Next Article in Journal
Visible-Light Photocatalytic Degradation of Methylene Blue by Yb3+-Doped 3D Nanosheet Arrays BiOI Anchored on High-Chloride Fly Ash Composites
Next Article in Special Issue
Synthesis of M-Doped MoSe2 (M = Fe, Co, Ni) via Chemical Vapor Deposition for an Electrocatalytic Hydrogen Evolution Reaction
Previous Article in Journal
Cadmium Complexes—A Novel Family in the Coordination Chemistry of 1,2-bis(arylimino)acenaphthenes
Previous Article in Special Issue
Boosting C-C Coupling for Electrochemical CO2 Reduction over Novel Cu-Cubic Catalysts with an Amorphous Shell
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Insights into Pd-Nb@In2Se3 Electrocatalyst for High-Performance and Selective CO2 Reduction Reaction from DFT

1
School of Physics and Electric Engineering, Anyang Normal University, Anyang 455000, China
2
College of Science, Nanjing Forestry University, Nanjing 210037, China
*
Author to whom correspondence should be addressed.
Inorganics 2025, 13(5), 146; https://doi.org/10.3390/inorganics13050146
Submission received: 10 April 2025 / Revised: 1 May 2025 / Accepted: 3 May 2025 / Published: 5 May 2025

Abstract

:
The electrochemical CO2 reduction reaction (eCO2RR), driven by renewable energy, represents a promising strategy for mitigating atmospheric CO2 levels while generating valuable fuels and chemicals. Its practical implementation hinges on the development of highly efficient electrocatalysts. In this study, a novel dual-metal atomic catalyst (DAC), composed of niobium and palladium single atoms anchored on a ferroelectric α-In2Se3 monolayer (Nb-Pd@In2Se3), is proposed based on density functional theory (DFT) calculations. The investigation encompassed analyses of structural and electronic characteristics, CO2 adsorption configurations, transition-state energetics, and Gibbs free energy changes during the eCO2RR process, elucidating a synergistic catalytic mechanism. The Nb-Pd@In2Se3 DAC system demonstrates enhanced CO2 activation compared to single-atom counterparts, which is attributed to the complementary roles of Nb and Pd sites. Specifically, Nb atoms primarily drive carbon reduction, while neighboring Pd atoms facilitate oxygen species removal through proton-coupled electron transfer. This dual-site interaction lowers the overall reaction barrier, promoting efficient CO2 conversion. Notably, the polarization switching of the In2Se3 substrate dynamically modulates energy barriers and reaction pathways, thereby influencing product selectivity. Our work provides theoretical guidance for designing ferroelectric-supported DACs for the eCO2RR.

Graphical Abstract

1. Introduction

Carbon dioxide (CO2), a principal climate-altering gas, drives critical environmental challenges including rising global temperatures, climatic instability, ocean acidification, and ecological carbon cycle disturbances [1,2]. Urgent action is required to achieve net-zero anthropogenic carbon dioxide equivalent emissions by 2050, as mandated by the Intergovernmental Panel on Climate Change (IPCC) to cap global temperature increases at 1.5 °C above preindustrial baselines [3]. This imperative is becoming increasingly critical given the persistent reliance on emission-generating technologies and escalating global energy demands. Addressing the dual challenge of atmospheric CO2 mitigation and sustainable utilization has emerged as a pivotal research frontier [4,5,6,7]. Carbon management encompasses three interconnected approaches to mitigating anthropogenic CO2 emissions. First, energy generation processes are being optimized through advanced combustion technologies, such as oxy-fuel combustion with pure oxygen injection and homogeneous charge compression ignition (HCCI) engines using biofuel blends [8,9]. Second, modern carbon capture technologies are advancing beyond conventional geological storage in saline aquifers or depleted oil fields by transforming captured CO2 into value-added products like construction aggregates, agricultural bio-fertilizers, and energy-efficient syngas, leveraging advanced separation processes such as pressure swing adsorption (PSA) and membrane systems [10,11]. Third, the chemical conversion pathways, particularly electrocatalytic CO2 reduction reaction (eCO2RR), utilize multi-electron transfer processes (2e to 12e) with tailored catalysts (e.g., Cu for hydrocarbons, Ag/Zn for CO) to produce value-added fuels and chemicals like CO, formate, methane, ethylene, and ethanol, while enabling renewable energy integration and high-pressure reactor designs for industrial scalability [12,13].
Among the above approaches, the eCO2RR stands out for its dual capacity to mitigate emissions and generate valuable chemical feedstocks [14,15,16]. And this technology has four key advantages. First, the eCO2RR leverages renewable electricity from solar, wind, or tidal energy to convert CO2 into chemicals, enabling carbon-neutral cycles while reducing reliance on fossil fuels [17]. Second, the process generates high-value chemicals like ethylene (feedstock for plastics) [18], methane (fuel) [19], and formate (industrial reagent) [20], directly replacing fossil-derived equivalents in existing supply chains. Third, unlike thermocatalytic methods requiring high temperatures/pressures, the eCO2RR operates efficiently at room temperature and atmospheric pressure, reducing energy demands and infrastructure complexity. Forth, modular flow cell designs and membrane electrode assemblies (MEAs) allow scalable deployment for industrial adoption [21].
The eCO2RR process involves intricate multi-stage reaction sequences: (1) CO2 adsorption at electrode interfaces; (2) proton=coupled electron transfer (PCET) steps, which forms C–H bonds with broken C=O bonds; and (3) detachment of synthesized compounds from active sites through desorption mechanisms and their subsequent transport into the bulk reaction medium via dissolution–diffusion pathways [22]. This sequential process becomes particularly consequential in water-containing systems where the parasitic hydrogen evolution reaction (HER) actively competes with the CO2RR for catalytic resources and electron transfer, fundamentally altering the thermodynamic landscape and resulting in suboptimal product distributions [23].
In the process of reducing CO2 with metal electrodes, there are many challenges, including the impact of the poor stability of the electrodes. Many metal electrodes, such as palladium (Pd), indium (In), and tin (Sn), have stability issues when reducing CO2. Take Pd as an example; CO has strong adsorption energy on its surface, which causes the electrode to be easily poisoned and deactivated, greatly shortening its service life [24]. Although In and Sn electrodes can achieve the reduction of CO2, they often require a higher voltage to drive the reaction [25,26,27], which not only causes a large waste of energy but may also trigger side reactions due to high voltage, further reducing the selectivity and efficiency of the reaction. At the same time, the stability of the electrode deteriorates further under high voltage, accelerating the electrode’s wear and tear. These problems seriously restrict the practical application of metal electrodes in the field of CO2 reduction [26]. Effective eCO2RR implementation hinges on catalytic systems being optimized across three critical dimensions, namely conversion efficiency under applied potentials, target product selectivity over competing reactions, and long-term structural and functional stability. These properties are intrinsically linked to atomic-scale catalyst architecture, particularly the coordination environment of active sites.
Since the conceptualization of single-atom catalysts (SACs) in 2011, these atomically dispersed metal systems have revolutionized catalytic research through maximum metal utilization efficiency (approaching 100%) [28,29,30,31,32,33], well-defined active site geometries for mechanistic studies [34,35,36], and enhanced selectivity profiles compared to nanoparticle analogues. However, SAC implementation faces two fundamental constraints: low active site density due to synthesis limitations and inadequate multi-intermediate stabilization in complex eCO2RR pathways [37,38]. This has prompted researchers to turn their attention to bimetallic atomic catalyst (DAC) systems with greater structural flexibility.
DACs exhibit distinct advantages over isolated SACs, combining the inherent merits of SACs with supplementary synergistic benefits: (1) DACs substantially elevate metal loading capacities while expanding the population of accessible reactive centers, a critical factor for scalable catalytic applications [39]; (2) DACs permit dynamic spatial arrangements of active sites, enabling coordinated interactions essential for multi-step CO2 reduction pathways [40,41]; (3) electronic modulation between dissimilar metal atoms allows precise regulation of intermediate binding strengths, creating tunable platforms for reaction pathway optimization and selectivity control [40,42,43,44,45,46,47]; (4) independent yet coexisting metallic centers enable parallel reaction channels, permitting simultaneous production of distinct value-added compounds through site-specific catalytic processes [48]; (5) and in DACs, one metallic component can modify support interactions that subsequently influence the catalytic behavior of adjacent metal sites, creating cascading activation mechanisms [30,49,50]. It is anticipated that further studies will concentrate on enhancing the performance of bimetallic atomic site catalysts.
Contemporary DACs predominantly utilize two-dimensional nitrogen-doped carbon substrates, as exemplified by Fe-Co [51], Ni-Fe [52], and Co-Pt [53] dual-atom configurations, which have been fabricated with dual-solvent, pyrolysis, and cyclic voltammetry methods, respectively. However, these conventional DACs exhibit inherent limitations in post-synthesis catalytic tunability due to their inability to leverage substrate property modifications for performance enhancement. Recent advancements in ferroelectric material engineering provide a promising solution. Substantial theoretical and experimental evidence demonstrates that polarization direction modulation in ferroelectrics not only governs molecular adsorption/desorption dynamics but also enables precise control over surface catalytic activity and product selectivity [54,55,56]. Capitalizing on this principle, we engineered a novel ferroelectric DAC system by anchoring niobium (Nb) and palladium (Pd) single atoms onto an α-In2Se3 monolayer substrate (Nb-Pd@In2Se3). The α-In2Se3 monolayer is a van der Waals material exhibiting coupled in-plane and out-of-plane ferroelectric polarization at room temperature [56,57,58]. It features a quintuple atomic-layer structure (Se-In-Se-In-Se) with spontaneous polarization arising from asymmetric Se atom displacements, enabling bidirectional electric field control of both polarization axes. This unique architecture allows for the dynamic regulation of metal–support interactions, addressing the static catalytic behavior inherent to traditional carbon-based DACs. The theoretical analysis revealed that the dual-metal atomic configuration significantly enhances CO2 activation capability during the eCO2RR process. In the synergistic catalytic mechanism, the Nb atomic sites predominantly function as catalytic centers for carbon reduction, while the adjacent Pd counterparts facilitate oxygen species removal through a complementary catalytic pathway. Notably, polarization switching of the ferroelectric substrate modulates both the energy barriers along the reaction coordinates and the dominant reaction pathways, ultimately dictating the selectivity of terminal products.

