Shuang Wang‡
a,
Xiaowa Nie‡*a,
Chunshan Song
*b and
Xinwen Guo
*a
aState Key Laboratory of Fine Chemicals, Frontier Science Center for Smart Materials, PSU-DUT Joint Center for Energy Research, School of Chemical Engineering, Dalian University of Technology, Dalian 116024, China. E-mail: niexiaowa@dlut.edu.cn; guoxw@dlut.edu.cn
bDepartment of Chemistry, Faculty of Science, The Chinese University of Hong Kong, Shatin, NT, Hong Kong 999077. E-mail: chunshansong@cuhk.edu.hk
First published on 23rd July 2025
In this work, density functional theory (DFT) calculations were conducted to investigate a series of metal node-modified Ti-MOF catalysts using transition metals (Mn, Fe, Co, Ni, Cu, Zn, Zr, Nb, Mo, Ru, Rh, Pd, Hf, Ta, W, Os, Ir, and Au) introduced into Ti-ATA (ATA = 2-aminoterephthalic acid) for the photocatalytic reduction of CO2 to C2 products. CO2 can be sufficiently activated on Ti(M)-ATA but the adsorption configuration depends on the nature of M. Over Ti(Nb)-ATA, Ti(Ta)-ATA, Ti(Zr)-ATA and Ti(Hf)-ATA, the two *CHO species undergo C–C coupling to form *CHOCHO, an important C2 intermediate. Ti(Nb)-ATA and Ti(Ta)-ATA tend to generate ethanol, while Ti(Zr)-ATA and Ti(Hf)-ATA are more selective to ethylene. Among the Ti(M)-ATA candidates studied, Ti(Nb)-ATA was identified as the most active catalyst for CO2 reduction to ethanol due to its smallest limiting free energy change (1.12 eV), over which the *CH2CH2O reduction to *CH2CH2OH was found to be the rate-determining step. The correlation curve analysis illustrates that the reduction activity of Ti(M)-ATA catalysts is highly dependent on the binding strength of CO2 and key reaction intermediates such as *OCHOH. The analysis of electronic and optical properties indicates that the altered energy band structure and charge transfer behavior around the bimetallic nodes of Ti(Nb)-ATA account for its excellent catalytic activity for CO2 reduction to ethanol.
Over the past few years, a new class of semiconductor-like crystalline porous materials known as metal–organic frameworks (MOFs) has attracted increasing attention in the fields of heterogeneous catalysis, gas storage, and separation.9,10 MOFs are composed of inorganic metal ions and organic ligands, featuring well-defined structures, high porosity, larger surface area, component diversity, and tailorability.11 Therefore, MOF-based materials possess rich physicochemical properties and unique structural advantages among porous catalytic materials. In particular, the large surface areas and specially designed active sites of MOFs could enrich CO2 efficiently with superior CO2 adsorption capacity and stabilize the reaction intermediates, making them promising platforms for the photocatalytic reduction of CO2 to chemicals and fuels.12,13
However, only a limited number of MOF-based photocatalytic systems have been developed for CO2 reduction and their efficiency to C2 products is still far from satisfactory. Enhancing the photocatalytic efficiency of MOFs for C2 formation mainly depends on their ability to generate and maintain photogenerated electrons as well as their ability to activate CO2 and accelerate C–C coupling.14 Over the past decade, researchers have designed strategies to improve the photocatalytic performance of MOFs, as summarized below:
(1) By constructing efficient active sites and using the chelating effect of ligands to immobilize atomically dispersed single/dual metals, the photocatalytic reaction can be promoted.15–17 For example, our group successfully synthesized a novel Ti-based MOF photocatalyst denoted as Fe/Ti-BPDC (BPDC = 2,2′-bipyridine-5,5′-dicarboxylic acid) with atomically dispersed Fe sites, which exhibited high activity and selectivity for CO2 reduction to HCOOH, with a yield of 703.9 μmol g−1 h−1 and a selectivity greater than 99.7%.18 Inspired by the experimental work, we proposed a new strategy for designing metal single-atom-modified and dual-atom-modified Ti-BPDC photocatalysts through structural and electronic modulation for CO2 reduction to C1 and C2 products using density functional theory (DFT) calculations.19,20 The computational results of structure–activity volcano curves show that metal single-atom modified Fe/Ti-BPDC (ΔGL = 0.40 eV) and Pd/Ti-BPDC (ΔGL = 1.17 eV) are optimal catalysts for the photoreduction of CO2 to HCOOH and CH3OH, respectively,19 due to their relatively small limiting free energy changes compared to those reported in the literature.21 Furthermore, metal dual-atom-modified Ti-BPDCs such as Cu–Sn/Ti-BPDC (ΔGL = 0.20 eV) and Cu–Os/Ti-BPDC (ΔGL = 0.70 eV) show enhanced activity toward the generation of HCOOH and CH3OH products from CO2 photoreduction compared to metal single-atom modified Fe/Ti-BPDC and Pd/Ti-BPDC. In particular, the dual-metal atom-modified Cu–In/Ti-BPDC (ΔGL = 1.37 eV) exhibits improved activity and selectivity towards the generation of C2 products (mainly C2H4).20
(2) Light absorption efficiency can be improved by choosing the right ligands. For example, in order to improve the optical properties of MIL-125 (Ti) MOFs, researchers successfully shifted MIL-125 (Ti) to the visible range by introducing the 2-aminoterephthalic acid (ATA) ligand with a broad spectral response, enhancing the visible light absorption of the MOF materials.22 Sun et al. reported that MIL-125-NH2(Ti) with {110}/{111}-heterojunction yields 10 and 18 times more CO and CH4 products, respectively, from CO2 photoreduction compared with the single {001} facet; DFT calculations identified energetically favorable pathways and rate-limiting steps for CO2 reduction on different low-index surfaces of MIL-125-NH2(Ti) in their studies.23
(3) In addition, metal substitution/doping has been demonstrated to be an effective method to improve the performance of MOF-based photocatalysts. One of the effective strategies is to construct bimetallic assemblies by partially replacing a node metal in MOFs with another metal. The bimetallic assemblies can harvest visible light and the doped metal cations can act as electronic mediators to promote charge transfer, facilitating the photocatalytic processes.24 In particular, partial substitution of metal cations in MOFs can lead to the formation of oxygen-bridged heterometallic structures within the framework, which could exhibit enhanced photocatalytic performance due to the structural flexibility and tunability of the designed MOFs.25 For example, Ti-substituted NH2-UiO-66 (Zr/Ti) was prepared by Li et al. by using the post-synthesis exchange (PSE) method which exhibits good catalytic performance for the photoreduction of CO2 to formic acid under visible light, with a yield of 5.8 mmol mol−1 for 10 h; DFT calculations and electron spin resonance (ESR) results indicate that the introduction of Ti substituents as a mediator promotes electron transfer, which improves the photocatalytic performance.25 In addition, Truhlar et al. used UiO-66 as the basic skeleton and replaced partial Zr atoms in the nodes with metals such as Hf, Th, Ti, U, Ce, etc.; DFT calculations show that UiO-66(Ce) effectively promotes electron–hole separation due to the fact that Ce4+ in the material has a low-energy 4f vacancy orbital, which can receive photogenerated electrons, thus more effectively driving the photocatalytic reaction.26
Based on the literature survey and prior studies in our group, the structures and properties of ligands in MOF materials have important impacts on the photocatalytic performance of CO2 reduction. On the one hand, metal single-atom and dual-atom assemblies can be stably anchored with N-containing ligands within the MOF framework to improve the efficiency of CO2 activation and conversion.18–20,27 On the other hand, metal-modified nodes can effectively regulate light absorption and electron–hole separation efficiency, and provide special active sites to further tune the activity and selectivity of MOF-based catalysts for CO2 photoreduction.25,26,28 However, the directional control of CO2 reduction products remains challenging, especially for the generation of C2 products. Due to the complex mechanisms of adsorption/activation of CO2 and cleavage/reconstruction of CO double bonds over the active sites of catalysts, the principles governing the formation and transformation of key intermediates and electron–proton transfer in the reaction processes are still not clear, which make it difficult to regulate the conversion paths and target products.
