Optimizing the reaction pathway of methane photo-oxidation over single copper sites

feature-image

Play all audios:

Loading...

ABSTRACT Direct photocatalytic conversion of methane to value-added C1 oxygenate with O2 is of great interest but presents a significant challenge in achieving highly selective product


formation. Herein, a general strategy for the construction of copper single-atom catalysts with a well-defined coordination microenvironment is developed on the basis of metal-organic


framework for selective photo-oxidation of CH4 to HCHO. We propose the directional activation of O2 on the mono-copper site breaks the original equilibrium and tilts the balance of radical


formation almost completely toward •OOH. The synchronously generated •OOH and •CH3 radicals rapidly combine to form HCHO while inhibiting competing reactions, thus resulting in ultra-highly


selective HCHO production (nearly 100%) with a time yield of 2.75 mmol gcat−1 h−1. This work highlights the potential of rationally designing reaction sites to manipulate reaction pathways


and achieve selective CH4 photo-oxidation, and could guide the further design of high-performance single-atom catalysts to meet future demand. SIMILAR CONTENT BEING VIEWED BY OTHERS SYNERGY


OF PD ATOMS AND OXYGEN VACANCIES ON IN2O3 FOR METHANE CONVERSION UNDER VISIBLE LIGHT Article Open access 25 May 2022 NEARLY 100% SELECTIVE AND VISIBLE-LIGHT-DRIVEN METHANE CONVERSION TO


FORMALDEHYDE VIA. SINGLE-ATOM CU AND WΔ+ Article Open access 10 May 2023 SELECTIVE LIGHT-DRIVEN METHANE OXIDATION TO ETHANOL Article Open access 01 December 2024 INTRODUCTION Methane (CH4)


comprises 70–90% of natural gas and is widely recognized as an essential feedstock in the manufacturing of chemicals1,2,3,4. However, due to the intense bond energy (439 kJ mol−1) and the


weak polarization of carbon-hydrogen (C–H) bonds in methane, selective conversion of methane remains a significant challenge5,6. Currently, the steam reforming reaction serves as the primary


approach for transforming methane into syngas. Unfortunately, this method requires extreme temperatures (>800 °C) and pressures (>30 bar), thereby hindering its application in


sustainable development7,8,9. Therefore, it becomes imperative to explore strategies that enable the conversion of methane to value-added products under mild conditions10,11. Recently,


photocatalytic oxidation of CH4 to HCHO or CH3OH under room temperature and mild pressure is considered to be one of the “holy grails” in C1 chemistry, but suffers from the challenge to


control the reaction pathway and product selectivity12,13. During the CH4 photo-oxidation, O2 and CH4 undergo initial activation into free radicals with subsequently combining. It is known


that the CH4 oxidation process involves multiple free radicals (•CH3, •OOH, and •OH) to produce a wide variety of products, such as HCHO and CH3OH, as well as overoxidized CO2, while the


distribution of final products is determined by the O2 activation and oxygen radical species14,15. $$\bullet {{{\rm{OOH}}}}({{{\rm{aq}}}})+{{{\rm{\bullet


}}}}{{{\rm{C}}}}{{{{\rm{H}}}}}_{3}({{{\rm{aq}}}})\to {{{\rm{C}}}}{{{{\rm{H}}}}}_{3}{{{\rm{OOH}}}}({{{\rm{aq}}}})\to


{{{\rm{HCHO}}}}({{{\rm{aq}}}})+{{{{\rm{H}}}}}_{2}{{{\rm{O}}}}({{{\rm{l}}}})$$ (1) $$\bullet {{{\rm{OH}}}}({{{\rm{aq}}}})+\bullet {{{\rm{C}}}}{{{{\rm{H}}}}}_{3}({{{\rm{aq}}}})\to


{{{\rm{C}}}}{{{{\rm{H}}}}}_{3}{{{\rm{OH}}}}({{{\rm{aq}}}})$$ (2) However, random reaction sites on the surface of traditional photocatalysts lead to the generation of series oxygen radicals,


and the overly oxidizing nature of •OH frequently results in multi-step oxidation or even over-oxidation of the product, thus influencing selectivity16,17,18. Hence, strategically managing


the O2 activation process to produce •OOH is advisable for achieving a highly selective HCHO synthesis19,20. Depending on the competitiveness of the reaction, constructing highly active and


selective reaction sites on the surface of randomly positioned catalysts can significantly break the balance between the main reaction and the competing reactions (Supplementary Fig. 1). If


the activity gap between the active site and the random site is sufficiently significant, the competitive equilibrium will continuously shift towards the main reaction, ultimately resulting


in the near-complete suppression of the competitive reaction. Under this guidance, manipulating the mono-metal site is expected to achieve selectivity in the activation of O2 to •OOH,


benefiting from the end-on O2 adsorption on it to minimize O-O bond breaking (Supplementary Fig. 2). Hence, to achieve high efficiency and selectivity for CH4 photo-oxidation, it is crucial


to construct single-atom catalysts with well-defined reaction microenvironments to control the types of active free radicals, which put forward higher requirements for the precise design of


photocatalysts. Metal-organic frameworks (MOFs) constructed from secondary building units of metal clusters and organic linkers provide a tunable platform for the locally precise design of


photocatalysts21,22,23,24. Metalizing the hydroxyl groups (μ3-OH) on the [Zr6(μ3-O)4(μ3-OH)4] cluster (Zr6) in Zr-MOF is expected to obtain single-atom sites with clear structures, which


allow precisely control of O2 activation process25, thereby achieving highly selective CH4 photo-oxidation. Herein, we support well-defined Cu active sites on deprotonated Zr6 connecting


nodes in the UIO66-NH2. The pores in MOFs act as microreactors, where O2 and CH4 can be activated on the Zr6 connecting node and organic linkers, respectively. Crucially, the entirely


coordinated Zr in the Zr-MOF leads to a deficiency in O2 adsorption sites, making the introduced mono-copper site serve as the primary effective O2 activation site (Supplementary Fig. 3),


significantly expediting the formation of •OOH while impeding the production of •OH. By employing meticulously designed in-situ spectroscopy and theoretical simulations, we illustrate that


the mono-copper site significantly enhances the transformation of O2 to •OOH, resulting in nearly 100% selectivity towards HCHO in the photo-oxidation of CH4. In addition, since the reaction


is confined in the porous microreactor, the steric confinement of the active species and the ultrafast local mass transfer efficiency greatly enhance the reactivity. In the judicious


designed reactive microenvironment, an exceptional time yield of 2.75 mmol gcat−1 h−1 for the production of HCHO from CH4 can be achieved. RESULTS SYNTHESIS AND CHARACTERIZATIONS UIO66-NH2


(UION) is synthesized via a solvothermal reaction between ZrCl4 and 2-Aminoterephthalic acid (H2ATA) in a mixture of acetic acid and DMF (1:10 v/v) at 120 °C for 12 h (Fig. 1a)26,27. The


UION is then treated with _n_-BuLi to deprotonate the μ3-OH at its Zr6 node, followed by metallization with CuCl2 in THF at room temperature to obtain UION-Cu(Cl) with node-supported copper


chloride. The deprotonation is confirmed by the disappearance of the vibrational bands of μ3-OH (3679 cm−1) in the Fourier transform infrared (FTIR) spectrum (Supplementary Fig. 4).


