Performance of hydrogenation of CO 2 to monocarboxylic acids. Ni–Zn alloy catalysts with different Ni/Zn molar ratios were synthesized via the coprecipitation of Ni(NO3)2 and Zn(NO3)2 precursors followed by calcination at different temperatures in the range of 500–900 °C for 6 h under air flow and subsequent reduction at 450 °C for 10 h under 5% H2/Ar flow. The CO2 hydrogenation reactions were performed over Ni–Zn alloy catalysts at 325 °C, 3.0 MPa, using a H2/CO2 ratio of 1:2, and gas hourly space velocity (GHSV) of 5400 mL g− 1 h− 1. The Ni–Zn alloy catalysts with Ni/Zn ratios of 1:3 calcined at 500, 600, 700, 800, and 900 °C, followed by reduction were denoted N1Z3-500, N1Z3-600, N1Z3-700, N1Z3-800, and N1Z3-900, respectively. Among these, the N1Z3-900 catalyst presented high AA and PA selectivities of 58.9% and 18.2%, respectively (excluding CO) at a CO2 conversion of 13.4% (Fig. 1a). The amounts of formic acid and C4+ acids (e.g., butyric acid, valeric acid) formed were negligible (< 1.5%). The total monocarboxylic acid selectivity of the reaction reached 77.1% (excluding CO). Conversely, the selectivities toward CH4 and C2–C4 were highly suppressed to 17.0% and 5.0%, respectively. Consequently, high space time yields (STYs) of 0.405 and 0.125 mmol g− 1 h− 1 were achieved for AA and PA, respectively (Supplementary Table 1). Although numerous reports have been published on the direct hydrogenation of CO2 to methanol5, 42–45, C5+ liquid hydrocarbons46–50, and aromatics19, 20, 51–53, studies dedicated to the direct CO2 conversion to monocarboxylic acids have been extremely rare. A single article, which described the possibility of forming AA over a Ag-promoted Rh/SiO2 catalyst, was published in the early 1990s34; however, the reported production rate of AA was very low (~ 0.035 mmol g− 1 h− 1). Furthermore, this is the first reported synthesis of PA via the direct catalytic hydrogenation of CO2. Even though ZnO is known to be an active catalyst for the synthesis of methanol via the hydrogenation of CO and CO2, the formation of methanol or higher alcohols over Ni–Zn alloy catalysts has not been reported6, 54, 55.
To analyze the effect of the calcination temperature on the CO2 hydrogenation, we used Ni–Zn alloy catalysts calcined at different temperatures in the range of 500–1000 °C (Fig. 1a). As the calcination temperature increased from 500 to 700 °C, the CO2 conversion slightly increased from 9.5–10.1%, whereas the AA, PA, and CH4 selectivities increased significantly from 39.5–61.7%, increased significantly from 8.4–15.2%, and decreased significantly from 51.0–21.8%, respectively. Further increasing the calcination temperature to 1000 °C resulted in the increase in CO2 conversion to 14.7%, slight decrease in AA selectivity to 56.1%, and slight increase in PA selectivity to 20.9%. For all catalysts calcined in the temperature range of 500–1000 °C, the C2–C4 selectivity was suppressed, and ranged between 1.0–7.1%. As discussed below, the formation of the Ni–Zn alloy, which played an important role in increasing the selectivity of the reaction toward the formation of monocarboxylic acids, occurred at calcination temperatures above 700 °C.
To further elucidate the role of Ni–Zn alloys for producing monocarboxylic acids with high selectivity, the performance of the N1Zn-900 (n = 2, 3, 4) catalysts was compared with that of monometallic ZnO, and the results are presented in Fig. 1b. The use of the monometallic ZnO catalyst prepared via precipitation resulted in a very low CO2 conversion of 2.0% and high methanol selectivity of 81.3%; moreover, no monocarboxylic acids were formed. Therefore, Ni and Zn presented a strong synergistic effect during the selective conversion of CO2 to monocarboxylic acids. To investigate the synergistic effect, Ni–Zn alloy catalysts with different Ni/Zn ratios were tested for the CO2 hydrogenation reaction. As the Ni/Zn ratios changed from 1:2 to 1:3, the AA and CH4 selectivities increased from 47.2–58.9% and decreased from 28.7–17.0%, respectively, which indicated that the methanation reaction over the Ni-deficient catalyst was suppressed to some extent. This suggested that the Ni/Zn ratios of the Ni–Zn alloy catalysts should be carefully adjusted to obtain catalysts with active sites that could lead to the selective production of monocarboxylic acids. As the Ni/Zn ratio further changed from 1:3 to 1:4, the changes in CO2 conversion and product selectivity were negligible.
The N1Z3-900 catalyst presented excellent long-term stability for the CO2 conversion reaction and remarkable selectivity toward monocarboxylic acids (Fig. 1c). The initial CO2 conversion activity of the N1Z3-900 catalyst was maintained for up to 216 h on-stream. After 216 h on-stream, the CH4 and AA + PA selectivities slightly increased from 17.0–21.6% and slightly decreased from 77.1–71.5%, respectively. However, even after the long-term catalytic run, the production of methanol and long-chain alcohols were not observed.
