Jia Wu,
Jianlong Zhang,
Zhixiang Zhai,
Zelong Sun,
Zhangyue Zheng,
Junxin Chen,
Zihui Ning,
Huan Wen* and
Shibin Yin*
Guangxi Key Laboratory of Electrochemical Energy Materials, School of Chemistry and Chemical Engineering, Guangxi University, 100 Daxue Road, Nanning 530004, China. E-mail: wenhuan@gxu.edu.cn; yinshibin@gxu.edu.cn
First published on 31st July 2025
Efficient electrooxidation of 5-hydroxymethylfurfural (HMF) is essential for the production of valuable chemicals from biomass. However, the sluggish interfacial proton transfer kinetics during the reaction significantly hinder its progress. This study proposes that the introduction of Brønsted bases on cobalt-based catalysts can enhance HMF oxidation by modulating interfacial proton transfer kinetics. Density functional theory calculations, in situ spectroscopy, and experimental results demonstrate that phosphate (Pi) groups on the cobalt surface shorten the distance between the proton donor and acceptor, effectively promoting interfacial proton transfer during the dehydrogenation of HMF. Consequently, the Pi-functionalized catalyst shows a 6.5-fold increase in current density compared with the unmodified cobalt catalyst and achieves near 100% selectivity for 2,5-furandicarboxylic acid, attaining a current density of 1000 mA cm−2 at 1.41 VRHE for efficient HMFOR. This work provides insights into designing efficient catalysts for industrial applications through surface site functionalization.
HMFOR occurs at the catalyst surface or interface through a dehydrogenation process involving the activation and break of C–H/O–H bonds and associated electron–proton transfer processes.17–19 Specifically, the catalyst (acting as a proton acceptor) facilitates proton transfer from the reactant molecules (proton donor) to the product side via the transition state, with associated electron transfer stabilizing the process,20 as illustrated in Fig. 1a. Owing to the significant kinetic disparity in transport between electrons and protons, electron transfer is generally faster than proton transfer, making the latter the rate-determining step (RDS).18,19 Consequently, understanding and optimizing the interfacial proton transfer kinetics is crucial for developing highly efficient HMFOR catalysts.
The kinetic core of the proton transfer process fundamentally depends on the overlap of the vibrational wave functions of the proton donor and acceptor, which is a quantum mechanical phenomenon that controls the probability of proton tunneling.21,22 The increase in the binding strength between the proton donor and acceptor leads to a shortness of the distance, which significantly increases the wave function overlap and enhances the proton tunneling effect, thereby reducing the proton transfer energy barrier and promoting the break of the chemical bond.22 According to Brønsted–Lowry acid–base theory, a Brønsted base can act as an effective proton acceptor by combining with proton to form conjugate acid, which is a process governed by its pKa value.23 Among various Brønsted base groups, the phosphate (PO43−, denoted as Pi) group, with its high pKa value,24,25 is particularly attractive for promoting rapid proton transfer kinetics during HMFOR processes.
Inspired by this idea, this work proposes a strategy to regulate the interfacial proton transfer kinetics over cobalt nanowires through Brønsted base (Pi) functionalization (Fig. 1b). Density functional theory (DFT) simulations and experimental approaches confirm that the functionalized Pi group provides additional equivalent hydrogen sites that act as effective carriers for proton transfer, shortening the distance between the proton donor (HMF) and acceptor (Pi-Co) and increasing the overlap of their proton vibration wave functions. This facilitates proton transfer from the reaction molecules to the catalyst surface, reducing the energy barrier for the dehydrogenation of the RDS and thereby improving the kinetics and selectivity of HMFOR. Consequently, the Pi-functionalized catalyst shows a 6.5-fold increase in current density compared with the unmodified cobalt catalyst and achieves near 100% selectivity for FDCA, attaining a current density of 1000 mA cm−2 at 1.41 VRHE for efficient HMFOR.