2. Results and Discussion

2.1. Geometric Structures of Nb-Pd@In2Se3

The 2D ferroelectric α-phase In2Se3 monolayer was chosen as the substrate material. Based on our previous work [56], we selected Pd and Nb transition metals as candidates for loaded metallic dual atoms, since they can stably form Pd@In2Se3 and Nb@In2Se3 SACs. According to the position of the intermediate Se layer offset (see Figure 1a,b), the α-In2Se3 monolayer has two stable polarization phases, namely polarization upward (P↑) and downward (P↓). We refer to the configuration where the Pd-Nb dual-atom pair is doped on the outer surface of two layers with a Se spacing of 3.06 Å as the Pd-Nb@P↓-In2Se3. Moreover, the configuration where the Pd-Nb dual-atom pair is doped on the outer surface of two layers of Se with a spacing of 3.84 Å is called Pd-Nb@P↑-In2Se3. As for the SACs [59,60,61], strong binding strength is essential for the stability of DACs. At the energetically most favorable site, the binding energy of the Nb-Pd dual-atom pair in Pd-Nb@P↓-In2Se3 and Pd-Nb@P↑-In2Se3 were −15.69 and −12.27 eV/unit, respectively. The low binding energies indicate that there is a strong interaction between the metal atoms and the substrate, as evidenced by the strong covalent Pd–Se and Nb-Se bonds (see Figure 1a,b), which are caused by the strong hybridization of Se p and Pd/Nb d orbitals near the Fermi level (see Figure S1a,b).
To further demonstrate the structural stability of Pd-Nb@In2Se3, ab initio molecular dynamics (AIMD) calculations were carried out on these two configuration at a temperature of 1000 K, with a time step of 1 fs and a total time length of 10 ps (see Figure 2a,b). During the AIMD simulations, the geometrical structure did not change significantly, and the total energy of the system tended to be stable, indicating that these configurations have good thermal stability. Overall, both the binding energy and molecular dynamics simulation results show the promising feasibility of fabricating Pd-Nb@In2Se3 DACs.

2.2. Activation of CO2 on Pd-Nb@In2Se3 DACs

The effective activation of CO2 molecules is a key step for the subsequent reduction reaction. The linear structure of CO2 molecules contributes to their pronounced chemical inertness, primarily due to the symmetry-induced cancellation of dipole moments despite the presence of polar C=O bonds. This linear geometry stabilizes the molecule thermodynamically, as the two C=O bonds exhibit dissociation energies of approximately 750 kJ·mol−1. Additionally, the large energy gap (13.7 eV) between the highest occupied molecular orbital (HOMO) and lowest unoccupied molecular orbital (LUMO) creates a significant kinetic barrier to electron transfer, necessitating substantial energy input for chemical transformations. In catalytic systems employing isolated atomic site catalysts (SACs), CO2 adsorption typically occurs via either oxygen– or carbon–metal (O-M or C-M) interactions at the active metal centers. However, the limited electron donating capacity and geometric constraints of isolated atomic sites restrict the extent of CO2 activation. Consequently, SACs often require elevated temperatures and pressures to achieve moderate reaction rates, further reducing reaction efficiency and selectivity toward multi-carbon products.
Dual-site catalysts can demonstrate enhanced CO2 adsorption via coordination involving both metallic dual atoms and support atoms, requiring lower activation energy than the SAC to activate CO2 [62]. As shown in Table S1, by separately comparing the adsorption energy of CO2 on the Pd and Nb sites of Pd-Nb@In2Se3 DACs (see Figure S2), we conclude that the adsorption of CO2 on the Pd site had more negative adsorption energy in both P↑ and P↓ configurations, which means the CO2 tends to adsorb on the Pd site of Pd-Nb@In2Se3 DACs. The ∠OCO values of CO2 adsorbed on Pd-Nb@P↓-In2Se3 and Pd-Nb@P↑-In2Se3 were significantly reduced to 131.38° and 125.18° from an initial value of 180°, respectively. Based on the structural analysis, it could be stated that the Pd-Nb@In2Se3 catalyst could effectively activate CO2 molecules. Moreover, according to our previous report [56] that the ∠OCO values of CO2 adsorbed on Pd@P↓-In2Se3 and Pd@P↑-In2Se3 were 149.2° and 151.2°, respectively, it can be concluded that the synergistic effect of a bimetallic atomic site causes a more obvious deformation of adsorbed CO2 molecules (17.82° and 26.02° for P↓ and P↑, respectively), indicating a higher activation of CO2. The similar synergistic effects in the Pt-Ni@CeO2 bimetallic system were extensively validated through comprehensive characterization and computational studies [62]. Temperature-programmed reduction (TPR) and desorption (TPD) analyses revealed that Pt-Ni synergy significantly enhances the generation of reactive oxygen intermediates (e.g., surface peroxides and hydroxyl radicals) and strengthens metal–support interactions via charge transfer mechanisms. These modifications collectively optimize CO2 chemisorption capacity, as evidenced by the higher adsorption efficiency compared to the monometallic counterparts (Pt@CeO2 and Ni@CeO2). Density functional theory (DFT) simulations further demonstrated a 35–40% reduction in CO2 activation energy (Ea = 10.1 kcal mol−1 for Pt-Ni@CeO2 vs. 32.9 and 36.5 kcal mol−1 for Ni@CeO2 and Pt@CeO2, respectively), which is attributed to electronic structure modulation at Pt-Ni-CeO2 interfacial sites.
In order to explore the activation mechanism, we employed a charge density difference calculation, which demonstrated that there is both a loss and gain of electrons on the adsorbed CO2 molecule and the bimetallic atomic sites (see Figure 3a,b). This is because the empty Pd/Nb d orbital can accept electrons from the highest occupied molecular orbitals (1πg orbital) of the CO2 molecule, while the occupied Pd/Nb d orbital can back-donate electrons to the lowest unoccupied molecular orbitals (2πu orbital) of the CO2 molecule (see Figure 3c). The synergy of electron donation and backdonation-dependent d orbital occupation ensures that CO2 can be efficiently activated [56].
As mentioned before, the CO2 adsorbed on Pd-Nb@P↑-In2Se3 (∠OCO = 125.18°) had a more obvious deformation than that absorbed on Pd-Nb@P↓-In2Se3 (∠OCO = 131.38°), which means Pd-Nb@P↑-In2Se3 has higher CO2 activation. In order to investigate the reason for this, we calculated the Pd and Nb d orbitals in Pd-Nb@P↑-In2Se3 and Pd-Nb@P↓-In2Se3 DACs, respectively. Then, based on the partial electronic orbitals (see Figure 4a), we obtained the d band center of the Pd and Nb atoms for both kinds of Pd-Nb@In2Se3 DACs. As displayed in Figure 4b, both d band centers of the Pd (−2.09 eV) and Nb (0.11 eV) atoms in Pd-Nb@P↑-In2Se3 were closer to the Fermi level than the ones of Pd (−2.68 eV) and Nb (0.44 eV) atoms in Pd-Nb@P↓-In2Se3. And it has been reported that the exceptional electrocatalytic performance of Co2 anchored graphdiyne in nitrogen fixation originates from the presence of localized electronic states adjacent to the Fermi level, which optimize charge transfer dynamics during the reaction [63]. Therefore, in the case of P↑, the proximity of the d band center of Pd and Nb atoms to the Fermi level induces a high CO2 activation performance. Note that the different d band centers of bimetallic atomic sites in Pd-Nb@In2Se3 under different polarization states implies that both the reaction path and the limiting potential of the subsequent CO2 hydrogenation process may be affected by the polarization switch.

