Open Access Article
Shi
Li
a,
Tianyue
Gao
b,
Yupo
Lin
b,
Christopher G.
Arges
*b and
Rajeev Surendran
Assary
*a
aMaterials Science Division, Argonne National Laboratory, Lemont, IL, USA. E-mail: assary@anl.gov
bApplied Materials Division, Argonne National Laboratory, Lemont, IL, USA. E-mail: carges@anl.gov
First published on 2nd October 2025
Bipolar membrane (BPM) electrochemical processes are a promising platform for carbon dioxide (CO2) separations, but the molecular level thermodynamic and kinetic understanding of CO2-to-bicarbonate (HCO3−) transformation remain poorly understood. This study employs a multiscale computational approach to systematically explore the adsorption and reactive transformation of CO2 in five anion exchange ionomer systems. Classical molecular dynamics (MD) simulation results demonstrate that polymers with imidazolium groups significantly reduce CO2 diffusion and enhance (OH−)–CO2 interactions due to stronger electrostatic and π-interactions. Compared to the commonly used quaternary ammonium ionomers, imidazolium-functionalized ionomers show improved CO2 proximity and interaction strength. Ab initio MD and density functional theory (DFT) calculations reveal that the benzyl-substituted imidazolium (IM-Ben) substantially reduces the energy barrier for HCO3− formation (∼72 meV lower) compared to the alkyl-substituted IM-nBu, while also mitigating imidazolium deprotonation under moderate hydration conditions. Transition state analysis shows IM-Ben forms more extensive hydrogen-bonding networks, which stabilize the transition state structure and contribute to a lower energy barrier for bicarbonate formation. These findings highlight the advantage of the adjacent benzyl moiety in enabling efficient CO2-to-bicarbonate transformation via hydrated hydroxide ion counterions, offering mechanistic insights and clear molecular design principles for optimizing anion exchange ionomers at bipolar membrane interfaces for electrochemical CO2 separation applications.
New conceptsThis work presents a multiscale computational approach to uncover how polymer chemistry and hydration structure govern the molecular transformation of CO2 to bicarbonate (HCO3−) in anion exchange ionomers - key materials in electrochemical CO2 separation technologies. Using classical MD, ab initio MD, and DFT-based transition state calculations, we reveal how imidazolium-based functional groups promote closer CO2 association and stabilize reaction intermediates via directional hydrogen bonding leading to lower energy barriers in converting CO2 and hydroxide to HCO3−. A key conceptual advance is the identification of competing reaction pathways - bicarbonate formation versus imidazolium deprotonation and how solvation and ionomer architecture influence the preferred route. These insights clarify structure–function relationships critical for designing membranes or porous conductors that not only transport but also activate CO2. This work moves beyond static models of CO2 binding by elucidating reactive pathways at the molecular level, offering design principles for next-generation materials that integrate ion conduction with selective CO2 transformation. |
The in situ pH adjustment efficacy in BPM electrochemical processes largely depends on the BPM design, such as the anion and cation exchange ionomer chemistry, layer thickness, water dissociation and proton–hydroxide ion recombination catalyst, and the quality of the bipolar junction interface. Recent studies have shown that functional groups within ionomers influence membrane hydration, ion selectivity and ionic conductivity, all of which affect BPM polarization as well as membrane and electrode performance (when used as a binder) in electrochemical processes.10–13
The most conventional method for electrochemical CO2 separation and concentration is BPM electrodialysis. However, membrane materials for BPM electrodialysis may not significantly impact cell polarization and energy efficiency when one of the process streams (i.e., the diluate process stream) has a low concentration of ionic species (e.g., carbonate species from the interaction of CO2 and hydroxide ions). The diluate chamber ohmic resistances can be ameliorated by utilizing a closely related process called electrodeionization. This process features a porous ionic conductor in the diluate chamber to augment the ionic conductivity and to curtail ohmic resistances. Lin and co-workers have used resin wafer BPM electrodeionization for electrochemical CO2 separation and concentration.14 Recently, Arges and co-workers have also developed BPM capacitive deionization for pH-assisted selective ion separations.15–17 Like electrodialysis, the ohmic resistances in the process fluid channel (i.e., spacer channel) of membrane capacitive deionization can also be quite large at low dissolved salt concentrations.18 Arges and co-workers reduced the ohmic resistances in membrane capacitive deionization systems by using ionomer coated nylon meshes (i.e., a type of porous ionic conductor).19–21
Most electrochemical CO2 separation process employ energy intensive faradaic reactions that often use platinum group metals in the electrode layers. Additionally, many of these systems operate with liquid process streams, which can limit CO2 separation efficiency due to the inherently low solubility of CO2 in aqueous solutions. The BPM capacitive deionization process proposed in Fig. 1 is being developed for electrochemical CO2 separations. This process uses a humidified gas stream containing CO2. This process utilizes a humidified gas stream containing CO2 and avoids both faradaic reactions and the use of platinum group metals, enabling a more energy-efficient and material-sustainable pathway for CO2 handling and potential downstream utilization. To realize this process, it is important to design appropriate ionomers in the porous ionic conductor in the spacer channel and for the anion exchange layer in the BPM. Arges and Lin and co-workers have shown that imidazolium ionomers, processed into anion exchange membranes and resin wafers, are effective for selective removal of lactate and p-coumarate (organic acid anions with carbonyl groups) from process streams.22,23 Their work, and others,24,25 showed that imidazolium-based functional groups are effective for electrochemical CO2 separations due to their intrinsic affinity toward CO2 molecules – which is attributed largely to the planar aromatic structure that facilitates favorable electrostatic and π–π interactions with CO2. Despite these promising characteristics, detailed mechanistic insights into HCO3− formation from CO2–hydroxide counterions in close vicinity to a tethered cation, as well as competing chemical pathways, remain insufficiently understood at the molecular level.
To address the said challenges in streamlining materials development for electrochemical CO2 separations, advanced computational methods, including molecular dynamics (MD) simulations and density functional theory (DFT) calculations, are considered because they offer valuable molecular scale insights that can a priori elucidate the interactions and reaction energetics underlying the chemical transformation of CO2 in ionomers. Previous computational studies have explored polymer–CO2 interactions, identifying factors such as backbone chemistry and ionomer functionalization as critical parameters influencing CO2 diffusivity and interaction strength.12,26,27 For instance, quaternary benzyl ammonium polysulfone (QAPSf) and quaternary benzyl imidazolium polysulfone (QIPSf) have demonstrated distinct differences in ion selectivity and CO2 affinity, underscoring the importance of structural features such as backbone rigidity and functional-group positioning (vide infra, structures of the ionomers are schematically shown in Fig. 2).22,24 Motivated by these findings, we investigate five distinct anion exchange ionomer systems to systematically evaluate the impact of backbone architecture (polysulfone) with cations at the benzyl position vs. poly(m-terphenyl phenylene) with cations at the end of an n-hexyl side chain and functional group chemistry (quaternary ammonium (trimethyl ammonium and n-methyl pyrrolidinium) vs. imidazolium) on CO2 separation performance and HCO3− formation efficiency. By integrating classic MD simulations and accurate DFT and AIMD calculations, this work reveals critical molecular-scale phenomena—including CO2 diffusion, spatial organization, hydration effects, and competing reaction pathways—thus providing a detailed molecular understanding of CO2 separation and HCO3− formation in anion exchange ionomers – which can comprise the anion exchange layer of the BPM as well as the porous ionic conductor in Fig. 1. The outcomes from this computational investigation not only clarify fundamental reaction mechanisms but also guides the rational design for next-generation ionomers in electrochemical CO2 separation processes.
To elucidate deeper structural interactions and energetic characteristics, AIMD simulations and static DFT calculations were further performed on representative subsystems extracted from the polymer systems. These subsystems were specifically chosen to encompass critical interactions between the functional groups and a representative segment of the polymer backbone. This multi-scale modeling approach enabled accurate characterization of key chemical and physical interactions within the polymeric environment. Detailed chemical structures of the complete polymer systems used in the MD simulations and their corresponding subsystem models are presented in Fig. 2(a). The geometries of all models are presented in Fig. S1 and S2.
Five distinct monomer polymer repeat units were initially optimized at the B3LYP/6-31g(d) level of theory before being assembled into polymer chains for subsequent simulations, each single chain consists of 20 repeat units, using the mbuild software package from the Molecular Simulation Design Framework (MoSDef).53 The total number of atoms per polymer chain was maintained between approximately 1200 and 1500, to maintain a balance between chemical fidelity and computational efficiency. A representative initial structure of a single polymer chain is shown in Fig. 2(b), while the complete set of initial polymer structures is provided in Fig. S1.
To simulate conditions representative of anion exchange ionomer environments and ensure statistically meaningful sampling, fifty of these single polymer chains were randomly packed into a 50 nm × 50 nm × 20 nm simulation box. This initial box size was selected to accommodate all chains with sufficient spacing to prevent atomic overlap during packing. It serves as a starting configuration that allows for effective rearrangement and densification during the subsequent compression and relaxation process. Following equilibration, the system compresses into a thinner and denser configuration, maintaining an appropriate environment for polymer–solvent–ion interactions, while providing enough area and volume to enable multiple CO2 and OH− encounters.
To maintain charge neutrality and support realistic hydration conditions, 1000 OH− ions and 5000 water molecules were added to the system, resulting in a hydration number (λ) of 5 (i.e., five water molecules per OH−). This hydration level is adopted in the literature to emulate moderate hydration conditions within anion exchange membranes and was selected based on the works by Luo et al.,54 which demonstrated that λ in the range of 3–7 is a representative and stable hydration state for evaluating ion transport and membrane morphology under confinement.
To obtain polymer membranes with realistic experimental densities and structures, we employed a multi-step compression and relaxation procedure adapted from Larsen et al.,55 which is a refined technique previously developed by Hofmann and Karayiannis et al.52,56 Specifically, the polymer–water–hydroxide system underwent a carefully structured, 21-step thermal cycling protocol. Each thermal cycle consisted of one NPT (constant number, N, pressure, P, temperature, T) compression step at a low temperature, followed by two NVT (constant number, N, volume, V, temperature, T) equilibration steps, first at a high temperature and then at a lower temperature. A detailed description of this compression and relaxation procedure is provided in Table S1. Recent studies by Risko et al.30 have demonstrated that this methodology reliably yields condensed polymer structures and densities consistent with experimental observations. After the 21-steps compression and relaxation process, 500 carbon dioxide molecules were randomly inserted into the simulation box, and an initial energy minimization was performed to obtain the ground-state structure. The Nose–Hoover thermostat57,58 was used to stabilize the system temperature at 300 K with a time constant of 1 ps and the Parrinello–Rahman barostat59,60 was used to maintain the pressure at 1 bar with a time constant of 10 ps. Three-dimensional periodic boundary condition (PBC) was applied, and a spherical cut-off of 1.2 nm was used for all van der Waals (vdW) interactions. The particle-mesh Ewald (PME)61 with 1.2 nm cutoff for long-range electrostatic interactions were used throughout the simulations. Each system was simulated for 20 ns under the NPT conditions, with the final 10 ns used for analysis with MDAnalysis package.62,63 The diffusion coefficients were extracted by fitting the mean-squared displacement within the 12–18 ns window using GROMACS tools.
An in-house Python script was utilized to extract smaller sub-systems from the final MD simulation trajectories; a sample script is included in the SI. Each extracted sub-system contains the polymer functional group, a portion of the polymer backbone, a OH− ion with its surrounding water solvation shell, and a CO2 molecule located within 5 Å of this solvation shell. The total atom in each sub-systems was in the range of 35 to 45 atoms after removing the additional fragments. These sub-systems served as the basis for subsequent AIMD and DFT simulations to provide deeper insights into the structural and energetic characteristics associated with HCO3− formation.
For the AIMD simulations, the initial extracted sub-systems were positioned at the center of a 2 nm × 2 nm × 2 nm simulation box. A temperature of 300 K was maintained for a duration of 80 ps simulation with a time step of 1 fs. The NVT was adopted, utilizing the Nose–Hoover thermostat.69 Thermal energies and trajectories were recorded every 1 fs. An example of the input file used for the AIMD simulation is included in the SI. The MDAnalysis package62,63 was used to analyze the AIMD trajectories and to calculate the atomic distance and radial distribution functions.
The transition-state (TS) geometries were identified using the quadratic synchronous transit (QST3) method at the ωB97xD/def2-SVP level of theory.71 Sub-system structures extracted from classical MD simulations served as initial geometries for the reactant states. The TS and final product geometries were obtained by strategically repositioning one water molecule from the solvation shell to the vicinity of CO2. To accurately capture dispersion interactions, reactant, product, and TS structures were subsequently optimized at the ωB97xD/def2-SVP level of theory. Single-point calculations using the ωB97xD/def2-TZVP level of theory were then conducted to refine the energetic profiles and improve the accuracy of calculated reaction energies. Additionally, frequency analyses performed at the ωB97xD/def2-SVP level confirmed the nature of each stationary point, where TS structures exhibited one imaginary frequency aligned with the reaction coordinate. All TS calculations incorporated the Polarizable Continuum Model (PCM) with a dielectric constant (ε) of 78.3 to correspond to the aqueous conditions.
000 to 50
000 bar) for each polymer system. This approach, adapted from the work of Larsen55 and Risko et al.,30 enables densification of initially low-density polymer configurations while allowing structural relaxation at intermediate and low temperatures. For each pressure condition, five independent simulations were conducted to assess reproducibility and statistical consistency of the final densities.
Based on the analysis of the resulting equilibrium densities, a Pmax value of 40
000 bar was selected for all subsequent MD and DFT studies. This choice ensures that the polymer matrices used in downstream simulations are representative of equilibrated bulk systems with consistent densities, while avoiding the added computational cost associated with higher compression pressures. Additional discussion of the density trends and statistical variation across pressure conditions is provided in Fig. S3.
:
CO2 close-contact ratio (17.6%), notably surpassing the ratios (12–16%) observed in other polymer systems. This elevated ratio suggests that QIPSf provides more robust and effective interaction sites to stabilize CO2, significantly enhancing its local concentration and immobilization within the polymer matrix. Although HCO3− formation was not directly modeled, the proximity and strong association between CO2 and OH− ions observed in QIPSf highlight its potential for facilitating HCO3− formation relative to other polymers in this study.
To further evaluate the robustness of these observations, we conducted additional simulations in which all atomic charges were uniformly scaled by factors of 0.6, 0.7, 0.8, and 1.0, in addition to the default value of 0.75. The results, presented in Fig. S11, show that while the absolute values of the CO2 diffusion coefficients decrease with increasing charge magnitude, as expected due to stronger electrostatic interactions, the qualitative trends among polymers remain consistent. Notably, QIPSf consistently exhibits the lowest CO2 diffusion coefficient across all scaling factors, confirming that its strong interaction with CO2 is not an artifact of the chosen scaling scheme but rather an intrinsic feature of its chemistry. This reinforces our conclusion that QIPSf provides the most effective environment for stabilizing CO2 and potentially promoting HCO3− formation.
QAPSf and QIPSf, which feature a quaternary benzyl polysulfone backbone, exhibit notably lower diffusion coefficients (0.009–0.0126 × 10−3 nm2 ps−1) than their counterparts (QTMTTf, QATTf and QITTf) containing m-terphenyl units and alkyl chains (0.0208–0.0308 × 10−3 nm2 ps−1). Additionally, polymer containing the imidazolium functional group displayed the lowest CO2 diffusion coefficient within each backbone category. The detailed diffusion coefficient value and OH−
:
CO2 contact ratio are included in the Table S2. A plausible explanation for this is that the quaternary benzyl polysulfone (PSf) backbone provides enhanced noncovalent interactions—such as π–π stacking or electrostatic forces—between CO2 and the polymer matrix. These interactions may increase the likelihood of CO2 being transiently associated with the polymer environment, thereby reducing its effective diffusion. Furthermore, the presence of imidazolium-based functionality (particularly in QIPSf and QITTf) further augments these interactions, as imidazolium's planar ring can engage in robust π-character and electrostatic interactions with CO2. By contrast, QTMTTf, QATTf and QITTf with their phenyl and alkyl chain structures, appear to offer relatively less affinity towards CO2, leading to higher MSD values and faster diffusion (Fig. 2(c)). These results underscore how minor structural modifications – such as switching from m-terphenyl plus alkyl groups to a quaternary benzyl polysulfone backbone – significantly influence the polymer's ability to stabilize CO2 and promote its transformation into carbonate species.
From a practical perspective, lower CO2 diffusivity may be advantageous if it aligns with longer residence times that enhance the conversion of CO2 to carbonate species; however, it also raises considerations regarding gas treatment throughput. Overall, QIPSf was identified as a promising ionomer material for use in the BPM electrochemical separation unit because its polymer backbone and tethered imidazolium group work cooperatively to effectively coordinate with CO2 and potentially promote subsequent carbonate formation reactions.
In Fig. 3(a), the RDFs from classical MD simulations show that QIPSf and QITTf, both of which incorporate imidazolium-based cations, exhibit earlier rising peaks (indicated by the red arrows) beginning around 3.5–4.0 Å, compared to the ammonium-based systems (QAPSF, QATTf, and QTMTTf), whose RDFs rise later and peak at slightly longer distances (∼5.3 Å). This early rise in RDFs for imidazolium systems reflects closer and more immediate N–C(CO2) contacts, suggesting a stronger electrostatic and directional interaction between CO2 and the imidazolium ring. Similar behaviors are also observed in the AIMD-derived RDFs for selected subsystems as shown in Fig. 3(b). Both IM-Ben and IM-nBu show sharp increases at shorter distances compared to AM-Ben, the ammonium containing subsystems. These results further support the idea that the planar aromatic structure of imidazolium enhances CO2 localization near the cation, likely through π–quadrupole and electrostatic interactions. In contrast, ammonium-based cations tend to stabilize CO2 at slightly greater distances, consistent with their more spherical and less delocalized charge distribution. The consistency between AIMD and classical MD RDFs further validates the reliability of the simulations performed and supports the choice of the OPLS-AA force field, as discussed in the computational approach section.
The RDF data indicate that imidazolium-functionalized polymers provide a structural advantage in bringing CO2 into closer proximity to the nitrogen site, a feature that could lower the entropic cost of reactant alignment and enhance the likelihood of CO2 activation or conversion in HCO3− transformation reactions.
Using DFT, we mapped out the relative energy levels where one solvation–shell water molecule reorients to facilitate CO2 conversion into HCO3−. The results (Fig. 4(a)) indicate that IM-Ben exhibits a substantially lower reaction barrier (∼252 meV) compared to both IM-nBu (∼324 meV) and AM-Ben (∼875 meV). This finding suggests that the aromatic benzyl substitution on the imidazolium ring stabilizes the transition state more effectively, thus promoting a more favorable pathway to HCO3− formation.
![]() | ||
| Fig. 4 (a) Relative energies for bicarbonate (HCO3−) formation from CO2 and OH− across three polymeric model subsystems (IM-Ben, IM-nBu, and AM-Ben, see Fig. 2). The energy profiles represent initial states (IS), transition states (TS), and final states (FS), computed at the ωB97xD/def2-TZVP//ωB97xD/def2-SVP level of theory with PCM solvation (ε = 78.3). (b) Corresponding optimized geometries highlighting hydrogen-bonding interactions (depicted as dashed lines) during the reaction pathway. Energy values (in meV) are relative to the respective IS. Notably, IM-Ben exhibits more extensive hydrogen bonding in the TS structure compared to IM-nBu and AM-Ben (6 vs. 5 interactions), correlating with a lower activation energy barrier and enhanced stabilization during CO2-to-bicarbonate conversion. | ||
The smaller TS energy barrier (0.25 eV) in IM-Ben indicates a readily accessible route for proton transfer and CO2 incorporation into the HCO3− moiety. Such enhanced reactivity may arise from a synergistic interplay between the ring's π-conjugation and the relatively open environment offered by the benzyl group, allowing optimal hydrogen-bonding networks among OH−, water molecules, and CO2. The representative configurations of these subsystems showing the hydrogen-bonding networks in each reaction state are included in Fig. 4(b).
To account for entropic contributions at room temperature (298 K), we also computed Gibbs free energy (ΔG) profiles for the bicarbonate formation reaction (Fig. S9). The inclusion of entropy slightly shifts the energy barriers compared to the enthalpy-only results, but the overall trends remain consistent. Specifically, IM-Ben maintains the lowest barrier (0.15 eV) compared to IM-nBu (0.40 eV) and AM-Ben (0.93 eV). The final states remain exergonic for IM-Ben (−0.56 eV) and IM-nBu (−0.57 eV), while AM-Ben shows a less favorable stabilization (−0.35 eV). To further examine this mechanistic assignment, we analyzed electron-density/ESP maps and NBO charges for representative subsystems (Fig. S10 and Table S3). These results provide direct electronic-level evidence for transition-state stabilization in IM-Ben. Specifically, for the IM-Ben subsystem, the CO2 carbon becomes more positive at the TS (e.g., IM-Ben: +1.05 → +1.08 e), while the OH− oxygen becomes less negative (−1.24 e→ −1.07 e), consistent with nucleophilic attack and C–O bond formation. Moreover, the O atom from H2O (that eventually bonds to CO2) transiently becomes more negative (IM-Ben: −1.03 e → −1.16 e), reflecting strengthening of the H-bonding network at the TS. Simultaneously, the imidazolium unit in IM-Ben exhibits a transient reduction in net positive charge (IS → TS → FS: +0.67 → +0.66 → +0.68 e), reflecting enhanced charge accommodation relative to IM-nBu and AM-Ben analogues. The ESP maps (Fig. S10) further highlight polarization of the benzyl-imidazolium π system, which delocalizes the positive potential and buffers developing charge separation. Together, these data reinforce that the benzyl-substituted imidazolium lowers the CO2-to-bicarbonate reaction barrier not only through hydrogen-bonding network strengthening but also through π-driven electronic stabilization of the transition state, providing a more favorable kinetic and thermodynamic pathway.
At low hydration (1
:
1 ratio of OH−
:
H2O), both IM-Ben and IM-nBu display favorable deprotonation energies of −30.5 and −39.7 kJ mol−1, respectively, indicating that the imidazolium C−H bond is readily cleaved in strongly alkaline or underhydrated conditions. When the water content doubles (1
:
2 ratio), IM-Ben becomes slightly unfavorable for deprotonation (+0.5 kJ mol−1), whereas IM-nBu remains favorable (−5.0 kJ mol−1). The disparity grows with increased hydration: at 1
:
3 ratio, IM-Ben requires +22.0 kJ mol−1vs. +10.9 kJ mol−1 for IM-nBu, and at 1
:
4 ratio, the deprotonation energy barrier climbs further, reaching +34.2 kJ mol−1 for IM-Ben and +22.7 kJ mol−1 for IM-nBu. Consequently, IM-Ben is substantially more resistant to proton loss at moderate to high hydration.
These findings underscore the delicate balance between polymer structure, local pH, and the availability of water molecules in the membrane's solvation shell. Under intensely alkaline or poorly hydrated conditions (1
:
1 or 1
:
2), both imidazolium systems are at risk of deprotonation, which would compromise HCO3− formation by consuming OH− in water-producing side reactions. By contrast, increased hydration (≥1
:
3) stabilizes the imidazolium ring, thus preserving available OH− for efficient CO2 conversion to HCO3−. The marked contrast between IM-Ben and IM-nBu at these ratios suggests that benzyl substitution confers greater structural resilience against deprotonation, favoring the formation of HCO3− intermediates.
In practical terms, these energy profiles emphasize that controlling membrane hydration (or effectively moderating pH) is critical for sustaining imidazolium functionality. Excessively high pH and elevated temperatures (i.e., >40 °C) can accelerate ring deprotonation and reduce the efficiency of CO2 conversion to bicarbonate and carbonate species, while more moderate alkalinity provides an optimal operating window for robust CO2 uptake and HCO3− generation. Mitigating C–H deprotonation in the imidazolium cation can be achieved by adding an ‘R’ group to the C2 between the nitrogen atoms (e.g., 1,2-dimethyl imidazolium74 or trimethoxyphenyl75). In concentrated alkaline environments, some researchers have a methyl on all the carbon atoms in the imidazolium ring.76 Overall, these insights will be useful in down selecting anion exchange ionomers for the ionomer coated porous fabric in the spacer channel and the anion exchange membrane layer in the bipolar membrane because judicious selection of the material is vital for CO2 conversion effectiveness.
Distinctly different energetic preferences for HCO3− formation versus imidazolium deprotonation in IM-Ben and IM-nBu systems were observed, particularly at moderate hydration levels (OH−
:
H2O = 1
:
2). To elucidate this behavior, we calculated the PES for both reaction pathways in each subsystem. To compare the imidazolium deprotonation energy with the HCO3− formation, we used a subsystem that contains two water molecules in the OH− solvation shell, as well as a CO2 in the outer shell as shown in the insert of Fig. 6. Each geometry was optimized at B3LYP/6-31+g(d,p) level of theory for the PES calculations.
Our calculations reveal that IM-Ben exhibits a higher energy barrier for imidazolium deprotonation (approximately 10 kJ mol−1), whereas HCO3− formation remains significantly more energetically favorable with a barrier of 6 kJ mol−1. Conversely, in IM-nBu, the deprotonation pathway has a low reaction barrier of 6 kJ mol−1 and a slightly favorable deprotonated state (−0.3 kJ mol−1), while a high barrier for the HCO3− formation (23 kJ mol−1), resulting in a competitive reaction route for proton transfer to OH− to produce water and thereby diminishes the availability of OH− for HCO3− generation.
The underlying structural differences between IM-Ben and IM-nBu were identified to have a profound impact on OH− and CO2 conversion to HCO3−. The benzyl substitution in IM-Ben provides enhanced electronic and steric stabilization of the imidazolium ring, reducing its susceptibility to proton loss. This stabilization encourages CO2 to preferentially engage in HCO3− formation rather than promoting proton transfer to OH−. In contrast, the alkyl substituent in IM-nBu lacks this stabilizing effect, rendering the system more prone to undesirable proton loss. Furthermore, IM-Ben is more effective for HCO3− formation under moderately alkaline conditions because the local solvation environment adequately stabilizes OH− without inducing excessive ring deprotonation. In summary, the results underscore the importance of polymer structure and hydration control in anion exchange ionomer design, suggesting that benzyl-substituted imidazolium systems could deliver enhanced stability and superior CO2 conversion performance compared to their alkyl-substituted counterparts.
:
OH− interactions compared to quaternary m-terphenyl trifluoromethyl systems. The enhanced interaction in imidazolium-based systems was confirmed by radial distribution function analyses, which indicated closer spatial proximity of CO2 and OH− to the aromatic nitrogen centers.
AIMD and DFT calculations further validated these findings, highlighting the benzyl-substituted imidazolium (IM-Ben) as particularly effective in stabilizing the transition state and thus significantly lowering the energy barrier (by approximately 72 meV relative to alkyl-substituted IM-nBu) for HCO3− formation. Additionally, systematic exploration of hydration conditions revealed that benzyl-substituted imidazolium groups provide superior resistance against undesirable deprotonation at moderate to high hydration levels, preserving the availability of OH− ions for effective CO2 conversion to HCO3−.
The competing reaction pathways, illustrated by the PES analysis, underscore the advantages of IM-Ben systems over IM-nBu, as the former preferentially favor HCO3− formation over imidazolium deprotonation under realistic operating conditions. Collectively, these computational insights indicate that carefully engineered polymer structures, particularly benzyl-substituted imidazolium-functionalized membranes, hold considerable promise for improving the efficiency and stability of bipolar membranes in electrochemical CO2 separation and HCO3− formation applications. Future experimental validation guided by these computational predictions could further refine polymer design criteria and optimize operational parameters for enhanced CO2 separations technologies.
Additional input files, output files, and scripts can be made available on reasonable request to the corresponding author.
| This journal is © The Royal Society of Chemistry 2026 |