Subsequently, the coordination environment of the mono-copper site is adjusted from -Cu(Cl) to -Cu(OH), which facilitates the initiation of the O2 reduction reaction (Fig. 1b). The XRD


patterns of UION, UION-Cu(Cl) and UION-Cu(OH) are similar to that of pristine UIO as well as simulated UION-Cu(OH) MOF, confirming the retention of UIO-type structure constructed from Zr6


nodes and H2ATA linkers (Supplementary Fig. 5). Scanning electron microscopy (SEM) and transmission electron microscopy (TEM) further confirmed that the modification process will not destroy


the morphology and crystal form of UION-Cu(OH), and no agglomerated Cu species can be observed (Supplementary Figs. 6–10). Considering the electron beam sensitive characteristics of MOF


materials, the low-dose high-angle annular dark-field-scanning transmission electron microscopy (HAADF-STEM) images of samples have been collected to further character the structure


information. Figure 1c illustrates the regularly arranged Zr6 clusters in UION-Cu(OH), which were consistent with the simulated results (Fig. 1d, e). Moreover, the high contrast spots of


UION-Cu(OH) match well with the high-resolution structure of UION (Supplementary Fig. 11), implying that Cu species are uniformly distributed on Zr6 clusters. Energy dispersive X-ray


spectroscopy (EDX) mapping further indicates the uniform distribution of Cu throughout the UION-Cu(OH) particles (Fig. 1f). The Cu loading content is determined by inductively coupled plasma


mass spectrometry (ICP-MS) to be 3.48 Cu per Zr6 node, corresponding to complete metallation of four μ3-OH sites (Supplementary Table 1). Nitrogen adsorption experiments give a


Brunauer-Emmett-Teller surface area of 414.9 m2 g−1 for UION-Cu(OH), which is smaller than that of the unmetallated UION sample (763.7 m2 g−1), as expected (Supplementary Fig. 12 and Table 


2). The X-ray photoelectron spectroscopy (XPS) of UION-Cu(Cl) and UION-Cu(OH) displays binding energy peaks for Cu 2p3/2 and 2p1/2 along with satellite peaks (Supplementary Figs. 13 and 14),


indicating that the oxidation state of Cu is +222. The +2 oxidation state of Cu species is also assigned by Cu LMM spectra (Supplementary Fig. 15), which shows characteristic peaks at 572.5


 eV28. Moreover, electron paramagnetic resonance (EPR) of UION-Cu(OH) exhibits two signals at _g_ = 2.003 and _g_ = 2.320, respectively (Fig. 2a). The Lorentzian EPR signal at _g_ = 2.003


can be assigned to the intrinsic signal of MOF, which is also observed on UIO and UION29. The additional hyperfine peak assigned to the hybridized CuII species is observed at _g_ = 2.320,


further demonstrating the highly dispersed copper species (CuII)30,31. The local coordination environments of mono-copper site in prepared samples have been probed by X-ray absorption fine


structure (XAFS) spectra at the Cu K edge. The rising edge at 8985.5 eV in the X-ray absorption near edge structure (XANES) is assigned to the 1s→4p transition of CuII (Fig. 2b)32. The


_k_2-weighted Fourier transform EXAFS spectra of UION-Cu(Cl) and UION-Cu(OH) exhibit only one peak at around 1.53 Å, with no Cu–Cu (2.23 Å) or Cu-O-Cu (2.54 Å) observed, suggesting that the


Cu species are atomically anchored (Fig. 2d). To achieve enhanced resolution in both R and k spaces, an exhaustive wavelet transform (WT) analysis of Cu K-edge EXAFS oscillations has been


conducted. As depicted in Fig. 2c, no discernible extended Cu-Cu and Cu-O-Cu features are evident in either UION-Cu(Cl) or UION-Cu(OH), providing additional confirmation of the atomic


dispersion of Cu species (Supplementary Fig. 16). The EXAFS spectra are fitted within the 1.0–3.0 Å range based on the density functional theory (DFT) optimized structure model, extracting


coordination details for the first shell. The Cu site in UION-Cu(Cl) is stabilized by three oxygens on the Zr6 cluster and connected to a Cl atom with a Cu-Cl bond length of 2.18 Å


(Supplementary Fig. 17 and Supplementary Table 3). In UION-Cu(OH), the Cl atom is substituted by OH, resulting in a four-oxygen coordination environment for Cu. This coordination involves


three Cu-O bonds with oxygen atoms at a distance of 1.94 Å and one Cu-O bond with an oxygen atom at a distance of 1.97 Å originating from the -OH group (Fig. 2e and Supplementary Table 3).


The preceding results clarify the coordination structure of the mono-copper site and lay the groundwork for investigating the reaction mechanism. PHOTOCATALYTIC PERFORMANCE The


photocatalytic performance is assessed through methane conversion in a top-irradiation high-pressure batch reactor, in which 5 mg of photocatalysts are suspended in 20 mL of distilled water


in a mixture of 9 bar CH4 and 1 bar O2 and subjected to a 5-h irradiation period at 20 °C (Supplementary Fig. 18). Figure 3a present the oxygenates production including HCHO and CH3OH


together with overoxidized CO2 over prepared photocatalysts. UIO exhibits trace yields of C1 oxygenates in the reaction, consistent with its limited light absorption and severe charge


recombination. UION and UION-Cu(Cl) demonstrate increased production of C1 oxygenates, yet the resultant products lack selectivity. After adjusting the coordination environment of the


mono-copper site into -Cu(OH), there is a substantial enhancement in selectivity, with the target product HCHO reaching almost 100%. The highest HCHO yield is attained at 13.76 mmol g−1 over


UION-Cu(OH)-3.48, at which point the Cu site is approaching saturation. This outperforms most state-of-the-art photocatalysts operating at similar conditions or even higher pressure (Fig. 


3b)6,16,33,34,35,36. The ultra-high selectivity of the product is further confirmed through NMR 1H and 13C spectra, only HOCH2OH (the major species in aqueous solutions of HCHO) can be


detected over UION-Cu(OH) (Fig. 3c, d). It is important to note that due to the uncontrolled self-conversion of CH3OOH to HCHO in the product, we quantified HCHO only after CH3OOH was


completely converted to enhance accuracy (Fig. 3c, with no CH3OOH in the 1H NMR spectrum). To detect the unstable intermediate species CH3OOH, the reaction solution was stored at low


temperature and subjected to NMR testing as soon as possible. The characteristic signal of CH3OOH was observed at 3.68 ppm, providing clear evidence of CH3OOH formation (Supplementary Fig. 


