Open Access Article
This Open Access Article is licensed under a Creative Commons Attribution-Non Commercial 3.0 Unported Licence

Unveiling the mechanism of the electroreduction of CO2 to CO over ZIF-8, ZIF-67, and ZIF-90: a DFT perspective

Imtiaz Hussaina, Umar Farooqa and Khurshid Ayub*b
aSchool of Chemistry, University of the Punjab, Lahore, Pakistan
bDepartment of Chemistry, COMSATS University Islamabad, Abbottabad Campus, KPK, 22060, Pakistan. E-mail: khurshid@cuiatd.edu.pk; Fax: +0092 992 383591-5

Received 27th July 2025 , Accepted 23rd November 2025

First published on 11th December 2025


Abstract

The carbon dioxide reduction reaction (CO2RR) is an important solution for reducing carbon levels. Achieving high selectivity and low cost for the catalyst are the main challenges. In this study, computational approaches at the density functional theory (DFT) level are employed to investigate the catalytic activity and mechanism of the conversion of CO2 to CO by zeolitic imidazolate framework (ZIF)-based catalysts (ZIF-8, ZIF-67, ZIF-90). Analysis of the activation energy (Ea) barrier reveals that ZIF-8 is the most efficient (Ea = 0.39 eV) among the considered catalysts. The quantum theory of atoms in molecules (QTAIM) and noncovalent interaction index (NCI) analyses during the activation step of the CO2 molecule reveal the partial covalent and strong electrostatic interactions between the analyte (CO2 molecule) and the catalysts (ZIF-8, ZIF-67, and ZIF-90) at various sites. The natural bonding orbitals (NBO) and electron density difference (EDD) analyses confirm the charge redistribution and bending of the CO2 molecule as it interacts with the catalyst surface during the activation step. Overall, the study's findings indicate ZIF-8 to be the most stable and effective electrocatalyst surface for the CO2RR.


1. Introduction

The consumption of fossil fuels, viz., coal, natural gas, and petroleum, has led to a gradual increase in CO2 emissions, leading to disasters such as global warming,1,2 flooding, droughts, and heat waves.3–7 Since the start of the 21st century, the CO2 emission rate has been increasing at ≈3% per year.8 To achieve the goal of net-zero CO2 emissions (reducing by 80–95% by 20509,10), many chemical efforts are being introduced to overcome the problem via thermal catalysis,11,12 photocatalysis,13,14 and chemical catalysis.15 The electrochemical conversion of CO2 into value-added chemicals such as CO, CH4, CH3OH, HCOOH, and C2H5OH has been employed as an eco-friendly method.16–22 The direct conversion of CO2 to CO demonstrates significant promise, as it involves only two electron transfer steps along with proton transfer.23 The as-made CO can be further reduced by the Fischer–Tropsch process24 to value-added hydrocarbons (C1 and C2 products) and their oxygen derivatives.

As CO2 is inert under most conditions, the electrochemical CO2RR is thermodynamically and kinetically unable to activate C[double bond, length as m-dash]O bonds, hindering the formation of C–C and C–H bonds.25–29 The CO2RR process also suffers from drawbacks, such as poor selectivity toward the target products, inferior catalytic activity, and low energy. It demands the development of efficient electrocatalysts. Many efforts have been devoted to the electrochemical catalysts for the CO2RR, including metals,30–32 metal alloys,33–35 oxides,36,37 and chalcogenides38,39 of transition metals, carbon-based materials,40–42 and single-atom catalysts (SACs).43–45 The noble metals, including Ag, Au, Cu, and Pd, demonstrate outstanding efficacy in the electrocatalytic conversion of CO2 to CO.46–54 Nonetheless, the limitations of these catalysts, including scarcity, elevated costs, and susceptibility to poisoning, significantly hinder their widespread commercial utilization.

Recently, heteroatom-doped carbon materials have emerged as effective catalysts for the electrochemical CO2RR.42,55–58 Particularly, new carbon catalysts doped with atomic metals have garnered a great deal of interest due to their effective electrochemical reduction of CO2 to CO.42,59–62 Metal–organic frameworks (MOFs), porous materials consisting of metal and organic linkers, have superb capability for CO2 capture and conversion due to their tunable bandgap, adjustable pore size, well-defined crystalline structure, and large specific surface area.63–65

Among the various MOFs, ZIFs exhibit superior CO2RR selectivity toward CO. This CO is used for coupling to form C–C, which is necessary for obtaining multi-carbon value-added products.66–68 ZIF-8, ZIF-67, and ZIF-90 were selected for this study due to their electrocatalytic performance, high surface area and porosity, chemical stability, and selective CO2 adsorption. ZIF-8—which consists of Zn2+ linked to 2-methylimidazole (2-mIm) through N moieties in a quadrilateral geometry—exhibits superior electrochemical CO2RR activity among the ZIF family.69–71 ZIF-67, which consists of Co2+ linked to 2-mIm by N moieties in a quadrilateral geometry, has wide applications in various fields, including catalysis.72 ZIF-90, which consists of Zn2+ linked to imidazole-2-carboxyaldehyde (ICA), has wider applications in catalysis, drug delivery, etc.73

Density functional theory (DFT) simulations are considered to be one of the most efficient and advanced tools for theoretical exploration of the pathways of catalysis.74–76 Here, to explore the mystery of the electronic properties of the reactants, products, and intermediates involved, and the pathway of the CO2RR to give CO, a DFT-based comparative study of structurally related ZIFs (ZIF-8, ZIF-67, and ZIF-90) is performed. The aim of the study is to find an efficient and a more-selective electrocatalyst for the CO2RR. We have explored the energetics and the stabilities of various species involved in the conversion of CO2 to CO. The electronic properties and interactions of the analyte, the catalyst, and the intermediates are characterized using frontier molecular orbitals (FMO), natural bonding orbitals (NBO) charge, electron density differences (EDD), noncovalent index interactions (NCI), and quantum theory of atoms in molecules (QTAIM) analyses. This mechanistic study will enable us to design and understand more efficient catalysts for CO2RR.

2. Computational methodology

The Gaussian 09®77 program package was used for DFT calculations. The B3LYP functional with the 6-31G(d,p) basis set was used during geometry optimizations for all non-metallic elements (C, H, O, and N), while the B3LYP/SDD (Stuttgart–Dresden effective core potential) basis set was employed for the Zn and Co atoms. The SDD basis set has been widely used and validated for describing the electronic structure and coordination environment of transition metals in metal–organic frameworks and related catalytic systems. Its use ensures a reliable balance between computational efficiency and accuracy in modeling metal–ligand interactions. The hybrid DFT functional B3LYP is well known for its ability to calculate the thermodynamic parameters and electronic properties of compounds in a quantum-chemical manner.78–80 It is frequently adopted for both geometry optimizations and mechanistic studies as a reliable functional.81–83 For structure visualization, GaussView® (version 5), Visual Molecular Dynamics® (VMD), Multiwfn package® (version 3.8) and Chemcraft® packages were used.84–86

Vibrational frequency analysis was performed for all states at the DFT-optimized geometries to confirm the stationary points as true minima or transition states as saddle points. True minima were confirmed by the absence of imaginary frequencies, while transition states were characterized by a single imaginary frequency. Three ZIF-based catalysts (ZIF-8, ZIF-67, and ZIF-90) were employed in the mechanistic investigation of the CO2 reduction reaction (CO2RR) after being optimized based on their experimentally available structural data. Various spin states were considered for geometry optimization to obtain the most stable spin state (Table S1).

The adsorption energy (Eads) of a species is the amount of energy released when it is adsorbed on a surface or that is required for it to desorb from that surface. Eqn (1) is used to determine the energy for CO2 adsorption (Eads) over various catalysts, where ECO2@Cat stands for the electronic energy of the CO2 adsorbed on the catalyst, ECO2 for the electronic energy of a single CO2 molecule, and ECat for the electronic energy of the specified catalyst.

 
Eads = ECO2@Cat − (ECat + ECO2) (1)

Eqn (2) provides the activation energy (Ea), where ETS and ER are energies of the transition state and reactants, respectively.

 
Ea = ETSER (2)

The reaction energy (Er) corresponds to the heat released or absorbed during a reaction. It helps to find whether a reaction is endothermic (+ve value of Er) or exothermic (−ve value of Er) and is given by eqn (3).

 
Er = EPER (3)
where, EP and ER are the energies of the products (P) and reactants (R), respectively.

EDD and NBO analyses help evaluate the mechanism of the CO2RR by the designed complexes by investigating the nature of donor–acceptor interactions. To explain the nature of the bonding in terms of orbital interaction and charge transfer, NBO analysis was performed at B3LYP/6-31G(d,p) level of theory. Additionally, topological studies such as QTAIM and NCI were carried out using the Multiwfn package (version 3.8) for a better understanding of the nature of interatomic interactions.