2.3. eCO2RR on Pd-Nb@In2Se3 DACs

Considering the complexity of the CO2 reaction (see Figure S3), we have summarized the minimum-energy reaction paths of the CO2RR over Pd-Nb@In2Se3 catalysts (see Figure 5). The CO2RR pathways over Pd-Nb@In2Se3 catalysts exhibit stark polarization-dependent selectivity, governed by the structural and electronic asymmetry induced by the ferroelectric In2Se3 substrate’s polarization orientation (P↓ and P↑). The CO2 reduction reaction begins with the adsorption of CO2 onto the Pd-Nb@P↓-In2Se3 catalyst surface, forming a CO2 intermediate (CO2 + * → CO2*). This adsorbed CO2 undergoes a proton-coupled electron transfer (PCET) step, where one hydrogen atom binds to an O atom (CO2* + H+ + e → OCHO*), leading to the formation of an OCHO intermediate (bent configuration with O–CH–O bonding). Subsequently, a second protonation occurs at the adjacent oxygen atom (OCHO* + H+ + e → HOCHO*), stabilizing the HOCHO intermediate (adsorbed formate-like structure). Finally, the HOCHO intermediate desorbs from the catalyst surface (HOCHO* → HOCHO + *), releasing formic acid (HOCHO) as the end product. Due to the high energy barrier associated with stabilizing the second protonation step, the rate-determining step (PDS) is identified as the transformation of OCHO* → HOCHO*, requiring an applied potential of −0.45 V vs. RHE. In contrast, in the CO2RR on a Pd-Nb@P↑-In2Se3 catalyst, the reaction course is significantly different from that on Pd-Nb@P↓-In2Se3, which undergoes an alternative protonation pathway. The first PCET step preferentially protonates the carbon atom, forming the OCOH intermediate (CO2* + H+ + e → OCOH*) (linear configuration with O–C–O–H bonding). Further reduction cleaves the C–O bond. A second protonation occurs at the hydroxyl group of OCOH intermediate. The hydroxyl group binds a hydrogen ion to generate H2O and desorb (OCOH* + H+ + e → CO* + H2O), producing a CO intermediate adsorbed on the Pd site of the catalyst. Sequential hydrogenation steps follow: CO* + H+ + e → CHO* (formyl), CHO* + H+ + e → CH2O* (formaldehyde-like) and CH2O* + H+ + e → CH2* + OH* via C–O bond scission. The critical C–O bond scission step (CH2O* + H+ + e → CH2* + OH*) in the methane production pathway over the Pd-Nb@P↓-In2Se3 catalyst is driven by the synergistic interplay between the Pd and Nb dual active sites. The CH2 intermediate absorbs the Pd site, and the OH intermediate absorbs the Nb site. CH2* undergoes progressive hydrogenation (CH2* + OH* + H+ + e → CH3* + OH*), followed by a critical step where CH3 reacts with a co-adsorbed OH group to form CH3 and H2O (CH3* + OH* + H+ + e → CH3* + H2O*). Then, the H2O molecule desorbs from the Nb site (CH3* + H2O* → CH3* + H2O). This step is identified as the PDS, with a limiting potential of −0.66 V vs. RHE, reflecting the substantial energy required for H2O dissociation. The final hydrogenation of CH3 yields methane (CH4) (CH3* + H+ + e → CH4*), which desorbs from the catalyst surface (CH4* → CH4 + *). Consequently, the Ul, catalytic pathway, and the ultimate products during the eCO2RR, are governed by polarization orientation and are adjustable via ferroelectric switching. Moreover, compared to the Ul of Pd@In2Se3 SACs for the eCO2RR (−0.77 V under the P↓ state and −0.87 V under the P↑ state) [56], each corresponding Pd-Nb@In2Se3 DAC for the eCO2RR has a smaller Ul, indicating that the synergistic effect of the bimetallic atomic site could effectively lower the eCO2RR barrier. According to the reduction reaction path of CH2O* → CH2* + OH* → CH3* + OH* → CH3*, we found that carbon-containing intermediates (e.g., CH*, CH2*) tend to undergo hydrogenation reduction at Pd atomic sites, while oxygen-containing intermediates (e.g., OH*) prefer to be hydrogenated and reduced into water at Nb atomic sites (see Figure S4). Therefore, the synergistic effect of Nb-Pd dual-site metal atoms can be described as such: Pd sites act as functional sites for carbon reduction, whereas Nb sites serve as functional sites for oxygen removal. Notably, the Ul of Pd@In2Se3 SACs for the eCO2RR (−0.45 V under the P↓ state and −0.66 V under the P↑ state) are markedly superior to those of conventional eCO2RR catalysts, such as Cu (211) (−0.8 V) [64], Co@Cu (−0.87 V) [65], and WC (0001) (−1.0 V) [66]. Since the equilibrium potential (E°) for the formation of HCOOH and CH4 from CO2, respectively, are −0.25 V vs. RHE and 0.17 V vs. RHE at pH = 0 [67], the corresponding overpotentials are −0.2 V vs. RHE and −0.83 V vs. RHE, which are lower than the ones of many eCO2RR electrocatalysts synthesized experimentally, such as Pd2 (−0.85 V vs. RHE for generating CO) [68], NiCu-NC (−1.07 V vs. RHE for generating CO) [69], FeNi-NSC (−1.0 V vs. RHE for generating CO) [70], Cu/Ni-NC (−0.9 V vs. RHE for generating CO) [71], and Cu-N2 (−0.84 V vs. RHE for generating CH4) [72]. Therefore, the Pd-Nb@In2Se3 DAC is a promising candidate for a tunable and efficient eCO2RR, particularly in achieving high selectivity for final products through controlled polarization modulation.

2.4. Selectivity for HER vs. eCO2RR on Pd-Nb@In2Se3 DACs

The HER is a prevalent parasitic process in electrochemical systems. It poses a significant challenge to the Faradaic efficiency of the eCO2RR by diverting critical proton and electron resources from the electrolyte. This competition arises through the HER’s multi-step mechanisms—Volmer adsorption, Heyrovsky/Tafel recombination, and H2 desorption—which operate synergistically to deplete reactive intermediates (H*) and impose additional energy barriers for CO2 reduction pathways. Particularly in acidic media, the HER’s thermodynamic favorability and low activation overpotential exacerbate its dominance over the eCO2RR, resulting in difficulty in achieving an efficiency of the eCO2RR that meets the needs of industrial production [73,74]. To check if the eCO2RR is more favorable, we calculated the Gibbs free energy changes (ΔG) at the first hydrogenation step of the eCO2RR (* + CO2 → CO2*) and the HER (* + H+ + e → H*). As displayed in Figure 6, on the Pd-Nb@P↓-In2Se3 DAC, the ΔG for the formation of H* is 0.61 eV, while that of CO2* is −0.22 eV. As for the case of the Pd-Nb@P↑-In2Se3 DAC, the ΔG for the formation of H* is −0.11 eV, while that of CO2* is −1.04 eV. According to the Brønsted–Evans–Polanyi relation, reactions with lower ΔG have smaller reaction barriers and consequently are kinetically more favorable. The Gibbs free energy changes indicate that the Pd-Nb@In2Se3 DACs have a higher selectivity toward the CO2RR, instead of the HER. These results demonstrate the feasibility of using Pd-Nb@In2Se3 catalysts as cathodes for the eCO2RR with a high Faradaic efficiency.