19). In addition, when the reaction time is extended to 10 h (Fig. 3e), the yield of HCHO increased linearly, while the selectivity remained almost unchanged. Under visible light


irradiation, UION-Cu(OH) maintains its high efficiency in converting CH4 to HCHO, underscoring its excellent catalytic activity (Supplementary Fig. 20). The catalytic performance under


varied conditions is investigated with the absence of catalyst, light, CH4, and O2, respectively (Fig. 3f). In all these cases no C1 oxygenates could be detected, thus confirming that the


reaction is a photocatalytic CH4 oxidation process driven by UION-Cu(OH). Remarkably, in the absence of solvent (H2O) within the system, CH4 undergoes over-oxidation to CO2. This underscores


the significance of the free radical reaction process in the liquid phase as a crucial factor in preventing over-oxidation. Furthermore, the isotope-labeling experiments are conducted to


verify that HCHO is indeed produced via photocatalytic CH4 oxidation. As shown in Fig. 3g, the H13CHO and HCH18O peaks prove that the product is derived from CH4 and O2. The control


experiment in D2O demonstrates that the solvent does not participate directly in the reaction, it merely serves as a medium for the generation and transfer of free radicals (Supplementary


Fig. 21). Notably, a high apparent quantum yield (AQY) of 3.52% at 375 nm has been achieved on UION-Cu(OH), and the calculated AQY at 400, 420, and 450 nm are 1.26%, 0.83%, and 0.22%,


respectively, which resembles well with the UV-vis absorption (Fig. 3h). To investigate the stability of the optimized photocatalyst, the cycling test experiment is carried out over


UION-Cu(OH) photocatalyst (Fig. 3i). No obvious decrease of oxygenate yield and selectivity can be observed under 25 h reaction (five cycles), demonstrating the good stability of


UION-Cu(OH). Meanwhile, the XRD, XPS, TEM, and XAFS comparison of the used UION-Cu(OH) is carried out (Supplementary Figs. 22–25). These characterizations of UION-Cu(OH) after reaction


remain almost the same as the fresh samples, confirming the stable topology of catalyst. Nitrogen adsorption experiment has confirmed that the pore structure of the catalyst remained


unblocked even after undergoing multiple runs (Supplementary Fig. 26 and Table S2). In addition to its stable structure, the used catalyst does not exhibit significant carbon deposition or


other adsorbed species, which ensures the activity of the mono-copper site (Supplementary Fig. 27). CHARGE EXCITATION AND MIGRATION PROPERTIES The optical absorption ability of prepared


samples is first inspected by UV-vis diffuse reflectance spectroscopy (DRS). As shown in Fig. 4a, the absorption in UIO is induced by O to Zr charge transfer in Zr6 inorganic clusters.


However, the introduction of amino functional groups initiates an additional charge transfer from the linker to the Zr6 node, corresponding to strong light absorption in the visible region,


indicating superb spatial separation of photogenerated carriers between the linker and node27,37. According to the Tauc-plot curves, the mono-copper sites further induce a subtle red shift


in the bandgap (Supplementary Fig. 28), implying that these sites optimize the photoexcitation process. To understand the role of mono-copper sites in photocatalytic reaction, the


photoelectric properties and kinetic pathways of photocarriers are systematically investigated. The photoluminescence (PL) spectra of the samples are performed to study the separation of


photogenerated charge carriers (Supplementary Fig. 29). An intensive PL emission peak at 400-500 nm is detected over the bare UION samples, which is remarkably decreased after the


introduction of mono-copper sites, indicating that Cu species favorably prevents the recombination of charge carriers generated on UION MOFs. The time-resolved PL (TRPL) spectra illustrate


the dynamic changes in charge transfer induced by the Cu sites. The UION-Cu(OH) sample exhibits the longest fluorescence decay lifetime, corresponding to its optimal photoinduced charge


carrier separation efficiency (Supplementary Fig. 30). Furthermore, the photocurrent density increases after the introduction of mono-copper sites, further confirming the enhancement of


photoelectron separation and migration (Supplementary Fig. 31b). Similarly, the charge migration improvement is also demonstrated by electrochemical impedance spectroscopy analysis. Nyquist


plots (Supplementary Fig. 31a) show a smaller impedance radius of UION-Cu(OH) compared to the bare UION samples, demonstrating the superior separation and transport of photoinduced charges


in UION-Cu(OH), which accounts well for its superior photocatalytic activity38. To thoroughly clarify the regulation of transient photoelectron transfer kinetics facilitated by mono-copper


sites, femtosecond transient absorption (fs-TA) measurements are conducted. As illustrated in Fig. 4b, c, the fs-TA spectra of bare UION and UION-Cu(OH) are presented in the 390–690 nm


region. Evidently, the fs-TA spectra of UION and UION-Cu are primarily characterized by a photoinduced absorption (PA) signal centered at 625 nm, which can be ascribed to the secondary light


absorption by excited state electrons. The PA signal intensifies as photogenerated electrons accumulate in the conduction band and subsequently diminishes as a result of recombination


between photogenerated electrons and holes (Supplementary Fig. 32). Therefore, by fitting the decay process of the PA signal, the kinetic behavior of photogenerated carrier recombination can


be obtained. As shown in Fig. 4d, e, the kinetic plots with typical fitting curves illustrate the decay of both UION and UION-Cu(OH) followed a biexponential model, exhibiting one fast


component (τ1) with time constants of 17.87 and 23.10 ps, respectively, as well as a slower component (τ2) with a time constant extending to 183.2 and 400.4 ps. The noteworthy increase in


both τ1 (charge-carrier trapping) and τ2 (excitonic recombination) lifetime suggests that the mono-copper species can serve as electron storage sites, which can facilitate the extraction of


photoexcited electron and delay the recombination of excited state electrons. The charge transfer from UION to mono-copper sites is proved by in situ XPS spectra (Fig. 4f). Under light


illumination, the XPS characteristic peak of Cu moves toward the low binding energy direction accompanied by the weakening of the CuII satellite peak, confirming the role of mono-copper site


as an electron acceptor in the photocatalytic processes by UION-Cu(OH). In situ EPR spectra of the UION and UION-Cu(OH) are performed to further explore the details of the photoexcitation


and charge transfer mechanisms. Under illumination, the EPR signal of UION and UION-Cu(OH) at _g_ = 2.003 gradually enhanced, suggesting the generation of photoexcited electrons


(Supplementary Fig 33)39. By contrast for the UION-Cu(OH) sample, the CuII EPR signal weakened under the light irradiation, indicating that the content of CuII species decreased, as the CuI


and Cu0 are EPR silent, proving that photogenerated electrons will move and gather to Cu sites (Supplementary Fig 34)17. The XAFS test under illumination reveals a consistent trend. As


illustrated in Fig. 4g, the Cu k-edge XANES spectra of UION-Cu(OH) under illumination shift towards a lower energy direction, suggesting a decrease in the valence state of the Cu site,


confirming the transfer of photogenerated electrons to the mono-copper sites. It is worth noting that the peak intensity in the R-space decreases under illumination (Fig. 4h), suggesting a


reduction in the coordination number of the mono-copper site, which may originate from the detachment of -OH groups during the photoactivation process. Hence, further fitting of the data


under illumination was conducted (Fig. 4i). The results indicate that the Cu species undergo a transition to tri-oxygen coordination after illumination, exposing the activated mono-copper


sites. In summary, the mono-copper site accumulates photogenerated electrons and eliminates coordinated hydroxyl groups, creating an electron-rich exposed atomic active site and serving as a


platform for selective methane oxidation. MECHANISTIC INVESTIGATIONS The core issue regarding the reaction mechanism is the generation of radicals, which involves the band structure of the


photocatalyst and the surface reaction sites. We first determined the bandgap and band positions of the UION-Cu(OH) photocatalyst using UV-VIS DRS and XPS valence band spectra (Supplementary