The limited efficiency of MOF-based photocatalysts in the reduction of CO2 to C2 products (e.g., ethanol and ethylene) stems from several intrinsic challenges, among which insufficient activation of CO2, slow kinetics of C–C coupling, and generation of key intermediates required for C–C coupling are the major bottlenecks.29 On the one hand, many MOFs lack robust adsorption sites (e.g., metal active sites or Lewis base sites) to activate the linear structure of CO2, which prevents the reduction of CO2 to the critical *HCOO or *COOH intermediate. On the other hand, C2 product generation requires multiple electrons, but MOFs usually lack efficient charge transport channels to deliver electrons quickly. In addition, the lack of bimetallic site nodes in some MOF materials leads to low stability of C–C coupling intermediates. To summarize, the synergistic optimization of CO2 adsorption/activation, electron transfer, intermediate stability, and C–C coupling kinetics inside MOF materials is a crucial factor in overcoming the inefficiency of MOF-based photocatalysts for the production of C2 products.
Inspired by our previous experimental studies on Ti-based MOFs and the potential application of 2-aminoterephthalic acid (ATA) ligands in MOFs for photocatalysis,30,31 the present work systematically investigates the adsorption/activation and reduction of CO2 to C2 products (mainly ethanol and ethylene) on a series of metal node-modified Ti(M)-ATA MOFs (ATA = 2-aminoterephthalic acid, M = Mn, Fe, Co, Ni, Cu, Zn, Zr, Nb, Mo, Ru, Rh, Pd, Hf, Ta, W, Os, Ir, Au) by means of DFT calculations. By screening these heterometallic node structures via structure–activity correlation, Ti(Nb)-ATA is identified as the most active photocatalyst for CO2 reduction to ethanol (C2H5OH), with a limiting free energy change of 1.12 eV associated with the *CH2CH2O → *CH2CH2OH step. Electronic property analysis further demonstrates that the computationally designed Ti(Nb)-ATA MOF catalyst via node modification is more favorable for the photocatalytic reduction of CO2 to C2 compared to unmodified Ti-ATA and other Ti-based MOFs such as Ti/BPDC and MIL-125-NH2(Ti). In addition, we also find that node-modified metals from different groups of the periodic table lead to different C2 products. For example, Ti(Ta)-ATA and Ti(Nb)-ATA tend to generate ethanol, while Ti(Zr)-ATA and Ti(Hf)-ATA are more selective to ethylene.
An all-electronic approach implemented in the Vienna Ab initio Simulation Package (VASP) program was used to perform spin-polarized DFT calculations.32,33 The projector-augmented wave (PAW) pseudopotentials were used to describe the electron–ion interactions. The Perdew–Burke–Ernzerhof (PBE) functional in the generalized gradient approximation (GGA) was used to calculate the exchange–correlation energies of electrons.34 The Coulomb and exchange interactions were corrected by setting the Ueff parameter (Ueff = Coulomb U − exchange J) for each transition metal.35,36 The Ueff parameter for titanium atom was set to 3.0 eV, and those for remaining metals are given in Table S1 in the ESI.† To include van der Waals (vdW) interactions, the PBE + D3 method was used.37,38 The valence electrons were described using a plane wave basis set with a cutoff energy of 450 eV. The convergence criteria for all calculated electronic energies and atomic forces were set to 10−4 eV and 0.03 eV Å−1, respectively. Both Ti-ATA and Ti(M)-ATA structures were optimized using a 4 × 1 × 4 k-point mesh for sampling the Brillouin zone in the VASP calculations.
The adsorption energy (Eads) of an adsorbate onto the catalyst was calculated using the equation Eads = Etotal – Ecatalyst – Eadsorbate, where Etotal, Ecatalyst, and Eadsorbate represent the total energy of the system containing the adsorbate and catalyst, the energy of the bare catalyst, and the energy of the adsorbate in the gas phase, respectively. In order to evaluate the catalytic activity of CO2 photoreduction over these Ti(M)-ATA catalysts, the free energy change (ΔG) was calculated for each elementary step involved in the CO2 reduction process using the formula ΔG = ΔE + ΔEZEP – TΔS,39,40 where ΔE is the total electronic energy change obtained through DFT calculations, ΔEZEP is the zero-point energy contribution, T is the temperature at 298. 15 K, and ΔS is the entropy change. The zero-point energy and entropy were obtained from vibrational frequency calculations using DFT. The zero-point energies of all adsorbed species are provided in Table S2.† In this work, we investigated the reaction pathways for the formation of different C2 products (mainly C2H4 and C2H5OH) over Ti-ATA and Ti(M)-ATA catalysts, in which protons and photogenerated electrons are added to the reaction intermediates progressively in the presence of a catalyst, ultimately resulting in the formation of C2 products.