3. Results and discussion

3.1. Geometry optimization of catalysts

The geometries of the designated catalysts (ZIF-8, ZIF-67, and ZIF-90) were optimized by taking their single tetrahedron unit before using them as electrocatalysts in CO2RR. In the optimized geometries of the catalysts, the linker (2-mIm for ZIF-8 and ZIF-67 and ICA for ZIF-90) surrounds the central metal (Zn for ZIF-8 and ZIF-90 and Co for ZIF-67) tetrahedrally through nitrogen sites (Fig. 1), which is consistent with the literature.87 The approximate metal–linker (M–L) bond lengths for ZIF-8, ZIF-67, and ZIF-90 are 2.03 Å, 2.01 Å, and 2.06 Å, respectively. These bond lengths are indicative of the strength of the M–L bond and how the linker influences the coordination environment.
image file: d5ra05431a-f1.tif
Fig. 1 Optimized geometries of ZIF-8, ZIF-67, and ZIF-90 with the M–L bond lengths (through nitrogen sites) determined at the B3LYP/6-31G(d,p) (H, C, N, and O) and SDD (Zn and Co) levels of theory.

To understand the influence of changing the central metal ion (in the case of ZIF-8 and ZIF-67) and linkers (in case of ZIF-8 and ZIF-90) on the electronic properties of the catalysts, FMO analysis of the designed catalysts was performed (Fig. 2). The analysis showed that the gap between the highest occupied molecular orbital (HOMO) and the lowest unoccupied molecular orbital (LUMO) is increased by changing the metal from Zn in ZIF-8 to Co in ZIF-67, whereas this gap is reduced by changing the linkers from 2-mIm in ZIF-8 to ICA in ZIF-90. The HOMO–LUMO gap for ZIF-8, ZIF-67, and ZIF-90 is 6.50 eV, 6.56 eV, and 4.77 eV, respectively. In the case of ZIF-67, its stronger M–L bonding, compared with that of ZIF-8, leads to stabilization of the HOMO and destabilization of the LUMO, and increases the gap between them. In the case of ZIF-90, the electron-withdrawing formyl/aldehyde group adds new unoccupied orbitals of low energy and reduces electron density at the linker side, resulting in a reduced HOMO–LUMO gap. The energy values of the HOMO, LUMO and HOMO–LUMO energy gap for ZIF-8, ZIF-67, and ZIF-90 are listed in Table S2.


image file: d5ra05431a-f2.tif
Fig. 2 HOMO–LUMO isosurfaces for ZIF-8, ZIF-67, and ZIF-90.

3.2. ZIF-catalyzed mechanism for the CO2RR

To investigate the mechanism of the CO2RR to give CO,88,89 the geometries of the transition states and intermediates were studied. Initially, the CO2 molecule is weakly physisorbed on the surface of the catalyst, resulting in the formation of a van der Waals complex consisting of a tetrahedral unit of the catalyst and a linear CO2 molecule. During this step, the CO2 molecule orients towards the active site on the catalyst before the activation step. This does not affect the geometries of either of the reactants relative to their free states. The next step is the activation of CO2 by the electrocatalyst, yielding ZIF–COO as an intermediate. This step is followed by the first protonation step, which produces ZIF–COOH. Finally, the second protonation, accompanied by electron transfer, leads to the formation of water gas (CO and H2O), which immediately desorbs from the catalyst surface.90 This is due to the polarity of CO, as its negatively charged C atom does not remain bonded to the C atom of the linker, which is also an electron donor site of the electrocatalyst. To evaluate the catalytic performance for the reduction of CO2 to CO, the energy barrier for the activation step was studied. The activation energy or energy barrier (Ea) values for CO2 using ZIF-8, ZIF-67, and ZIF-90 are listed in Table S3.

3.3. Mechanism of the CO2RR catalyzed by ZIF-8

The energy diagram of the DFT-based mechanistic study of the CO2RR catalyzed by ZIF-8 is shown in Fig. S1. Initially, the reactant CO2 is physiosorbed at the surface of the catalyst, forming a van der Waals complex (vdW) comprising a tetrahedral unit of ZIF-8 and a linear CO2 molecule with a relative energy of −0.27 eV with respect to the energy of the reactants in their free states. In this step, the reactants are properly oriented toward each other and are stabilized relative to their free states by 0.27 eV before activation of the CO2 molecule for the CO2RR. Electron transfer then occurs to form a complex (Int1) comprising the ZIF-8, CO2, and an electron with an electron affinity value of 5.27 eV. This leads to the activation of the CO2 molecule, which results in the formation of a ZIF8–COO intermediate (Int2) by passing through the transition state TS. In TS, CO2 interacts with the catalyst through its C atom at the sp2-hybridized C atom of the linker and is then chemisorbed at the catalyst to give Int2.91 The formation of TS with a relative energy of −5.06 eV must overcome an activation energy barrier of 0.48 eV (11.07 kcal mol−1) to give Int2 with a relative energy of −5.61 eV. The structural parameters, such as vibrational frequencies, bond lengths and bond angles, of the reacting species are critical in the activation step, which demonstrate the stability of the transition state and in turn facilitate protonation. In the case of ZIF-8, the C[double bond, length as m-dash]O bond lengths of the CO2 molecule extend from 1.17 Å in Int1 to 1.27 Å and 1.21 Å in TS and finally to 1.29 Å and 1.22 Å in Int2. During this activation, the C–C bond distance between the interacting carbon atoms of CO2 and the linker is reduced to 2.18 Å in TS and 1.58 Å in Int2, from a distance of 3.39 Å in Int1 (Fig. S1). The O–C–O bond angle also changes from 180° in Int1 to 135° in TS and finally becomes 128° in Int2. The L–M–L bond angle also increases to accommodate the attachment of CO2. In the gas phase, CO2 shows two main vibrational modes, asymmetric stretching at 2436 cm−1 and symmetric stretching at 1372 cm−1. During the activation step, the frequencies of these modes decrease, indicating bending of the CO2 molecules and weakening of the carbon–oxygen double bonds. The asymmetric stretching frequency changes from 2436 cm−1 in the gas phase to 1788 cm−1 in Int1, 1873 cm−1 in TS, and 1805 cm−1 in Int2. This elongation of the C[double bond, length as m-dash]O bond lengths, reduction in the C–C bond distance, and bending of the CO2 molecule upon adsorption, as evident from the bond angle reduction and reduction in vibrational frequencies of the C[double bond, length as m-dash]O bond, confirm the activation of the CO2 molecule over the ZIF-8. This activation step leads to the formation of Int2, releasing 0.17 eV of energy overall in this step. In the next step, the first protonation occurs, giving the ZIF8–COOH intermediate (Int3) with a relative energy of −14.69 eV. The stability of intermediate Int3 is evident by the release of 9.08 eV as proton affinity energy and is in accordance with the value found by Lias et al.92 Int3 undergoes decarboxylation upon the second protonation, accompanied by the addition of an electron to give a product complex comprising ZIF-8 and water gas, i.e., CO and H2O, with a relative energy of −15.72 eV. During this decarboxylation step, 1.03 eV of energy is released, as supported by the study of Hu et al. on the decarboxylation mechanism in the gas phase.93 The enthalpy of the product complex indicates that it is more stable than the reactants.

To ensure that the located transition states truly connect the desired reactants and products, intrinsic reaction coordinate (IRC) calculations were performed for a representative case: the CO2 activation step on ZIF-8. The IRC analysis was carried out using the B3LYP/SDD (for Zn) and 6-31G(d,p) (for non-metal atoms) level of theory. The successful IRC path confirms that the optimized transition state for CO2 activation over ZIF-8 is a true first-order saddle point connecting the correct reactant and product minima. Based on this representative validation, the same TS optimization protocol was applied to other catalysts (ZIF-67 and ZIF-90).