Fig. 35). Given that the valence band and conduction band positions of the UION-Cu(OH) photocatalyst are 2.18 eV and −0.65 eV, respectively, the generated photoinduced carriers can drive


the activation of CH4 and O2 but cannot convert H2O into •OH6,16,40. Therefore, the activation process of O2 at the reaction sites determines the final product selectivity. Subsequently, a


series of in situ characterizations, including in situ XAFS, in situ EPR, in situ TRPL, and in situ diffuse reflectance infrared Fourier transform spectroscopy (DRIFTS), have been employed


to elucidate the evolution of the electronic states at the reaction sites and the surface species during the photocatalytic process. As shown in Fig. 5a, a shift of the Cu K edge to higher


energy levels can be observed as O2 is introduced under dark condition, indicating the increase in the oxidation state of the Cu species, presumably due to the adsorption of O2 on the


mono-copper site. Upon commencement of illumination, a reversal in the Cu K-edge absorption energy change was noted, suggesting a reduction in the oxidation state of Cu, potentially


attributed to the transfer of photogenerated electrons to mono-copper site and the photoreduction of adsorbed O2. As CH4 is additionally introduced into the reaction system, a discernible


shift in the Cu K-edge absorption energy towards higher values is observed, possibly because the consumption of active oxygen species by CH4 thus facilitates the adsorption of O2 onto


mono-copper sites. To further elucidate the local coordination structure evolution of the mono-copper active sites during the reaction, the Cu K-edge EXAFS spectra are investigated (Fig. 5b,


c and Supplementary Fig. 36). Following the introduction of O2, the coordination bond length of the Cu site expands, suggesting that the adsorption of O2 reduces the charge density of the


Cu site and weakens the coordination bond. With the introduction of light and CH4, the bond length gradually recovered, confirming the light-driven reaction between CH4 and reactive oxygen


species. In situ EPR measurements have also been conducted to unveil the charge transfer behavior during the photocatalytic O2 activation over UION-Cu(OH) (Supplementary Fig. 37). The CuII


EPR peak displays an evident decrease under light irradiation, and then these EPR signals are partially recovered after the introduction of O2, clearly revealing that the photogenerated


electrons first transfer to the mono-copper site and then migrate to O2, giving rise to its subsequent reduction41. Moreover, the excited-sate lifetimes of pure UION-Cu(OH) and those of


UION-Cu(OH) in the presence of O2, CH4 or O2 + CH4 are investigated by TRPL measurements (Supplementary Fig. 38). A decrease in the excited-state lifetime of UION-Cu(OH) has been observed


with either O2 or CH4, suggesting the involvement of both activation of O2 and CH4 in the reaction mechanism. Notably, the presence of O2 + CH4 decreased the excited-state lifetime further,


indicating a rapid reaction between the activated species of O2 and CH442. Furthermore, in situ DRIFTS measurements have been carried out to investigate the adsorption state of the reactants


and reaction intermediates during the reaction process. With the extension of the reaction time, the signals of the reaction intermediate *CH3 (1356 cm−1) and product *HCHO (1738 cm−1)


steadily rise43,44, and no distinctive peaks related to CH3OH are detected, which is consistent with the ultra-high HCHO selectivity of UION-Cu(OH) (Fig. 5d). However, a hydroxyl signal


possibly attributed to HCHO is observed on UION, corresponding to poor selectivity in the absence of a single copper site (Fig. 5e)45. Clearly, both UION and UION-Cu(OH) exhibit distinct CH4


adsorption and *CH3 intermediate signals46. Given that the linker can enrich photogenerated holes and CH4 (Supplementary Fig. 39), it can reasonably be speculated that CH4 undergoes


oxidation to form *CH3 active species through the action of photogenerated holes on the linker. Interestingly, a distinct O2 end-on adsorption signal (1080 cm−1) is observed on UION-Cu(OH),


but this phenomenon is not found on UION16, suggesting that the difference in product selectivity may arise from a change in the O2 adsorption state (Supplementary Fig. 40). As distinct


modes of O2 adsorption can influence its reduction process, consequently yielding different types of free radicals to affect the selectivity (Supplementary Fig. 41)16,25,47, EPR testing has


been conducted to confirm the existence of free radicals during the reaction48,49. The DMPO-OOH signal displays twice the intensity on UION-Cu(OH) compared to UION, implying that the


formation of •OOH is more favorable over the mono-copper site (Supplementary Fig. 42). Contrastingly, the DMPO-OH signal of UION-Cu(OH) almost disappeared after the introduction of


mono-copper sites (Fig. 5f). Based on the above in situ analysis, it is determined that O2 exhibits a preferential activation to form •OOH at mono-copper sites. The copper site has an


absolute advantage in competing for O2, and the rapid O2 conversion causes near-total inhibition of the reaction at random sites, thus promoting the significant tilt of the competitive


equilibrium and resulting in highly selective •OOH generation (Fig. 5k). Therefore, the predominant active species are •OOH and •CH3 in the UION-Cu(OH)-mediated photocatalytic CH4 oxidation


process (Fig. 5g), resulting in the ultra-high selectivity for the production of HCHO. Drawing from the aforementioned experimental results, the reaction pathway for the highly selective


photocatalytic oxidation of methane to formaldehyde on the engineered catalyst can be unequivocally determined. Initially, the UION-Cu(OH) catalyst is excited when exposed to illumination,


leading to the migration of photogenerated electrons from the linker to the mono-copper site. Following that, methane is adsorbed by the amino group and subsequently undergoes oxidation and


dehydrogenation facilitated by photogenerated holes, leading to the formation of •CH3 (Supplementary Fig. 43a). Simultaneously, the mono-copper site is activated under light irradiation,


leading to the removal of the original -OH and providing O2 adsorption sites. Notably, due to the end-on adsorption of O2 on the mono-copper site, a highly selective reduction occurs,


leading to the formation of •OOH (Supplementary Fig. 43b). Then, DFT calculations have been conducted to substantiate both the experimental findings and the derived reaction mechanism. In


both UION and UION-Cu, the highest occupied molecular orbital (HOMO) is situated on the linker (Fig. 5h and Supplementary Fig. 44a), affirming that the amino group-containing linker serves


as the primary excitation region for photoelectrons50. However, in contrast to the uniform distribution of the lowest unoccupied molecular orbital (LUMO) across the UION framework


(Supplementary Fig. 44b), the LUMO level on UION-Cu is predominantly concentrated on the mono-copper site (Fig. 5i), which strongly indicates that the mono-copper sites function as


electron-rich region (Supplementary Fig. 45). Partial density of states calculations provide a more intuitive demonstration of this conclusion, as they reveal that Cu provides the lowest


empty orbital just above the Fermi level (Supplementary Fig. 46). To gain further insights into the reaction mechanism at the molecular level, the catalytic processes for methane and oxygen


molecule activation were computed. The obtained energy distribution diagrams indicate that the linker with amino groups can facilitate the adsorption of CH4, with an adsorption energy of


−0.22 eV. More importantly, the activation barrier for methane activation to •CH3 on this linker is significantly reduced, being only 0.93 eV, thereby substantially enhancing the activation


of CH4 (Supplementary Fig. 47). On the other hand, the exposed Cu sites serve as open platforms for O2 adsorption and activation (Supplementary Fig. 48). Calculations indicate that the Cu


site significantly promotes the reduction process of O2 to •OOH. Although the protonation of *O2 is exothermic on both UION-Cu and UION, the protonation step from O2 to •OOH on UION-Cu is


more thermodynamically and kinetically favorable than on UION, with the Gibbs activation free energies of −1.11 eV and −0.48 eV, respectively (Fig. 5j). Moreover, the •OOH generated at the


Cu site needs to overcome a barrier of 0.8 eV to be further reduced to •OH, which directly contributes to the high selectivity of •OOH in the O2 reduction process. As a result, the generated