It is worth noting that DFT energy and substitution energy alone are not sufficient to determine the optimal doping model for metal modification. Therefore, we further explored all possible adsorption configurations of CO2 on the two different node models, considering four different initial adsorption configurations, including *COO, *OCO, O*CO, and O*OC (* indicates adsorption at the doped metal active site). With model structure 1, as shown in Fig. 1(d), after comprehensive structural screening and optimization, it was found that CO2 is adequately activated on Ti(M)-ATA (M = Nb, Zr, Ta, Hf), with one O atom of CO2 bonded to the doped metal site in the node and the C atom of CO2 bonded with the N atom in the ATA ligand, forming a stable *O(M)–C(N)–O angular adsorption structure, as shown in Fig. 2(a). However, on the catalyst models of Ti(M)-ATA (M = Mn, Fe, Co, Ni, Cu, Zn, Mo, Ru, Rh, Pd, W, Os, Ir, Au), CO2 shows a linear adsorption configuration and is not bonded to other atoms in the MOF catalysts, as illustrated in Fig. S3.† With model structure 2 as shown in Fig. 1(d), CO2 exhibits the *O(Zr)–C(N)–O adsorption mode merely on Ti(Zr)-ATA whereas it shows a linear adsorption pattern on the rest of the candidate catalysts, as shown in Fig. S4.† The adsorption energies of CO2 on pristine Ti-ATA and metal-modified Ti(M)-ATA (model structure 1 and model structure 2) were calculated, as shown in Fig. 2(b–d) and S5,† while the specific values of adsorption energies are provided in Table S4.† The calculation results show that the adsorption energy values of CO2 on Ti(M)-ATA (M = Nb, Zr, Ta, Hf) are 0.15, 0.31, −0.10, and 0.26 eV, respectively, which are significantly lower than that obtained over the pristine Ti-ATA catalyst (1.45 eV) and those obtained over other metal-modified Ti(M)-ATA candidate catalysts. Therefore, the Ti(M)-ATA (M = Nb, Zr, Ta, Hf) candidates were chosen for subsequent exploration of the reaction pathways and reactivities for CO2 photoreduction to C2 products. It is worth noting that the adsorption energies of CO2 on model structure 1 are generally lower than those on model structure 2, as shown in Fig. S5.† Based on the calculated CO2 adsorption configurations and adsorption energies, the metal-modified model structure 1 exhibits good structural stability, sufficient interactions with CO2, and better ability to activate CO2; therefore it was chosen as the representative node-modification model for subsequent studies.
![]() | ||
Fig. 3 The calculated Gibbs free energy diagrams for CO2 reduction steps before C–C coupling on (a) Ti-ATA, (b) Ti(Zr)-ATA, (c) Ti(Hf)-ATA, (d) Ti(Nb)-ATA, and (e) Ti(Ta)-ATA catalysts. |
Starting with the formed *CHO species, all possible pathways and intermediates for the reduction of CO2 to C2H4 or C2H5OH products on pristine Ti-ATA and metal-modified Ti(M)-ATA (M = Zr, Hf, Nb, Ta) were considered, as illustrated in Fig. 4. By comparing the reaction Gibbs free energies, the optimal pathway for CO2 reduction to C2H4 or C2H5OH over each of the examined catalyst was identified, as shown in Fig. 5(a) and (b), and the optimized structures of relevant intermediates are given in Fig. 5(c) and (d). Other pathways considered and investigated are provided in Fig. S6.† In this scenario, we focus on discussing the reaction mechanisms starting from the C–C coupling step, as detailed below:
![]() | ||
Fig. 4 Possible reaction pathways considered for the reduction of CO2 to C2 (C2H5OH and C2H4) products on (a) Ti-ATA, Ti(Zr)-ATA, and Ti(Hf)-ATA and (b) Ti(Nb)-ATA and Ti(Ta)-ATA. |
