Roghayeh Azizi,
Mojtaba Shamsipur*,
Avat (Arman) Taherpour,
Maryam Miri and
Afshin Pashabadi
*
Department of Chemistry, Razi University, Kermanshah, Iran. E-mail: mshamsipur@yahoo.com; apashabadi@razi.ac.ir
First published on 30th July 2025
The sluggish rate of proton transfers during proton-coupled electron transfer is one of the challenges in designing a holistic blueprint for complete biomimicry. Here, we present a one-pot, facile method for the refrigerated synthesis of two different covalent organic polymers by the separate copolymerization of diphenylamine-4-sulfonic acid and diphenylamine with para-aminophenol. Substituting diphenylamine with its porous sulfonate analog (PCOP) significantly alters the structural design and proton shuttling characteristics, resulting in remarkable efficiency in the uncommon non-concentrated proton-coupled electron transfer (n-PCET) during OER in both alkaline and neutral media. The functionalized carbon fabric (FCF)/PCOP requires an overpotential of 234 mV at 10 mA cm−2, comparable to metal-based electrocatalysts. A striking semi-reflective boundary condition in the Nyquist plot in acidic media introduces a net capacitive behavior upon protonation of –SO3−, revealing an ideal capacitor, which has been targeted in the design of a supercapacitor in 0.1 M H2SO4 with capacitance Csp = 670.79 F g−1, retaining 89.52% even after 12000 charge–discharge cycles. Further experimental evidence was obtained by D/H isotope studies, proton inventory, pH dependence analysis, PCOP-water ATR spectroscopy and Gerischer impedance spectroscopy. The results of the DFT studies were used to further explore the structural engineering driven by water clusters and SDS molecules during the cooled synthesis of the COP.
This work experimentally tracks the additional functional role tied to the structure-directing agent (SDA, e.g., surfactant), where interfacial perturbation of the water cluster HBN, as the substrate, is characterized via attenuated total reflective (ATR) spectroscopy. Covalently-bonded molecular building has been of great interest for the synthesis of adaptable materials with distinct features in recent decades. Emphasizing covalent bonds in fabricating porous covalent structures endows reliable and adaptable chemistry beyond discrete molecules, aiming at diverse applications. These materials are known as covalent organic polymers (COPs). COPs have amorphous or crystalline nature and are basically characterized by highly cross-linked porous structures that feature robust covalent bonds.6–8 COPs have been impressively investigated as promising materials for gas storage, proton conduction, and energy applications.9–20 The abundant pore volume of COPs facilitates mass ion transfer in the electrolyte, allowing impressive processes and fast electrochemical response in energy harnessing and storage devices.21 However, their capability in PT, specifically multistep proton coupled electron transfer (PCET) reactions in heterogeneous electrocatalysis, has been overlooked compared to competitors such as conductive polymers4 or the metal-free organic frameworks3 developed recently by our research groups.
We propose crystalline COP as an advanced-generation, metal-free PCET catalyst, coupled with the unique stability of polymeric structures as we studied in the case of poly-arginine-based guanidine proton relay4 and our proposed orderly elongated structure of crystalline zwitterionic hydrogen bonded organic framework (ZHOF)3 in the harsh medium of water oxidation reaction (WOR). Recent advances in the design of covalent organic frameworks (COFs) functionalized with metal nanoparticles have demonstrated enhanced catalytic activity in the oxygen evolution reaction (OER), highlighting the synergistic effects of covalent frameworks and metal sites in electrocatalysis.22 This approach allows regulating the functionality and investigating the role of diverse functional groups in the crystalline structure, extended through rigidly strong covalent bonds with superior yield of proton hopping via water-included organic pores (proton wires).23 In the O4-water chain within PSII, hydrophobic residues such as D1-Asp61 and Glu354 rigidly fix the water molecules in the H-bond network and serve as hydrogen bond acceptors, which promote the Grotthuss-like mechanism, thereby forming a low-barrier proton wire. Fabricating rigidly arranged reticular COPs fixes proton donors and acceptors in the HBN, reduces the wandering paths of proton motion and minimizes wobbling H-bond directions, thus forming the pre-/post-PT arrays to shuttle protons.
In PSII, the presence of HBN-containing hydrophilic moieties in proton channels intensifies rigidity and strongly latches each group into the other; meanwhile, the hydrophobic motifs in the channel enable water-molecule mobility.2 OER electrocatalysts often face a PCET imbalance, requiring an interfacial proton acceptor, usually a base solution, which is limited by natural diffusion. This imbalance has led to exploring OER sideways, through proton relays instead of just pure electrocatalysts.
Recent studies emphasize that rational interface engineering through heterostructure formation and strategic doping can significantly modulate interfacial charge distribution and enhance catalytic performance. Such approaches, while often explored in other electrocatalyst contexts, provide valuable conceptual guidance for advancing OER catalysts.24 In the past decades, the lagged evolution of PT—the sluggish part of PCET—is one of the weaknesses of designing a holistic blueprint for biomimicked artificial leaves. This substantial problem, besides the large proton/electron mass ratio, requires utilizing a solution proton acceptor (OH−) for the efficient cleavage of O–H bond and liberation of molecular oxygen. Such an intrinsically dictated pathway suffers from the constrained kinetics of the natural diffusion of protonic species (H2O, OH, H+) within the diffusion layer and corrosion problems of metal-based electrocatalysts in harsh mediums.25 The lagging of water delivery and proton shuttling to/from the current collector may also cause incomplete WOR and the formation of reactive oxygen species, OH˙, which damage the catalyst frameworks.23 These limitations are great challenges to achieving full biomimicry, with the ideal being analogous to the active core of the PSII oxygen-evolving complex, which liberates O2. The advanced deprotonation encourages “redox leveling,” which allows for reaching high oxidation states within a limited potential range.
We have successfully developed a straightforward molecular-level design and a controlled method for synthesizing various morphologies, with the apparent distinction of reticularity, for a covalent organic polymer using an oxidative refrigerated synthetic procedure. The overall synthesis involves the combination of diphenylamine-4-sulfonic acid or diphenylamine with p-aminophenol. The results elucidate the self-catalytic role of p-aminophenol polymer during the oxidative polymerization process. The morphology and functionalities of the product can be directly influenced by reaction parameters such as temperature (refrigeration vs. room temperature), the impact of p-aminophenol, and the inclusion of surfactants like SDS and the role of water clusters, which has been studied through a detailed DFT study. The existence of a super-hydrophilic SO3 that can tightly control the rigidity and interfacial reactivity directly drives the OER activity and also energy storage properties of the carbon fabric/PCOP electrode. Eliminating SO3 from the monomer structure completely changed the reticularity and aborted OER activity at pH ≥ 7 and capacitive behavior at pH ≈ 1. Various electrochemical studies, such as proteo/deuteron isotopic effects, pH studies, and Gerischer impedance studies, unraveled the detailed mechanistic information of the PCET. An interfacial ATR study was performed to investigate possible agitation of water's HBN, providing mechanistic information. DFT studies were employed to elucidate the formed structures and consider the relative energies of the system.
The morphology of the various synthesized samples was examined using FE-SEM. As illustrated, the morphologies of FCF@RCOP, FCF@SCOP-SO3, and FCF@PCOP-SO3 differ from one another. Fig. 1(A) depicts RCOP, which exhibits a stacked polymer morphology. For SCOP, the polymerization of diphenylamine-4-sulfonic acid and p-aminophenol in an aqueous solution resulted in a distinctive spherical structure of varying sizes (Fig. 1(B)), attributed to the formation of the COP influenced by water clusters (as detailed in the computational section). This observation is further supported by the notably difficult drying of the residues in the beaker in the case of SCOP. Fig. 1(C) illustrates that the FCF is entirely enveloped by PCOP-SO3. The magnified view reveals that the PCOP-SO3 components interconnect to create a porous structure with a honeycomb-like appearance. Additionally, the presence of SDS inhibits the formation of water-cluster-driven SCOPs, significantly influencing the structure's morphology.27
The mapping analyses indicate that SCOP-SO3 (Fig. 1(E)) and PCOP-SO3 (Fig. 1(F)) contain oxygen, nitrogen, sulfur, and carbon. The presence of sulfur in the polymer, as confirmed by the results, validates the successful synthesis. In contrast, RCOP is composed solely of nitrogen, oxygen, and carbon elements (Fig. 1(D)). The hydrophilicity of the synthesized COP was assessed through contact angle measurements. The relevant image can be found in Fig. 1(G), along with the Supporting Videos VS1 and VS2 (ESI†). Fig. S2 (ESI†) displays the bare FCF at different magnifications, showing no changes in the morphology of the carbon fiber.
Herein, to investigate the function of water clusters during synthesis in aqueous environments, a model was created using 150 water molecules that are interconnected through hydrogen bonds. In Fig. 2, from left to right, we tracked the steps of scavenging of the water cluster (nH2O) by the modeled polymer. In Step 1, the modeled polymer has started folding around the water cluster through H-bond via functional groups (–OH, –NH−, –NH2, –SO3−), eventually curving around and completely capturing the water cluster (step 2). These two stages are depicted by a hand-ball illustration. Outside the dashed rectangle, we consider the polymer further unfolding from around the cluster upon SDS treatment. The modeled SDS (shown as a spoon) withdraws the H2O molecules from the main water cluster (nH2O) by its H-bond accepting agents (–SO3− functional group at the end of aliphatic chain). In the next steps (steps 2 and 3), the polymer has completed unfolding around the water cluster (nH2O), and some of the SDS molecules separated H2O molecules from the main water cluster and some other parts made a great micelle. In step 4, the modeled polymer has completely opened and unfolded, releasing the water cluster. The micelles formed between the SDS + water cluster can direct the growth of the uncapped (opened) polymer. These micelles have an interfacial hydrophobicity that can impede the spherical growth and establish a porous structure.
Fig. 3(A) and (B) show the differences in free energies (in kcal mol−1) between the unfolded model polymer and the water cluster with the model polymer folded around the water cluster by the discussed H-bonding agents. The energy gap was calculated to be 60.6 kcal mol−1. In the presence of SDS, the differences in free energies were calculated to be about 14.7 kcal mol−1 in this modeling. In the next step, to reach the unfolded modeled polymer and the water cluster (nH2O), the difference of free energy is 45.9 kcal mol−1. It is worth expressing that the formation of the mega complex (polymer + water cluster), as an exothermic reaction, releases energy. This phenomenon is consistent with the occurrence of the exothermic reaction under refrigeration, as explained in the experimental section.
Fig. S13 (ESI†) shows the modeled structure for the discussed polymer and the calculated IR spectrum of the discussed modeled polymer by B3LYP/6-31G* method. The vibrations of the functional groups are shown (3500–4000 cm−1 for H-bonding agents –OH, –NH–, –NH2, –SO3−). The additional details of the vibrational frequencies for the functional groups are shown in the calculated IR spectrum (1250–1450 cm−1 for –SO3H functional group). The stretching O–H groups in the discussed structure are shown on the right side of the IR spectrum. The signals between 2700–3700 cm−1 of the experimental ATR results could be compared with the calculated IR spectrum. In Fig. 3(C), the 1H-NMR spectrum of the modeled SCOP calculated by the B3LYP/6-31G* method shows good agreement with the results of the experimental 1H-NMR spectrum. The signals between 6.89 and 7.55 ppm in the experimental and between 6.87 and 7.40 ppm in the calculated spectrum concern the H-aromatics. The broad signal between 5.90 and 7.90 ppm concerns the N–H; its extension probably is related to the N–H quadrupole moment of the N atom. The doublet and triplet (at 7.28 and 7.55 ppm, respectively) agree well with the calculated 1H-NMR spectrum at 7.38 and 7.40 ppm for the HA and HB of the aromatic rings. The signals of the experimental 1H-NMR spectrum between 8.30 and 9.51 ppm are related to the H-bonding agents (–OH, –NH–, –NH2, –SO3− functional groups located on the aromatic rings), which match the theoretical 1H-NMR spectrum (8.36–9.44 ppm). Fig. 3(D) and (E) show the experimental 1H-NMR and 13C-NMR spectra of PCOP, with good agreement between the experimental and theoretical spectra. The signals of DMSO and DHO (solvent signals) appeared at 2.43 and 3.39 ppm, respectively. The signals between 6.04 and 7.13 ppm concern the H-aromatics in the porous form of the discussed polymer. These signals were shifted to 6.89–7.55 ppm in the spherical form (SCOP). In Fig. 3(D), the broad signal between 7.00 and 7.35 ppm relates to the N–H; its expansion probably relates to the N–H quadrupole moment of the N atom, which shows a broad shifted signal between 5.90 and 7.90 ppm. The doublet and triplet (at 7.28 and 7.55 ppm, respectively; as A, B and C signals) in SCOP agree well with the calculated 1H-NMR spectrum at 7.38 and 7.40 ppm for the HA and HB of the aromatic rings.
The signals (Fig. 3(C)-experimental 1H-NMR spectrum) between 8.30 and 9.51 ppm concern the H-bond-making agents (–OH, –NH–, –NH2, –SO3− functional groups located on the aromatic rings), which match with the theoretical Fig. 3(D) results of 1H-NMR spectrum (8.36–9.44 ppm). This type of signal in the 1H-NMR spectrum disappeared in the porous form of the discussed polymer. Fig. 3(E) shows the H-spin decoupled from the 13C-NMR spectrum of the porous form of the discussed polymer. The strong and broad signal of DMSO (solvent signals) appeared at 40.6 and 42.7 ppm. The signal at 129.6 ppm is related to CA atoms of the aromatic ring in the porous form of the discussed polymer. The signals of CB and CD atoms of the aromatic ring in the porous form of the discussed polymer appeared at 107.3 ppm. The signals of CC atoms of the aromatic ring in PCOP appeared at 105.1 ppm, and the signals of CE and CF atoms of the other part of the aromatic ring in appeared at 145.5 and 158.0 ppm, respectively. The signals of CK and CL atoms of the cyclohexa-2,5-diene-1,4-diimine moiety appeared at 103.1 ppm. Moreover, the weak signals between 150–155 ppm and 160–165 ppm concern the 4 °C atoms (G, H, I and J) of the aromatic and cyclohexa-2,5-diene-1,4-diimine parts of the discussed polymer. These 4 °C atoms are under the nuclear Overhauser effect, so they appeared as weak signals. The results show good agreement with the calculated data.
To compare the relative signal stability after two weeks, the measurements were repeated under the same conditions, and the corresponding voltammograms are presented in Fig. S14 (ESI†). Also exhibited is excellent long-term electrochemical stability under OER conditions (Fig. S12, ESI†).
![]() | ||
Fig. 5 (A) LSVs over the pH range of 12.5 to 13.7 for both FCF and FCF@PCOP-SO3; (B) the plot of log![]() |
![]() | (1) |
Electrochemical measurements were carried out by obtaining LSV plots in H2O and D2O media at the same pH(D), where the pD value was calculated from the equation pD = pHread + 0.41. KOH and K3PO4 basic solutions were utilized to adjust pH or pD. The resulting LSVs expressed the same isotopic behaviors for CF (Fig. 6(A)) and FCF (Fig. 6(B)) during the WOR. Over all potentials, both electrodes showed lower current density in the deuteron solution compared to those in the proton solution. The KIE values over the potential range of 1.7 to 1.9 V vs. RHE ranged from ca. 1.8 to 2.4, these primary KIE values (>1.5) indicate an O–H bond cleavage in the RDS.3,4,31 A slight decrease observed at the FCF, ranging from 3% to 15% (at different potentials), may be related to the participation of interfacial functions on the FCF, indicating that the CF may perturb the PCET process during low-yield WOR undertaken at the FCF.
The lagging of JD at higher driving forces may be related to the acceleration of ET, revealing a difference in the bond breaking of O–D/O–H, due to the shorter bond length of O–D being energetically unfavorable in PCET compared to the O–H bond. Therefore, in an energy threshold, this difference is revealed, resulting in a KIE >1. Fig. 6(C) shows the equivalent LSV at FCF@PCOP-SO3 with a dramatic decrease of KIE, ranging from 1.23 to 1.49 over the potential window of 1.7 to 1.9 V. Increasingly positive potentials cause the observable differences, although still, the KIE falls into the SKIE region with no PT in RDS, which can be related to the relatively delayed PT upon acceleration of ET. However, despite the increase in ET rate at higher overpotential, O–H bond cleavage does not occur in the RDS, implying the high capability of FCF@PCOP-SO3 in proton transport. Comparing the LSVs at FCF@PCOP-SO3 with the two former cases reveals an increased decoupling driving force (DDF) of 1.62 V, signifying that the proposed proton relay facilitates an effective PT during the PCET processes. The shift of the DDF from 1.43 for CF to 1.62 for FCF@PCOP-SO3 indicates a coupling of ET/PT over a greater sweep potential, where the higher kinetics of ET does not lead to delay of PT when protons are substituted with deuterium.
We investigated the role of the robust interfacial sulfonates as in situ heterogeneous stimuli on water's HB network, where the successful adsorption events of protonic species can be controlled. Using attenuated total reflection infrared spectroscopy (ATR-IR), we compared the broad OH-stretching band of water, 2800–3700 cm−1, to explore the extensive HB agitation by COP, a critical event that may directly predominate the vibrational modes and the different degrees of the freedom for the water molecules beyond the adsorption of substrate, as the first step of the heterogeneous WOR. After recording the ATR spectra for CF + water and FCF@PCOP-SO3 + water, the OH-stretching bands were deconvoluted to the four sub-bands of water, denoting the multiple degrees of HB modes.32,33 Fig. 6(D) shows the results of the curve-fitting analysis of the FTIR spectra for CF + water and FCF@PCOP-SO3 + water, singly. The resolved spectra displayed four peaks, from ca. 3200 cm−1 to 3600 cm−1, corresponding to triply, doubly, and singly hydrogen-bonded and free (non-hydrogen bonded) H2O, respectively.34,35 The degree of breaking of the HB network (DBHB), as the ratio of calculated peak area of the free (pink) and singly (blue) bond water spectra around 3600 cm−1 to that of the hydrogen-bonded water molecules around 3200 cm−1 (green curve), were respectively found to be 7.9–12.7% and 12.5–30.98%, respectively, for CF + water and FCF@PCOP-SO3 + water. The increased DBHB for FCF@PCOP-SO3 compared to CF + water elucidates the increased delivery of free water molecules and efficient capability of sulfonate residues in the perturbation of the water HBN, which directly improves the success of adsorption events during heterogeneous WOR.34
![]() | ||
Fig. 7 Energy storage properties of FCF@PCOP-SO3. (A) The comparative CVs. (B) Stability of shape and area for FCF@PCOP-SO3 after 2000 successive CVs. (C) Calculated Csp as a function of the cycle number for the FCF@PCOP-SO3 electrode with a 2.44% loss of the initial value. (D) CVs as a function of potential, ranging from −0.87 to 0.01 V. (E) The calculated Csp against the applied potential window. (F) GCD curves recorded at 1, 2.5, 5, 7.5 and 10 A g−1. (G) The plot of ED versus PD. (H) Ragone plot related to ED and PD of the FCF@PCOP-SO3 electrode compared with some recent values in literature.40–50 (I) Long-term stability indexed with Csp and coulombic efficiency after 12![]() |
Long-term cyclic stability is another vital factor that should be considered to estimate the capacitance constant alterations (Fig. 7(B)). The curve of the CV data for the FCF@PCOP-SO3 is maintained until around 1000 cycles without a change in the loop area, with excellent reversibility within the operating potential window. As illustrated in Fig. 7(C), it shows remarkable long-term CV stability, maintaining over 97.56% specific capacitance performance (Csp) even after 2000 cycles. The Csp of the FCF@PCOP-SO3 can be estimated from the CV curves using the following equation:36
![]() | (2) |
![]() | (3) |
From the GCD curves, the Csp of the FCF@PCOP-SO3 electrode is found to be 670.79 F g−1 at the J value of 1 A g−1. The approximately symmetric curve-like shapes of charge–discharge curves rather confirm the electric double layer capacitor energy storage mechanism. Even at the high J of 10 A g−1, the discharge curves of FCF@PCOP-SO3 exhibit only a partially low voltage drop (IR drop), showing low total resistance and excellent electrochemical reversibility. The impressive charge storage capabilities of the FCF@RCOP-SO3 can be attributed to its porous structure, well-organized ion transport pathways, interconnected cavities, and superior wettability. The high hydrophilicity of the as-prepared COP is seen in the contact angle shown in Fig. 1(G) and the corresponding Supporting Videos VS1 and VS2 (ESI†). Moreover, the symmetry of the charge/discharge curves demonstrates the excellent capacitance characteristics of FCF@PCOP-SO3.
Rate capability is an important factor in the selection of SCs for power applications. The SC capacity decreases slowly with the rise in current density, which is usually related to the ion diffusion resistance. Higher J values requires faster ion diffusion to make all the active materials available. The obstruction–diffusion effect limits the penetration of ions due to the time limitation, eventually resulting in some active surface areas being out of reach for charge storage. The relationship between energy density (ED, Wh kg−1) and power density (PD, W kg−1) was investigated to evaluate the overall performance of the energy storage source. The ED generally decreases with increasing PD. Fig. 7(G) displays the Ragone plot of FCF@PCOP-SO3 at different J values (1 to 10 A g−1) with the ΔV of 0.88 V. We calculated the PD and ED using the following equations:37
![]() | (4) |
![]() | (5) |
The maximum ED of the FCF@PCOP-SO3 is 288.59 Wh kg−1, which decreased to 87.520 Wh kg−1 as the PD increased from 1760 to 17600 W kg−1. The large increase in ED may be due to the high V and the increase in capacitance due to faradaic reactions.38 Further, the Ragone plot highlights the superior performance of FCF@PCOP-SO3 toward other SCs in previous reports (Fig. 7(H)). The energy density of FCF@PCOP-SO3 is comparable to or higher than that of most SCs. Another important parameter to be measured in the evaluation of the changed electrode is the stability of the SC over repeated charge–discharge cycles. The symmetrical two-electrode configuration device with the FCF@PCOP-SO3 electrode was subjected to a long-term cyclic stability test at a high J of 11.5 A g−1 with 89.52% retention of the initial Csp after 12
000 cycles (Fig. 7(I)). Especially, the almost-symmetrical GCD curves showed a partial shape change with long-time cycling owing to: (i) an interconnected, predominantly porous structure for the rapid entry and extraction of ions, (ii) structural defects and active sites leading to high reactivity, and (iii) a functionalized porous framework with hopping mechanism for the transport of ions. In Fig. 7(I), the coulombic efficiency (CE) was obtained from the following equation:39
![]() | (6) |
![]() | ||
Fig. 8 The Nyquist plots are recorded in (A) 0.1 M H2SO4, (B) 0.1 NaOH, and (C) 0.1 M PBS (pH ≈ 7) at FCF@PCOP-SO3. The inset of the middle plot displays the Gerischer element-included equivalent circuit. In an acidic environment, the protonated SO3H effectively creates a blocking surface that aligns with the limited scenario of FLW, which describes a 'blocking' surface across most frequencies. This assumption leads to a semi-vertical line in the complex plane, signifying an ideal capacitor, a finding that is strongly reinforced by the data presented in Fig. 7. (D) A standard model for a CEC process defined by Gerischer impedance occurs during proton hopping within the walls of the PCOP. |
The TLM effectively defines the natural transport coupled with a chemical reaction occurring within the diffusion layer. From a physical perspective, the high-frequency TLM indicates a chemical–electrochemical–chemical (CEC) mechanism. The high-frequency Warburg portion is not affected by the particular type of boundary (reflective or adsorptive) at the end of the diffusion zone. On this basis, there are also multiple features, namely, (i) the aborted chemical reaction, i.e., a semi-reflective boundary condition; (ii) the rate of chemical reaction being lower than the natural transport (Rs < Rex), so that the Nyquist plot in the high-frequency region shows a small Warburg part; and (iii) when the Rs ≫ Rex, the natural transport is practically blocked and proton hopping predominates the interfacial PCET reaction running at the applied DC potential. In Fig. 8(A), the blue curve is related to the FCF/PCOP interface where the protonation of the sulfonate groups in acidic medium blocks the possible exchange reaction when sweeping the frequency to the medium- and low-frequency regions (a semi-reflective boundary case, capacitive behavior), while the two non-sulfonated surfaces, CF and FCF, did not cause the behavior. In alkaline medium (Fig. 8(B)), the Nyquist plot displays a semicircle in the low-frequency region, although it continues through a 17° transmission line to the high frequencies instead of turning to the horizontal axis. The depressed transmission line instead of the conventional 45° beyond the perturbation of natural diffusion by exchange reaction reveals a fractal Gerischer (FG) impedance.3,4 In systems with fractal geometry (e.g., porous electrodes or rough surfaces), the diffusion process may demonstrate non-integer (fractal) dimensions, leading to a modified impedance response:28
![]() | (7) |
In terms of Gerischer impedance, Rex and Rs are represented by absolute values, referred to as the product values (RG), which cannot be distinguished separately.51 As schematically demonstrated in Fig. 8(D), the dominance of the exchange reaction over the inherent diffusion process is denoted as proton sinking.3 This phenomenon suggests natural diffusion within a quasi-homogeneous medium is blocked by a hierarchical arrangement of functional groups. The time required for the exchange reaction involving sulfonate residues is less than that for their diffusion through COP cavities. The observed efficient exchange reaction between protonic species and internal residues in crystalline channels, combined with a KIE of less than 1.5, indicates a low potential barrier height for hydrogen bonds, which facilitates the proton transfer process. During the electrochemical step, the proposed active sites for the OER on non-metallic surfaces include heteroatom dopants (e.g., –N, –S), polarized carbon atoms adjacent to polar groups, and surface defects, which facilitate the adsorption and activation of water molecules and reaction intermediates.52–54 We hypothesize that OER proceeds via the adsorption of H2O, OH, and OOH on nitrogen sites or polarized carbon atoms (e.g., electron-deficient C–SO3−), forming surface hydrogen-bond networks (HBN) that promote the deprotonation of OH* to adjacent –SO3− groups.
Experimental evidence suggests that sulfonated porous covalent organic polymers (PCOP-SO3−) play three key roles in enhancing OER performance: (1) proton transport and PCET mechanism modification: the –SO3H groups facilitate efficient proton-coupled electron transfer (PCET) through hydrogen-bond networks. The observed kinetic isotope effect (KIE) values (<1.5, suggesting a secondary KIE) indicate that proton transfer is not involved in the rate-determining step (RDS), while pH-dependent studies confirm a non-concerted PCET mechanism. (2) Enhanced hydrophilicity: Improved wettability, as evidenced by contact angle measurements (Supporting Videos VS1 and VS2, ESI†), ensures better electrolyte accessibility to active sites. (3) Water cluster disruption: the –SO3− groups perturb bulk water clusters, increasing the availability of free water molecules and accelerating adsorption kinetics, as supported by attenuated total reflectance (ATR) spectroscopy (Fig. 6(D)).
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ma00451a |
This journal is © The Royal Society of Chemistry 2025 |