3. Methods

All density functional theory (DFT) calculations were performed using the Vienna Ab Initio Simulation Package (VASP) (Hanger team, University of Vienna, version 5.3) [75,76]. The calculations were based on the electron exchange correlation of the PBE generalized function calculation under the generalized gradient approximation (GGA) [77,78,79]. The van der Waals (vdW) interactions were described by applying Grimme’s DFT-D3 method [80], with the colinear spin turned off, the truncation energy set to 500 eV, and the convergence thresholds for force and total energy set to 10−2 eV/Å and 10−5 eV. The catalytic system was constructed by distributing one Pd atom and one Nb atom on adjacent hexagonal hollow sites of an 4 × 4 α-In2Se3 supercell. For the sampling of the dimensional Brillouin zone, 1 × 1 × 1 gamma-centered Monkhorst-pack k-mesh was chosen for performing structural relaxation and frequency calculations. To minimize periodic image interactions, a vacuum layer exceeding 20 Å was introduced along the non-periodic direction. For Gibbs free energy calculations of the eCO2RR, solvent effects were accounted for by using the implicit water model in VASPsol (version 5.3) [55,81]. Charge transfer analysis was performed using the Bader charge partitioning method.
The binding energy for the Nb-Pd dual-atom pair (Eb-(Nb-Pd)) added onto the substrate was calculated using the following equation [63,82]:
E b - ( N b - P d ) = E t o t E s u b E N b E P d
where E s u b and E t o t are the total energies of the substrate without and with the added Nb-Pd dual-atom pair, respectively. E N b and E P d are the total energy of the isolated Nb and Pd atoms.
The charge density difference ( ρ ) for the CO2 molecule adsorbed on Pd-Nb@P↓-In2Se3 and Pd-Nb@P↑-In2Se3 was calculated using the following equation:
ρ = ρ C O 2 - P d - N b - s u b ρ C O 2 ρ P d - N b - s u b
where ρ C O 2 - P d - N b - s u b and ρ P d - N b - s u b are the charge density of Pd-Nb@In2Se3 with and without CO2 molecules, respectively. ρ C O 2 is the charge density of the CO2 molecule.
In order to obtain the individual orbital components, we employed a code named “splitdos.dos” to process the output file of density of states (DOS). The average d band (d band center) shifts were calculated for the surface metal atoms for both the total d partial DOS and the orbital-resolved d partial DOS. The d band center ( ε d ) was calculated as [83]
ε d = n d ( ε ) ε d s n d ( ε ) d ε
where ε is the electronic energy of states, and n d ( ε ) is the electronic density of states.

4. Conclusions

Our study presents a theoretical investigation of a novel Pd-Nb@In2Se3 DAC for the eCO2RR, leveraging DFT calculations to elucidate its structural, electronic, and catalytic properties. The Pd-Nb@In2Se3 system, anchored on a ferroelectric α-In2Se3 monolayer, combines Nb and Pd single atoms to exploit synergistic interactions, while the substrate’s switchable polarization enables dynamic control over catalytic activity and selectivity. Structural analyses confirm the stability of Pd-Nb@In2Se3, with binding energies of −15.69 eV/unit (P↓ polarization) and −12.27 eV/unit (P↑ polarization) and ab initio molecular dynamics simulations demonstrating thermal stability at 1000 K. The CO2 adsorption studies revealed enhanced activation on Pd sites, where the linear CO2 molecule bends significantly, which is attributed to electron donation/backdonation between CO2 π orbitals and the Pd/Nb d orbitals. The d band centers of Pd and Nb atoms, which are closer to the Fermi level in the P↑ configuration, correlate with stronger CO2 activation, emphasizing the role of electronic structure in catalytic performance. Reaction pathway analyses showed polarization-dependent product selectivity. The P↓ polarization favors formic acid production via a Ul of −0.45 V vs. RHE, while the P↑ polarization directs the reaction toward methane with a Ul of −0.66 V vs. RHE. The dual-site mechanism involves Pd atoms driving carbon reduction and Nb atoms facilitating oxygen removal, synergistically lowering energy barriers compared to their SAC counterparts. Additionally, Gibbs free energy calculations indicate superior selectivity for the CO2RR over the HER, with CO2 adsorption energies significantly more favorable than H* formation. The ferroelectric substrate’s polarization switching dynamically modulates metal–support interactions, adjusting d orbital occupancy and intermediate binding strengths, thereby controlling reaction pathways and product outcomes. This work not only advances the understanding of dual-metal catalysis and polarization–catalysis coupling but also provides a strategy to design tunable electrocatalysts for sustainable energy conversion and CO2 utilization.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/inorganics13050146/s1, Table S1: Adsorption energy of CO2 molecules adsorbed at the adsorption sites of Pd-Nb@In2Se3; Table S2: Zero-point energy correction (EZPE) and entropy contribution (TS, T=298.15 K) of molecules; Table S3: Zero-point energy correction (EZPE), entropy contribution (TS, T=298.15 K), total energy (E), and the Gibbs free energy (G) of molecules and adsorbates along the reaction pathway on Pd-Nb@P↓-In2Se3, where * represents the adsorption site; Table S4: Zero-point energy correction (EZPE), entropy contribution (TS, T=298.15 K), total energy (E), and the Gibbs free energy (G) of molecules and adsorbates along the reaction pathway on Pd-Nb@P↑-In2Se3, where * represents the adsorption site; Figure S1: Pd d orbitals and Nb d orbitals from Pd-Nb@In2Se3; Figure S2: The selected adsorption sites in Pd-Nb@In2Se3; Figure S3: The search process for the minimum energy reaction pathways of the CO2 reduction reactions on Pd-Nb@In2Se3; Figure S4: Three configurations on the optimal CO2RR reduction path of Pd-Nb@P↑-In2Se3; The calculation details of the reaction free energy change. References [84,85] are cited in the supplementary materials.

Author Contributions

Conceptualization, X.T.; data curation, Y.Z.; formal analysis, L.J., X.T., M.C., S.L. and C.L.; funding acquisition, L.J.; investigation, L.J. and X.T.; software, Y.Z.; supervision, L.J.; validation, Y.Z. and M.C.; writing—original draft, L.J., X.T., Y.Z. and S.L. All authors have read and agreed to the published version of the manuscript.

Funding

This study was funded by the Young Scientist Project of Henan Province (Grant No. 225200810103), the Program for Science & Technology Innovation Talents in Universities of Henan Province (Grant No. 24HASTIT013), the College Students Innovation Fund of Anyang Normal University (Grant No. 202410479017), and the Natural Science Foundation of Henan Province (Grant No. 232300420128).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

The original contributions presented in this study are included in the article/supplementary material. Further inquiries can be directed to the corresponding author.

Conflicts of Interest

The authors declare no conflict of interest.