(1) According to the free energy diagram shown in Fig. 5(a), over pristine Ti-ATA, metal-modified Ti(Zr)-ATA and Ti(Hf)-ATA, the two *CHO species undergo C–C coupling to form the important C2 intermediate *CHOCHO (Fig. 5(b)), releasing the energies of 0.98, 1.91, and 1.99 eV, respectively on the three catalysts. The subsequent conversion of *CHOCHO species leads to the formation of four possible intermediates, namely *CHOCHOH, *CHOCH2O, *CHOHCHO, or *CH2OCHO. The calculated Gibbs free energies in Fig. S6(a–c)† show that the formation of these four intermediates is thermodynamically unfavorable due to the uphill free energy changes, with the *CHOCHOH species formation involving a relatively smaller energy increase compared to the other three species (ΔG = 0.31 (Ti-ATA), 1.17 (Ti(Zr-ATA)), and 1.20 eV (Ti(Hf)-ATA)). The formed *CHOCHOH intermediate is then reduced by proton-coupled electron transfer to form *CHOHCHOH, *CHOCH + H2O, *CHOCH2OH, or *CH2OCHOH species. Notably, the ΔG values for the formation of *CHOCH + *H2O on Ti-ATA, Ti(Zr)-ATA, and Ti(Hf)-ATA are −0.78, −1.00, and −1.07 eV, respectively, which are lower than those for the formation of the other three intermediates (Fig. S6(a–c)†), suggesting that *CHOCH + *H2O is the preferred intermediate from *CHOCHOH reduction. In the next step, further reduction of *CHOCH leads to the formation of *CH2OCH, *CHOCH2, or *CHOHCH. As demonstrated in Fig. 5(a) and S6(a–c),† the formation of *CH2OCH, *CHOCH2, and *CHOHCH species is thermodynamically unfavorable due to uphill free energy changes, but *CHOCH2 is relatively more stable than the other two species, with ΔG values of 0.67 (Ti-ATA), 0.53 (Ti(Zr)-ATA), and 0.63 eV (Ti(Hf)-ATA), respectively. Subsequently, the *CHOCH2 intermediate undergoes further reduction to either *CH2OCH2 or *CHOHCH2 species, with the formation of *CHOHCH2 species being thermodynamically more favorable, as shown in Fig. S6(a–c).† In contrast, the reduction of *CHOCH2 to *CH2OCH2 species is thermodynamically unfavorable, with ΔG values of −0.26 (Ti-ATA), −0.47 (Ti(Zr)-ATA), and −0.48 eV (Ti(Hf)-ATA), as shown in Fig. 5(a). For the generated *CHOHCH2 intermediate, two possible hydrogenation pathways lead to different C2 intermediates: either the formation of *CHCH2 species by releasing a H2O molecule (ΔG = 0.21 (Ti-ATA), 0.37 (Ti(Zr)-ATA), and 0.38 eV (Ti(Hf)-ATA)), which is then reduced to *CH2CH2 (ΔG = −0.44 (Ti-ATA), −0.54 (Ti(Zr)-ATA), and −0.61 eV (Ti(Hf)-ATA)), or the generated *CHOHCH2 intermediate continues to be hydrogenated to form *CH2OHCH2 species which is further reduced to *CH2OHCH3. Finally, the formed *CH2OHCH3 (*C2H5OH) or *CH2CH2 (*C2H4) desorbs from the catalysts. Apparently, the reduction of CO2 to *C2H4 over Ti-ATA, Ti(Zr)-ATA, and Ti(Hf)-ATA catalysts is more selective (energetically favorable) than to *C2H5OH according to the calculation results given in Fig. S6(a–c).† However, the desorption of formed *CH2CH2 from the catalyst proceeds slowly due to the substantial free energy changes on the three catalysts examined, as shown in Fig. 5(a). Based on the above results, the optimal pathway for the photocatalytic reduction of CO2 to the preferred C2H4 product on pristine Ti-ATA, metal-modified Ti(Zr)-ATA and Ti(Hf)-ATA catalysts can be summarized as *CO2 → *OCHO → *OCHOH → *CHO → *CHOCHO → *CHOCHOH → *CHOCH + H2O → *CHOCH2 → *CHOHCH2 → *CHCH2 → *CH2CH2 → *+C2H4. According to the free energy diagram in Fig. 5(a), the reduction step *OCHOH → *CHO (ΔG = 1.48 (Ti-ATA), 2.04 (Ti(Zr)-ATA), and 2.10 eV (Ti(Hf)-ATA)) has the largest free energy change and is considered to be the rate-determining step for the overall reaction. On these three catalysts, C–C coupling is not difficult to achieve with a substantially downhill free energy change; however the hydrogenation reactions before C–C coupling proceed slowly, especially the generation of the key *CHO intermediate.