X-ray photoelectron spectroscopy (XPS) is performed to study the chemical properties of the samples.26 The XPS survey spectrum confirms the presence of Co, P, and O in Pi-Co (Fig. S4). The Co 2p spectra are deconvoluted into two peaks corresponding to Co0 (778.0/793.4 eV) and Co2+ (781.1/796.5 eV) (Fig. 2g), respectively, and the oxidation state is attributed to the inevitable surface oxidation.10,27 Compared with that of the Co sample, the intensity of the Co0 peak is lower because of the introduction of Pi groups on the Co surfaces. The appearance of a P 2p3/2 peak (133.6 eV) and a P 2p1/2 peak (134.4 eV) in the P 2p spectra corresponds to PO43− (Fig. S5a).28 Notably, the PO43− ligand has a strong coordination effect and can form P–O–M bonds with metal sites, which is confirmed by the P–O–Co linkage (531.2 eV) of O 1s spectra (Fig. S5b) and the P–O–Co vibrations (1055 cm−1) of Fourier transform infrared spectroscopy (FTIR) (Fig. S6).29,30 Furthermore, the chemical composition of the catalyst at different depths is examined via XPS in conjunction with Ar+ sputtering. Fig. 2h–j shows that the peak intensity of Co0 in the Co 2p spectra notably increases with an Ar+ sputtering depth of 5 nm. Conversely, the peak intensities of PO43− and P–O–Co linkages in the P 2p and O 1s spectra almost disappear. By combining the above results, it can be inferred that the Pi groups are successfully functionalized on the Co nanowires surface.
The electrocatalytic study of HMFOR is then performed in a three-electrode electrolyzer with 1.0 M KOH/50 mM HMF. Compared with the Pi-Co and Co samples, the contribution of bare NF to the electrocatalytic activity is negligible and will not be discussed later (Fig. 3a). Pi-Co catalyst only needs a small potential of 1.08/1.16/1.41 VRHE to reach a current density of 10/100/1000 mA cm−2, respectively, which is significantly lower than that of the Co sample (1.17/1.36/1.65 VRHE) and competitive with the reported catalysts (Fig. 3b and Table S1).31–38 Notably, the current density of Pi-Co is 6.5 times that of Co catalyst at 1.40 VRHE (Fig. 3c), confirming that the functionalization of Pi is beneficial for HMFOR. Additionally, the HMFOR LSV curves are normalized by the electrochemical surface area (ECSA), which is determined by the double-layer capacitance (Cdl) method, and the results indicate the better intrinsic activity of Pi-Co than that of Co sample (Fig. S7).
During the electrooxidation process of HMF, the OER emerges as the primary competitive reaction. Therefore, a comparative analysis of the HMFOR and OER processes is conducted. Fig. S8 shows that the addition of 50 mM HMF results in a substantially lower onset potential and a higher current density compared to 1.0 M KOH, demonstrating that the oxidation of HMF is dominant, while OER occurs at more positive potentials than 1.45 VRHE. Subsequently, the HMF conversion tests are conducted in the potential of 1.25 to 1.45 VRHE over the Pi-Co catalyst. The results indicate that 1.35 VRHE is appropriate for electrooxidizing HMF to FDCA since it simultaneously achieves high FDCA selectivity, formation rate, and energy conservation (Fig. S9 and S10). After electrolysis for 3600 s, the Pi-Co sample exhibits greater FDCA selectivity, faradaic efficiency (FE), and formation rate (98.8%, 98.5%, and 0.0274 μmol s−1) than Co sample (16.1%, 50.3% and 0.00387 μmol s−1) (Fig. 3d and S11). This suggests that the introduction of Pi leads to a significant improvement in the reaction rate and FDCA selectivity. The electrolyte could be acidified to generate a precipitate and isolate FDCA (Fig. S12), allowing for effective product separation. Additionally, Pi-Co exhibits good conversion rates, product selectivity, and FE for furfuryl alcohol (FFA) and furfural (FF) (Fig. S13), respectively, demonstrating its potential application prospects in biomass oxidation.
To elucidate the underlying reaction mechanisms, the pathways on a Pi-Co catalyst are investigated (Fig. S14 and S15). Fig. 3e and S16 illustrate that the concentration of HMFCA is consistently greater than that of 2,5-diformylfuran (DFF) throughout the electrolysis process, suggesting that the oxidation of HMF on the Pi-Co catalyst predominantly follows the HMFCA pathway. To verify this distinction, in situ attenuated total reflectance FTIR (ATR-FTIR) spectroscopy is employed to identify the reaction intermediates (Fig. 3f).39 As the potential increases, the intensities of the bands at 1460, 1507, and 1558 cm−1 gradually decrease, whereas the intensities of the bands at 1358, 1544, 1627, and 1734 cm−1 increase significantly. This trend demonstrates the consumption of HMF and the simultaneous generation of 5-hydroxymethyl-2-furan carboxylic acid (HMFCA) and FDCA, confirming that the HMFOR is carried out through the HMFCA pathway.