Process optimization. The aforementioned experiments demonstrated that the use of the N1Zn-900 (n = 2, 3, 4) catalysts produced monocarboxylic acids directly via the hydrogenation of CO2 with high selectivity. For the process optimization experiments, CO2 hydrogenation was performed under different reaction conditions to optimize the CO2 conversion and monocarboxylic acid selectivity. First, the effect of temperature on the CO2 conversion over the N1Z3-900 catalyst was analyzed (Supplementary Fig. 3a). As the temperature was increased from 280 to 325 °C, the CO2 conversion increased from 8.8–13.4% and the AA + PA selectivity (excluding CO) increased from 64.3–77.1%. As the temperature was further increased to 350 °C, the CO2 conversion, AA + PA selectivity, and CH4 selectivity slightly increased to 13.6%, decreased significantly to 55.4%, and increased significantly from 17.0–38.5%, respectively. The increase in CH4 selectivity with the temperature could be attributed to the high availability of dissociated H on the surface of the catalyst. When the pressure was increased from 1.5 to 3.0 MPa, the CO2 conversion increased from 8.6–13.4% and the change in AA + PA selectivity was negligible (Supplementary Fig. 3b). At a high pressure of 4.0 MPa, the AA + PA and CH4 selectivities decreased to 60.8% and increased to 35.1%, respectively. The increase in CH4 formation at high H2 partial pressure was ascribed to the increase in the availability of dissociated H on the surface of the catalyst, which increased the possibility of hydrogenation of the *CO species (* denotes surface-adsorbed species). As the H2/CO2 ratio increased from 1:1 to 2:1, the CO selectivity decreased from 86.5–63.4% and the CO2 conversion and AA + PA selectivity remained almost the same (Supplementary Fig. 3c). The increase in the H2/CO2 ratio to 3:1 resulted in the decrease in the AA + PA selectivity to 45.3% and significant increase in the CH4 selectivity to 45.7%, which was caused by the enhanced methanation reaction under a sufficient H2 supply. As the GHSV increased from 2800 to 8000 mL g− 1 h− 1, the CO2 conversion, AA + PA selectivity, and CO selectivity decreased from 18.6–11.1%, decreased from 80–65.1%, and increased from 63.7–72.6%, respectively (Supplementary Fig. 3d). The enhanced desorption of *CO adsorbed on the catalyst surface at high GHSV could be responsible for the decrease in AA + PA selectivity.
Characterization of active phases. To analyze the active sites for the effective conversion of CO2 to monocarboxylic acids, the crystallinity and phase structure of the catalysts were examined using XRD analysis. We assumed that a new alloy-based active phase was formed during the calcination and reduction of the catalysts because neither metallic Ni0 nor ZnO were the active sites for the formation of the monocarboxylic acids. The XRD patterns of the N1Z3 catalysts calcined at temperatures in the range of 500–900 °C under air flow, followed by reduction at 450 °C under pure H2 flow, are presented in Supplementary Fig. 4. The peaks at 36.5°, 57.2°, and 63.1°, which were associated with the respective (101), (110), and (103) planes of pure ZnO, and the 44.7° peak, which was ascribed to the (111) plane of pure metallic Ni0, shifted toward lower 2θ angles in the XRD spectra of the reduced N1Z3 catalysts (Supplementary Figs. 4c–e). In addition, a new peak at 43.0° appeared in the XRD spectrum of the N1Z3-900 catalyst. This suggested that the intermetallic diffusion of the Zn and Ni atoms occurred during the high-temperature calcination and subsequent reduction, which led to the formation of a Ni–Zn alloy phase. The migration of Ni atoms into the ZnO phase during calcination formed the Zn-rich NixZnyO phase56. The XRD peaks associated with the Zn-rich NixZnyO phase became sharper and narrower as the calcination temperature increased. The decrease in the full-width at half-maximum indicated that the crystallinity of the Zn-rich NixZnyO phase increased at high calcination temperatures, and the shifting of the peak positions toward lower 2θ angles suggested that the Ni atoms in the Zn-rich NixZnyO phase were located in the interstitial sites of ZnO57 − 59. Further, the shifting of the Ni0 peak toward a lower 2θ angle by approximately 0.5° indicated the partial dissolution of Zn0 in the Ni0 crystalline structure60. The diffusion of Zn into the NiO phase during calcination followed by reduction could produce the Ni–Zn alloy phase. The crystal structure of the Ni–Zn alloy with a Ni/Zn ratio of 1:3, which was predicted using the Reflex plus program of the Material Studio software, was estimated to be Ni4Zn22 (Supplementary Table 2, Supplementary Fig. 4b, Supplementary Fig. 5b)
To further investigate the formation mechanism of the Ni–Zn alloys, in situ XRD patterns of the N1Z3 precursor collected under calcination and reduction conditions were analyzed. First, the intermetallic diffusion of Ni and Zn into the ZnO and NiO phases, respectively, was investigated by collecting XRD patterns during the calcination of the co-precipitated N1Zn3 precursor at temperatures in the range of 27–900 °C under air flow (Supplementary Fig. 6). As the calcination temperature increased to approximately 230 °C, dehydration of the co-precipitated Ni(CO3)3 and Zn(CO3)2 species and subsequent formation of the NiO and ZnO phases were observed. As the calcination temperature further increased to 900 °C, the positions of the peaks of the pure ZnO and NiO phases progressively shifted toward lower 2θ angles, which indicated that Zn and Ni diffused into the interstitial sites of NiO and ZnO, respectively. During calcination, the Zn atoms in the four-coordinated tetrahedral sites migrated into the tetrahedral NiO4 sites and formed a Ni-rich NixZnyO phase. Because the ionic radius of Zn2+ (74 pm) was larger than that of Ni2+ (69 pm), the unit cell of NiO was expanded, which resulted in the shifting of the peaks ascribed to the (111), (200), and (220) planes of pure NiO toward lower 2θ angles. The unit cell expansion of the Zn-rich NixZnyO phase implied that Ni atoms could be present in the interstitial lattice sites of the ZnO phase.