(2) In contrast, the mechanisms of CO2 reduction to C2 products on Ti(Nb)-ATA and Ti(Ta)-ATA catalysts are different. According to the free energy diagrams given in Fig. 5(c), for Ti(Nb)-ATA and Ti(Ta)-ATA, the two *CHO species undergo C–C coupling to form *CHOCHO (Fig. 5(d)), an important C2 intermediate, releasing the energies of 1.02 and 1.45 eV, respectively on the two catalysts. Further reduction of *CHOCHO species proceeds via different pathways. As illustrated in Fig. S6(d) and (e),† among the four possible intermediates produced from *CHOCHO reduction, the formation of *CHOCH2O is thermodynamically the most favorable, with ΔG values of 0.45 and 0.46 eV, on Ti(Nb)-ATA and Ti(Ta)-ATA, respectively. The formed *CHOCH2O species can be further reduced to *CH2OCH2O, *CHOHCH2O or *CHOCH2OH. From Fig. S6(d) and (e),† it can be seen that *CHOHCH2O formation has a lower Gibbs free energy change compared to *CH2OCH2O and *CHOCH2OH. In the next step, the formed *CHOHCH2O species can be further reduced to three possible intermediates including *CH2OHCH2O, *CHCH2O + H2O or *CHOHCH2OH. Based on the Gibbs free energy diagrams given in Fig. 5(c) and S6(d), (e),† it can be observed that the formation of the *CH2OHCH2O intermediate is thermodynamically much more favorable (ΔG = −0.60 (Ti(Nb)-ATA) and −0.42 eV (Ti(Ta)-ATA)), whereas the formation of *CHCH2O + H2O and *CHOHCH2OH intermediates is associated with large free energy barriers. Subsequently, the resulting *CH2OHCH2O species can either be reduced to *CH2OHCH2OH species which is thermodynamically unfavorable, or to *CH2CH2O and a H2O molecule which is thermodynamically favorable. The ΔG values for *CH2OHCH2O reduction to *CH2CH2O + *H2O on Ti(Nb)-ATA and Ti(Ta)-ATA are calculated to be 0.40 and 0.74 eV, respectively. After the release of H2O molecule, the remaining *CH2CH2O species can be further reduced by proton coupled electron transfer to produce the *CH2CH2OH species. As shown in Fig. 5(b), remarkably, the Gibbs free energy change for this hydrogenation process increases by 1.12 and 1.62 eV for Ti(Nb)-ATA and Ti(Ta)-ATA, respectively. In the last step, further reduction of *CH2CH2OH generates either *CH2CH2 + H2O or *CH3CH2OH. Unlike Ti-ATA, Ti(Zr)-ATA, and Ti(Hf)-ATA, it is clear that the generation of *CH3CH2OH (*C2H5OH) on Ti(Nb)-ATA and Ti(Ta)-ATA is energetically more favorable than the formation of *CH2CH2 (*C2H4), as shown in Fig. S6(d) and (e),† and the corresponding ΔG values for *C2H5OH formation on these two catalysts are −0.78 and −0.54 eV, respectively. It is worth noting that there is a small uphill energy change involved in the desorption of *CH3CH2OH species on Ti(Nb)-ATA; however, the *CH3CH2OH desorption is not a highly energy-consuming step compared to the preceding step of *CH2CH2O reduction to *CH2CH2OH with a large positive ΔG value, as illustrated in Fig. 5(c). To sum up, for the reduction of CO2 to the preferred C2H5OH product on Ti(Nb)-ATA and Ti(Ta)-ATA catalysts, the optimal pathway is identified as *CO2 → *OCHO → *OCHOH → *CHO → *CHOCHO → *CHOCH2O → *CHOHCH2O → *CH2OHCH2O → *CH2CH2O + H2O → *CH2CH2OH → *CH3CH2OH → *+C2H5OH, in which the rate-determining step is found to be the reduction of *CH2CH2O → *CH2CH2OH as it involves the highest free energy change (ΔG = 1.12 (Ti(Nb)-ATA) and 1.62 eV (Ti(Ta)-ATA)) among all of the elementary steps.