3.4. Mechanism of the CO2RR catalyzed by ZIF-67

The energy diagram of the DFT-based mechanistic study of the CO2RR catalyzed by ZIF-67 is shown in Fig. S2. In this case, the van der Waals complex (vdW′) comprising the tetrahedral unit of ZIF-67 and linear CO2 molecule is formed with a relative energy of −0.11 eV. Following the addition of an electron, an intermediate (Int1′) comprising ZIF-67 and CO2 is formed, with an electron affinity of 6.59 eV. The intermediate ZIF-67–COO (Int2′) is then formed, resulting in the activation of the CO2 molecule by passing through a transition state (TS′). Here, CO2 interacts through its C atom with the ZIF-67 at the sp2-hybridized C atom of 2-mIm. The formation of TS′ with a relative energy of −5.40 eV must overcome an activation energy barrier of 1.30 eV (29.98 kcal mol−1) to form Int2′ with a relative energy of −5.75 eV. In the case of ZIF-67, the C[double bond, length as m-dash]O bond lengths of the CO2 molecule extend from 1.17 Å in Int1′ to 1.23 Å and 1.20 Å in TS′ and finally to 1.30 Å and 1.22 Å in Int2′. During this activation, the C–C bond distance between the carbon atom of CO2 and that of the linker is reduced from 4.46 Å in Int1′ to 2.18 Å in TS′ and 1.57 Å in Int2′ (Fig. S2). The O–C–O bond angle changes from 180° in Int1′ to 143° in TS′ and finally becomes 128° in Int2′. The L–M–L bond angle also increases to accommodate the attachment of CO2. Frequency analysis of the activation step revealed a redshift in the values of the CO2 molecule, indicating its bending and weakening of its carbon–oxygen bonds. The asymmetric stretching frequency of CO2 changes from 2436 cm−1 in the gas phase to 2427 cm−1 in Int1, 1996 cm−1 in TS, and 1788 cm−1 in Int2. These changes confirm the activation of the CO2 molecule over ZIF-67. This activation step leads to the formation of Int2′ with the absorption of an energy of 0.50 eV. In the next step, the first protonation occurs, giving a ZIF67–COOH intermediate (Int3′) with a relative energy of −14.42 eV. The stability of intermediate Int3′ is evident from the release of 8.67 eV as proton affinity energy.92 Int3′ undergoes decarboxylation upon the second protonation, accompanied by the addition of an electron, giving a product complex comprising ZIF-67 and water gas, i.e. CO and H2O, with a relative energy of −16.23 eV.

3.5. Mechanism of the CO2RR catalyzed by ZIF-90

The energy diagram of the DFT-based mechanistic study of the CO2RR catalyzed by ZIF-90 is shown in Fig. S3. For ZIF-90, the van der Waals complex (vdW″) is formed and stabilized with a relative energy of −0.40 eV. The activation of the CO2 molecule is initiated by the addition of an electron to give a complex (Int1″) including ZIF-90, CO2 and an electron with an electron affinity of 6.53 eV. The activation of the CO2 molecule results in the formation of the ZIF90–COO intermediate (Int2″) by passing through a transition state (TS″). CO2 interacts through its C atom with the sp2-hybridized C atom of the ICA in the ZIF-90. The formation of TS″ with a relative energy of −5.42 eV must overcome an activation energy barrier of 1.51 eV (34.82 kcal mol−1) to give Int2″ with a relative energy of −5.52 eV. In the case of ZIF-90, the C[double bond, length as m-dash]O bond lengths of the CO2 molecule extend from 1.17 Å in Int1″ to 1.23 Å and 1.20 Å in TS″ and finally become 1.28 Å and 1.22 Å in Int2″. During this activation, the C–C bond distance between the interacting carbon atoms of CO2 and the linker is reduced to 1.90 Å in TS″ and 1.61 Å in Int2″, from a distance of 4.05 Å in Int1″ (Fig. S3). The O–C–O bond angle changes from 180° in Int1″ to 142° in TS″ and finally becomes 131° in Int2″. The L–M–L bond angle also increases to accommodate the attachment of CO2. Frequency analysis of the activation step revealed a redshift for CO2, which indicated its bending and weakening of its carbon–oxygen bonds. The asymmetric stretching frequency of CO2 changes from 2436 cm−1 in the gas phase to 2433 cm−1 in Int1, 1999 cm−1 in TS, and 1829 cm−1 in Int2. These changes indicate the activation of the CO2 molecule over the ZIF-90. This activation step leads to the formation of Int2″ by absorbing an energy of 1.41 eV overall in this step. In the next step, the first protonation occurs, giving a ZIF90–COOH intermediate (Int3″) with a relative energy of −14.04 eV. The stability of intermediate Int3″ is evident by the release of 8.52 eV as proton affinity energy.92 Int3″ undergoes decarboxylation upon second protonation, accompanied by the addition of an electron, giving a product complex comprising ZIF-90 and water gas, i.e., CO and H2O, with a relative energy of −16.39 eV.

3.6. Effect of solvent

To evaluate the solvent effect on the energetics of the reaction mechanism, single-point SMD calculations of the involved intermediates and transition states were performed using water as a solvent. This systematically modified the energetics of the intermediates, but did not alter the sequence of the mechanism observed in the gas phase.

Transition states for all catalysts were stabilized by the addition of the solvent effect, as evident from the lowering of their activation barriers. In the cases of ZIF-8, ZIF-67, and ZIF-90, Ea changes from 0.48 eV, 1.30 eV, and 1.51 eV to 0.11 eV, 0.83 eV, and 0.95 eV, with the activation barriers being lowered by 0.37 V, 0.47 V and 0.56 V, respectively.

The solvent stabilization of the transition state structures (TS) implies that transition states with a polar nature will be more stabilized in the aqueous environment of the electrochemical experiment, thus lowering the kinetic barriers for the reaction to occur. The solvent causes some intermediates like Int3 to gain a significant amount of stability (for instance, ZIF-8 Int3: −9.08 → −12.98 eV), which hints that polar/charged species are intensively solvated. Conversely, the very early intermediates (Int1 to be exact) are the least stabilized by solvation in the case of ZIF-8 (ZIF-8 Int1: −5.27 → −1.75 eV), which implies that the effects caused by solvation are relatively dependent on the charge distribution and shape of each species.

While the energies were altered in absolute terms, the relative positions of the intermediates and the identity of the probable rate-determining step remained nearly the same in qualitative terms in both the gas and implicit-solvent data. A good example is the situation in which the energies of all the TSs decrease, but ZIF-90 retains the highest TS energy (0.95 eV) among the three in the SMD calculations, which is in line with the gas-phase trend.

The net effect of solvation is the lowering of the activation barriers, which is seen as additional stabilization of the polar intermediates; this is in line with lower overpotentials and changed selectivity in aqueous electrochemical cells, which are consistent with the experimental observations. Thus, the gas-phase investigation gives a mechanistic view, and the SMD results indicate that solvent inclusion changes the specific barrier heights but not the proposed mechanism.

3.7. Comparison of catalytic activity

The activation energy barriers for the CO2RR over the proposed catalysts are between 0.48 and 1.51 eV, as shown in Table S3. Overall, ZIF-8 exhibits the lowest activation energy barrier among the catalysts considered in this study (0.39 eV), which can be readily overcome under mild conditions. This low activation barrier for ZIF-8, as compared to the other ZIFs, can be explained by comparing the geometries of their transition states for the activation step. The geometries of the transition state for the CO2 reduction reaction over ZIF-8, ZIF-67, and ZIF-90 are quite complex and differ from each other in accordance with their respective activation barriers. In the frequency analyses of the respective species, a progressive redshift is observed in the asymmetric stretching mode from the reactant (Int1) via the transition state (TS) and the intermediate (Int2). This confirms the stepwise activation and the gradual bending of CO2. ZIF-8 shows the strongest frequency reduction among all the studied catalysts; this is an indication of stronger catalyst –CO2 interaction and more effective activation (Table S4). A correlation exists between this trend and the higher catalytic activity of ZIF-8 that has been reported experimentally for CO2 reduction reactions. In the transition state of ZIF-8, the bond lengths of C[double bond, length as m-dash]O are 1.27 Å and 1.21 Å, which are greater than that of 1.17 Å in free CO2 and suggest a great deal of charge transfer from the catalyst to CO2. In contrast, the ZIF-67 and ZIF-90 transition states have less pronounced C[double bond, length as m-dash]O elongation with distances of 1.23 Å and 1.20 Å, respectively, which indicates lower charge transfer. Another critical factor influencing activation is the O–C–O bond angle; its significant reduction to 135° in the case of ZIF-8 strongly favors CO2 activation. For ZIF-67 and ZIF-90, the angles are greater than those for ZIF-8, being 143° and 142°, respectively, indicating reduced bending of CO2 and therefore weaker activation. The C–C bond length between carbon dioxide and the sp2-hybrid carbon of the linker also offers insight regarding the interaction strength. Shorter bond lengths indicate stronger interactions and more-efficient charge transfer. ZIF-90 possesses a shorter C–C distance (1.90 Å) than ZIF-8 (2.18 Å) and ZIF-67 (2.18 Å). However, ZIF-90 has the highest activation barrier of 1.51 eV due to the lack of charge transfer from the imidazole-2-carboxaldehyde linker, which is poor at transition state stabilization because of its lower effectiveness. The trend in the observed activation barriers (ZIF-8 < ZIF-67 < ZIF-90) is a result of the CO2 distortion, linking strength, and charge transfer from the complex. Regarding these ZIFs, ZIF-8 has the most changeable transition state shape, thus leading to the smallest amount of energy being needed for activation. This indicates that, in comparison to the other catalysts under study, ZIF-8 is the best candidate for the CO2RR because it offers the most favorable transition state geometry together with the lowest energy needed for activation.