To investigate the practical application of the Pi-Co catalyst, an anion exchange membrane (AEM)-based electrolyzer is assembled for the oxidation of HMF (Fig. 3g and S17). In this setup, Pi-Co and NF are employed as the anode and cathode respectively. For an electrolyzer with an active area of 1 cm2 (1 × 1), the onset potential for the overall water splitting is measured at 1.50 V in a 1.0 M KOH solution at 30 °C. After adding 50 mM HMF, current densities of 500 and 1000 mA cm−2 are attained at potentials of 1.53 and 1.79 V (Fig. S18), respectively. Notably, when the temperature is elevated to 80 °C, the 1 cm2 device achieves a remarkable current density of 2000 mA cm−2 at a potential of only 1.72 V (Fig. 3h). Additionally, the system operates stably for 70 h at a current density of 200 mA cm−2 while maintaining high FDCA selectivity (>90%) (Fig. S19), showing promising industrial application potential.
To gain insight into the structural evolution of catalysts, in situ Raman spectroscopy is performed at different potentials. Fig S20 reveals that both Pi-Co and Co catalysts exhibit a Co3+–O peak at 503 cm−1 within the potential range of 1.05–1.45 VRHE,40,41 indicating the formation of CoOOH during the HMFOR. The structural integrity of the Pi-Co catalyst after HMF conversion tests is also investigated (Fig. S21). SEM images show that the nanowire structure is well preserved, and its surface becomes rough (Fig. S22). XPS characterization in conjunction with Ar+ sputtering indicate that the Co0 peak on the surface of the Pi-Co catalyst disappears, and a new Co3+ peak emerges,42,43 while the P–O peak remains prominent. When the sputtering depth reaches 5 nm, the Co0 peak reappears, and the P–O peak almost vanishes (Fig. S23). These results indicate that the metallic Co on the surface undergoes reconstruction, while the Pi component remains basically stable in the surface, and the internal metallic Co still exists in the catalyst.
Then, to verify the effects of CoOOH on the catalytic activity of HMFOR, multi-potential step experiments are employed. The experimental procedure is as follows: first, a potential of 1.40 VRHE is applied in 1.0 M KOH to generate Co3+ on the catalyst surface. Subsequently, HMF solution is added to the system under open-circuit potential (OCP) conditions. Finally, a reduction potential of 0.6 VRHE is applied, and the change in reduction current density is monitored. Fig. S2 reveals that the chemical reaction between the activated catalyst and HMF proceeds relatively slowly below 1.40 VRHE (Fig. S24(a)). In addition, electrochemical tests of the Co sample after activation show that the formation of CoOOH on the Co sample slightly improves the catalytic activity and FDCA selectivity (Fig. S24(b)). These results collectively demonstrate that CoOOH is not the key active species promoting the efficient conversion of HMF to FDCA and thus is not further discussed below.
To elucidate the internal reasons for the increase in HMFOR activity, the deuterium kinetic isotope effects (KIEs) are investigated.44,45 When the KIEs value is greater than 1.5, it is generally regarded as clear evidence that proton transfer is involved in or affects the RDS.18 The KIEs experiments are performed in a 1.0 M NaOD/10 mM HMF D2O solution (Fig. S25). The Co sample has an H/D effect of 3.36–4.42 (Fig. 4a), indicating that the RDS of HMFOR over the Co sample involves C–H/O–H bond cleavage.46 For the Pi-Co sample, the H/D effect (1.96–3.07) becomes weaker (Fig. 4b). The overlap of proton vibration wave functions plays an important role in determining the rate of proton transfer reactions and KIEs, and the degree of this overlap strongly depends on the distance between the proton donor and acceptor.18 The smaller KIEs value of Pi-Co indicates that the Pi groups near the catalytic site shorten the proton donor–acceptor distance, thereby promoting the break of C–H/O–H bonds and improving the kinetics and selectivity of HMFOR.
When the RDS of HMFOR involves proton transfer, the Brønsted base in the solution can act as a proton acceptor to reduce the energy barrier of the reaction, affecting the overall reaction kinetics.47,48 The relationships between the catalytic activities of the Pi-Co and Co samples and the concentration of the additional base (K2CO3) are studied. As shown in Fig. 4c and d, the current density of the Co sample increases with K2CO3 concentration increasing, whereas the activity of Pi-Co is nearly unaffected. In addition, the additional base order of Pi-Co (0.04) is significantly smaller than that of the Co sample (0.57) (Fig. 4c and d inset), indicating that the proton transfer process in the additional base solution is inhibited. This is because the proton affinity of the phosphate group is greater than the carbonate group, which effectively offsets the deprotonation of K2CO3 in the electrolyte and greatly improves the efficiency of proton transfer on the catalyst surface. These results demonstrate that the Pi groups, as efficient carriers of proton transfer, accelerate the proton transfer kinetics of HMFOR.