References

  1. Obama, B. The irreversible momentum of clean energy. Science 2017, 355, 126–129. [Google Scholar] [CrossRef] [PubMed]
  2. Lüthi, D.; Le Floch, M.; Bereiter, B.; Blunier, T.; Barnola, J.-M.; Siegenthaler, U.; Raynaud, D.; Jouzel, J.; Fischer, H.; Kawamura, K. High-resolution carbon dioxide concentration record 650,000–800,000 years before present. Nature 2008, 453, 379–382. [Google Scholar] [CrossRef] [PubMed]
  3. Metz, B.; Davidson, O.; De Coninck, H.; Loos, M.; Meyer, L. Intergovernmental panel on climate change. In IPCC Special Report on Carbon Dioxide Capture and Storage; Cambridge University Press: New York, NY, USA, 2005; 442p. [Google Scholar]
  4. Chu, S.; Cui, Y.; Liu, N. The path towards sustainable energy. Nat. Mater. 2017, 16, 16–22. [Google Scholar] [CrossRef] [PubMed]
  5. Yuan, Z.; Zhu, X.; Jiang, Z. Recent Advances of Constructing Metal/Semiconductor Catalysts Designing for Photocatalytic CO2 Hydrogenation. Molecules 2023, 28, 5693. [Google Scholar] [CrossRef]
  6. Zhu, X.; Zong, H.; Pérez, C.J.V.; Miao, H.; Sun, W.; Yuan, Z.; Wang, S.; Zeng, G.; Xu, H.; Jiang, Z. Supercharged CO2 photothermal catalytic methanation: High conversion, rate, and selectivity. Angew. Chem. 2023, 135, e202218694. [Google Scholar] [CrossRef]
  7. Zhu, X.; Zhou, E.; Tai, X.; Zong, H.; Yi, J.; Yuan, Z.; Zhao, X.; Huang, P.; Xu, H.; Jiang, Z. g-C3N4 S-Scheme Homojunction through Van der Waals Interface Regulation by Intrinsic Polymerization Tailoring for Enhanced Photocatalytic H2 Evolution and CO2 Reduction. Angew. Chem. Int. Ed. 2025, 64, e202425439. [Google Scholar] [CrossRef]
  8. Yap, D.; Megaritis, A. Applying Forced Induction to Bioethanol HCCI Operation with Residual Gas Trapping. Energy Fuels 2005, 19, 1812–1821. [Google Scholar] [CrossRef]
  9. Kang, Z.; Wu, Z.; Zhang, Z.; Deng, J.; Hu, Z.; Li, L. Study of the Combustion Characteristics of a HCCI Engine Coupled with Oxy-Fuel Combustion Mode. SAE Int. J. Engines 2017, 10, 908–916. [Google Scholar] [CrossRef]
  10. Berstad, D.; Straus, J.; Gundersen, T. CO2 capture and enhanced hydrogen production enabled by low-temperature separation of PSA tail gas: A detailed exergy analysis. Energies 2024, 17, 1072. [Google Scholar] [CrossRef]
  11. Reynolds, S.P.; Ebner, A.D.; Ritter, J.A. Stripping PSA cycles for CO2 recovery from flue gas at high temperature using a hydrotalcite-like adsorbent. Ind. Eng. Chem. Res. 2006, 45, 4278–4294. [Google Scholar] [CrossRef]
  12. Hoffert, M.I.; Caldeira, K.; Benford, G.; Criswell, D.R.; Green, C.; Herzog, H.; Jain, A.K.; Kheshgi, H.S.; Lackner, K.S.; Lewis, J.S. Advanced technology paths to global climate stability: Energy for a greenhouse planet. Science 2002, 298, 981–987. [Google Scholar] [CrossRef] [PubMed]
  13. Duan, X.; Xu, J.; Wei, Z.; Ma, J.; Guo, S.; Wang, S.; Liu, H.; Dou, S. Metal-free carbon materials for CO2 electrochemical reduction. Adv. Mater. 2017, 29, 1701784. [Google Scholar] [CrossRef] [PubMed]
  14. Qiao, J.; Liu, Y.; Hong, F.; Zhang, J. A review of catalysts for the electroreduction of carbon dioxide to produce low-carbon fuels. Chem. Soc. Rev. 2014, 43, 631–675. [Google Scholar] [CrossRef]
  15. Centi, G.; Quadrelli, E.A.; Perathoner, S. Catalysis for CO2 conversion: A key technology for rapid introduction of renewable energy in the value chain of chemical industries. Energy Environ. Sci. 2013, 6, 1711–1731. [Google Scholar] [CrossRef]
  16. Ross, M.B.; De Luna, P.; Li, Y.; Dinh, C.-T.; Kim, D.; Yang, P.; Sargent, E.H. Designing materials for electrochemical carbon dioxide recycling. Nat. Catal. 2019, 2, 648–658. [Google Scholar] [CrossRef]
  17. Wang, G.; Zhong, S.; Xiong, X.; Li, J.; Wang, F.; Huo, L.; Wu, D.; Han, X.; Wang, Z.; Chen, Q. Plasma induced grain boundaries to boost electrochemical reduction of CO2 to formate. J. Energy Chem. 2024, 95, 636–643. [Google Scholar] [CrossRef]
  18. Wei, Z.; Wang, W.; Shao, T.; Yang, S.; Liu, C.; Si, D.; Cao, R.; Cao, M. Constructing Ag/Cu2O interface for efficient neutral CO2 electroreduction to C2H4. Angew. Chem. Int. Ed. 2025, 64, e202417066. [Google Scholar] [CrossRef]
  19. Yue, M.; Xie, W.; Zhong, Z.; Li, M.; Zhang, T.; Shao, M.; Li, H.; Wang, Q. Isolating and stabilizing active copper species in layered double hydroxide to enhance electrocatalytic CO2 reduction to CH4. J. Catal. 2025, 443, 115959. [Google Scholar] [CrossRef]
  20. Dong, J.; Liu, Y.; Pei, J.; Li, H.; Ji, S.; Shi, L.; Zhang, Y.; Li, C.; Tang, C.; Liao, J. Continuous electroproduction of formate via CO2 reduction on local symmetry-broken single-atom catalysts. Nat. Commun. 2023, 14, 6849. [Google Scholar] [CrossRef]
  21. Chatterjee, S.; Dutta, I.; Lum, Y.; Lai, Z.; Huang, K.-W. Enabling storage and utilization of low-carbon electricity: Power to formic acid. Energy Environ. Sci. 2021, 14, 1194–1246. [Google Scholar] [CrossRef]
  22. Li, M.; Idros, M.N.; Wu, Y.; Burdyny, T.; Garg, S.; Zhao, X.S.; Wang, G.; Rufford, T.E. The role of electrode wettability in electrochemical reduction of carbon dioxide. J. Mater. Chem. A 2021, 9, 19369–19409. [Google Scholar] [CrossRef]
  23. Möller, T.; Ju, W.; Bagger, A.; Wang, X.; Luo, F.; Ngo Thanh, T.; Varela, A.S.; Rossmeisl, J.; Strasser, P. Efficient CO2 to CO electrolysis on solid Ni–N–C catalysts at industrial current densities. Energy Environ. Sci. 2019, 12, 640–647. [Google Scholar] [CrossRef]
  24. Guo, S.; Liu, Y.; Murphy, E.; Ly, A.; Xu, M.; Matanovic, I.; Pan, X.; Atanassov, P. Robust palladium hydride catalyst for electrocatalytic formate formation with high CO tolerance. Appl. Catal. B Environ. 2022, 316, 121659. [Google Scholar] [CrossRef]
  25. Detweiler, Z.M.; White, J.L.; Bernasek, S.L.; Bocarsly, A.B. Anodized indium metal electrodes for enhanced carbon dioxide reduction in aqueous electrolyte. Langmuir 2014, 30, 7593–7600. [Google Scholar] [CrossRef]
  26. Ikeda, S.; Takagi, T.; Ito, K. Selective formation of formic acid, oxalic acid, and carbon monoxide by electrochemical reduction of carbon dioxide. Bull. Chem. Soc. Jpn. 1987, 60, 2517–2522. [Google Scholar] [CrossRef]
  27. Hori, Y.; Wakebe, H.; Tsukamoto, T.; Koga, O. Electrocatalytic process of CO selectivity in electrochemical reduction of CO2 at metal electrodes in aqueous media. Electrochim. Acta 1994, 39, 1833–1839. [Google Scholar] [CrossRef]