Moreover, we also calculated the kinetic barrier of the rate-determining step in the optimal energy pathway for CO2 reduction to ethanol on Ti(Nb)-ATA and Ti(Zr)-ATA, as shown in Fig. S7.† In addition to considering the energy barrier of the rate-determining step, the corresponding reaction rate constant was also calculated and is provided in Table S5.† The results show that the CO2 reduction has an obvious kinetic advantage on Ti(Nb)-ATA compared to Ti(Zr)-ATA based on both the Gibbs free energy barrier and reaction rate constant calculations, further consolidate the above screening results.
Since the adsorption and activation of CO2 is an extremely important initial step in CO2 reduction, as demonstrated in the previous studies,19,20,41 here we correlated the adsorption energy of CO2 with the limiting free energy change for the generation of C2 products (C2H4 and C2H5OH) on the four metal-modified Ti(M)-ATA (M = Zr, Hf, Nb, Ta) catalysts, and plotted the corresponding volcano type curve as displayed in Fig. 6(a). The calculation results show that the Ti(Nb)-ATA catalyst has the lowest limiting free energy change and a moderate CO2 adsorption intensity, positioning it at the top of the volcano-type curve. In addition, we note that the stability of the *OCHOH intermediate affects the next step of dehydration to form the key C–C coupling intermediate of *CHO, and thus we plotted the volcano type curve of G*OCHOH and the limiting free energy change on the four metal-modified Ti(M)-ATA (M = Zr, Hf, Nb, Ta) catalysts, as displayed in Fig. 6(b). The results show that the Ti(Nb)-ATA catalyst remains at the top of the volcano curve, further indicating that the moderate stability of the *OCHOH intermediate over Ti(Nb)-ATA is quite favorable for the subsequent generation of the *CHO intermediate. Therefore, the Ti(Nb)-ATA catalyst would be the best candidate for C2 production among the four candidates studied. These results further clarify that the adsorption strength of the CO2 reactant and key reaction intermediates (e.g.*OCHOH) directly affects the CO2 reduction activity, and the relatively mild bonding of *CO2 and *OCHOH over Ti(Nb)-ATA is responsible for the highest activity of Ti(Nb)-ATA among the screened candidate catalysts. This finding is consistent with our previous studies on CO2 photoreduction over other Ti-MOF-based catalytic materials.19,20
![]() | ||
Fig. 6 Volcano-type curves between (a) CO2 adsorption energy, (b) G*OCHOH and limiting free energy change for C2 products generation on Ti(M)-ATA (M = Zr, Hf, Nb, Ta) catalysts. |
In addition, we note that CO2 molecules can be adequately activated when the doped nodal metals belong to the IVB and VB groups and could provide favorable active sites for CO2 adsorption and activation. For IVB group metal-modified Ti-ATA catalysts, such as Ti(Zr)-ATA and Ti(Hf)-ATA, CO2 photocatalytic reduction tends to generate the C2H4 product and follows similar reaction pathways. In contrast, for VB group metal-modified Ti-ATA catalysts, such as Ti(Nb)-ATA and Ti(Ta)-ATA, CO2 photoreduction to the C2H5OH product is energetically more preferred. The calculated limiting free energy change (ΔGL = 1.12 eV) over Ti(Nb)-ATA for C2H5OH generation is even smaller than that (ΔGL = 1.37 eV) over the dual metal-atom-modified Cu–In/Ti-BPDC catalyst which was identified to be a good catalyst for C2 (C2H4) generation in our previous DFT work.20 Therefore, Ti(Nb)-ATA can be proposed as a promising candidate catalyst for the photocatalytic reduction of CO2 to C2 products, especially C2H5OH. Overall, the calculation results indicate that the ATA ligand with a Ti–O–Nb dual-metal-node structure as well as with a broad absorption spectrum effectively improves the photocatalytic efficiency for CO2 reduction to C2 products, which are also essential for regulating the catalytic activity and product selectivity of Ti-MOF-based photocatalysts.