After the N1Zn3 sample calcined at 900 °C was cooled to 25 °C, in situ XRD patterns were collected as the temperature was increased to 450 °C under H2 flow (Supplementary Fig. 7). The peaks of pure metallic Ni0 (44.8°, 52.1°, and 76.9°, Supplementary Fig. 3) were not observed in the XRD pattern of the reduced N1Zn3-900 catalyst, which indicated that pure Ni0 phase did not form during the reduction process. Instead, the peaks associated with the (111) and (220) planes of the Ni-rich NixZnyO phase shifted toward lower 2θ angles, from 37.1° to 36.8° and from 62.9° to 62.7°, respectively. In addition, the 43.1° peak ascribed to the (200) plane of the Ni-rich NixZnyO phase shifted to 42.7° in the XRD pattern of the Ni–Zn alloy. The peaks associated with the (101) and (103) planes of the Zn-rich NixZnyO phase shifted toward lower 2θ angles—from 36.4° to 36.2° and from 62.5° to 62.2°, respectively.
The local chemical environments of Ni and Zn in the N1Z3-700 and N1Z3-900 catalysts were analyzed using X-ray absorption spectroscopy (XAS). The Ni K-edge X-ray absorption near edge structure (XANES) spectra of N1Z3-700 and N1Z3-900 were similar to that of a reference Ni foil, which confirmed the presence of Ni0 in the structure of the catalysts (Fig. 2a). A close inspection of the XANES profile revealed that the onset of the adsorption edge downshifted slightly compared to that of the Ni foil (inset of Fig. 2a). In addition, the height of the white line peak of the N1Z3-900 catalyst was lower than that of the Ni foil. These results indicated that Ni in the N1Z3-900 catalyst presented a slightly negative charge compared to that of Ni in the Ni foil, which was attributed to the electron transfer from Zn to Ni. The Fourier-transform magnitudes of the Ni K-edge extended X-ray absorption fine structure (EXAFS) spectra of the catalysts were similar to that of Ni foil (Fig. 2b). This was ascribed to the similar scattering parameters of Ni and Zn owing to the close proximity of Ni and Zn in the periodic table61. The two main peaks of NiO at 1.65 and 2.57 Å, which were assigned to the Ni–O and Ni–Ni bonds in the rocksalt crystal structure, respectively62, were not observed in the profile of the N1Z3 catalysts. Moreover, a peak at 2.2 Å appeared in the spectrum of the catalysts, which indicated the presence of Ni or Zn in the nearest Ni neighbors. The Zn K-edge XANES spectra of the N1Z3-700 and N1Z3-900 catalysts are illustrated in Fig. 2c. The onset of the adsorption edge downshifted and the height of the white line peak decreased compared to that of the ZnO reference, which indicated that the valence of Zn in the catalysts was lower than that in ZnO. The Zn K-edge EXAFS spectra presented two main peaks centered at 1.55 and 2.93 Å, which corresponded to the Zn–O and Zn–Ni(Zn) bond lengths. A new peak centered at 4.08 Å was observed in the spectra of the N1Z3-700 and N1Z3-900 catalysts and was assigned to the Zn–Ni coordination.
The morphology of the N1Z3 catalysts calcined at different temperatures was analyzed using field-emission scanning electron microscopy (FE–SEM) analysis, and the images are presented in Supplementary Figs. 8 and 9. Prior to calcination, the coprecipitated Ni(CO3)3 and Zn(CO3)2 particles were needle-shaped (Supplementary Fig. 8a). At a low calcination temperature of 500 °C, agglomerated spherical-shape particles with an average particle size of 25 nm formed (Supplementary Fig. 8b), and the degree of interparticle agglomeration increased at 900 °C (Supplementary Fig. 8c). The elemental mapping and energy-dispersive X-ray spectroscopy (EDX) analysis results (Supplementary Figs. 8d–g) revealed that the Ni and Zn species were well-distributed in the co-precipitated precursor and the Ni/Zn ratio was approximately 1:3. After the reduction of the calcined catalysts at 450 °C under H2 flow, no noticeable changes were observed in the morphology of the catalyst (Supplementary Fig. 8).The primary particle size of the catalysts increased from 25–30 to 100–400 nm as the calcination temperature increased from 500 °C (N1Z3-500) to 900 °C (N1Z3-900). For the N1Z3-900 catalyst, the primary particles were highly agglomerated and converted to secondary micron-sized particles (Supplementary Fig. 9e). Furthermore, as the calcination temperature increased from 500 to 900 °C, the Brunauer–Emmet–Teller (BET) surface area of the catalysts decreased from 42.1 to 14.6 m2 g− 1 (Supplementary Fig. 10 and Supplementary Table 3).
The diffusion of Zn into the NiO phase during calcination and formation of the Ni–Zn alloy during the reduction were analyzed using high-resolution transmission electron microscopy (HR–TEM) and high-angle angular dark field–scanning transmission electron microscopy (HAADF–STEM), and the results are presented in Fig. 3. The formation of uniform nano-sized 10–30 nm particles was observed at the surface of the 100–400 nm spherical ZnO particles. Some nano-sized particles were in close contact with each other, which suggested that the nanoparticles migrated during calcination owing to their high surface energy. The Zn and O species were well-distributed throughout the micron- and nano-sized particles. In contrast, the Ni species were predominantly distributed throughout the nano-sized particles, and only a few Ni species were distributed throughout the Zn-rich micron-sized particles. This indicated that most micron- and nano-sized particles of the N1Z3-900 catalyst were low-level Ni-doped ZnO and Ni–Zn alloy, respectively. Two different lattice parameters of the nano-sized Ni–Zn alloy particles, 7.71 and 2.29 Å, were observed in the HR–TEM images of the N1Z3-900 catalyst. The presence of the enlarged interplane spacing of 7.71 Å, which was approximately three times larger than that of pure ZnO (2.5 Å63), suggested the formation of a Ni–Zn alloy64. In addition, the interlayer spacing of 2.29 Å of the N1Z3-900 catalyst was larger than that of pure metallic Ni (2.03 Å), which agreed well with the XRD results. To further investigate the formation of the Ni–Zn alloy, the elemental compositions of the nano-sized particles of the N1Z3-900 catalyst were analyzed using EDX line scanning (Supplementary Fig. 11). The elemental concentrations of Zn, Ni, and O measured over the lateral distance of 40 nm confirmed that the nano-sized particles on the surface of ZnO were O-deficient Ni–Zn alloy species. Unlike the Ni and Zn species in N1Z3-900, those in N1Z3-700 were well distributed throughout the catalyst (Supplementary Figs. 12a–l). The segregation of the nano-sized Ni–Zn alloy phases was first observed for the N1Z3-800 catalyst (Supplementary Figs. 12g–i). Therefore, highly active nano-sized Ni–Zn alloy particles were formed via calcination at the high temperatures of 800 and 900 °C followed by reduction at 450 °C.