  28. Ding, S.; Hülsey, M.J.; Pérez-Ramírez, J.; Yan, N. Transforming energy with single-atom catalysts. Joule 2019, 3, 2897–2929. [Google Scholar] [CrossRef]
  29. Su, X.; Yang, X.-F.; Huang, Y.; Liu, B.; Zhang, T. Single-atom catalysis toward efficient CO2 conversion to CO and formate products. Acc. Chem. Res. 2018, 52, 656–664. [Google Scholar] [CrossRef]
  30. Liu, K.; Fu, J.; Zhu, L.; Zhang, X.; Li, H.; Liu, H.; Hu, J.; Liu, M. Single-atom transition metals supported on black phosphorene for electrochemical nitrogen reduction. Nanoscale 2020, 12, 4903–4908. [Google Scholar] [CrossRef]
  31. Lu, F.; Yi, D.; Liu, S.; Zhan, F.; Zhou, B.; Gu, L.; Golberg, D.; Wang, X.; Yao, J. Engineering platinum–oxygen dual catalytic sites via charge transfer towards highly efficient hydrogen evolution. Angew. Chem. 2020, 132, 17865–17871. [Google Scholar] [CrossRef]
  32. Nguyen, T.; Salehi, M.; Le, Q.; Seifitokaldani, A.; Dinh, C. Fundamentals of electrochemical CO2 reduction on single-metal-atom catalysts. ACS Catal. 2020, 10, 10068–10095. [Google Scholar] [CrossRef]
  33. Sun, J.-F.; Wu, J.-T.; Xu, Q.-Q.; Zhou, D.; Yin, J.-Z. CO2 electrochemical reduction using single-atom catalysts. Preparation, characterization and anchoring strategies: A review. Environ. Chem. Lett. 2020, 18, 1593–1623. [Google Scholar] [CrossRef]
  34. Asset, T.; Garcia, S.T.; Herrera, S.; Andersen, N.; Chen, Y.; Peterson, E.J.; Matanovic, I.; Artyushkova, K.; Lee, J.; Minteer, S.D. Investigating the nature of the active sites for the CO2 reduction reaction on carbon-based electrocatalysts. ACS Catal. 2019, 9, 7668–7678. [Google Scholar] [CrossRef]
  35. Zhang, Q.; Guan, J. Single-atom catalysts for electrocatalytic applications. Adv. Funct. Mater. 2020, 30, 2000768. [Google Scholar] [CrossRef]
  36. Zhang, Y.; Xia, B.; Ran, J.; Davey, K.; Qiao, S.Z. Atomic-level reactive sites for semiconductor-based photocatalytic CO2 reduction. Adv. Energy Mater. 2020, 10, 1903879. [Google Scholar] [CrossRef]
  37. Luo, G.; Jing, Y.; Li, Y. Rational design of dual-metal-site catalysts for electroreduction of carbon dioxide. J. Mater. Chem. A 2020, 8, 15809–15815. [Google Scholar] [CrossRef]
  38. Ouyang, Y.; Shi, L.; Bai, X.; Li, Q.; Wang, J. Breaking scaling relations for efficient CO2 electrochemical reduction through dual-atom catalysts. Chem. Sci. 2020, 11, 1807–1813. [Google Scholar] [CrossRef]
  39. Li, Y.; Chen, C.; Cao, R.; Pan, Z.; He, H.; Zhou, K. Dual-atom Ag2/graphene catalyst for efficient electroreduction of CO2 to CO. Appl. Catal. B Environ. 2020, 268, 118747. [Google Scholar] [CrossRef]
  40. Li, Y.; Su, H.; Chan, S.H.; Sun, Q. CO2 Electroreduction Performance of Transition Metal Dimers Supported on Graphene: A Theoretical Study. ACS Catal. 2015, 5, 6658–6664. [Google Scholar] [CrossRef]
  41. Li, X.; Sun, Y.; Xu, J.; Shao, Y.; Wu, J.; Xu, X.; Pan, Y.; Ju, H.; Zhu, J.; Xie, Y. Selective visible-light-driven photocatalytic CO2 reduction to CH4 mediated by atomically thin CuIn5S8 layers. Nat. Energy 2019, 4, 690–699. [Google Scholar] [CrossRef]
  42. Li, S.; Nagarajan, A.V.; Alfonso, D.R.; Sun, M.; Kauffman, D.R.; Mpourmpakis, G.; Jin, R. Boosting CO2 electrochemical reduction with atomically precise surface modification on gold nanoclusters. Angew. Chem. Int. Ed. 2021, 60, 6351–6356. [Google Scholar] [CrossRef] [PubMed]
  43. Wen, G.; Lee, D.U.; Ren, B.; Hassan, F.M.; Jiang, G.; Cano, Z.P.; Gostick, J.; Croiset, E.; Bai, Z.; Yang, L. Orbital interactions in Bi-Sn bimetallic electrocatalysts for highly selective electrochemical CO2 reduction toward formate production. Adv. Energy Mater. 2018, 8, 1802427. [Google Scholar] [CrossRef]
  44. Lin, L.; Li, H.; Yan, C.; Li, H.; Si, R.; Li, M.; Xiao, J.; Wang, G.; Bao, X. Synergistic catalysis over iron-nitrogen sites anchored with cobalt phthalocyanine for efficient CO2 electroreduction. Adv. Mater. 2019, 31, 1903470. [Google Scholar] [CrossRef]
  45. Hu, K.; Li, Z.; Bai, L.; Yang, F.; Chu, X.; Bian, J.; Zhang, Z.; Xu, H.; Jing, L. Synergetic subnano Ni-and Mn-Oxo clusters anchored by chitosan oligomers on 2D g-C3N4 boost photocatalytic CO2 reduction. Sol. Rrl 2021, 5, 2000472. [Google Scholar] [CrossRef]
  46. Bok, J.; Lee, S.Y.; Lee, B.-H.; Kim, C.; Nguyen, D.L.T.; Kim, J.W.; Jung, E.; Lee, C.W.; Jung, Y.; Lee, H.S. Designing atomically dispersed Au on tensile-strained Pd for efficient CO2 electroreduction to formate. J. Am. Chem. Soc. 2021, 143, 5386–5395. [Google Scholar] [CrossRef]
  47. Ding, T.; Liu, X.; Tao, Z.; Liu, T.; Chen, T.; Zhang, W.; Shen, X.; Liu, D.; Wang, S.; Pang, B. Atomically precise dinuclear site active toward electrocatalytic CO2 reduction. J. Am. Chem. Soc. 2021, 143, 11317–11324. [Google Scholar] [CrossRef]
  48. He, Q.; Liu, D.; Lee, J.H.; Liu, Y.; Xie, Z.; Hwang, S.; Kattel, S.; Song, L.; Chen, J.G. Electrochemical conversion of CO2 to syngas with controllable CO/H2 ratios over Co and Ni single-atom catalysts. Angew. Chem. Int. Ed. 2020, 59, 3033–3037. [Google Scholar] [CrossRef]
  49. Darif, B.; Ojala, S.; Pirault-Roy, L.; Bensitel, M.; Brahmi, R.; Keiski, R.L. Study on the catalytic oxidation of DMDS over Pt-Cu catalysts supported on Al2O3, AlSi20 and SiO2. Appl. Catal. B Environ. 2016, 181, 24–33. [Google Scholar] [CrossRef]
  50. Winter, L.R.; Gomez, E.; Yan, B.; Yao, S.; Chen, J.G. Tuning Ni-catalyzed CO2 hydrogenation selectivity via Ni-ceria support interactions and Ni-Fe bimetallic formation. Appl. Catal. B Environ. 2018, 224, 442–450. [Google Scholar] [CrossRef]
  51. Wang, J.; Huang, Z.; Liu, W.; Chang, C.; Tang, H.; Li, Z.; Chen, W.; Jia, C.; Yao, T.; Wei, S.; et al. Design of N-Coordinated Dual-Metal Sites: A Stable and Active Pt-Free Catalyst for Acidic Oxygen Reduction Reaction. J. Am. Chem. Soc. 2017, 139, 17281–17284. [Google Scholar] [CrossRef]
  52. Ren, W.; Tan, X.; Yang, W.; Jia, C.; Xu, S.; Wang, K.; Smith, S.C.; Zhao, C. Isolated Diatomic Ni-Fe Metal–Nitrogen Sites for Synergistic Electroreduction of CO2. Angew. Chem. Int. Ed. 2019, 58, 6972–6976. [Google Scholar] [CrossRef] [PubMed]