Our DFT study, which revealed the mechanistic trends, showed a strong relationship to the experimentally observed CO2R activity and selectivity in ZIF-based catalysts. Specifically, the extent of CO2 activation (indicated by the ν3 reduction in the frequencies of IR modes, charge transfer, and transition state stabilization) correlates with higher CO selectivity and lower overpotential in the experimental data. Compared to those obtained using Ag,94 Cu and Cu2O,95 and Pd3Au96 surfaces as the electrocatalyst for the CO2RR, the computed CO2RR barrier for ZIF-8 is substantially smaller. ZIF-8 is a potential catalyst for the CO2RR owing to its highest activity among all catalysts studied here.

3.8. QTAIM and NCI analyses

The nature of the interactions between the analyte and the catalyst was analyzed using a topological analysis, quantum theory of atoms in molecules (QTAIM). The values of the topological parameters, viz., the electron density (ρ), Laplacian of electron density (∇2ρ), sum of electron densities (H), kinetic energy (G), potential energy (V), and interaction energy (Eint) of individual bonds, were used to gain deep insight into the nature of interactions. Values of ρ less than 0.1 a.u. indicate the presence of weak interactions (for van der Waals <0.02 a.u., for partially covalent 0.02–0.2 a.u.), whereas values greater than 0.2 a.u. represent strong covalent interactions. Positive values of ∇2ρ and H represent noncovalent interactions, whereas negative values indicate covalent interactions. Values of Eint less negative than −3 kcal mol−1 for individual bonds represent weak interactions such as van der Waals forces; values of −3 to −10 kcal mol−1 represent strong electrostatic interactions like hydrogen bonding, and values more negative than −10 kcal mol−1 represent strong interactions, for example, covalent bond values are −50 kcal mol−1 or more negative.84,97 The Eint of individual bonds was calculated by using eqn (4).
 
Eint = V/2 (4)

The ratio −V/G is another parameter that indicates the nature of interactions. It should be less than 1 a.u. for noncovalent interactions and 1–2 a.u. for partially covalent interactions. At bond critical points (BCPs), the values of G and V remain positive and negative, respectively. The QTAIM analysis results for intermediate and transition states are given in Table S5, whereas BCPs are shown in Fig. 3.


image file: d5ra05431a-f3.tif
Fig. 3 QTAIM analysis results of respective intermediates and transition states for ZIF-8, ZIF-67, and ZIF-90.

Through QTAIM analysis, 3 and 5 BCPs were seen for Int1 and TS on the ZIF-8 surface, 4 and 5 BCPs were discovered for Int1′ and TS′ on the ZIF-67 surface, and 3 and 4 BCPs were discovered for Int1″ and TS″ on the ZIF-90 surface. An increase in the number of BCPs for the respective transition states relative to that of intermediates for each catalytic surface indicates the stabilization of the transition state relative to the reactants. At all the BCPs of the intermediates (Int1, Int1′ and Int1″), the values of ρ and –V/G are less than 0.02 and 1, respectively (Table S5). The values of H and ∇2ρ are positive, and the values of Eint are less negative than −3 kcal mol−1. These values of the structural parameters indicate the presence of weak noncovalent interactions such as van der Waals interactions. In the TS for ZIF-8, at the BCPs C50–C14 and O51–Zn32, the values of all these parameters are in ranges indicating partially covalent interactions, while at the BCP O51–H40, the value of Eint is −5.648 kcal mol−1, suggesting the presence of hydrogen bonding. In the TS′ for ZIF-67, at two BCPs, C49–C14 and O50–Co52, the values of all these parameters indicate partially covalent interaction, while at the BCP O50–H39, the value of Eint is −4.706 kcal mol−1, indicating the presence of hydrogen bonding. In the TS″ for ZIF-90, at the two BCPs, C46–C3 and O47–Zn1, the values of all these structural parameters are in the range suggesting partial covalent interaction. These QTAIM parameters at the remaining BCPs for the TS indicate noncovalent interactions.

QTAIM analysis of the transition states for ZIF-8, ZIF-67, and ZIF-90 confirmed a distinct trend concerning stability. ZIF-8 has the electron density (ρ) at the bond critical points (BCPs), with 0.062 a.u. for C50–C14 and 0.055 a.u. for O51–Zn32, indicating the strongest bonding interactions. ZIF-67 comes next with moderate values of 0.052 a.u. for C49–C14 and 0.038 a.u. for O50–Co52, while ZIF-90 shows the weakest interactions of 0.038 a.u. for O47–Zn1 and 0.022 a.u. for C46–C3. Similarly, the most negative value of potential energy density (V), which indicates the level of bond stabilization, was recorded for ZIF-8 (−0.050 a.u. for C50–C14 and −0.108 a.u. for O51–Zn32), ZIF-67 (−0.036 a.u. for C49–C14 and −0.074 a.u. for O50–Co52) and the least negative value for ZIF-90 (−0.050 a.u. for C46–C3 and −0.064 a.u. for O47–Zn1). The interaction energy (Eint = V/2) shows the same trend, with ZIF-8 having the most stabilizing values (−15.688 kcal mol−1 for C50–C14 and −33.885 kcal mol−1 for O51–Zn32), followed by ZIF-67 (−11.295 kcal mol−1 for C49–C14 and −23.218 kcal mol−1 for O50–Co52), and ZIF-90 having the least stabilizing interactions (−15.688 kcal mol−1 for C46–C3 and −20.080 kcal mol−1 for O47–Zn1). The bond character ratio (−(V/G)) supports this, with ZIF-8 demonstrating the highest values that show strong to partially covalent bonding, ZIF-67 showing some covalent and ionic character, and ZIF-90 showing more closed-shell bonding interactions.

The reason that ZIF-8 has a comparatively higher transition state stability than ZIF-67 and ZIF-90 can be explained by considering the bonding interactions. ZIF-8 has a higher electron density, more negative potential energy density, and larger interaction energy, suggesting greater charge transfer efficiency, which yields a more stable transition state. The cobalt in ZIF-67 tends to alter the electronic environment, weakening the interactions a bit compared to those in ZIF-8, hence its moderate stability. Comparatively, ZIF-90 has the least amount of those bonds because of the presence of ICA, which decreases charge transfer efficiency and stablilizes bonds, which in turn explains the observed progression towards weaker transition state stability: that of ZIF-8 is greater than ZIF-67, which is greater than ZIF-90, where ZIF-8 is most favorable for CO2 activation, and ZIF-90 is the least stable.

Noncovalent interaction index (NCI) analysis is another tool for studying the nature of the different attractive as well as repulsive interactions, viz., H-bonding, weak van der Waals interactions, and steric repulsion interactions between an analyte and catalyst. It assists in finding the stability and reactivity of the transition states and intermediates of interest. The NCI analysis is based on the reduced density gradient (RDG) maps (2D and 3D).98,99 These RDG maps and isosurfaces are further based on ρ and ∇2ρ values, as shown in eqn (5).

 
image file: d5ra05431a-t1.tif(5)

Both the 2D maps and 3D isosurfaces depend upon the values of sign(λ2)ρ. In 2D maps, sign(λ2)ρ is plotted at the abscissa while RDG is plotted at the ordinate. The factors sign(λ2) and ρ represent the type and strength of interaction, respectively. Positive, slightly negative, and highly negative values of sign(λ2)ρ indicate steric repulsions, weak van der Waals interactions, and strong electrostatic interactions, and appear as red-, green-, and blue-colored areas or patches in the maps, respectively. These RDG maps and isosurfaces of the respective intermediates (Int1, Int1′ and Int1″) and transition states (TS, TS′ and TS″) for ZIF-8, ZIF-67, and ZIF-90 are shown in Fig. 4.


image file: d5ra05431a-f4.tif
Fig. 4 NCI isosurfaces of the transition states and intermediates for ZIF-8, ZIF-67, and ZIF-90.

The green patches in the 3D RDG isosurfaces of the intermediates (Int1, Int1′ and Int1″) indicate van der Waals interactions between the interacting moieties of fragments. The appearance of green areas in the 2D RDG maps in the range of 0.002 to −0.018 a.u. indicates weak van der Waals interactions between the fragments. For the transition states, the blue donut-shaped patches in the 3D RDG isosurfaces indicate strong electrostatic interactions, e.g. partially covalent interactions and H-bonding, and thick green patches indicate weak van der Waals interactions between the interacting moieties of the fragments. In the 2D RDG maps, the appearance of green areas in the range of 0.002 to −0.018 a.u. correspond to weak van der Waals interactions and the blue areas in the range of −0.018 to −0.050 a.u. indicate the strong electrostatic interactions, e.g. H-bonding and partially covalent interactions between interacting moieties of the fragments. This analysis of the isosurfaces indicates the stability of the transition states as compared to the complexes. In case of TS for ZIF-8, the strong attractive forces (as shown by the donut-shaped blue patches in Fig. 4) between the carbon atoms of CO2 and the linker group of the catalyst correspond to the greater charge transfer capability of ZIF-8 and lead to a low activation barrier. This barrier increases in TS′ of ZIF-67 compared to that in TS of ZIF-8 due to the increased repulsive forces operating (as evident from the red-colored patches in Fig. 4) between O of CO2 and N atom of the linker group of the catalyst. In the case of ZIF-90, TS″ has the highest energy barrier due to the presence of only weak attractive forces (as evident from the cyan colored patch in Fig. 4) between the C atoms of CO2 and the linker group of the catalyst and increased repulsive interactions.