To gain a deeper understanding of the promotional effect of Pi on proton transfer, an electrochemical proton inventory study is conducted. This method explores the number of exchangeable hydrogen sites and their impact on proton transfer kinetics by measuring the reaction rate as a function of D2O concentration in a D2O/H2O mixed solvent (Fig. S26).49,50 Fig. 4e reveals that the Pi-Co sample exhibits a distinct nonlinear arc pattern with a large Z value (1.78) and a small ϕ value (0.22). This finding indicates that the Pi groups improve the efficiency of proton transfer by providing additional equivalent hydrogen sites, which is closely related to the polymer reverse isotope effect in the RDS. In contrast, the Co sample without the Pi groups has a smaller Z value (0.85) and a slightly larger ϕ value (0.24) (Fig. 4f) and has a negligible effect on the proton transfer kinetics. These findings suggest that Pi groups affect proton transfer in the RDS through their equivalent hydrogen sites, thereby improving the proton transfer efficiency of the Co catalytic center.
The proton transfer energy barrier in the step of HMF dehydrogenation on Co and Pi-Co surfaces is further investigated via DFT simulations (Fig. 4g). For the Co sample, the hydrogen proton in the HMF is transferred to the Co site. However, after the introduction of the Pi groups, the hydrogen proton is preferentially transferred to the O position of Pi. This leads to a reduction in the proton donor–acceptor distance from 5.86 Å for Co to 4.56 Å for Pi-Co. Consequently, the energy barrier for proton transfer from the HMF molecule to the surfaces of Co and Pi-Co is significantly decreased from 2.70 eV to 1.75 eV (Fig. 4h). This confirms that the introduction of Pi shortens the distance between the proton donor and acceptor and promotes the proton transfer kinetics during the dehydrogenation of HMF.
The specific reaction process of HMFOR is studied to reveal the RDS step of the dehydrogenation of HMF to FDCA (Fig. 5a). The calculated results indicate that the RDS of both the Pi-Co and Co samples are the dehydrogenation of HMF* to HMFCA* (Fig. 5b). For the Co sample, the Gibbs free energy for the formation of HMFCA* is 0.65 eV. The presence of Pi on the Co surface significantly decreases the Gibbs free energy for HMFCA* (0.35 eV), suggesting that introducing Pi on the Co surface is more energetically favorable for the dehydrogenation process. To elucidate the mechanism underlying this phenomenon, the adsorption energies of key reactant molecules are calculated. The results reveal that the adsorption energy of HMF* on Pi-Co (−2.29 eV) is significantly negative than that on Co (−1.61 eV) (Fig. S27), demonstrating that Pi functionalization effectively enhances the adsorption of HMF. Furthermore, crystal orbital Hamilton population (COHP)51 calculations show that the –ICOHP value of the C–H bond on the Co sample (6.13) is more positive than that on Pi-Co (5.46) (Fig. 5c), suggesting a relatively weak C–H strength of HMF* on Pi-Co. These results indicate that the introduction of Pi promotes the adsorption of the reaction molecules and the break of the C–H bonds in the RDS process, which is beneficial for improving the kinetics and selectivity of HMFOR.
The oxidation rate constants of HMF, HMFCA, and FFCA over the catalysts are examined to validate the RDS (Fig. S28).52 Fig. 5d shows that the rate constant of HMF is significantly lower than those of HMFCA and FFCA, confirming that the RDS is the conversion of HMF to HMFCA. Notably, the HMF, HMFCA, and FFCA rates of Pi-Co are 5.33, 4.43, and 2.84 times greater than Co sample, respectively, further demonstrating that the introduction of Pi groups improves the reaction rates and optimizes the RDS during HMFOR.
Furthermore, to obtain additional evidence for the vital role of Pi during HMF oxidation, zinc ions (Zn2+) are introduced into the solution to block the surface Pi groups through strong interactions between the Zn2+ and PO43− sites (Fig. 5e).53 Fig. 5f and the inset reveal a notable decrease in the current density and an increase in the Tafel slope after the addition of Zn2+. KIEs experiments show that the current density of the Pi-Co blocked with Zn2+ decreases, and the H/D effect increases to 3.48–4.02 (Fig. 5g and S29), demonstrating that interfacial proton transfer is inhibited when surface Pi sites are blocked. HMF conversion tests indicate that the FDCA selectivity, FE, and formation rate significantly decrease to 24.0%, 66.5%, and 0.00691 μmol s−1 (Fig. 5h and S30), respectively. These findings provide further evidence for the crucial role of surface Pi species in promoting HMF oxidation.
Supplementary information is available. See DOI: https://doi.org/10.1039/d5sc04035c.
This journal is © The Royal Society of Chemistry 2025 |