We used the aforementioned results to propose a schematic of the Ni–Zn alloy formation (Fig. 4). During the calcination of the co-precipitated product at temperatures of up to 230 °C, Ni(CO3)3 and Zn(CO3)2 were transformed into NiO and ZnO, respectively. As the temperature further increased up to 900 °C, the intermetallic diffusion was enhanced and Ni- and Zn-rich NixZnyO phases formed. Furthermore, the degree of intermetallic diffusion increased with increasing the calcination temperature. The amount of Zn that could diffuse into NiO could reach 40 mol% when a mixture of ZnO and NiO with a ZnO/NiO ratio of 3:1 was calcined at 900 °C56. During the reduction reaction in the presence of H2, the Ni-rich NixZnyO phase lost its O either via the formation of water or the generation of O vacancies to form Ni–Zn alloy species. The structural rearrangement of the Ni and Zn atoms in the unit cell of the catalyst was examined using the Reflex plus program of the Material Studio software utilizing experimental XRD data (Supplementary Fig. 4b). The predicted atomic arrangement on the Ni–Zn alloy phase indicated that Ni-centered pentagonal and hexagonal structures could be present in the alloy phase, and majority of the Ni–Zn alloy species presented (111) and (200) facets (Fig. 4b).
To gain insight regarding the synthesis mechanisms of monocarboxylic acids, the chemical state of the N1Z3 catalyst under CO2 hydrogenation conditions was monitored using NAP–XPS, and the results are presented in Fig. 5. All peaks were referenced to the C 1 s spectrum at 284.8 eV. In the presence of a H2 and CO2 mixture at 0.2 mbar and 25 °C, the high-resolution Zn 2p spectra presented Zn 2p3/2 and Zn 2p1/2 peaks at 1021.8 and 1045.0 eV, respectively, which indicated the presence of Zn(2+)O at the surface of the catalyst (Fig. 4a). The divalent oxidation state of Zn(2+)O shifted toward higher binding energy levels when the reaction was performed at 325 °C and after the evacuation of gases at 325 °C; this could be attributed to the catalyst surface interacting with the electronegative C and O species. The metallic Ni0 peak at 852.6 eV was the major Ni species identified in the Ni 2p spectrum of the reduced N1Z3 catalyst (Fig. 5b). In addition, the Ni2+ 2p3/2 and Ni 2p1/2 peaks at 856.1 and 873.0 eV, respectively, and satellite peak associated with the Ni2+ 2p3/2 peak were observed. The intensity of the Ni2+ peaks decreased during the hydrogenation of CO2, owing to the reduction reaction promoted by the presence of H2. Under ultra-high vacuum (UHV) conditions, several sub-peaks were observed in the Ni 2p spectra of N1Zn3-900, which indicated that different chemical states of Ni were present in this catalyst (Supplementary Fig. S13b). The Ni 2p3/2 peak collected under UHV conditions was deconvoluted into three peaks: Ni0, Ni+, and Ni2+ at 852.0, 853.3, and 855.4 eV, respectively. The peaks of the Ni species in the N1Zn3-900 catalyst downshifted compared with those of Ni0, Ni2+, and Ni3+ (852.6, 853.7, and 855.6 eV, respectively)65, 66. This could be caused by the interaction of less electronegative Zn with Ni. Additionally, the increase in the electron density on the surface of the Ni atoms could be caused by the electrons donated by the O vacancies in the ZnO lattice67. Under CO2 hydrogenation conditions, the peaks of metallic Ni0 and Ni2+ upshifted, which could be caused by interaction of Ni with the O-containing intermediates (Fig. 5b). The O 1 s ultra-high vacuum (UHV) XPS profiles of the catalysts were deconvoluted into three sub-peaks at 529.0, 530.9, and 531.8 eV, which corresponded to the lattice oxygen anions (O2−) in the wurtzite structure, oxygen defects caused by O vacancies (Ox−), and adsorbed oxygen species at the surface (Oa)68, 69, respectively, (Supplementary Fig. S13c). The %areas of the O vacancies of the N1Z3-700 and N1Z3-900 catalysts were 34.3% and 44.7%, respectively. In the presence of H2/CO2 under the reaction conditions (Fig. 5c), the peaks associated with the O2− species shifted to 529.8 eV. This upshift, which was caused by the oxidation of the catalyst surface, could be ascribed to the surface O* being produced either via the formation of CO* or further CO* dissociation. In addition, the formation of CO2-adsorbed intermediates (e.g., carbonate, bicarbonate, and formate, which will be discussed further) caused the O vacancy peak to upshift to 531.1 eV via increasing the electron density. The C 1 s spectrum of the fresh N1Z3 catalyst obtained under UHV conditions presented two peaks at 284.8 and 289.3 eV, which corresponded to the surface organic contaminants and residual carbonate species on the surface of catalysts (Supplementary Fig. S13d)70, 71. When the catalyst was exposed to a mixture of H2 and CO2 at 25 °C, an additional peak at 292.7 eV appeared, which could be attributed to the gaseous CO2 in the in situ cell (Fig. 5d)72; this peak almost disappeared after the cell was evacuated. In the presence of H2 and CO2 at 325 °C, several adsorption peaks of reaction intermediates, which were produced via CO2 hydrogenation, were observed. The peaks at 291.4 and 288.2 eV corresponded to the adsorbed CO and formate species, respectively.73. In addition, a carboxylate peak was observed at 291.1 eV74.