3.9. Natural bond orbitals (NBO) and electron density differences (EDD) analyses

The activation and reduction of CO2 over the designed catalysts were further confirmed by natural bond orbitals (NBO) and electron density differences (EDD) analyses. NBO analysis was performed to determine the exact amount of charge transfer from the donor (catalyst) to the acceptor (analyte, i.e., CO2 molecule). The calculated values of the NBO charges on the corresponding species, i.e., analyte and catalyst, are listed in Table S6. Study of the intermediates (Int1, Int1′ and Int1″) and transition states (TS, TS′ and TS″) in the NBO analysis indicated that the catalyst transfers its charge to the analyte species and activates the CO2 molecule for its subsequent reduction to CO. In the case of the intermediates, the NBO charges on the catalyst moiety are +1.003, 0.994, and 0.997, while those for the analyte are −0.003, 0.006, and 0.003 for ZIF-8, ZIF-67, and ZIF-90, respectively. In the transition states, the NBO charges on the catalysts are +1.612, 1.387, and 1.354, whereas those on the analytes are −0.612, −0.387, and −0.354 in the case of ZIF-8, ZIF-67, and ZIF-90, respectively. It clearly indicates the charge transfer from the catalyst to the analyte. This charge transfer follows the trend of highest to lowest for ZIF-8, ZIF-67, and ZIF-90, which explains the order of their activation barriers from lowest to highest for CO2 activation, respectively. The greater the charge transfer from the catalyst to CO2, the easier its activation and hence the smaller activation barrier will be. Overall, this charge transfer reveals that the catalyst shifts charge to the analyte, assisting the C–O bond elongation and activation of the CO2 molecule for further reduction to CO.

The electron density difference (EDD) contours for the relevant intermediates and transition states were used to investigate the interaction of CO2 molecules with the ZIF-8, ZIF-67, and ZIF-90 surfaces. Fig. 5 shows the contours for the interaction of CO2 on ZIF-8, ZIF-67, and ZIF-90. Purple represents the accumulation of charge, while pink represents the depletion of charge. The contours of the respective intermediates and transition states for each catalyst demonstrate that, upon adsorption of CO2 on the catalyst surface, the carbon atom of CO2 interacts with the sp2-hybridized carbon atom of the corresponding linker, converting its localized π-bond character into a delocalized, resonating form. During this interaction, the CO2 molecule bends, and both C–O bonds elongate. For ZIF-8, these bonds elongate from 1.17 Å to 1.27 Å and 1.20 Å, for ZIF-67 from 1.170 Å to 1.235 Å and 1.198 Å, and for ZIF-90 from 1.171 Å to 1.200 Å and 1.240 Å. This bending and bond elongation in CO2 causes the redistribution of the electronic density. It also reveals that the charge is being transferred from the catalyst to the analyte. The EDD results in Fig. 5 agree with the NBO analysis, indicating a strong interaction between the analyte and the catalytic surface.


image file: d5ra05431a-f5.tif
Fig. 5 EDD analysis showing the bending of the CO2 molecule, C–O bond elongation, and charge redistribution upon moving from the first intermediate to the transition state for ZIF-8, ZIF-67, and ZIF-90.

4. Conclusion

In this study, the catalytic reduction of CO2 to CO over a selection of ZIFs (ZIF-8, ZIF-67, and ZIF-90) was comprehensively investigated using quantum-chemical DFT investigations. Comparison of the CO2 activation barriers over these ZIFs indicated that the activation of CO2 is more thermodynamically favorable over ZIF-8, as indicated by its having the lowest activation barrier (0.48 eV) compared to those for ZIF-67 (1.30 eV) and ZIF-90 (1.51 eV). The better catalytic activity of ZIF-8 is attributed to its better electron transferring capability and stabilization of the carboxylate group produced during activation in comparison to ZIF-67 and ZIF-90. This facilitates the subsequent reduction to CO. QTAIM and NCI analyses were carried out to explore the interactions occurring during the activation step of the CO2 molecule. These reveal the partially covalent and strong electrostatic interactions at the respective interacting sites of the analyte (CO2 molecule) and the catalyst (ZIF-8, ZIF67 and ZIF-90). NBO and EDD analyses of the activation step also confirmed the charge redistribution and bending of the CO2 molecule as it interacts with the catalyst surface. During the protonation step of the carboxylate ion, the values of released energy for ZIF-8, ZIF-67, and ZIF-90 are 9.08 eV, 8.67 eV, and 8.52 eV, indicating the stabilization of the produced carboxyl group. Upon electron-coupled proton transfer, decarboxylation occurs, which releases an energy of 1.03 eV, 1.81 eV, and 2.35 eV, respectively. Overall, this study indicates that the ZIF-8 complex is an effective electrocatalyst for the CO2RR, which gives researchers a new opportunity to enhance the catalytic potential of ZIF-8 and create a novel electrocatalyst derived from ZIF-8 for effective CO2RR.

Conflicts of interest

The authors declare that they have no known competing financial interests or personal relationships that could have appeared to influence the work reported in this paper.

Data availability

Data will be made available upon request.

Supplementary information (SI) is available. See DOI: https://doi.org/10.1039/d5ra05431a.

Acknowledgements

The authors are grateful to Dr Sehrish Sarfaraz for her insightful recommendations and unwavering assistance during the research work of this manuscript. The authors are also thankful to the University of the Punjab, Lahore, Pakistan for providing facilities during this work as a host institution.