  53. Zhang, L.; Fischer, J.M.T.A.; Jia, Y.; Yan, X.; Xu, W.; Wang, X.; Chen, J.; Yang, D.; Liu, H.; Zhuang, L.; et al. Coordination of Atomic Co–Pt Coupling Species at Carbon Defects as Active Sites for Oxygen Reduction Reaction. J. Am. Chem. Soc. 2018, 140, 10757–10763. [Google Scholar] [CrossRef] [PubMed]
  54. Li, D.; Zhao, M.H.; Garra, J.; Kolpak, A.M.; Rappe, A.M.; Bonnell, D.A.; Vohs, J.M. Direct in situ determination of the polarization dependence of physisorption on ferroelectric surfaces. Nat. Mater. 2008, 7, 473–477. [Google Scholar] [CrossRef]
  55. Ju, L.; Shang, J.; Tang, X.; Kou, L. Tunable Photocatalytic Water Splitting by the Ferroelectric Switch in a 2D AgBiP2Se6 Monolayer. J. Am. Chem. Soc. 2019, 142, 1492–1500. [Google Scholar] [CrossRef]
  56. Ju, L.; Tan, X.; Mao, X.; Gu, Y.; Smith, S.; Du, A.; Chen, Z.; Chen, C.; Kou, L. Controllable CO2 electrocatalytic reduction via ferroelectric switching on single atom anchored In2Se3 monolayer. Nat. Commun. 2021, 12, 5128. [Google Scholar] [CrossRef]
  57. Ding, W.; Zhu, J.; Wang, Z.; Gao, Y.; Xiao, D.; Gu, Y.; Zhang, Z.; Zhu, W. Prediction of intrinsic two-dimensional ferroelectrics in In2Se3 and other III2-VI3 van der Waals materials. Nat. Commun. 2017, 8, 14956. [Google Scholar] [CrossRef]
  58. Cui, C.; Hu, W.J.; Yan, X.; Addiego, C.; Gao, W.; Wang, Y.; Wang, Z.; Li, L.; Cheng, Y.; Li, P.; et al. Intercorrelated In-Plane and Out-of-Plane Ferroelectricity in Ultrathin Two-Dimensional Layered Semiconductor In2Se3. Nano Lett. 2018, 18, 1253–1258. [Google Scholar] [CrossRef]
  59. Gao, G.; Jiao, Y.; Waclawik, E.R.; Du, A. Single Atom (Pd/Pt) Supported on Graphitic Carbon Nitride as an Efficient Photocatalyst for Visible-Light Reduction of Carbon Dioxide. J. Am. Chem. Soc. 2016, 138, 6292–6297. [Google Scholar] [CrossRef]
  60. Li, X.; Bi, W.; Zhang, L.; Tao, S.; Chu, W.; Zhang, Q.; Luo, Y.; Wu, C.; Xie, Y. Single-Atom Pt as Co-Catalyst for Enhanced Photocatalytic H2 Evolution. Adv. Mater. 2016, 28, 2427–2431. [Google Scholar] [CrossRef]
  61. Vile, G.; Albani, D.; Nachtegaal, M.; Chen, Z.; Dontsova, D.; Antonietti, M.; Lopez, N.; Perez-Ramirez, J. A stable single-site palladium catalyst for hydrogenations. Angew. Chem. Int. Ed. 2015, 54, 11265–11269. [Google Scholar] [CrossRef]
  62. Khatun, R.; Siddiqui, N.; Pal, R.S.; Bhandari, S.; Khan, T.S.; Singh, S.; Poddar, M.K.; Samanta, C.; Bal, R. Low temperature reforming of methane with CO2 over Pt/CeO2, Ni/CeO2 and Pt–Ni/CeO2 catalysts prepared by a solution-combustion method. Catal. Sci. Technol. 2023, 13, 6431–6445. [Google Scholar] [CrossRef]
  63. Ma, D.; Zeng, Z.; Liu, L.; Huang, X.; Jia, Y. Computational Evaluation of Electrocatalytic Nitrogen Reduction on TM Single-, Double-, and Triple-Atom Catalysts (TM = Mn, Fe, Co, Ni) Based on Graphdiyne Monolayers. J. Phys. Chem. C 2019, 123, 19066–19076. [Google Scholar] [CrossRef]
  64. Back, S.; Kim, H.; Jung, Y. Selective Heterogeneous CO2 Electroreduction to Methanol. ACS Catal. 2015, 5, 965–971. [Google Scholar] [CrossRef]
  65. Zhao, Z.; Lu, G. Cu-Based Single-Atom Catalysts Boost Electroreduction of CO2 to CH3OH: First-Principles Predictions. J. Phys. Chem. C 2019, 123, 4380–4387. [Google Scholar] [CrossRef]
  66. Wannakao, S.; Artrith, N.; Limtrakul, J.; Kolpak, A.M. Engineering Transition-Metal-Coated Tungsten Carbides for Efficient and Selective Electrochemical Reduction of CO2 to Methane. ChemSusChem 2015, 8, 2745–2751. [Google Scholar] [CrossRef]
  67. Elhenawy, S.E.M.; Khraisheh, M.; AlMomani, F.; Walker, G. Metal-organic frameworks as a platform for CO2 capture and chemical processes: Adsorption, membrane separation, catalytic-conversion, and electrochemical reduction of CO2. Catalysts 2020, 10, 1293. [Google Scholar] [CrossRef]
  68. Zhang, N.; Zhang, X.; Kang, Y.; Ye, C.; Jin, R.; Yan, H.; Lin, R.; Yang, J.; Xu, Q.; Wang, Y.; et al. A Supported Pd2 Dual-Atom Site Catalyst for Efficient Electrochemical CO2 Reduction. Angew. Chem. Int. Ed. 2021, 60, 13388–13393. [Google Scholar] [CrossRef]
  69. Hao, Q.; Zhong, H.; Wang, J.; Liu, K.; Yan, J.; Ren, Z.; Zhou, N.; Zhao, X.; Zhang, H.; Liu, D.; et al. Nickel dual-atom sites for electrochemical carbon dioxide reduction. Nat. Synth. 2022, 1, 719–728. [Google Scholar] [CrossRef]
  70. Huang, K.; Li, R.; Qi, H.; Yang, S.; An, S.; Lian, C.; Xu, Q.; Liu, H.; Hu, J. Regulating Adsorption of Intermediates via the Sulfur Modulating Dual-Atomic Sites for Boosting CO2RR. ACS Catal. 2024, 14, 8889–8898. [Google Scholar] [CrossRef]
  71. Chen, B.; Shi, D.; Deng, R.; Xu, X.; Liu, W.; Wei, Y.; Liu, Z.; Zhong, S.; Huang, J.; Yu, Y. Leveraging Atomic-Scale Synergy for Selective CO2 Electrocatalysis to CO over CuNi Dual-Atom Catalysts. ACS Catal. 2024, 14, 16224–16233. [Google Scholar] [CrossRef]
  72. Roy, S.; Li, Z.; Chen, Z.; Mata, A.C.; Kumar, P.; Sarma, S.C.; Teixeira, I.F.; Silva, I.F.; Gao, G.; Tarakina, N.V.; et al. Cooperative Copper Single-Atom Catalyst in 2D Carbon Nitride for Enhanced CO2 Electrolysis to Methane. Adv. Mater. 2024, 36, e2300713. [Google Scholar] [CrossRef]
  73. Xue, H.; Zhao, Z.H.; Liao, P.Q.; Chen, X.M. “Ship-in-a-Bottle” Integration of Ditin(IV) Sites into a Metal-Organic Framework for Boosting Electroreduction of CO2 in Acidic Electrolyte. J. Am. Chem. Soc. 2023, 145, 16978–16982. [Google Scholar] [CrossRef]
  74. Monteiro, M.C.O.; Dattila, F.; Hagedoorn, B.; García-Muelas, R.; López, N.; Koper, M.T.M. Absence of CO2 electroreduction on copper, gold and silver electrodes without metal cations in solution. Nat. Catal. 2021, 4, 654–662. [Google Scholar] [CrossRef]
  75. Kresse, G.; Furthmüller, J. Efficient iterative schemes for ab initio total-energy calculations using a plane-wave basis set. Phys. Rev. B 1996, 54, 11169. [Google Scholar] [CrossRef]
  76. Kresse, G.; Furthmüller, J. Efficiency of ab-initio total energy calculations for metals and semiconductors using a plane-wave basis set. Comput. Mater. Sci. 1996, 6, 15–50. [Google Scholar] [CrossRef]
  77. Blöchl, P.E. Projector augmented-wave method. Phys. Rev. B 1994, 50, 17953–17979. [Google Scholar] [CrossRef]
  78. Kresse, G.; Joubert, D. From ultrasoft pseudopotentials to the projector augmented-wave method. Phys. Rev. B 1999, 59, 1758. [Google Scholar] [CrossRef]
  79. Perdew, J.P.; Burke, K.; Ernzerhof, M. Generalized gradient approximation made simple. Phys. Rev. Lett. 1996, 77, 3865. [Google Scholar] [CrossRef]