•CH3 and •OOH swiftly react within the reaction microcavity of the MOF, resulting in the high-selective production of HCHO (Fig. 5l). DISCUSSION In summary, we have constructed a model of


highly reactive mono-copper sites on the secondary structural units of Zr6-MOF to precisely tailor the selectivity and reactivity of CH4 photo-oxidation. The mono-copper sites can


efficiently convert O2 into •OOH, maintaining an absolute advantage in competition with low-active random sites, consequently suppressing the formation of •OH. The •OOH and •CH3 formed in


the micro-reaction chamber provided by the MOF combine rapidly to promote reaction efficiency while achieving high HCHO selectivity. Fundamentally, substrate adsorption and radical


generation during CH4 photo-oxidation have been observed through a series of in situ spectroscopy and assigned to the oxidation-reduction active zones integrated in MOF pores. The operando


investigations elucidated the highly selective •OOH generation caused by end-on O2 adsorption at the mono-copper site, thus inducing a highly single CH4 oxidation pathway. As an outcome, the


UION-Cu(OH) exhibited near 100% product selectivity for CH4 photo-oxidation to HCHO with a rate of 2.75 mmol gcat−1 h−1. Our study offers guidance for the selective oxidation of CH4 to C1


compounds facilitated by single-atom catalysts and underscores the potential of using MOFs as innovative supports for the design of metal-centered catalysts. METHODS CHEMICALS The following


chemicals were purchased and used as-received without further purification. Zirconium tetrachloride (ZrCl4, ≥99.9%, Sigma Aldrich), 2-aminoterephthalic acid (H2ATA, ≥99%, Sigma Aldrich),


1,4-benzenedicarboxylic acid (≥98%, Sigma Aldrich), copper(II) chloride (CuCl2, ≥99%, Sigma Aldrich), Formic acid (HCOOH, reagent grade, CS Pharm Chemical), n-Butyllithium (n-BuLi, 2.5 M


solution in hexanes, Sigma Aldrich), sodium triethylborohydride (NaEt3BH, 0.1 M in tetrahydrofuran solution, Sigma Aldrich), dimethylformamide (DMF, HPLC grade, Sigma Aldrich),


tetrahydrofuran (THF, HPLC grade, Sigma Aldrich). CATALYST PREPARATION SYNTHESIS OF UION UIO-66-NH2 (abbreviated as UION) is synthesized via a solvothermal reaction of H2ATA and ZrCl4 in DMF


and water. H2ATA (0.2 mmol, 36 mg) is dissolved in 1.2 mL DMF containing 0.1 mL of HCOOH in a 15 mL glass vial and kept on stirring for 15 min. In a separate glass vial, ZrCl4 (0.2 mmol, 47


 mg) is dissolved in 0.2 mL water and transferred to the vial containing a solution of H2ATA. The mixture is kept on heating for 12 h at 120 °C. After cooling to room temperature, the light


yellow solid is collected via centrifugation, which is washed with DMF, acetone, and ethanol several times. The as-synthesized UiO-66-NH2 is immersed in the acetone to exchange DMF, followed


by drying at 80 °C under vacuum for 12 h. For comparison, UiO-66 (abbreviated as UIO) is synthesized by using 1,4-benzenedicarboxylic acid instead of 2-aminoterephthalic acid, whereas other


parameters remain the same. SYNTHESIS OF UION-CU(CL) n-BuLi (50 μL, 2.5 M in hexanes) is added to the slurry of UION (20 mg) in 1 mL THF in a 5 mL glass vial, and the mixture is kept for 1 


h at room temperature inside the glovebox. The solid is washed with THF several times to remove the excess n-BuLi. A THF solution of CuCl2 (1 mg/mL) is added to the vial, and the mixture is


kept overnight at room temperature. The UION-Cu(Cl) is centrifuged out of the suspension, and followed by washing with THF several times. The sample, UION-Cu(Cl), is further dried at 80 °C


under vacuum for 12 h. SYNTHESIS OF UION-CU(OH) A weighted amount of UION-Cu(Cl) (20 mg) is added into an anhydrous THF solution (50 mL) of the reducing agent sodium triethylborohydride


(NaEt3BH, 0.1 M, 1 mL). The mixture is stirred for 1 h at room temperature, and the obtained suspension of UION-CuH is separated via centrifugation followed by washing with THF several


times. UION-CuH is taken outside of the glovebox and kept in water for 30 min to afford UION-Cu(OH). The obtained sample is dried under vacuum at 80 °C for 12 h. The ratio of Cu to Zr in the


sample is determined by ICP-MS. CHARACTERIZATIONS Powder XRD are characterized by a powder X-ray diffraction instrument (Bruker D8 Advanced A25 diffractometer) with a Cu Kα target (λ = 


1.54056 Å) at 40 kV and 40 mA. FTIR spectra are obtained on a Nicolet iS10 FTIR spectrometer. SEM is performed on an FEI Teneo VS SEM. TEM images are obtained on a Titan ST microscope from


Thermo Fisher Scientific. HAADF-STEM and elemental mapping are performed on a Titan Themis Z microscope. XPS measurements are performed on a Thermo Scientific K-Alpha spectrometer with a


monochromatic Al Kα X-ray source. EPR measurements are obtained at room temperature using a Bruker EMX-10/12 EPR spectrometer operated in the X-band frequency. The XAFS spectra (Mn K-edge)


are collected at the beamline 1W1B of the Beijing Synchrotron Radiation Facility (BSRF, Beijing) in a fluorescence mode at room temperature. The optical absorption properties of the samples


are determined using the diffuse reflection method on a UV-visible light near-infrared spectrometer (Lambda 950). Steady-state PL spectra are recorded on a Carry Eclipse fluorescence


spectrometer. N2 adsorption isotherms were operated using a Micromeritics ASAP 2420 at 77 K. CH4 adsorption isotherms were operated using a Micromeritics ASAP 2420 at 273 K. TRPL decay


curves are obtained on FLS980 fluorescence spectrometers. In situ DRIFT spectra measurements are performed on a Nicolet 6700 Harrick spectrometer with an MCT detector. FEMTOSECOND TRANSIENT


ABSORPTION MEASUREMENT The fs-TA measurements are conducted using a commercial fs-TA system51. A fundamental 800 nm pulse generated by a Coherent Astrella regenerative amplifier served as


the pump source. This pulse is used to pump an optical parametric amplifier (Coherent, OperA Solo), producing a frequency-tunable pump beam spanning the visible light region. The pump beam,


with a wavelength of 350 nm, is then focused onto the sample. A white-light continuum probe beam is generated by focusing a weaker portion of the fundamental 800 nm beam onto a sapphire


window. The sample is positioned at the overlap of the pump beam and the white-light continuum probe beam. All measurements are conducted with samples in water solution using 1 mm quartz


cuvettes. IN SITU XAFS MEASUREMENTS The in situ XAFS measurements are collected at the beamline 1W1B of the (BSRF, Beijing) in a fluorescence mode at room temperature. Placing the sample and


vacuum-degassed deionized water into the reaction vessel ensures uniform dispersion. Continuously introduce argon gas into the reaction vessel as a protective atmosphere. Collect data


separately under dark and illuminated conditions to investigate the activation of mono-copper sites under light exposure. For in-situ testing of methane oxidation, reaction conditions are


introduced step by step to study the reaction process. Firstly, subject the sample to thorough illumination in an argon atmosphere to preliminarily activate mono-copper sites. Subsequently,


in the absence of light, collect spectroscopic information as the initial state of the reaction. Introduce O2 into the reaction vessel, stabilize, and collect spectroscopic data to obtain


material information during O2 adsorption. Resume illumination to gather information during O2 activation. Finally, introduce CH4 to complete the methane oxidation reaction. IN SITU DRIFTS


MEASUREMENTS The in situ DRIFTS measurements were conducted on a Nicolet 6700 IR spectrophotometer (Thermo Scientific) equipped with a Harrick Praying Mantis DRIFTS gas cell. After loading


the catalyst and assembling the reaction cell, He gas (20 mL/min) was introduced to purge the system for 30 min to remove air. The background was collected after purging. Subsequently, the


catalyst was irradiated using a xenon lamp light source, and infrared spectra collection was initiated. Each spectrum was scanned 32 times at a resolution of 4 cm−1. After collecting spectra


for 10 min, a mixed gas of CH4 and O2 (CH4/O2 = 1/1, total flow rate of 20 mL/min) was introduced, and data collection continued for 50 min. Throughout the experiment, the system was


temperature-controlled at 20 °C using a recirculating water cooling system to prevent potential hazards. TRPL MEASUREMENTS The TRPL spectra were conducted on a FLS980 fluorescence


spectrometer. Specifically, the catalyst was dispersed in deionized water and sonicated to form a uniform dispersion (1 mg/mL). Then, 1 mL of the dispersion was placed into a cuvette,


excited at 380 nm, and decay transients were recorded at an emission wavelength of 610 nm. For TRPL testing with different atmospheres, the corresponding gas was bubbled through the


dispersion until saturation before the test, and the gas was continuously purged over the liquid surface during the test. All other testing conditions remained the same. EPR MEASUREMENTS The


EPR measurements were carried out at X-band on a Bruker EMX-10/12 EPR spectrometer. For the solid EPR tests, the MOF sample was prepared by loading 25 mg into a J. Young quartz tube (outer


diameter, 4 mm; inner diameter, 2.8 mm). The gas in the tube was replaced with Ar, and measurements were taken, including EPR spectra under dark conditions and after 2 min and 5 min of


illumination. The EPR spectra under an oxygen atmosphere were measured by replacing Ar with O2. Additionally, the radical signals were tested in a dispersion system. In these experiments,


DMPO was used as the spin-trapping reagent. As a nitrone spin trap, DMPO is widely employed to capture short-lived radicals by forming more stable radicals, thereby generating a significant


EPR signal. Specifically, mix 50 µL of the catalyst dispersion (1 mg/mL) with 10 µL of DMPO solution (10%) in the dark. Introduce the reaction gas (O2 or CH4) into the system, and then


collect in-situ EPR spectra under visible light. For the detection of •OOH, methanol was used as the solvent for the system, while deionized water was used as the solvent for •OH and •CH3.


All reagents and test tubes were deoxygenated with Ar gas before use. COMPUTATIONAL DETAILS The first-principles calculation of spin polarization based on DFT implemented by CP2K/Quickstep


package52 was performed with the plane wave cutoff of 350 Ry. The Gaussian basis set consisting of a double-ζ with one set of polarization functions (DZVP)53 was used to optimize structures.


The Perdew-Burke-Ernzerhof exchange-correlation functional54 with the approach of Grimme (DFT-D3)55 was adopted. Due to the large size of the model, single gamma point grid sampling was


used. The analysis of the excited state was finished via the Multiwfn code56. DATA AVAILABILITY All the data that support the findings of this study are available within the paper and its


Supplementary Information files. Source data are provided with this paper. REFERENCES * Ravi, M., Ranocchiari, M. & van Bokhoven, J. A. The direct catalytic oxidation of methane to


methanol—a critical assessment. _Angew. Chem. Int. Ed._ 56, 16464–16483 (2017). Article  CAS  Google Scholar  * Schwach, P., Pan, X. & Bao, X. Direct conversion of methane to value-added


chemicals over heterogeneous catalysts: challenges and prospects. _Chem. Rev._ 117, 8497–8520 (2017). Article  CAS  PubMed  Google Scholar  * Gan, Y. et al. Highly selective photocatalytic


methane oxidation to methanol using CO2 as a soft oxidant. _ACS Sustain. Chem. Eng._ 11, 5537–5546 (2023). Article  CAS  Google Scholar  * Saputera, W. H., Yuniar, G. & Sasongko, D.


Light-driven methane conversion: unveiling methanol using a TiO2/TiOF2 photocatalyst. _RSC Adv._ 14, 8740–8751 (2024). Article  ADS  CAS  PubMed  PubMed Central  Google Scholar  * Luo, L. et


al. Water enables mild oxidation of methane to methanol on gold single-atom catalysts. _Nat. Commun._ 12, 1218 (2021). Article  ADS  CAS  PubMed  PubMed Central  Google Scholar  * Fan, Y.


et al. Selective photocatalytic oxidation of methane by quantum-sized bismuth vanadate. _Nat. Sustain._ 4, 509–515 (2021). Article  Google Scholar  * Sushkevich, V. L., Palagin, D.,


Ranocchiari, M. & Van Bokhoven, J. A. Selective anaerobic oxidation of methane enables direct synthesis of methanol. _Science_ 356, 523–527 (2017). Article  ADS  CAS  PubMed  Google


Scholar  * Ravi, M. et al. Misconceptions and challenges in methane-to-methanol over transition-metal-exchanged zeolites. _Nat. Catal._ 2, 485–494 (2019). Article  CAS  Google Scholar  *


Kwon, Y., Kim, T. Y., Kwon, G., Yi, J. & Lee, H. Selective activation of methane on single-atom catalyst of rhodium dispersed on zirconia for direct conversion. _J. Am. Chem. Soc._ 139,


17694–17699 (2017). Article  CAS  PubMed  Google Scholar  * Yuniar, G. et al. Recent advances in photocatalytic oxidation of methane to methanol. _Molecules_ 27, 5496 (2022). Article  CAS 


PubMed  PubMed Central  Google Scholar  * Liu, Z. et al. Photocatalytic conversion of methane: current state of the art, challenges, and future perspectives. _ACS Environ. Au_ 3, 252–276


(2023). Article  CAS  PubMed  PubMed Central  Google Scholar  * Li, X., Wang, C. & Tang, J. Methane transformation by photocatalysis. _Nat. Rev. Mater._ 7, 617–632 (2022). Article  ADS 


CAS  Google Scholar  * Song, H., Meng, X., Wang, Z.-J., Liu, H. & Ye, J. Solar-energy-mediated methane conversion. _Joule_ 3, 1606–1636 (2019). Article  CAS  Google Scholar  * Song, S.


et al. A selective Au-ZnO/TiO2 hybrid photocatalyst for oxidative coupling of methane to ethane with dioxygen. _Nat. Catal._ 4, 1032–1042 (2021). Article  CAS  Google Scholar  * Zheng, K. et


al. Room-temperature photooxidation of CH4 to CH3OH with nearly 100% selectivity over hetero-ZnO/Fe2O3 porous nanosheets. _J. Am. Chem. Soc._ 144, 12357–12366 (2022). Article  CAS  PubMed 


Google Scholar  * Jiang, Y. et al. Enabling specific photocatalytic methane oxidation by controlling free radical type. _J. Am. Chem. Soc._ 145, 2698–2707 (2023). Article  CAS  PubMed 


Google Scholar  * Luo, L. et al. Nearly 100% selective and visible-light-driven methane conversion to formaldehyde via. single-atom Cu and Wδ+. _Nat. Commun._ 14, 2690 (2023). Article  ADS 


CAS  PubMed  PubMed Central  Google Scholar  * Song, H. et al. Atomically dispersed nickel anchored on a nitrogen‐doped carbon/TiO2 composite for efficient and selective photocatalytic CH4


oxidation to oxygenates. _Angew. Chem. Int. Ed._ 135, e202215057 (2023). Article  Google Scholar  * Jiang, Y., Fan, Y., Li, S. & Tang, Z. Photocatalytic methane conversion: insight into


the mechanism of C(sp3)-H bond activation. _CCS Chem._ 5, 30–54 (2023). Article  CAS  Google Scholar  * Li, Q., Ouyang, Y., Li, H., Wang, L. & Zeng, J. Photocatalytic conversion of


methane: recent advancements and prospects. _Angew. Chem. Int. Ed._ 134, e202108069 (2022). Article  Google Scholar  * An, B. et al. Cooperative copper centres in a metal-organic framework


for selective conversion of CO2 to ethanol. _Nat. Catal._ 2, 709–717 (2019). Article  CAS  Google Scholar  * 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._ 145, 6156–6165 (2023). Article  CAS  PubMed  Google Scholar  * An, B. et al. Direct


photo-oxidation of methane to methanol over a mono-iron hydroxyl site. _Nat. Mater._ 21, 932–938 (2022). Article  ADS  CAS  PubMed  Google Scholar  * Feng, C., Wu, Z. P., Huang, K. W., Ye,


J. & Zhang, H. Surface modification of 2D photocatalysts for solar energy conversion. _Adv. Mater._ 34, 2200180 (2022). Article  CAS  Google Scholar  * Teng, Z. et al. Atomically


dispersed antimony on carbon nitride for the artificial photosynthesis of hydrogen peroxide. _Nat. Catal._ 4, 374–384 (2021). Article  CAS  Google Scholar  * Garibay, S. J. & Cohen, S.


M. Isoreticular synthesis and modification of frameworks with the UiO-66 topology. _Chem. Commun._ 46, 7700–7702 (2010). Article  CAS  Google Scholar  * Ma, X. et al. Modulating coordination


environment of single-atom catalysts and their proximity to photosensitive units for boosting MOF photocatalysis. _J. Am. Chem. Soc._ 143, 12220–12229 (2021). Article  CAS  PubMed  Google


Scholar  * Dong, J. et al. Continuous electroproduction of formate via CO2 reduction on local symmetry-broken single-atom catalysts. _Nat. Commun._ 14, 6849 (2023). Article  ADS  CAS  PubMed


  PubMed Central  Google Scholar  * He, Y., Li, C., Chen, X.-B., Shi, Z. & Feng, S. Visible-light-responsive UiO-66 (Zr) with defects efficiently promoting photocatalytic CO2 reduction.


_ACS Appl. Mater. Inter._ 14, 28977–28984 (2022). Article  CAS  Google Scholar  * Han, Y. et al. Control of the pore chemistry in metal-organic frameworks for efficient adsorption of benzene


and separation of benzene/cyclohexane. _Chem_ 9, 739–754 (2023). Article  CAS  Google Scholar  * Bruzzese, P. C. et al. 17O-EPR determination of the structure and dynamics of copper


single-metal sites in zeolites. _Nat. Commun._ 12, 4638 (2021). Article  ADS  CAS  PubMed  PubMed Central  Google Scholar  * Ikuno, T. et al. Methane oxidation to methanol catalyzed by


Cu-oxo clusters stabilized in NU-1000 metal-organic framework. _J. Am. Chem. Soc._ 139, 10294–10301 (2017). Article  CAS  PubMed  Google Scholar  * Feng, N. et al. Efficient and selective


photocatalytic CH4 conversion to CH3OH with O2 by controlling overoxidation on TiO2. _Nat. Commun._ 12, 4652 (2021). Article  ADS  CAS  PubMed  PubMed Central  Google Scholar  * Luo, L. et


al. Synergy of Pd atoms and oxygen vacancies on In2O3 for methane conversion under visible light. _Nat. Commun._ 13, 2930 (2022). Article  ADS  CAS  PubMed  PubMed Central  Google Scholar  *


Fu, L. et al. Highly selective conversion of CH4 to high value‐added C1 oxygenates over Pd loaded ZnTi‐LDH. _Adv. Energy Mater._ 13, 2301118 (2023). Article  CAS  Google Scholar  * Jiang,


Y. et al. Elevating photooxidation of methane to formaldehyde via TiO2 crystal phase engineering. _J. Am. Chem. Soc._ 144, 15977–15987 (2022). Article  CAS  PubMed  Google Scholar  * Fu, Y.


et al. An amine‐functionalized titanium metal-organic framework photocatalyst with visible-light-induced activity for CO2 reduction. _Angew. Chem. Int. Ed._ 51, 3364–3367 (2012). Article 


CAS  Google Scholar  * Zhang, R. et al. Direct photocatalytic methane oxidation to formaldehyde by N doping Co-decorated mixed crystal TiO2. _ACS Nano_ 18, 12994–13005 (2024). Article  CAS 


PubMed  Google Scholar  * Tan, H. et al. Photocatalysis of water into hydrogen peroxide over an atomic Ga-N5 site. _Nat. Synth._ 2, 557–563 (2023). Article  ADS  Google Scholar  * Amano, F.


et al. Photoelectrochemical homocoupling of methane under blue light irradiation. _ACS Energy Lett._ 4, 502–507 (2019). Article  CAS  Google Scholar  * Li, J. et al. Self-adaptive


dual-metal-site pairs in metal-organic frameworks for selective CO2 photoreduction to CH4. _Nat. Catal._ 4, 719–729 (2021). Article  CAS  Google Scholar  * Parvatkar, P. T. et al. A tailored


COF for visible-light photosynthesis of 2, 3-dihydrobenzofurans. _J. Am. Chem. Soc._ 145, 5074–5082 (2023). Article  CAS  PubMed  PubMed Central  Google Scholar  * Bai, S. et al.


High-efficiency direct methane conversion to oxygenates on a cerium dioxide nanowires supported rhodium single-atom catalyst. _Nat. Commun._ 11, 954 (2020). Article  ADS  CAS  PubMed  PubMed


Central  Google Scholar  * Cheng, Q. et al. Maximizing active Fe species in ZSM-5 zeolite using organic-template-free synthesis for efficient selective methane oxidation. _J. Am. Chem.


Soc._ 145, 5888–5898 (2023). Article  CAS  PubMed  PubMed Central  Google Scholar  * He, Y. et al. In situ identification of reaction intermediates and mechanistic understandings of methane


oxidation over hematite: a combined experimental and theoretical study. _J. Am. Chem. Soc._ 142, 17119–17130 (2020). Article  CAS  PubMed  Google Scholar  * Li, X. et al. Efficient hole


abstraction for highly selective oxidative coupling of methane by Au-sputtered TiO2 photocatalysts. _Nat. Energy_ 8, 1013–1022 (2023). Article  ADS  CAS  Google Scholar  * Zhang, X. et al.


Developing Ni single-atom sites in carbon nitride for efficient photocatalytic H2O2 production. _Nat. Commun._ 14, 7115 (2023). Article  ADS  CAS  PubMed  PubMed Central  Google Scholar  *


Zhou, X.-K., Li, Y., Luo, P.-P. & Lu, T.-B. Synergy of surface phosphates and oxygen vacancies enables efficient photocatalytic methane conversion at room temperature. _ACS Appl. Mater.


Inter._ 15, 36280–36288 (2023). Article  CAS  Google Scholar  * Feng, C. et al. Cooperative tungsten centers in a polymetric carbon nitride for efficient overall photosynthesis of hydrogen


peroxide. _Energy Environ. Sci._ 17, 1520–1530 (2024). Article  CAS  Google Scholar  * Feng, C. et al. Synthesis of leaf‐vein‐like g‐C3N4 with tunable band structures and charge transfer


properties for selective photocatalytic H2O2 evolution. _Adv. Funct. Mater._ 30, 2001922 (2020). Article  CAS  Google Scholar  * Feng, C. et al. Regulating photocatalytic CO2 reduction


kinetics through modification of surface coordination sphere. _Adv. Funct. Mater._ 34, 2309761 (2024). Article  CAS  Google Scholar  * VandeVondele, J. et al. Quickstep: fast and accurate


density functional calculations using a mixed Gaussian and plane waves approach. _Comput. Phys. Commun._ 167, 103–128 (2005). Article  ADS  CAS  Google Scholar  * VandeVondele, J. &


Hutter, J. Gaussian basis sets for accurate calculations on molecular systems in gas and condensed phases. _J. Chem. Phys._ 127, 114105 (2007). Article  ADS  PubMed  Google Scholar  *


Perdew, J. P., Burke, K. & Ernzerhof, M. Generalized gradient approximation made simple. _Phys. Rev. Lett._ 77, 3865–3868 (1996). Article  ADS  CAS  PubMed  Google Scholar  * Grimme, S.,


Antony, J., Ehrlich, S. & Krieg, H. A consistent and accurate ab initio parametrization of density functional dispersion correction (DFT-D) for the 94 elements H-Pu. _J. Chem. Phys._


132, 154104 (2010). Article  ADS  PubMed  Google Scholar  * Lu, T. & Chen, F. Multiwfn: a multifunctional wavefunction analyzer. _J. Comput. Chem._ 33, 580–592 (2012). Article  PubMed 


Google Scholar  Download references ACKNOWLEDGEMENTS This work was supported by King Abdullah University of Science and Technology and Center of Excellence for Renewable Energy and Storage


Technologies under award number 5937, the National Key Research and Development Program of China (2022YFE0113800), and the National Natural Science Foundation of China (22122505, 22075250,


21771161). AUTHOR INFORMATION AUTHORS AND AFFILIATIONS * Center for Renewable Energy and Storage Technologies (CREST), Physical Science and Engineering Division, King Abdullah University of


Science and Technology, Thuwal, Saudi Arabia Chengyang Feng, Shouwei Zuo, Miao Hu, Yuanfu Ren, Chen Zou & Huabin Zhang * KAUST Catalysis Center (KCC), Division of Physical Science and


Engineering, King Abdullah University of Science and Technology (KAUST), Thuwal, Saudi Arabia Chengyang Feng, Shouwei Zuo, Miao Hu, Yuanfu Ren, Chen Zou, Magnus Rueping & Huabin Zhang *


Center for Electron Microscopy, Institute for Frontier and Interdisciplinary Sciences, State Key Laboratory Breeding Base of Green Chemistry Synthesis Technology and College of Chemical


Engineering, Zhejiang University of Technology, Hangzhou, Zhejiang, China Liwei Xia & Yihan Zhu * State Key Laboratory of Featured Metal Materials and Life-cycle Safety for Composite


Structures, MOE Key Laboratory of New Processing Technology for Nonferrous Metals and Materials, and School of Resources, Environment and Materials, Guangxi University, Nanning, China Jun


Luo * State Key Laboratory of Photocatalysis on Energy and Environment, College of Chemistry, Fuzhou University, Fuzhou, China Sibo Wang * Electron Microscopy Center, South China University


of Technology, Guangzhou, China Yu Han Authors * Chengyang Feng View author publications You can also search for this author inPubMed Google Scholar * Shouwei Zuo View author publications


You can also search for this author inPubMed Google Scholar * Miao Hu View author publications You can also search for this author inPubMed Google Scholar * Yuanfu Ren View author


publications You can also search for this author inPubMed Google Scholar * Liwei Xia View author publications You can also search for this author inPubMed Google Scholar * Jun Luo View


author publications You can also search for this author inPubMed Google Scholar * Chen Zou View author publications You can also search for this author inPubMed Google Scholar * Sibo Wang


View author publications You can also search for this author inPubMed Google Scholar * Yihan Zhu View author publications You can also search for this author inPubMed Google Scholar * Magnus


Rueping View author publications You can also search for this author inPubMed Google Scholar * Yu Han View author publications You can also search for this author inPubMed Google Scholar *


Huabin Zhang View author publications You can also search for this author inPubMed Google Scholar CONTRIBUTIONS C.F., M.R., and H.Z. constructed and planned the whole project. C.F. carried


out the synthesis of the samples and photocatalytic experiments. S.Z. carried out the XANES and EXAFS characterizations. M.H. performed the EPR tests. Y.R. performed electron microscope


imaging. J.L. conducted the DFT calculations. L.X. and C.Z. performed the ultrahigh-resolution HAADF STEM imaging under the guidance of Y.Z. and Y.H. S.W. helped to discuss the reaction


mechanism and improve the manuscript. C.F. wrote the manuscript. H.Z. reviewed and edited the manuscript. All authors discussed the results and commented on the manuscript. CORRESPONDING


AUTHOR Correspondence to Huabin Zhang. ETHICS DECLARATIONS COMPETING INTERESTS The authors declare no competing interests. PEER REVIEW PEER REVIEW INFORMATION _Nature Communications_ thanks


Wibawa Hendra Saputera, Hai-Ying Jiang and the other, anonymous, reviewers for their contribution to the peer review of this work. A peer review file is available. ADDITIONAL INFORMATION


PUBLISHER’S NOTE Springer Nature remains neutral with regard to jurisdictional claims in published maps and institutional affiliations. SUPPLEMENTARY INFORMATION SUPPLEMENTARY INFORMATION


PEER REVIEW FILE SOURCE DATA SOURCE DATA RIGHTS AND PERMISSIONS OPEN ACCESS This article is licensed under a Creative Commons Attribution-NonCommercial-NoDerivatives 4.0 International


License, which permits any non-commercial use, sharing, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the


source, provide a link to the Creative Commons licence, and indicate if you modified the licensed material. You do not have permission under this licence to share adapted material derived


from this article or parts of it. The images or other third party material in this article are included in the article’s Creative Commons licence, unless indicated otherwise in a credit line


to the material. If material is not included in the article’s Creative Commons licence and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will


need to obtain permission directly from the copyright holder. To view a copy of this licence, visit http://creativecommons.org/licenses/by-nc-nd/4.0/. Reprints and permissions ABOUT THIS


ARTICLE CITE THIS ARTICLE Feng, C., Zuo, S., Hu, M. _et al._ Optimizing the reaction pathway of methane photo-oxidation over single copper sites. _Nat Commun_ 15, 9088 (2024).


https://doi.org/10.1038/s41467-024-53483-z Download citation * Received: 21 May 2024 * Accepted: 14 October 2024 * Published: 21 October 2024 * DOI:


https://doi.org/10.1038/s41467-024-53483-z SHARE THIS ARTICLE Anyone you share the following link with will be able to read this content: Get shareable link Sorry, a shareable link is not


currently available for this article. Copy to clipboard Provided by the Springer Nature SharedIt content-sharing initiative