References

  1. S. Chu and A. Majumdar, Opportunities and challenges for a sustainable energy future, Nature, 2012, 488(7411), 294–303 CrossRef CAS PubMed.
  2. S. J. Davis, K. Caldeira and H. D. Matthews, Future CO2 emissions and climate change from existing energy infrastructure, Science, 2010, 329(5997), 1330–1333 CrossRef CAS.
  3. M. McNutt, Time's up, CO2, Science, 2019, 365(6452), 411 CrossRef CAS.
  4. H. Mikulčić, I. R. Skov, D. F. Dominković, S. R. W. Alwi, Z. A. Manan, R. Tan and X. Wang, Flexible Carbon Capture and Utilization technologies in future energy systems and the utilization pathways of captured CO2, Renew. Sustain. Energy Rev., 2019, 114, 109338 CrossRef.
  5. Y. Ma, X. Wang, Y. Jia, X. Chen, H. Han and C. Li, Titanium dioxide-based nanomaterials for photocatalytic fuel generations, Chem. Rev., 2014, 114(19), 9987–10043 CrossRef CAS.
  6. P. F. Sui, C. Xu, M. N. Zhu, S. Liu, Q. Liu and J. L. Luo, Interface-Induced Electrocatalytic Enhancement of CO2-to-Formate Conversion on Heterostructured Bismuth-Based Catalysts, Small, 2022, 18(1), 2105682 CrossRef CAS PubMed.
  7. J. M. Ali, A. M. Mohammed and Y. S. Mekonnen, Mechanistic study on the coupling reaction of CO2 with propylene oxide catalyzed by (CH3)4PI· MgCl2, J. Comput. Chem., 2022, 43(14), 961–971 CrossRef CAS PubMed.
  8. C. S. Galik, A continuing need to revisit BECCS and its potential, Nat. Clim. Change, 2020, 10(1), 2–3 CrossRef.
  9. S. J. Davis, N. S. Lewis, M. Shaner, S. Aggarwal, D. Arent, I. L. Azevedo and K. Caldeira, Net-zero emissions energy systems, Science, 2018, 360(6396), eaas9793 CrossRef PubMed.
  10. R. Meys, A. Kätelhön, M. Bachmann, B. Winter, C. Zibunas, S. Suh and A. Bardow, Achieving net-zero greenhouse gas emission plastics by a circular carbon economy, Science, 2021, 374(6563), 71–76 CrossRef CAS.
  11. G. D. Fao and J. C. Jiang, Theoretical investigation of CO2 conversion on corrugated g-C3N4 Surface decorated by single-atom of Fe, Co, and Pd, Mol. Catal., 2022, 526, 112402 CAS.
  12. F. Raziq, A. Aligayev, H. Shen, S. Ali, R. Shah, S. Ali and L. Qiao, Exceptional photocatalytic activities of rGO modified (B, N) Co-doped WO3, coupled with CdSe QDs for one photon Z-scheme system: a joint experimental and dft study, Adv. Sci., 2022, 9(2), 2102530 CrossRef CAS PubMed.
  13. F. Raziq, K. Khan, S. Ali, S. Ali, H. Xu, I. Ali and L. Qiao, Accelerating CO2 reduction on novel double perovskite oxide with sulfur, carbon incorporation: synergistic electronic and chemical engineering, Chem. Eng. J., 2022, 446, 137161 CrossRef CAS.
  14. F. Wahid, S. Ali, P. M. Ismail, F. Raziq, S. Ali, J. Yi and L. Qiao, Metal single atom doped 2D materials for photocatalysis: current status and future perspectives, Prog. Energy, 2023, 5, 012001 CrossRef CAS.
  15. S. Wang, G. Li, L. Chen, Z. Li, Z. Wu, X. Kong and Z. Liu, The emerging applications of metal phosphides in carbon dioxide reduction reaction, Funct. Mater. Lett., 2021, 14(07), 2130017 CrossRef CAS.
  16. W. Zhang, Y. Hu, L. Ma, G. Zhu, Y. Wang, X. Xue and Z. Jin, Progress and perspective of electrocatalytic CO2 reduction for renewable carbonaceous fuels and chemicals, Adv. Sci., 2018, 5(1), 1700275 CrossRef PubMed.
  17. D. Gao, R. M. Arán-Ais, H. S. Jeon and B. Roldan Cuenya, Rational catalyst and electrolyte design for CO2 electroreduction towards multicarbon products, Nat. Catal., 2019, 2(3), 198–210 CrossRef CAS.
  18. D. U. Nielsen, X. M. Hu, K. Daasbjerg and T. Skrydstrup, Chemically and electrochemically catalysed conversion of CO2 to CO with follow-up utilization to value-added chemicals, Nat. Catal., 2018, 1(4), 244–254 CrossRef CAS.
  19. S. Ali, P. M. Ismail, F. Wahid, A. Kumar, M. Haneef, F. Raziq and H. Xu, Benchmarking the two-dimensional conductive Y3(C6X6)2 (Y= Co, Cu, Pd, Pt; X= NH, NHS, S) metal-organic framework nanosheets for CO2 reduction reaction with tunable performance, Fuel Process. Technol., 2022, 236, 107427 CrossRef CAS.
  20. S. Ali, R. Iqbal, F. Wahid, P. M. Ismail, A. Saleem, S. Ali and L. Qiao, Cobalt coordinated two-dimensional covalent organic framework a sustainable and robust electrocatalyst for selective CO2 electrochemical conversion to formic acid, Fuel Process. Technol., 2022, 237, 107451 CrossRef CAS.
  21. S. Ali, G. Yasin, R. Iqbal, X. Huang, J. Su, S. Ibraheem and H. Xu, Porous aza-doped graphene-analogous 2D material a unique catalyst for CO2 conversion to formic-acid by hydrogenation and electroreduction approaches, Mol. Catal., 2022, 524, 112285 CAS.
  22. R. Iqbal, M. B. Akbar, A. Ahmad, A. Hussain, N. Altaf, S. Ibraheem and S. Ali, Exploring the Synergistic Effect of Novel Ni-Fe in 2D Bimetallic Metal-Organic Frameworks for Enhanced Electrochemical Reduction of CO2, Adv. Mater. Interfaces, 2022, 9(1), 2101505 CrossRef CAS.
  23. C. Cometto, L. Chen, P. K. Lo, Z. Guo, K. C. Lau, E. Anxolabéhère-Mallart and M. Robert, Highly selective molecular catalysts for the CO2-to-CO electrochemical conversion at very low overpotential. Contrasting Fe vs. Co quaterpyridine complexes upon mechanistic studies, ACS Catal., 2018, 8(4), 3411–3417 CrossRef CAS.
  24. G. Yun, S. Y. Hwang, J. Y. Maeng, Y. J. Kim, C. K. Rhee and Y. Sohn, Electrochemical CO2/CO reduction on Ag/Cu electrodes and exploring minor Fischer–Tropsch reaction pathways, Appl. Surf. Sci., 2024, 649, 159179 CrossRef CAS.
  25. B. Hu, C. Guild and S. L. Suib, Thermal, electrochemical, and photochemical conversion of CO2 to fuels and value-added products, J. CO2 Util., 2013, 1, 18–27 CrossRef CAS.
  26. Y. Zhou, R. Zhou, X. Zhu, N. Han, B. Song, T. Liu and Y. Li, Mesoporous PdAg nanospheres for stable electrochemical CO2 reduction to formate, Adv. Mater., 2020, 32(30), 2000992 CrossRef CAS PubMed.
  27. P. Su, W. Xu, Y. Qiu, T. Zhang, X. Li and H. Zhang, Ultrathin bismuth nanosheets as a highly efficient CO2 reduction electrocatalyst, ChemSusChem, 2018, 11(5), 848–853 CrossRef CAS PubMed.
  28. C. Costentin, S. Drouet, G. Passard, M. Robert and J. M. Savéant, Proton-coupled electron transfer cleavage of heavy-atom bonds in electrocatalytic processes. Cleavage of a C–O bond in the catalyzed electrochemical reduction of CO2, J. Am. Chem. Soc., 2013, 135(24), 9023–9031 CrossRef CAS.
  29. C. Yang, J. Chai, Z. Wang, Y. Xing, J. Peng and Q. Yan, Recent progress on bismuth-based nanomaterials for electrocatalytic carbon dioxide reduction, Chem. Res. Chin. Univ., 2020, 36(3), 410–419 CrossRef CAS.
  30. M. T. Tang, H. Peng, P. S. Lamoureux, M. Bajdich and F. Abild-Pedersen, From electricity to fuels: descriptors for C1 selectivity in electrochemical CO2 reduction, Appl. Catal., B, 2020, 279, 119384 CrossRef CAS.
  31. D. Y. Jo, H. C. Ham and K. Y. Lee, Facet-dependent electrocatalysis in the HCOOH synthesis from CO2 reduction on Cu catalyst: a density functional theory study, Appl. Surf. Sci., 2020, 527, 146857 CrossRef CAS.
  32. Z. Li, D. He, X. Yan, S. Dai, S. Younan, Z. Ke and J. Gu, Size-dependent nickel-based electrocatalysts for selective CO2 reduction, Angew. Chem., 2020, 132(42), 18731–18736 CrossRef.
  33. Y. Li, Z. Tian and L. Chen, Theoretical understanding of the interface effect in promoting electrochemical CO2 reduction on Cu–Pd alloys, J. Phys. Chem. C, 2021, 125(39), 21381–21389 CrossRef CAS.