  80. Grimme, S. Semiempirical GGA-type density functional constructed with a long-range dispersion correction. J. Comput. Chem. 2006, 27, 1787–1799. [Google Scholar] [CrossRef]
  81. Mao, X.; Kour, G.; Zhang, L.; He, T.; Wang, S.; Yan, C.; Zhu, Z.; Du, A. Silicon-doped graphene edges: An efficient metal-free catalyst for the reduction of CO2 into methanol and ethanol. Catal. Sci. Technol. 2019, 9, 6800–6807. [Google Scholar] [CrossRef]
  82. Cui, X.; An, W.; Liu, X.; Wang, H.; Men, Y.; Wang, J. C2N-graphene supported single-atom catalysts for CO2 electrochemical reduction reaction: Mechanistic insight and catalyst screening. Nanoscale 2018, 10, 15262–15272. [Google Scholar] [CrossRef] [PubMed]
  83. Nørskov, J.K.; Studt, F.; Abild-Pedersen, F.; Bligaard, T. Fundamental Concepts in Heterogeneous Catalysis. Focus Catal. 2015, 2015, 7. [Google Scholar] [CrossRef]
  84. Nørskov, J.K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J.R.; Bligaard, T.; Jonsson, H. Origin of the overpotential for oxygen reduction at a fuel-cell cathode. J. Phys. Chem. B 2004, 108, 17886–17892. [Google Scholar] [CrossRef]
  85. Peterson, A.A.; Abild-Pedersen, F.; Studt, F.; Rossmeisl, J.; Nørskov, J.K. How copper catalyzes the electroreduction of carbon dioxide into hydrocarbon fuels. Energ. Environ. Sci. 2010, 3, 1311–1315. [Google Scholar] [CrossRef]
Figure 1. Side views of the optimized configuration of (a) Pd-Nb@P↓-In2Se3 and (b) Pd-Nb@P↑-In2Se3 DACs.
Figure 1. Side views of the optimized configuration of (a) Pd-Nb@P↓-In2Se3 and (b) Pd-Nb@P↑-In2Se3 DACs.
Inorganics 13 00146 g001
Figure 2. The ab initio molecular dynamics simulations of (a) Pd-Nb@P↓-In2Se3 and (b) Pd-Nb@P↑-In2Se3 configurations at 1000 K, showing the fluctuations in the total energy (upper) and temperature (lower) over a time step of 1 fs and a duration of 10 ps.
Figure 2. The ab initio molecular dynamics simulations of (a) Pd-Nb@P↓-In2Se3 and (b) Pd-Nb@P↑-In2Se3 configurations at 1000 K, showing the fluctuations in the total energy (upper) and temperature (lower) over a time step of 1 fs and a duration of 10 ps.
Inorganics 13 00146 g002
Figure 3. The charge differential density plots of CO2 adsorbed on (a) Pd-Nb@P↓-In2Se3 and (b) Pd-Nb@P↑-In2Se. The charge accumulation is indicated by areas that are yellow. The charge depletion is indicated by areas that are cyan. The value of the isosurface is 0.01 e Å−3. (c) Simplified schematic diagrams of CO2 activation on transition metals.
Figure 3. The charge differential density plots of CO2 adsorbed on (a) Pd-Nb@P↓-In2Se3 and (b) Pd-Nb@P↑-In2Se. The charge accumulation is indicated by areas that are yellow. The charge depletion is indicated by areas that are cyan. The value of the isosurface is 0.01 e Å−3. (c) Simplified schematic diagrams of CO2 activation on transition metals.
Inorganics 13 00146 g003
Figure 4. (a) Pd d orbitals (in light red) from Pd-Nb@P↓-In2Se3 and Pd-Nb@P↑-In2Se3 DACs. Nb d orbitals (in light blue) from Pd-Nb@P↓-In2Se3 and Pd-Nb@P↑-In2Se3 DACs. (b) The d band center of Pd-Nb@In2Se3, with orange columns for Pd atoms and blue columns for Nb atoms.
Figure 4. (a) Pd d orbitals (in light red) from Pd-Nb@P↓-In2Se3 and Pd-Nb@P↑-In2Se3 DACs. Nb d orbitals (in light blue) from Pd-Nb@P↓-In2Se3 and Pd-Nb@P↑-In2Se3 DACs. (b) The d band center of Pd-Nb@In2Se3, with orange columns for Pd atoms and blue columns for Nb atoms.
Inorganics 13 00146 g004
Figure 5. The free-energy profile for the eCO2RR along the minimum energy path at 0 V (vs. RHE). Energy-minimized structural arrangements of transient species are displayed in the supplementary panels. Gold-shaded region delineates the catalytic process occurring on the Pd-Nb@P↓-In2Se3 DAC, while the cyan-tinted zone corresponds to the electrochemical transformation happening on the Pd-Nb@P↑-In2Se3 DAC. * stands for the adsorption site at the surface of catalyst.
Figure 5. The free-energy profile for the eCO2RR along the minimum energy path at 0 V (vs. RHE). Energy-minimized structural arrangements of transient species are displayed in the supplementary panels. Gold-shaded region delineates the catalytic process occurring on the Pd-Nb@P↓-In2Se3 DAC, while the cyan-tinted zone corresponds to the electrochemical transformation happening on the Pd-Nb@P↑-In2Se3 DAC. * stands for the adsorption site at the surface of catalyst.
Inorganics 13 00146 g005
Figure 6. The blue bar indicates the H* formation reaction energy barrier of the optimal active site in the catalyst, and the purple bar indicates the CO2 adsorption reaction energy barrier. A negative reaction barrier indicates a spontaneous exothermic reaction. * stands for the adsorption site at the surface of catalyst.
Figure 6. The blue bar indicates the H* formation reaction energy barrier of the optimal active site in the catalyst, and the purple bar indicates the CO2 adsorption reaction energy barrier. A negative reaction barrier indicates a spontaneous exothermic reaction. * stands for the adsorption site at the surface of catalyst.
Inorganics 13 00146 g006
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Ju, L.; Tang, X.; Zhang, Y.; Chen, M.; Liu, S.; Long, C. Insights into Pd-Nb@In2Se3 Electrocatalyst for High-Performance and Selective CO2 Reduction Reaction from DFT. Inorganics 2025, 13, 146. https://doi.org/10.3390/inorganics13050146

AMA Style

Ju L, Tang X, Zhang Y, Chen M, Liu S, Long C. Insights into Pd-Nb@In2Se3 Electrocatalyst for High-Performance and Selective CO2 Reduction Reaction from DFT. Inorganics. 2025; 13(5):146. https://doi.org/10.3390/inorganics13050146

Chicago/Turabian Style

Ju, Lin, Xiao Tang, Yixin Zhang, Mengya Chen, Shuli Liu, and Chen Long. 2025. "Insights into Pd-Nb@In2Se3 Electrocatalyst for High-Performance and Selective CO2 Reduction Reaction from DFT" Inorganics 13, no. 5: 146. https://doi.org/10.3390/inorganics13050146

APA Style

Ju, L., Tang, X., Zhang, Y., Chen, M., Liu, S., & Long, C. (2025). Insights into Pd-Nb@In2Se3 Electrocatalyst for High-Performance and Selective CO2 Reduction Reaction from DFT. Inorganics, 13(5), 146. https://doi.org/10.3390/inorganics13050146

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop