xu Y.; Chen H.Emerging Dynamic Structure of Electrocatalysts Unveiled by in Situ X-Ray Diffraction/Absorption Spectroscopy.Energy Environ. Sci.2021, 14, 1928–1958. Google Scholar
    • 77. Bergman S. L.; Dahlin S.; Mesilov V. V.; Xiao Y.; Englund J.; Xi S.; Tang C.; Skoglundh M.; Pettersson L. J.; Bernasek S. L.In-Situ Studies of Oxidation/Reduction of Copper in Cu-CHA SCR Catalysts: Comparison of Fresh and SO2-Poisoned Catalysts.Appl. Catal. B-Environ.2020, 269, 118722. Google Scholar
    • 78. Farberow C. A.; Wegener E. C.; Kumar A.; Miller J. H.; Dupuis D. P.; Kim S.; Ruddy D. A.Connecting Cation Site Location to Alkane Dehydrogenation Activity in Ni/BEA Catalysts.J. Catal.2022, 413, 264–273. Google Scholar
    • 79. Hall J. N.; Kropf A. J.; Delferro M.; Bollini P.Kinetic and X-Ray Absorption Spectroscopic Analysis of Catalytic Redox Cycles over Highly Uniform Polymetal Oxo Clusters.ACS Catal.2023, 13, 5406–5427. Google Scholar
    • 80. Sushkevich V. L.; Artsiusheuski M.; Klose D.; Jeschke G.; van Bokhoven J. A.Identification of Kinetic and Spectroscopic Signatures of Copper Sites for Direct Oxidation of Methane to Methanol.Angew. Chem. Int. Ed.2021, 60, 15944–15953. Google Scholar
    • 81. Gorlin Y.; Lassalle-Kaiser B.; Benck J. D.; Gul S.; Webb S. M.; Yachandra V. K.; Yano J.; Jaramillo T. F.In Situ XAFS Study of a Modified TS-1 Framework for Carbonyl Formation.J. Phys. Chem. C2021, 125, 16483–16488. Google Scholar
    • 82. Kim I. S.; Li Z.; Zheng J.; Platero-Prats A. E.; Mavrandonakis M. A.; Pellizzeri S.; Ferrandon M.; Vjunov A.; Gallington L. C.; Webber T.E.; Vermeulen N. A.; Penn R.L.; Getman R. B.; Cramer C. J.; Chapman K.W.; Camaioni D. M.; Fulton J. L.; Lercher J. A.; Farha O. K.; Hupp J. K.; Martinson A. B. F.Sinter-Resistant Platinum Catalyst Supported by Metal-Organic Framework.Angew. Chem. Int. Ed.2018, 57, 909–913. Google Scholar
    • 83. Hannemann S.; Casapu M.; Grunwaldt J. D.; Haider P.; Trussel P.; Baiker A.; Welter E.A Versatile in Situ Spectroscopic Cell for Fluorescence/Transmission EXAFS and X-Ray Diffraction of Heterogeneous Catalysts in Gas and Liquid Phase.J. Synchrotron Rad.2007, 14, 345–354. Google Scholar
    • 84. Qi L.; Babucci M.; Zhang Y.; Lund A.; Liu L.; Li J.; Chen Y.; Hoffman A. S.; Bare S. R.; Han Y.; Gates B. C.; Bell A. T.Propane Dehydrogenation Catalyzed by Isolated Pt Atoms in ≡SiOZn-OH Nests in Dealuminated Zeolite Beta.J. Am. Chem. Soc.2021, 143, 21364–21378. Google Scholar
    • 85. Yasumura S.; Ide H.; Ueda T.; Jing Y.; Liu C.; Kon K.; Toyao T.; Maeno Z.; Shimizu K. I.Transformation of Bulk Pd to Pd Cations in Small-Pore CHA Zeolites Facilitated by NO.JACS Au2021, 1, 201–211. Google Scholar
    • 86. Pappas D. K.; Martini A.; Dyballa M.; Kvande K.; Teketel S.; Lomachenko K. A.; Baran R.; Glatzel P.; Arstad B.; Berlier G.; Lamberti C.; Bordiga S.; Olsbye U.; Svelle S.; Beato P.; Borfecchia E.The Nuclearity of the Active Site for Methane to Methanol Conversion in Cu-Mordenite: A Quantitative Assessment.J. Am. Chem. Soc.2018, 140, 15270–15278. Google Scholar
    • 87. Abdel-Mageed A. M.; Rungtaweevoranit B.; Parlinska-Wojtan M.; Pei X.; Yaghi O. M.; Behm R. J.Highly Active and Stable Single-Atom Cu Catalysts Supported by a Metal-Organic Framework.J. Am. Chem. Soc.2019, 141, 5201–5210. Google Scholar
    • 88. Kvande K.; Pappas D.; Borfecchia E.; Lomachenko K. A.Advanced X-Ray Absorption Spectroscopy Analysis to Determine Structure-Activity Relationships for Cu-Zeolites in the Direct Conversion of Methane to Methanol.ChemCatChem2020, 12, 2385–2405. Google Scholar
    • 89. Negri C.; Selleri T.; Borfecchia E.; Martini A.; Lomachenko K. A.; Janssens T. V. W.; Cutini M.; Bordiga S.; Berlier G.Structure and Reactivity of Oxygen-Bridged Diamino Dicopper(II) Complexes in Cu-Ion-Exchanged Chabazite Catalyst for NH3-Mediated Selective Catalytic Reduction.J. Am. Chem. Soc.2020, 142, 15884–15896. Google Scholar
    • 90. Gorlin Y.; Lassalle-Kaiser B.; Benck J. D.; Gul S.; Webb S. M.; Yachandra V. K.; Yano J.; Jaramillo T. F.In Situ X-Ray Absorption Spectroscopy Investigation of a Bifunctional Manganese Oxide Catalyst with High Activity for Electrochemical Water Oxidation and Oxygen Reduction.J. Am. Chem. Soc.2013, 135, 8525–8534. Google Scholar
    • 91. Xu Y.; Xie M.; Zhong H.; Cao Y.In Situ Clustering of Single-Atom Copper Precatalysts in a Metal-Organic Framework for Efficient Electrocatalytic Nitrate-to-Ammonia Reduction.ACS Catal.2022, 12, 8698–8706. Google Scholar
    • 92. Kozyr E. G.; Njoroge P. N.; Chapek S. V.; Shapovalov V. V.; Skorynina A. A.; Pnevskaya A. Y.; Bulgakov A. N.; Soldatov A. V.; Pellegrino F.; Groppo E.; Bordiga S.; Mino L.; Bugaev A.Operando Laboratory X-Ray Absorption Spectroscopy and UV–Vis Study of Pt/TiO2 Photocatalysts During Photodeposition and Hydrogen Evolution Reactions.Catalysts2023, 13, 414. Google Scholar
    • 93. Yang S.; Pattengale B.; Lee S.; Huang J.Real-Time Visualization of Active Species in a Single-Site Metal–Organic Framework Photocatalyst.ACS Energy Lett.2018, 3, 532–539. Google Scholar
    • 94. Petrov A. W.; Ferri D.; Krumeich F.; Nachtegaal M.; van Bokhoven J. A.; Krocher O.Stable Complete Methane Oxidation over Palladium Based Zeolite Catalysts.Nat. Commun.2018, 9, 2545. Google Scholar
    • 95. Lomachenko K. A.; Molokova A. Y.; Atzori C.; Mathon O.Quantification of Adsorbates by X-Ray Absorption Spectroscopy: Getting TGA-Like Information for Free.J. Phys. Chem. C2022, 126, 5175–5179. Google Scholar
    • 96. Molokova A. Y.; Borfecchia E.; Martini A.; Pankin I. A.; Atzori C.; Mathon O.; Bordiga S.; Wen F.; Vennestrom P. N. R.; Berlier G.; Janssens T. V. W.; Lomachenko K. A.SO2 Poisoning of Cu-CHA deNO(x) Catalyst: The Most Vulnerable Cu Species Identified by X-Ray Absorption Spectroscopy.JACS Au2022, 2, 787–792. Google Scholar