Bo Lia,
Wenke Liua,
Mengqi Dua,
Anan Lia,
Xuzhuo Sun*a,
Jiaxun Fenga,
Jing Chena,
Dongjin Wan*b and
Haibo Zhang
*c
aCollege of Chemistry and Chemical Engineering, Henan University of Technology, Zhengzhou 450001, P.R. China. E-mail: sunxuzhuo@haut.edu.cn
bSchool of Environmental Engineering, Henan University of Technology, Zhengzhou 450001, P.R. China. E-mail: djwan@haut.edu.cn
cCollege of Chemistry and Molecular Sciences, Wuhan University, Wuhan 430072, PR China. E-mail: haibozhang1980@gmail.com
First published on 16th July 2025
The structures of interfacial water and key HER intermediates significantly influence the electrochemical performance of the HER. Herein, a doping strategy was applied to induce charge redistribution within the substrate, which changed the local electron field of the entire catalyst. Boron organic polymers were selected as B and N sources to fabricate Ru NPs dispersed on a BCN substrate (Ru/BCN), which were assembled through Cs2[closo-B12H12] and 4,4′-bipyridine. In situ surface-enhanced infrared absorption spectroscopy results and density functional theory calculations revealed that the local electron field of the substrate was redistributed via the synergistic effect of B and N, which not only modulated the electron density on Ru sites to decrease the OH− adsorption but also increased the flexibility of the water hydrogen network for aiding OH− transfer. Moreover, the charge redistribution increased the metal-support interaction, boosting the charge transfer efficiency in the HER process. As a result, Ru/BCN required an overpotential of only 17 mV to achieve 10 mA cm−2. This work provides a new insight into engineering charge redistribution to modulate the structures of interfacial water and key HER intermediates in alkaline HER.
The structures of interfacial water and key HER intermediates significantly influence the electrochemical performance of the HER.12–19 For the Volmer step, the water adsorbed on the active sites dissociates into H (Had) and OH (OHad) intermediates, which is the key step for the alkaline HER.20 During the process, the water adsorbed onto the catalyst's active site, as the initial reactant, is strongly affected by the hydrogen bond network of water. It tends to become more rigid in alkaline environments, hindering the reorientation of free water molecules.21,22 Moreover, it is more difficult to facilitate the transfer of OH− with the rigid hydrogen bond network of interfacial water because of the presence of large OH− transfer barriers.23 Thus, it is crucial to decrease the rigidity of the water network to ensure the increased availability of free water molecules, which in turn would facilitate the Volmer step. Water, being a typical polar molecule, features an oxygen atom with a partial negative charge and hydrogen atoms with partial positive charges. This causes a change in the H2O orientation under the influence of electrostatic interactions.24,25 These charges could also be readily influenced by electric fields, resulting in the alignment of the hydrogen atoms towards the negatively charged surfaces, which can disrupt the hydrogen bond network.26–30 A strategic approach to manipulate the local electronic field on the surface of the catalyst is to encourage water molecules to reorient towards the surface, thereby avoiding the disruption of the hydrogen bond network.
Doping is considered an efficient strategy for modulating the electron density redistribution of a catalyst.11,31,32 It can affect the intrinsic catalytic activity at the atomic level. For example, B is a promising dopant because it is electron-deficient and has only three valence electrons, which can easily polarize the surrounding atoms, resulting in electronic density changes in the HER catalysts.33–37 In contrast to B, N is more electronegative (3.04), drawing electrons from other low electronegative elements; thus, it is more stronger in introducing electron distribution after doping.38–43 Compared with single-atom doping, binary- or multiple-heteroatom doping is more versatile to modulate the electronic structure, which can be attributed to the coexisting active sites and synergistic effects. In particular, combining the opposing electronegativities of N and B (2.04) to that of carbon, carbon nanomaterials simultaneously doped with N and B could activate nearby carbon atoms through the synergistic coupling effect between N and B atoms, leading to an increased local electron density at the surface of carbon-based materials.44–47 However, researches on the effect of doping on reorganizing interface water is still lacking.
Herein, a doping strategy was applied to induce charge redistribution within the substrate, which changed the local electron field of the entire catalyst. Accordingly, we fabricated an HER catalyst, namely, B and N co-doped carbon matrix supported Ru nanoparticles (denoted as Ru/BCN). Benefiting from the synergistic effect of B and N, electron transfer was enhanced between Ru sites and BCN substrate. It not only modulated the electron density at the Ru site to decrease the desorption barrier of the intermediate OHad but also disrupted the rigid hydrogen bond networks of interfacial water to increase the migration of OH−. The as-synthesized Ru/BCN demonstrated an excellent alkaline HER performance and durability, which required an overpotential of only 17 mV to achieve 10 mA cm−2.
The morphological evolution from BOPs to Ru/BCN was analyzed using scanning electron microscopy (SEM). As shown in Fig. 1a and S1–S3,† all the three samples existed as regular nanorods with a loose layered sheet structure. This may be attributed to the presence of C–H⋯H–B and B–H⋯π intermolecular interactions between 4,4′-bipyridine and [B12H12]2− clusters, which can be assembled to form a layered structure and then stacked.48 After in situ reduction, the morphology did not change significantly but became looser, which was attributed to the partially breaking intermolecular interactions between 4,4′-bipyridine and [B12H12]2− during the reduction process. The morphology of the calcined sample Ru/BCN did not significantly change compared with Ru/BOPs. In contrast, Ru/CN without [B12H12]2− showed an irregular nanorod morphology. (Fig. S4†) TEM images further demonstrated a graphene-like nanosheet structure of Ru/BOPs and Ru/BCN (Fig. 1b, S3 and S5†). In contrast to Ru/BOPs, which showed no discernible Ru nanoparticles, Ru/BCN exhibited well-dispersed Ru nanoparticles with an average diameter of approximately 2.2 nm, which were uniformly distributed on the BCN nanosheet support. (Fig. S6†). High-resolution transmission electron microscopy (HRTEM) image revealed distinct lattice stripes on the nanoparticles, exhibiting a spacing of approximately 0.202 nm, which corresponded to the (101) crystal plane of hexagonal Ru. This indicated the high crystallinity of Ru NPs after the pyrolysis process (Fig. 1c). In addition, the high-angle annular dark-field scanning TEM (HAADF-STEM) energy dispersive X-ray (EDX) spectroscopy measurements proved that Ru, B, N and C were uniformly distributed in the catalyst (Fig. 1d).
FT-IR spectroscopy and X-ray powder diffraction (XRD) analyses were used to verify the morphological changes from BOPs to Ru/BCN. As illustrated in Fig. S7,† the characteristic peaks of Ru/BOPs were similar to those of BOPs, indicating that only a small portion of the closo-[B12H12]2− was consumed in the process of reducing Ru3+ ions. After calcination, the characteristic peaks of B–C (1415 cm−1), B–N (1252 cm−1), and C–N (1158 cm−1) appeared (Fig. 1e), suggesting the successful formation of the BCN structure.49 Based on the XRD patterns, aa new set of peaks was observed in Ru/BOPs, which were quite different from those observed in BOPs, suggesting a new crystal arrangement due to the partial Ru NP-occupied closo-[B12H12]2− sites (Fig. S9†). However, no XRD peaks of the Ru phase were found, likely owing to the low crystallinity or small size of Ru NPs in Ru/BOPs, which further verified the formation of metallic Ru as observed in the HRTEM image and XPS data (Fig. S10 and S11†). In contrast, the XRD peak appeared at 43.1° in Ru/BCN (Fig. 1f) corresponded to the (101) crystal plane of the hexagonal metal Ru (JCPDS 06-663). Compared with Ru/BCN, the corresponding peaks of Ru in Ru/CN were more clearly revealed, indicating a larger size of Ru particles. These results suggested that the synergistic effect of B and N induced a stronger metal-support interaction, which enabled Ru to be uniformly dispersed with a smaller particle size.
To gain further insight into the effect of additional B doping on the metal-support interaction and local charge distribution, Ru/BCN and Ru/CN were analyzed using X-ray photoelectron spectroscopy (XPS). The XPS survey spectra (Fig. S12 and S13†) revealed the presence of carbon and nitrogen elements in the Ru/BCN and Ru/CN samples, with an additional peak corresponding to the B element observed in the Ru/BCN spectra, indicating the successful doping of B atoms. In the high-resolution B 1s spectrum (Fig. 1g), the binding energy peaks can be deconvoluted into three components corresponding to B–N (190.5 eV), B–C (188.5 eV), and B–O (192.0 eV) bonds.8 By comparing with the spectrum of uncalcined sample Ru/BOPs, it can be further demonstrated that B was successfully doped into the carbon skeleton (Fig. S14†). The high solution of N 1s spectra of Ru/BCN revealed a negative shift for the peaks at 400.1 eV and 398.9 eV, which were assigned to pyrrolic-N and pyridinic-N. Compared with the corresponding spectrum of Ru/CN in Fig. 1h, this shift suggested that the introduction of B enabled N to gain more electrons, further promoting the charge distribution of the carbon substrate. Notably, an additional peak at 396.2 eV was emerged, which belonged to the N–B bond.32 In the high-resolution Ru 3p XPS spectra of Ru/BCN, the characteristic peaks at 462.1 eV (3p3/2) and 484.3 eV (3p1/2) were attributed to the shift of metal Ru to a higher binding energy than that of Ru/CN (Fig. 1i). This suggested that more electrons were transferred from Ru to the BCN substrate than to the CN substrate owing to the synergistic coupling effect between N and B. The electronic properties of Ru/BCN and Ru/CN were further investigated using differential charge density analysis. As shown in Fig. 1j and k, increased electron transfer from Ru to BCN (0.852 e−) relative to CN (0.694 e−), suggesting introduction of B promoting the redistribution of localized charge. Thus, as expected, the introduction of B atoms into the CN substrate promoted the charge distribution of BCN, leading to stronger metal-support interactions between Ru and BCN substrate, which further modulated the electron density of Ru, thereby enhancing the HER performance.
The HER catalytic activities of Ru/BCN and Ru/CN were evaluated in a 1 M KOH solution using a conventional three-electrode system. As illustrated in Fig. 2a, Ru/BCN exhibited the highest electrocatalytic activity. It attained a current density of 10 mA cm−2 with an overpotential of only 17 mV.
This was markedly better than that of Ru/BOPs (with an overpotential of 196 mV) and Ru/CN (with an overpotential of 76 mV), and it even surpassed that of commercial 20% Pt/C catalyst. Moreover, Ru/BCN achieved a significantly lower overpotential at the industrially relevant current density of 1 A cm−2, approximately 1.5 times lower than that of commercial Pt/C and half of that of Ru/CN (Fig. S15†). This suggested that the synergistic coupling effect between N and B atoms improved the HER performance. In addition, the Tafel slope of 43 mV dec−1 for Ru/BCN was smaller than that of Ru/BOPs (180 mV dec−1), Ru/CN (108 mV dec−1), and Pt/C (47 mV dec−1), suggesting that the Heyrovsky step (slow water dissociation), as the rate-determining step, was enhanced (Fig. 2b). Moreover, the activation energies confirmed the better HER performance of Ru/BCN than that of Ru/CN. As shown in Fig. 2c and S16,† the calculated activation energy of Ru/BCN was 8.74 kJ mol−1, which was much lower than that of Ru/CN (24.49 kJ mol−1), which suggested that B doping played a key role in the enhancement of the HER activity. As shown in Fig. S17,† the double-layer capacitance of Ru/BCN reached 33.5 mF cm−2, which was significantly larger than those of Ru/CN and Ru/BOPs, indicating that Ru/BCN was exposed to more active sites. The ECSA-normalized LSV curves of Ru/BCN and Ru/CN (Fig. 2d and S18†) revealed that the intrinsic activity of Ru/BCN at 100 mV was −2.75 mA cm−2ECSA, which was approximately four times higher than that of Ru/CN (−0.73 mA cm−2ECSA). This suggests that the higher activity of Ru/BCN not only came from a larger ECSA but also depended on the intrinsic activity of the catalytic site. To further examine the intrinsic HER activity, the turnover frequency (TOF) values of the catalysts were calculated. As shown in Fig S19,† Ru/BCN catalyst offered a superior TOF of 1.08 s−1 at −0.1 V vs. RHE, which was significantly higher than that of Ru/CN (0.27 s−1). Moreover, the faradaic efficiencies for both HER and OER were calculated by comparing the experimentally measured gas volumes with the theoretically predicted values using Ru/BCN as the cathode and RuO2 as the anode.50 As shown in Fig. S20, S21 and Table S1,† a good agreement was observed between the experimentally generated and theoretically calculated amounts of H2 to O2, which demonstrated a near 100% faradaic efficiency for both the half-reactions, with the consistently observed 2:
1 H2 to O2 ratio further confirming the stoichiometric water splitting behavior. Both the accelerated durability test and chronoamperometry experiment showed that the stability and durability of Ru/BCN were better than those of Ru/CN and Pt/C (Fig. 2e and f). The exceptional electrochemical stability of Ru/BCN stems from B–N synergistic electronic coupling that strengthens metal-support interactions, effectively suppressing both nanoparticle growth and interfacial detachment during prolonged operation.
To further understand the effect of B doping on the kinetics of the HER reaction, electrochemical impedance spectroscopy (EIS) measurements were carried out. The reduced semicircle diameters in the Nyquist plots of Ru/BCN demonstrate a significant decrease in interfacial charge-transfer resistance (Rct) following B incorporation (Fig. 2g). To further determine the effect of B doping on the reaction kinetics, Bode plots of Ru/BCN and Ru/CN were used to explain it through the variations in phase angle with frequency. The observed Bode plot region could be divided into two parts: a low-frequency region (light blue) and a high-frequency region (light pink). The high-frequency region represented the electronic conduction in the inner layer of the catalyst surface, while the low-frequency region indicated the predominance of charge transfer activity at the electrode–electrolyte interface.51,52 As shown in Fig. 2i, all the samples with high phase angles were located in the low-frequency region, suggesting that the resistance to electron transfer in the catalytic process was mainly dominated by the charge transfer process based on the electrode–electrolyte interface. The lower phase angle value of Ru/BCN in the high-frequency region indicated a rapid charge transfer process. In addition, in situ EIS measurements were performed to assess the kinetics of the electrode in the potential range from 0 mV to −30 mV. As shown in Fig. 2j–l, in the Bode plots of Ru/BCN and Ru/CN, the phase angles decreased when the potential became negative. In particular, the phase angle of Ru/BCN decreased more rapidly, further manifesting faster charge transfer. Therefore, doping B in the substrate can boost the charge transfer efficiency of the catalyst.
In addition, the zeta potential experiments confirmed that B doping modified the electronic structure of the CN substrate. As shown in Fig. 3a, the zeta potential values of both BCN and CN were negative. In particular, the value of BCN was significantly lower than that of CN, which manifested adsorption of more OH− on the surface of BCN than CN. This suggests that B doping enhanced the charge distribution. However, an opposite result was observed for Ru/BCN and Ru/CN catalysts. The zeta potential of Ru/BCN was more positive, which suggested the hindrance of OH− adsorption at Ru sites.53–55 These results further indicated that B coupling with N modulated the electronic structure of Ru and enhanced the charge redistribution of the catalyst to decrease OH− adsorption on the catalyst surface.
In alkaline media, water plays the dual role of a solvent and a reactant, which is critical for the Volmer step. As a polar molecule, water is highly sensitive to the local electric field on the catalyst surface. As previously discussed, B doping can effectively tune the electronic structure of the catalyst, consequently altering the interfacial water arrangement. To probe these structural changes, operando surface-enhanced infrared absorption spectroscopy (SEIRAS) was performed on Ru/BCN and Ru/CN in 0.1 M KOH under applied potentials ranging from 0.1 V to −0.2 V vs. RHE (Fig. 3c and d). In the IR spectra, a broad absorption peak at 3600–3000 cm−1 could be attributed to the O–H stretching vibration of water.56 The higher the frequency value, the greater the flexibility of the hydrogen bonding network of water. Typically, the characteristic bands located near 3600, 3400, 3200, and 2900 cm−1 were attributed to free water (K·H2O), trihedral water (2HB·H2O), tetrahedral water (4HB·H2O), and strongly hydrogen-bonded water (chemically adsorbed water), respectively.57,58 For Ru/CN, a broad peak was be observed between 2750 to 3700 cm−1, which could be fitted with three different components, namely, 4HB·H2O (3143 cm−1), 2HB·H2O (3463 cm−1) and K·H2O (3605 cm−1), using Gaussian fitting. This revealed that the surface of Ru/CN was predominantly covered by tetrahedral water, manifesting the rigid hydrogen bonding network of the interfacial water. In contrast, the O–H stretching peak of Ru/BCN shifted to a higher wave number at 3354 cm−1 upon B incorporation, suggesting an enhanced flexibility of the hydrogen bonding network of the interfacial water. Gaussian fitting further confirmed the dominance of 2HB·H2O. The introduction of B modified the surface charge distribution of the catalyst, disrupting the original hydrogen-bonding network of interfacial water and increasing its structural flexibility.
To elucidate the role of B in the catalytic HER process, density functional theory (DFT) calculations were systematically performed in Ru/BCN and Ru/CN (Fig. S22 and S23†). Fig. 3e and S24–S28† revealed adsorption energy of water (EH2O*) on different active sites. Compared with the B and N sites, the Ru site showed a relatively stronger adsorption energy of water, suggesting that the Ru site preferred water adsorption. Similar results were confirmed for Ru/CN. The oxophilicity of the catalysts also had an important effect on facilitating the HER in alkaline media.32 The adsorption energy of OHad on various active sites was calculated (Fig. S29–S33†). It was found that OHad inclined to adsorb at the Ru sites rather than at B and C sites (Fig. 3f). In contrast to Ru/CN, the adsorption energy of OHad became higher (−0.54 eV) after B doping, indicating OHad desorbed easily from the surface of Ru in Ru/BCN, which agreed well with the zeta potentials results. Based on these results, the Ru sites, as active sites, were selected to probe into the reaction pathways underlying the alkaline HER. As shown in Fig. 3g, although water dissociation on Ru/CN proceeded exothermically, subsequent reaction steps exhibited endothermic characteristics. Crucially, the desorption of OHad species encountered a significant energy barrier (ΔG = 0.75 eV), creating a kinetic bottleneck that severely impeded the active site regeneration. In contrast, the energy barrier of desorption of OHad was greatly decreased to 0.49 eV in Ru/BCN, indicating that OHad desorbed more easily on Ru sites after doping B in the CN substrate. Therefore, introducing B was beneficial for boosting OH− desorption, which was crucial for enhancing the HER in these systems. Furthermore, the formed H2 desorbed more easily from the surface of Ru/BCN than from Ru/CN. In addition to lowering the energy barrier for OHad desorption, the flexible water hydrogen network induced by B doping aided in the release of the adsorbed hydroxyl on the Ru site via a proton exchange route involving nearby H2O molecules to transfer OH− into the bulk phase.59
Based on these findings, the excellent HER performance of Ru/BCN can be attributed to three aspects: (1) enhanced exposure of catalytically active Ru sites through uniform dispersion of ultrasmall nanoparticles on well-ordered nanorod substrates with hierarchical porosity; (2) redistribution of local electron density through synergistic effect of B and N weakened the hydrogen bonding networks, facilitating hydroxyl desorption from active sites; and (3) optimized electronic structure of Ru through strong metal-support interactions fine-tuned the intermediate adsorption energy while accelerating charge transfer kinetics.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta03786g |
This journal is © The Royal Society of Chemistry 2025 |