To investigate the interactions between the surface Ni species and ZnO, the hydrogen temperature-programmed reduction and desorption (H2–TPR and H2–TPD, respectively) profiles of the N1Z3 catalysts were analyzed. The N1Z3 catalysts prepared at temperatures in the range of 500–800 °C exhibited three peaks at low, medium, and high temperatures (100–200, 200–500, and > 500 °C, respectively) in the H2–TPR profile (Supplementary Fig. 14a). Typically, the H2–TPR profile of bulk NiO particles, in the absence of supports, presents an asymmetric peak in the temperature range of 250–500 °C, which corresponds to the overlapped reduction of Ni3+ to Ni2+ and Ni2+ to Ni0 75–78. When NiO interacts strongly with a support (e.g., γ-Al2O3 or ZnO), the difficulty in reducing the NiO phase generates broad high-temperature peaks above 500 °C67, 79. The high-temperature reduction peaks could also be ascribed to the reduction of Zn2+ to metallic Zn in the presence of Ni, although bulk ZnO is barely reducible60, 80. For highly-dispersed NiO nanoparticles that interact weakly with their support, the reduction temperature range was lower (200–250 °C) than that of the bulk NiO particles79, 81. Thus, it was reasonable to assign the low-, medium-, and high-temperature peaks of the N1Z3 catalysts to the reduction of highly-dispersed NiO nanoparticles that interacted weakly with ZnO, bulk NiO phases, and NiO phases that interacted strongly with ZnO, respectively. The low-temperature peak at 230–290 °C, which were observed in the H2–TPR profiles of the N1Z3 catalysts prepared at low calcination temperatures of 500–800 °C shifted to higher temperatures in the range of 436–484 °C in the H2–TPR profile of N1Z3-900, and the peak intensity of the highly dispersed NiO nanoparticles at 175 °C decreased significantly. In addition, the amount of H2 consumed in the medium temperature range increased with increasing the calcination temperature (Supplementary Table 4). This indicated that the highly dispersed NiO nanoparticles were converted into Ni-rich NixZnyO phase via the enhanced diffusion of the Zn2+ cations into the NiO lattice during calcination. Further, high reduction temperatures were required to reduce the lattice O of the Ni-rich NixZnyO phase to form Ni–Zn alloy, because of the Zn2+ doping. The H2–TPD profiles of the N1Z3 catalysts presented a low-temperature desorption region (< 400 °C), which corresponded to the H atoms that were either weakly or moderately adsorbed on the surface of the catalyst, and a high-temperature desorption region (> 401–900 °C), which corresponded to the strongly adsorbed H atoms (Supplementary Fig. 14b). The N1Z3 catalysts prepared at calcination temperatures above 700 °C exhibited similar H2 desorption behavior.
To gain insight into the CO2 adsorption behavior and active sites for the CO2 conversion reaction, the O2temperature-programed desorption (O2–TPD) profiles of the catalysts were obtained (Supplementary Fig. 15). Depending on the number of gained electrons, the O species at the surface of metal oxides can be categorized into physically adsorbed oxygen () at temperatures below 100 °C, superoxide species () in the temperature range of 100–200 °C, monatomic oxygen () in the temperature range of 200–400 °C, lattice oxygen at the surface () in the temperature range of 400–700 °C, and lattice oxygen in the bulk () at temperatures above 700 °C82-84. The surface-adsorbed and species, which were associated with surface vacancies, were weakly bonded to the catalyst surfaces and thus, were relatively easy to desorb; however, the lattice O species were difficult to extract82, 83, 85. Broad peaks centered at 360–370 °C, which were associated with the desorption of species, were observed in the O2–TPD profiles of the N1Z3-500 and N1Z3-600 catalysts. A low-temperature peak at 310 °C appeared in the O2–TPD profile of the N1Z3-700 catalyst, and a low-temperature more intense peak at 321 °C was observed in the O2–TPD profile of the N1Z3-800 catalyst. This suggested the formation of more easily desorbed species on the surface of catalysts prepared at higher calcination temperatures. Conversely, the high-temperature peaks in the temperature range of 360–385 °C, which were the major desorption peaks in the O2–TPD profiles of the N1Z3-500, N1Z3-600, and N1Z3-700 catalysts calcined, were less prominent in the O2–TPD profile of the N1Z3-800 catalyst. A remarkable change in the desorption of O was observed in the O2-TPD profile of the N1Z3-900 catalyst; the removal of from the metal oxide lattice was significantly increased from 0.033–0.059 mmol g–1 (for the N1Z3-600, N1Z3-700, and N1Z3-800 catalysts) to 0.135 mmol g–1 (for the N1Z3-900 catalyst), as listed in Supplementary Table 5. The presence of O vacancies in metal oxides helps to create unsaturated metal centers, which renders the lattice O adjacent to O vacancies more active83. The metal oxide can reserve some outer electrons that can be donated to the antibonding orbital of the CO2 molecules on the metal surface to enhance CO2 adsorption. Moreover, the presence of such O vacancies on the catalyst surface increases the electron density at the metal centers, which can enhance the reactivity of CO286, 87. The creation of O vacancies can facilitate the surface restructuring of Ni and Zn to form Ni–Zn alloys because of the differences between the ionic radii of the O2– ion (140 pm) and metal ions (Zn2+, 74 pm and Ni2+, 69 pm) and different diffusion paths could be accessible to the O interstitials in the wurtzite lattice88, 89.
The CO2 temperature-programed desorption (CO2–TPD) profiles of the catalysts were used to examine the adsorption behavior of CO2 on the catalyst surfaces, and the results are illustrated in Supplementary Fig. 16 and Supplementary Table 6. The desorption peaks in the CO2–TPD profiles of the catalysts could be categorized into three regions: <250, 250–600, and >600 °C, which represented weak, medium, and strong basic sites. The main peak in the CO2–TPD profiles of the N1Z3-500, N1Z3-600, and N1Z3-700 catalysts occurred in the temperature range of 535–555 °C and did not change significantly as the calcination temperature of the catalysts increased, which indicated that strong interactions were maintained between the adsorbed CO2 and catalyst surfaces. However, the main desorption temperature of the N1Z3-900 catalysts shifted to 432 °C, which suggested that the surface basicity of the N1Z3-900 catalyst was significantly lower than those of the N1Z3-500, N1Z3-600, and N1Z3-700 catalysts. The favorable CO2 adsorption on the surface of the N1Z3-900 catalyst at low temperatures could be attributed to the increase in the number of sites at the calcination temperature of 900 °C. The amount of CO2 adsorbed at the medium basic sites of the N1Z3-900 catalyst was 0.158 mmol g−1, which was significantly higher than those of the N1Z3-800 catalyst (0.102 mmol g−1) and other lower-calcination-temperature catalysts (0.018–0.037 mmol g−1). The high CO2 hydrogenation activity of the N1Z3-900 catalyst suggested that the CO2 adsorbed at the medium basic sites effectively participated to the reaction and the CO2 adsorbed at the strong basic sites was less suitable for the subsequent conversion reaction. This hypothesis was supported by the strong interactions between the CO2 molecules adsorbed at the medium basic sites and the catalyst surface at the reaction temperatures and facile desorption of the reaction products from the catalyst surface after the reaction.
To elucidate the reaction mechanism for the formation of monocarboxylic acids, in situ DRIFTS experiments were performed to analyze the evolution of the reaction intermediates over Ni–Zn alloy catalysts. To identify the chemical species derived from the adsorption of CO2 on the surface of the N1Z3-900 catalyst, DRIFTS profiles were obtained by flowing CO2 at 3.0 MPa through the DRIFTS cell that contained an in situ reduced N1Z3-900 catalyst sample, as presented in Fig. 6a. The detailed peak assignment is listed in Supplementary Table 7. As the temperature of the cell increased from 30 to 325 °C, the intensities of the infrared (IR) bands of CH4 at 3015 and 1305 cm− 1 increased; however, further extending the adsorption time to 60 min at 325 °C led to a decrease in the intensity of CH4 bands. The reaction between the residual pre-adsorbed H2 on the catalyst surface after reduction and adsorbed CO2 species formed CH4 even at a low temperature of 30 °C. As the temperature increased from 30 to 325 °C the intensities of the bands of the CO2-adsorbed intermediate species (carbonate and bicarbonate) increased in the frequency range of 1800–1000 cm− 1. The bands at 1616 and 1293 cm− 1 were assigned to the asymmetric and symmetric vibrations of bidentate carbonate, respectively, and those at 1522 and 1327 cm− 1 corresponded to the asymmetric and symmetric vibrations of monodentate carbonate, respectively90, 91. The reaction between monodentate carbonate and adsorbed H2 produced bicarbonate, which generated the peaks at 1636 cm− 1 (νas(O − C−O)), 1419 cm− 1 (νs(O − C−O)), and 1219 cm− 1 (δ(–COH))90. In addition, the hydrogenation of bidentate carbonate produced formate, and its peaks were observed at 1577 cm− 1 (νas(O − C−O)), 1388 cm− 1 (δ(C − H)), and 1363 cm− 1 (νs(O − C−O))92, 93. The peaks of the CO formed via RWGS were observed at 2078 cm− 1 (linearly adsorbed CO, l-CO), 1933 and 1918 cm− 1 (bridge-adsorbed CO, b-CO), and 2055 cm− 1 (Ni(CO)4)94. Furthermore, the 1751 cm− 1 peak was ascribed to the formyl species produced via the hydrogenation of CO95. In addition, the IR bands at 1080, 1050, 977, and 950 cm− 1 were associated with the bending of the = C–H groups and those at 1540 and 1340 cm− 1 were ascribed to the carboxylate species92, 96. The intermediates and products formed via the adsorption of CO2 on the reduced N1Z3-900 catalyst (Fig. 6a) were different from those formed on the reduced N1Z3-500 catalyst (Fig. 6b). First, the formation of a large amount of CH4 was observed at 150 °C over the N1Z3-500 catalyst; however, the formation of CH4 was suppressed on the N1Z3-900 catalyst, and these results were in good agreement with the CO2 hydrogenation data (Fig. 1a). This indicated that the metallic Ni particles on the N1Z3-500 catalyst surface facilitated methanation97, 98, and the formation of the Ni–Zn alloy suppressed the methanation reaction. Second, the formation of CO2-adsorbed intermediates (i.e., carbonate, bicarbonate, and formate) on the surface of the N1Z3-500 catalyst was less favored than on the surface of the N1Z3-900 catalyst. This implied that the adsorption of CO2 on the Zn-rich NixZnyO phase was more favorable than that on the ZnO phase.
H2 at 325 °C and 3.0 MPa was used as flow gas in the DRIFTS cell immediately after CO2 flow, and the Fourier-transform infrared (FT–IR) spectra of the N1Z3-900 catalyst were collected to investigate its hydrogenation behavior (Fig. 7a). As the H2 flow time increased to 15 min, the formation of CH4 began to be favored, and the intensity of the IR band associated with the = C–H groups decreased significantly. In addition, the amounts of CO2-adsorbed intermediate species, RWGS-formed gaseous CO (2200–2100 cm− 1), and water (1653 cm− 1)99, which could be produced via the re-adsorption of OH from the decomposition of formate species100, 101, began to increase. This indicated that progressive CO2 hydrogenation occurred after 15 min of H2 flow, after changing the flow gas from CO2 to H2. After 140 min of H2 flow, the peaks of gaseous CO and l-CO almost disappeared from the FT–IR spectra of the N1Z3-900 catalyst, and the intensity of the CH4 peak began to decrease. Moreover, residual CO2-intermediate species were present on the surface of the N1Z3-900 catalyst. The CO2 hydrogenation behavior over the N1Z3-500 catalyst was similar (Fig. 7b).
A distinct difference in the catalytic behaviors of the N1Z3-900 and N1Z3-500 catalysts was observed using their operando DRIFTS profiles collected utilizing a mixture of H2 and CO2 with a H2/CO2 ratio of 2:1 at 325 °C and 3.0 MPa as flow gas. After 15 min of H2 and CO2 flow, the formation of gaseous CO and CH4 on the surface of N1Z3-500 was highly favored and the formation of CO and CH4 on the surface of N1Z3-500 was suppressed (Supplementary Fig. 17). As presented in Fig. 8, the time evolution of the formation of CO indicated that the amount of gaseous CO produced during the initial stage of reaction over the N1Z3-900 catalyst continuously decreased, which suggested that CO was actively transformed into next-stage hydrogenated products. After 90 min of reaction, the amount of gaseous CO remained almost unchanged, which indicated that the rates of the hydrogenation of CO and formation of CO reached steady-state. In contrast, the amount of gaseous CO reached a maximum after 30 min of reaction over the N1Z3-500 catalyst and progressively decreased as the reaction continued for 300 min. This indicated that the N1Z3-500 catalyst could not convert CO to next-stage hydrogenated products during the initial reaction stage. The CO2-adsorbed reaction intermediates (i.e., carbonate, bicarbonate, and formate) and double-bonded hydrocarbons reached steady-state peak intensities after 15 min of reaction over both catalysts (Supplementary Fig. 17).
To further elucidate the reaction behavior of the N1Z3-900 catalyst, CO adsorption on the in situ reduced catalyst followed by CO hydrogenation using a flow of H2 were performed. As presented in Supplementary Fig. 18a, during the CO adsorption reaction, as the temperature increased from 30 to 325 °C, the amount of formed CO2 increased, and that could be caused by the removal of the surface lattice O from the Zn-rich NixZnyO (CO + ZnO → CO2 + Zn). The produced CO2 could be adsorbed to form carbonate, bicarbonate, and formate species. In contrast to the CO2 adsorption, the formation of double-bonded hydrocarbons (1080–950 cm− 1) and CH4 (3105 and 1305 cm− 1) was not observed during the CO adsorption reaction, which could be ascribed to the low hydrogenation reactivity of the adsorbed CO. The increase in the amount of formed CO2 (which was generated via the water-gas shift reaction) was observed in the 30–40 min interval of H2 flow at 325 °C (Supplementary Fig. 18b); moreover, CH4 began to form simultaneously. In addition, CO2-adsorbed intermediates (i.e., carbonate, bicarbonate, and formate) and CO-hydrogenated intermediate (formyl) began to form after 30 min of H2 flow; this was attributed to the formation of CO2.
To further elucidate the adsorption behavior of hydrocarbon species formed on the N1Z3 catalysts, we used CH4 temperature-programed desorption (CH4–TPD) experiments; the results are presented in Supplementary Fig. 19 and Supplementary Table 8. The presence of the high-temperature desorption peaks at 450–541 °C in the CH4–TPD profiles of the N1Z3 catalysts suggested that dissociated CH4 adsorption occurred on the surface of the catalysts; furthermore, the methyl groups and H atoms generated by CH4 could be adsorbed on the Zn centers and surface O sites of the N1Z3 catalysts, respectively86. The activation of CH4 via dissociative adsorption on the Ni–Zn alloy catalyst suggested the participation of *CHx in the formation of monocarboxylic acids.
Reaction mechanism. Based on the TPD profiles of the N1Z3 catalysts, in situ NAP–XPS, and in situ DRIFTS, a plausible reaction pathway was proposed (Fig. 9a). First, molecular H2 was activated via heterolytic cleavage on the Ni centers in the Zn-rich NixZnyO and Ni–Zn alloy. Next, CO2 molecules were adsorbed on the surface of the Zn-rich NixZnyO to form carbonate, bicarbonate, and subsequently, formate species. Afterward, the formate species were hydrogenated to form gaseous CO. Subsequently, the produced CO was re-adsorbed on the Ni centers on the Ni–Zn alloy surface, and then it was hydrogenated to *CHx species via the formation of formyl species. Thereafter, the adsorbed *CHx species were hydrogenated to form *CH3. Next, a CO2 molecule was adsorbed on the Zn site of the Ni–Zn alloy surface, which was subsequently inserted into the adsorbed *CH3 species to form the *CH3COO intermediate. The subsequent hydrogenation of *CH3COO generated the final CH3COOH product. For PA, C–C coupling between the surface-adsorbed *CHx species could form *CHxCHx, and the subsequent hydrogenation of *CHxCHx could produce *CH2CH3 species. The surface-adsorbed CO2 inserted into the *CH2CH3 species formed *CH3CH2COO, and the subsequent hydrogenation of *CH3CH2COO produced CH3CH2COOH. Similarly, a small amount of butyric acid was produced over the N1Z3-900 and Ni1Zn4-900 (0.7% and 1.4% selectivity, respectively) catalysts via the CO2 insertion into *CH3CH2CH2 species to form CH3CH2CH2COO* species. The subsequent hydrogenation of CH3CH2CH2COO* could produce CH3CH2CH2COOH. The presence of alcohols, such as methanol and ethanol, was not observed in the reaction product, which implied that typical methanol pathways (e.g., formate and carboxylic pathways)102 might not be responsible for the production of monocarboxylic acids.
To further examine the possibility of direct C–C coupling of CO2 and the surface adsorbed *CH3 species, we used DFT simulation (Fig. 9b). The detailed steps of the hydrogenation of CO2 to AA are presented in Supplementary Fig. 20. Based on the experimental XRD data (Supplementary Fig. 4a), the (111) plane of Ni4Zn22 was used for the simulation. The electron transfer from the metal surface to CO2 facilitated the chemisorption of CO2δ− as a tridentate configuration on the Ni center and two adjacent Zn atoms. The Ni–C, Zn–O*, and Zn–O** bond lengths were 1.947, 1.992, and 2.039 Å, respectively. The adsorbed H* at the Ni center was subsequently reacted with the chemisorbed CO2δ− to form carboxylate species (*COOH) as bidentate configurations with bond lengths of 1.910 and 2.038 Å for Ni–C and Zn–O, respectively. The *COOH species were subsequently dissociated into CO and hydroxyl groups, which were adsorbed on the Ni and Zn sites, respectively. Subsequently, CO was hydrogenated to form *HCO by combining H with C or O atoms, and then, it was further hydrogenated and deoxygenated to form *CHx species. During the hydrogenation of CO, C was shifted from the Ni center to a Zn site, and the subsequent hydrogenation of Zn–CHx produced Zn–CH3 species. The bond length of Zn–C in the Zn–CH3 species was 1.956 Å. This was slightly longer than that of −[ZnCH3]+ (1.93 Å)103, which suggested that the Zn–C bonds at the Ni4Zn22 catalyst sites were slightly more activated. For the adjacent Zn–CH3 species, CO2 was adsorbed on the Zn sites and the adsorbed CO2 was tilted toward the Zn–CH3 bonds, and therefore, the Zn–CH3 bond length increased to 2.015 Å. Subsequently, the activated CO2 was inserted into the Zn–C bond of the Zn–CH3 species to produce surface acetate species (Zn–OOCCH3). Afterward, the surface Zn–OOCCH3 species were hydrogenated to produce AA, which desorbed from the catalyst surface. The total Gibbs free energy for the formation of AA from CO2 was − 1.03 eV, which indicated that the reaction was thermodynamically feasible on the surface of the Ni4Zn22 alloy. Similar alky group carboxylations using CO2 have been observed using homogeneous organometallic catalysts104, 105 and heterogeneous Zn-doped CeO286 and Zn-doped H-ZSM-5 catalysts106.
Stability of the N 1 Z 3 -900 catalyst. The stability of the crystal structure of the N1Z3-900 catalyst during the CO2 hydrogenation reaction was tested using in situ XRD, and the results are depicted in Supplementary Fig. 21. The phase structure of the N1Z3-900 catalyst did not change as the temperature increased up to 400 °C under a H2/CO2 flow, which indicated that the Ni–Zn alloy and Zn-rich NixZnyO phases were stable during the CO2 hydrogenation reaction. This suggested that the water molecules formed via the RWGS reaction did not alter the phase structure of the catalyst. The XRD pattern of the spent N1Z3-900 catalyst, which was collected after a 216 h on-stream test, was almost identical to that of the fresh catalyst. This demonstrated the robust structure of the N1Z3-900 catalyst under the reaction conditions (Supplementary Fig. 21b). In addition, the Ni–Zn alloy structure on the surface of the Zn-rich NixZnyO phase in the spent catalyst was maintained after the long-term test (Supplementary Fig. 21c). The high stability of the N1Z3-900 catalyst under the reaction conditions caused the CO2 conversion and product selectivity of the catalysts to remain unchanged (Fig. 1c).
In summary, we demonstrated that a Ni–Zn alloy/Zn-rich NixZnyO catalyst could be used to produce AA and PA with high selectivity (58.9% and 18.2%, respectively) via direct CO2 hydrogenation at a CO2 conversion of 13.4% and by suppressing the selectivity toward CH4 (17.0%) and C2–C4 (5.0%). The STYs of AA and PA were 0.405 and 0.125 mmol L− 1 h− 1, respectively. The mutual diffusion of Ni into the ZnO phase and Zn into the NiO phase during calcination at temperatures of up to 900 °C and formation of Ni–Zn alloy during the reduction produced the Ni–Zn alloy/Zn-rich NixZnyO catalyst. The numerous O vacancies in the Zn-rich NixZnyO phase facilitated the adsorption of CO2. The presence of CO2-adsorbed species (i.e., carbonate, bicarbonate, and formate) and gaseous CO indicated that the RWGS reaction converted CO2 to CO. Subsequently, CO was hydrogenated to formyl species and then to surface-adsorbed (*CH3)n species. AA and PA were produced via the direct C–C coupling of CO2 to the surface-adsorbed (*CH3)n species. The CO2 conversion and monocarboxylic acid selectivity were maintained for up to 216 h on-stream, and the morphology and phase structure of the Ni–Zn alloy/Zn-rich NixZnyO catalyst were maintained during the long-term stability test, which demonstrated the robust nature of the Ni–Zn alloy catalyst. Considering that, to date, direct CO2 conversion has mostly been used for the production of liquid fuels, methanol, and light olefins, the possibility of producing value-added monocarboxylic acids from CO2 demonstrated the novelty and usability of the Ni–Zn alloy catalyst.