The CO2 photoreduction was typically carried out in aqueous solution. Under the same conditions, *H species could be adsorbed on the active site and reduced by the proton–electron pair (H+ + e−), which would affect the photocatalytic efficiency of CO2 reduction. Here, we calculated the Gibbs free energy diagram for the hydrogen evolution reaction (HER) on the Ti(Nb)-ATA catalyst, as shown in Fig. S8(a),† which shows that the *H species is preferentially adsorbed on the Nb atom in the node of Ti(Nb)-ATA. It is worth noting that the limiting free energy change for the HER on Ti(Nb)-ATA is calculated to be 1.29 eV, which is higher than that (1.12 eV) for CO2 reduction; thus the competitive effect of HER is minor. In addition, we also calculated the Gibbs free energy diagrams for CO and HCOOH production from CO2 reduction on Ti(Nb)-ATA, as shown in Fig. S8(b),† and the results show that the stability of the *CHO intermediate is higher than that of CO and HCOOH products; thus the ethanol formation pathway through the CHO* intermediate is mainly considered in this work.
Finally, we calculated the optical absorption spectrum of Ti-ATA and Ti(Nb)-ATA, as shown in Fig. 8(a) and (b). The optical absorption of Ti-ATA and Ti(Nb)-ATA is mainly located in the ultraviolet region. However, the light absorption activity of the modified nodes was slightly improved. Due to the narrow band gap, the response of Ti(Nb)-ATA in visible and infrared bands is slightly stronger than that of Ti-ATA. The absorption spectra of other Ti(M)-ATA catalysts (M = Zr, Hf, Ta) were also calculated, as shown in Fig. S11.† In addition, the work function (Φ) is an important parameter to measure the electron gain/loss ability of materials, which is defined as the minimum energy required for electrons to escape from the interior of materials to the surface.42,43 In DFT calculations, the work function is obtained from the electrostatic potential distribution, which is calculated using the equation Φ = EV – Ef, where Ev and Ef are the potentials of the vacuum energy level and the Fermi energy level, respectively. As shown in Fig. 8(c) and (d), the work function of Ti-ATA is calculated to be 4.23 eV (Fig. 8(c)). After the introduction of Nb atoms into the Ti node, the work function is reduced to 2.91 eV (Fig. 8(d)). The work function of other Ti(M)-ATA catalysts is also lower than that of pristine Ti-ATA (Fig. S12†). The values of the work function, vacuum energy level and Fermi energy level for Ti-ATA and Ti(M)-ATA are provided in Table S6.† Further analysis reveals that the decrease in work function is caused by an up shift in the Fermi level. As mentioned above, the electrons from the Nb atom transfer to the O atom at the node, and the Fermi level of Ti(M)-ATA is enhanced due to Nb–O bonding. The reduced work function is conducive to the transfer of electrons from Ti(M)-ATA to the surface-adsorbed species in the photocatalytic reaction, thereby improving photocatalytic activity.
![]() | ||
Fig. 8 The optical absorption spectra for (a) Ti-ATA and (b) Ti(Nb)-ATA. The electrostatic potential profiles of (c) Ti-ATA and (d) Ti(Nb)-ATA. |
Remarkably, Ti(Nb)-ATA is the best photocatalyst for C2H5OH production among all the candidates studied, and the corresponding optimal reduction pathway is identified as: *CO2 → *OCHO → *OCHOH → *CHO → *CHOCHO → *CHOCH2O → *CHOHCH2O → *CH2OHCH2O → *CH2CH2O + H2O → *CH2CH2OH → *CH3CH2OH → *+C2H5OH, in which the hydrogenation of *CH2CH2O species is the rate-determining step for the overall reaction with a limiting free energy change of 1.12 eV. By analyzing the correlation, it is found that the catalytic activity of these metal node-modified catalysts is highly dependent on their binding strength to the CO2 reactant and key reaction intermediates (e.g. *OCHOH).
In addition, the analysis of electronic and optical properties indicates that the altered energy band structure and charge transfer characteristics at the bimetallic node of Ti(Nb)-ATA are responsible for its superior catalytic activity towards CO2 reduction to C2H5OH compared to pristine Ti-ATA and other metal-modified Ti(M)-ATA candidates. The facile substitution or doping of metals in Ti-MOFs and their structural flexibility and diversity enable us to design a variety of Ti-based MOF photocatalysts with desirable properties. Our findings will stimulate further in-depth experimental studies of Ti-based MOF materials and open up new avenues for developing Ti-MOF-based catalysts for CO2 photoreduction, especially for the synthesis of C2 chemicals.
Footnotes |
† Electronic supplementary information (ESI) available: Fig. S1–S8 and Tables S1–S4. See DOI: https://doi.org/10.1039/d5ta03415a |
‡ These authors contributed equally to this work. |
This journal is © The Royal Society of Chemistry 2025 |