  34. T. M. Suzuki, T. Ishizaki, S. Kosaka, N. Takahashi, N. Isomura, J. Seki and T. Morikawa, Electrochemical CO2 reduction over nanoparticles derived from an oxidized Cu–Ni intermetallic alloy, Chem. Commun., 2020, 56(95), 15008–15011 RSC.
  35. Y. Xu, C. Li, Y. Xiao, C. Wu, Y. Li, Y. Li and J. He, Tuning the selectivity of liquid products of CO2RR by Cu–Ag alloying, ACS Appl. Mater. Interfaces, 2022, 14(9), 11567–11574 CrossRef CAS.
  36. N. Kumar, N. Seriani and R. Gebauer, DFT insights into electrocatalytic CO2 reduction to methanol on α-Fe2O3 (0001) surfaces, Phys. Chem. Chem. Phys., 2020, 22(19), 10819–10827 RSC.
  37. B. Liu, X. Yao, Z. Zhang, C. Li, J. Zhang, P. Wang and C. Zhao, Synthesis of Cu2O nanostructures with tunable crystal facets for electrochemical CO2 reduction to alcohols, ACS Appl. Mater. Interfaces, 2021, 13(33), 39165–39177 CrossRef CAS PubMed.
  38. F. Y. Gao, Z. Z. Wu and M. R. Gao, Electrochemical CO2 reduction on transition-metal chalcogenide catalysts: recent advances and future perspectives, Energy Fuels, 2021, 35(16), 12869–12883 CrossRef CAS.
  39. X. Shao, X. Zhang, Y. Liu, J. Qiao, X. D. Zhou, N. Xu and J. Zhang, Metal chalcogenide-associated catalysts enabling CO2 electroreduction to produce low-carbon fuels for energy storage and emission reduction: catalyst structure, morphology, performance, and mechanism, J. Mater. Chem. A, 2021, 9(5), 2526–2559 RSC.
  40. T. Asset, S. T. Garcia, S. Herrera, N. Andersen, Y. Chen, E. J. Peterson and P. Atanassov, Investigating the nature of the active sites for the CO2 reduction reaction on carbon-based electrocatalysts, ACS Catal., 2019, 9(9), 7668–7678 CrossRef CAS.
  41. S. Fu, X. Liu, J. Ran, Y. Jiao and S. Z. Qiao, CO2 reduction by single copper atom supported on g-C3N4 with asymmetrical active sites, Appl. Surf. Sci., 2021, 540, 148293 CrossRef CAS.
  42. C. Zhang, S. Yang, J. Wu, M. Liu, S. Yazdi, M. Ren and J. M. Tour, Electrochemical CO2 reduction with atomic iron-dispersed on nitrogen-doped graphene, Adv. Energy Mater., 2018, 8(19), 1703487 CrossRef.
  43. Z. Feng, Y. Tang, Y. Ma, Y. Li, Y. Dai, H. Ding and X. Dai, Theoretical investigation of CO2 electroreduction on N (B)-doped graphdiyne mononlayer supported single copper atom, Appl. Surf. Sci., 2021, 538, 148145 CrossRef CAS.
  44. Z. Feng, G. Su, H. Ding, Y. Ma, Y. Li, Y. Tang and X. Dai, Atomic alkali metal anchoring on graphdiyne as single-atom catalysts for capture and conversion of CO2 to HCOOH, Mol. Catal., 2020, 494, 111142 CAS.
  45. Z. Feng, Y. Tang, Y. Ma, Y. Li, Y. Dai, W. Chen and X. Dai, Theoretical computation of the electrocatalytic performance of CO2 reduction and hydrogen evolution reactions on graphdiyne monolayer supported precise number of copper atoms, Int. J. Hydrogen Energy, 2021, 46(7), 5378–5389 CrossRef CAS.
  46. Y. Wang, J. Liu, Y. Wang, A. M. Al-Enizi and G. Zheng, Tuning of CO2 reduction selectivity on metal electrocatalysts, Small, 2017, 13(43), 1701809 CrossRef PubMed.
  47. M. Cho, J. M. Kim, B. Kim, S. Yim, Y. J. Kim, Y. S. Jung and J. Oh, Versatile, transferrable 3-dimensionally nanofabricated Au catalysts with high-index crystal planes for highly efficient and robust electrochemical CO2 reduction, J. Mater. Chem. A, 2019, 7(11), 6045–6052 RSC.
  48. Y. Zhao, C. Wang, Y. Liu, D. R. MacFarlane and G. G. Wallace, Engineering surface amine modifiers of ultrasmall gold nanoparticles supported on reduced graphene oxide for improved electrochemical CO2 reduction, Adv. Energy Mater., 2018, 8(25), 1801400 CrossRef.
  49. S. Zhao, Z. Tang, S. Guo, M. Han, C. Zhu, Y. Zhou and Z. Kang, Enhanced activity for CO2 electroreduction on a highly active and stable ternary Au-C Dots-C3N4 electrocatalyst, ACS Catal., 2018, 8(1), 188–197 CrossRef CAS.
  50. Y. Zhang, L. Hu and W. Han, Insights into in situ one-step synthesis of carbon-supported nano-particulate gold-based catalysts for efficient electrocatalytic CO2 reduction, J. Mater. Chem. A, 2018, 6(46), 23610–23620 RSC.
  51. J. Kim, J. T. Song, H. Ryoo, J. G. Kim, S. Y. Chung and J. Oh, Morphology-controlled Au nanostructures for efficient and selective electrochemical CO2 reduction, J. Mater. Chem. A, 2018, 6(12), 5119–5128 RSC.
  52. H. Q. Fu, L. Zhang, L. R. Zheng, P. F. Liu, H. Zhao and H. G. Yang, Enhanced CO2 electroreduction performance over Cl-modified metal catalysts, J. Mater. Chem. A, 2019, 7(20), 12420–12425 RSC.
  53. A. Seifitokaldani, Switching between CO2 Electroreduction Pathways. in Electrochemical Society Meeting Abstracts, The Electrochemical Society, Inc., 2019, vol. 235, 31, pp. 1588 Search PubMed.
  54. J. Wang, S. Kattel, C. J. Hawxhurst, J. H. Lee, B. M. Tackett, K. Chang and J. G. Chen, Enhancing activity and reducing cost for electrochemical reduction of CO2 by supporting palladium on metal carbides, Angew. Chem., Int. Ed., 2019, 58(19), 6271–6275 CrossRef CAS PubMed.
  55. H. Cui, Y. Guo, L. Guo, L. Wang, Z. Zhou and Z. Peng, Heteroatom-doped carbon materials and their composites as electrocatalysts for CO2 reduction, J. Mater. Chem. A, 2018, 6(39), 18782–18793 RSC.
  56. A. S. Varela, W. Ju and P. Strasser, Molecular nitrogen–carbon catalysts, solid metal organic framework catalysts, and solid metal/nitrogen-doped carbon (MNC) catalysts for the electrochemical CO2 reduction, Adv. Energy Mater., 2018, 8(30), 1703614 CrossRef.
  57. A. Vasileff, Y. Zheng and S. Z. Qiao, Carbon solving carbon's problems: recent progress of nanostructured carbon-based catalysts for the electrochemical reduction of CO2, Adv. Energy Mater., 2017, 7(21), 1700759 CrossRef.
  58. X. Duan, J. Xu, Z. Wei, J. Ma, S. Guo, S. Wang and S. Dou, Metal-free carbon materials for CO2 electrochemical reduction, Adv. Mater., 2017, 29(41), 1701784 CrossRef.
  59. X. Wang, Z. Chen, X. Zhao, T. Yao, W. Chen, R. You and Y. Li, Regulation of coordination number over single Co sites: triggering the efficient electroreduction of CO2, Angew. Chem., 2018, 130(7), 1962–1966 CrossRef.
  60. C. Yan, H. Li, Y. Ye, H. Wu, F. Cai, R. Si and X. Bao, Coordinatively unsaturated nickel–nitrogen sites towards selective and high-rate CO2 electroreduction, Energy Environ. Sci., 2018, 11(5), 1204–1210 RSC.
  61. H. B. Yang, S. F. Hung, S. Liu, K. Yuan, S. Miao, L. Zhang and B. Liu, Atomically dispersed Ni (i) as the active site for electrochemical CO2 reduction, Nat. Energy, 2018, 3(2), 140–147 CrossRef CAS.
  62. W. Ren, X. Tan, W. Yang, C. Jia, S. Xu, K. Wang and C. Zhao, Isolated diatomic Ni-Fe metal–nitrogen sites for synergistic electroreduction of CO2, Angew. Chem., Int. Ed., 2019, 58(21), 6972–6976 CrossRef CAS PubMed.
  63. G. Singh, J. Lee, A. Karakoti, R. Bahadur, J. Yi, D. Zhao and A. Vinu, Emerging trends in porous materials for CO2 capture and conversion, Chem. Soc. Rev., 2020, 49(13), 4360–4404 RSC.
  64. X. Wang, Y. Zou, Y. Zhang, B. Marchetti, Y. Liu, J. Yi and J. Zhang, Tin-based metal organic framework catalysts for high-efficiency electrocatalytic CO2 conversion into formate, J. Colloid Interface Sci., 2022, 626, 836–847 CrossRef CAS.
  65. T. Najam, S. S. A. Shah, M. S. Bashir, and A. U. Rehman, Covalent Organic Framework-Based Electrocatalysts for CO2 Reduction Reaction. in Noble Metal-Free Electrocatalysts: Fundamentals and Recent Advances in Electrocatalysts for Energy Applications, American Chemical Society, 2022, vol 1, pp. 257–274 Search PubMed.
  66. Y. Zheng, A. Vasileff, X. Zhou, Y. Jiao, M. Jaroniec and S. Z. Qiao, Understanding the roadmap for electrochemical reduction of CO2 to multi-carbon oxygenates and hydrocarbons on copper-based catalysts, J. Am. Chem. Soc., 2019, 141(19), 7646–7659 CrossRef CAS.
  67. V. H. Nguyen, B. S. Nguyen, Z. Jin, M. Shokouhimehr, H. W. Jang, C. Hu and Q. Van Le, Towards artificial photosynthesis: Sustainable hydrogen utilization for photocatalytic reduction of CO2 to high-value renewable fuels, Chem. Eng. J., 2020, 402, 126184 CrossRef CAS.
  68. T. P. Nguyen, D. L. T. Nguyen, V. H. Nguyen, T. H. Le, D. V. N. Vo, Q. T. Trinh and Q. V. Le, Recent advances in TiO2-based photocatalysts for reduction of CO2 to fuels, Nanomaterials, 2020, 10(2), 337 CrossRef CAS.
  69. Y. Wang, P. Hou, Z. Wang and P. Kang, Zinc imidazolate metal–organic frameworks (ZIF-8) for electrochemical reduction of CO2 to CO, ChemPhysChem, 2017, 18(22), 3142–3147 CrossRef CAS PubMed.
  70. X. Jiang, H. Li, J. Xiao, D. Gao, R. Si, F. Yang and X. Bao, Carbon dioxide electroreduction over imidazolate ligands coordinated with Zn (II) center in ZIFs, Nano Energy, 2018, 52, 345–350 CrossRef CAS.
  71. S. Dou, J. Song, S. Xi, Y. Du, J. Wang, Z. F. Huang and X. Wang, Boosting electrochemical CO2 reduction on metal–organic frameworks via ligand doping, Angew. Chem., Int. Ed., 2019, 58(12), 4041–4045 CrossRef CAS.
  72. J. Cao, S. Sun, X. Li, Z. Yang, W. Xiong, Y. Wu and Y. Zhang, Efficient charge transfer in aluminum-cobalt layered double hydroxide derived from Co-ZIF for enhanced catalytic degradation of tetracycline through peroxymonosulfate activation, Chem. Eng. J., 2020, 382, 122802 CrossRef CAS.
  73. I. S. Flyagina, E. M. Mahdi, K. Titov and J. C. Tan, Thermo-mechanical properties of mixed-matrix membranes encompassing zeolitic imidazolate framework-90 and polyvinylidine difluoride: ZIF-90/PVDF nanocomposites, APL Mater., 2017, 5(8), 086104 CrossRef.
  74. X. G. TIAN, Y. ZHANG and T. S. YANG, First-principles study of H2 dissociative adsorption reactions on WO3 surfaces, Acta Phys.-Chim. Sin., 2012, 28(5), 1063–1069 CAS.
  75. A. A. Latimer, A. R. Kulkarni, H. Aljama, J. H. Montoya, J. S. Yoo, C. Tsai and J. K. Nørskov, Understanding trends in C–H bond activation in heterogeneous catalysis, Nat. Mater., 2017, 16(2), 225–229 CrossRef CAS PubMed.
  76. Z. F. Huang, J. Song, Y. Du, S. Xi, S. Dou, J. M. V. Nsanzimana and X. Wang, Chemical and structural origin of lattice oxygen oxidation in Co–Zn oxyhydroxide oxygen evolution electrocatalysts, Nat. Energy, 2019, 4(4), 329–338 CrossRef CAS.
  77. M. Frisch, F. Clemente, G. Trucks, H. Schlegel, G. Scuseria, M. Robb, J. Cheeseman, G. Scalmani, V. Barone, B. Mennucci, G. A. Petersson, H. Nakatsuji, M. Caricato, X. Li, H. P. Hratchian, A. F. Izmaylov, J. Bloino, G. Zhe, Gaussian 09, Revision a. 01, 2009, pp. 20–44 Search PubMed.
  78. H. Tahir, N. Kosar, K. Ayub and T. Mahmood, Outstanding NLO response of thermodynamically stable single and multiple alkaline earth metals doped C20 fullerene, J. Mol. Liq., 2020, 305, 112875 CrossRef CAS.
  79. R. Ahmadi and S. Pourkarim, Investigation of fullerene (C), Int. J. Bio-Inorg. Hybr. Nanomater., 2015, 4(4), 249–254 Search PubMed.
  80. S. Osuna, M. Swart and M. Sola, Dispersion corrections essential for the study of chemical reactivity in fullerenes, J. Phys. Chem., 2011, 115(15), 3491–3496 CrossRef CAS.
  81. M. T. Baei, A. A. Peyghan, Z. Bagheri and M. B. Tabar, B-doping makes the carbon nanocones sensitive towards NO molecules, Phys. Lett. A, 2012, 377(1–2), 107–111 CrossRef CAS.
  82. S. F. Rastegar, A. A. Peyghan and H. Soleymanabadi, Ab initio studies of the interaction of formaldehyde with beryllium oxide nanotube, Phys. E Low-dimens. Syst. Nanostruct., 2015, 68, 22–27 CrossRef CAS.
  83. A. A. Peyghan, H. Soleymanabadi and M. Moradi, Structural and electronic properties of pyrrolidine-functionalized [60] fullerenes, J. Phys. Chem. Solids, 2013, 74(11), 1594–1598 CrossRef CAS.
  84. S. Sarfaraz, M. Yar, M. Ans, M. A. Gilani, R. Ludwig, M. A. Hashmi and K. Ayub, Computational investigation of a covalent triazine framework (CTF-0) as an efficient electrochemical sensor, RSC Adv., 2022, 12(7), 3909–3923 RSC.
  85. S. Sarfaraz, M. Yar and K. Ayub, Covalent triazine framework (CTF-0) surface as a smart sensing material for the detection of CWAs and industrial pollutants, Mater. Sci. Semicond. Process., 2022, 139, 106334 CrossRef CAS.
  86. S. Sarfaraz, M. Yar, A. A. Khan, R. Ahmad and K. Ayub, DFT investigation of adsorption of nitro-explosives over C2N surface: Highly selective towards trinitro benzene, J. Mol. Liq., 2022, 352, 118652 CrossRef CAS.
  87. X. Sun, M. Keywanlu and R. Tayebee, Experimental and molecular dynamics simulation study on the delivery of some common drugs by ZIF-67, ZIF-90, and ZIF-8 zeolitic imidazolate frameworks, Appl. Organomet. Chem., 2021, 35(11), e6377 CrossRef CAS.
  88. P. Saha, S. Amanullah and A. Dey, Selectivity in electrochemical CO2 reduction, Acc. Chem. Res., 2022, 55(2), 134–144 CrossRef CAS.
  89. G. M. Tomboc, S. Choi, T. Kwon, Y. J. Hwang and K. Lee, Potential link between Cu surface and selective CO2 electroreduction: perspective on future electrocatalyst designs, Adv. Mater., 2020, 32(17), 1908398 CrossRef CAS.
  90. F. Y. Gao, R. C. Bao, M. R. Gao and S. H. Yu, Electrochemical CO2-to-CO conversion: electrocatalysts, electrolytes, and electrolyzers, J. Mater. Chem. A, 2020, 8(31), 15458–15478 RSC.
  91. J. H. Cho, C. Lee, S. H. Hong, H. Y. Jang, S. Back, M. G. Seo and S. Y. Kim, Transition metal ion doping on ZIF-8 enhances the electrochemical CO2 reduction reaction, Adv. Mater., 2023, 35(43), 2208224 CrossRef CAS.
  92. S. G. Lias, J. F. Liebman and R. D. Levin, Evaluated gas phase basicities and proton affinities of molecules; heats of formation of protonated molecules, J. Phys. Chem. Ref. Data, 1984, 13(3), 695–808 CrossRef CAS.
  93. Y. Hu, L. Gao, Z. Dai, G. Sun, T. Zhang, S. Jia and X. Zhang, DFT investigation on the decarboxylation mechanism of ortho hydroxy benzoic acids with acid catalysis, J. Mol. Model., 2016, 22(3), 56 CrossRef.
  94. A. Seifitokaldani, C. M. Gabardo, T. Burdyny, C. T. Dinh, J. P. Edwards, M. G. Kibria and E. H. Sargent, Hydronium-induced switching between CO2 electroreduction pathways, J. Am. Chem. Soc., 2018, 140(11), 3833–3837 CrossRef CAS PubMed.
  95. J. Jiao, R. Lin, S. Liu, W. C. Cheong, C. Zhang, Z. Chen and Y. Li, Copper atom-pair catalyst anchored on alloy nanowires for selective and efficient electrochemical reduction of CO2, Nat. Chem., 2019, 11(3), 222–228 CrossRef CAS PubMed.
  96. M. Zheng, X. Zhou, Y. Zhou and M. Li, Theoretical insights into mechanisms of electrochemical reduction of CO2 to ethylene catalyzed by Pd3Au, Appl. Surf. Sci., 2022, 572, 151474 CrossRef CAS.
  97. Y. S. Al-Faiyz, S. Sarfaraz, M. Yar, S. Munsif, A. A. Khan, B. Amin and K. Ayub, Efficient detection of nerve agents through carbon nitride quantum dots: a DFT approach, Nanomaterials, 2023, 13(2), 251 CrossRef CAS.
  98. A. Mukhtar, S. Sarfaraz and K. Ayub, Organic transformations in the confined space of porous organic cage CC2; catalysis or inhibition, RSC Adv., 2022, 12(37), 24397–24411 RSC.
  99. M. Asif, S. Sarfaraz and K. Ayub, DFT investigation of M2F superalkali doped dodecafluorophenylene (C13H10F12) derivatives with remarkable static and dynamic NLO response, Phys. Scr., 2024, 99(6), 065108 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.