In situ surface reconstruction of silver leads to competent activity for the electrocatalytic hydrogenation of 5-hydroxymethylfurfural

Dalong Qu a, Nan Jiang *a and Yujie Sun *b
aDepartment of Chemistry and Chemical Engineering, Guizhou University, Guiyang 550025, Guizhou, China. E-mail: njiang@gzu.edu.cn
bDepartmt of Chemistry, University of Cincinnati, Cincinnati 45221, Ohio, USA. E-mail: yujie.sun@uc.edu

Received 7th April 2025 , Accepted 3rd June 2025

First published on 10th June 2025


Abstract

Electrochemical reconstruction is a well-documented phenomenon in electrocatalysis, yet it has rarely been purposefully harnessed to enhance catalyst performance in a controllable manner. In this study, we introduce a rational electrochemical reconstructing strategy using Ag as a model electrocatalyst and the electrocatalytic hydrogenation of 5-hydroxymethylfurfural (HMF) to 2,5-bis(hydroxymethyl)furan (BHMF) as a benchmark reaction. Through a simple yet effective electrochemical reconstruction approach, we generated preferentially oriented Ag nanoparticles (PO-Ag NPs) with more exposure of the Ag(110) facets, which exhibited significantly enhanced activity and selectivity compared to pristine Ag. The faradaic efficiency of PO-Ag NPs for HMF hydrogenation reached above 95% with a selectivity exceeding 98%. Comprehensive electrochemical, spectroscopic, and computational investigations identify the preferential adsorption of HMF on the Ag(110) facet as the primary factor contributing to the enhanced performance. Furthermore, PO-Ag NPs demonstrated excellent long-term stability, retaining high activity and selectivity over extended electrolysis cycles. These findings highlight electrochemical reconstruction as a powerful and controllable strategy for designing highly efficient electrocatalysts, with broad implications for a wide range of electrochemical transformations beyond hydrogenation.


image file: d5ta02755a-p1.tif

Nan Jiang

Nan Jiang earned her doctorate (2018) in Chemistry from Utah State University, United States. She then joined the University of Nevada, Las Vegas (UNLV) and Advanced Light Source (ALS) at Lawrence Berkeley National Laboratory (LBL) as a postdoctoral fellow. Currently, she is an Assistant Professor at Guizhou University. Her current research focuses on designing and developing electrode materials for various energy storage and conversion devices. She has published over 40 peer-reviewed papers in international journals. She was awarded by High-Level Talents Returning to China Program in 2024 and was included on the “World's Top 2% Scientists” list for 2023 by Mendeley Data.

Introduction

To reduce reliance on fossil resources, biomass upgrading using renewable energy has emerged as an appealing alternative for the sustainable production of fuels and green chemicals.1,2 5-Hydroxymethylfurfural (HMF), a versatile platform molecule derived from biomass, can be efficiently obtained through the hydrolysis of cellulose with high yields.3,4 Given the presence of aldehyde and alcohol functional groups anchored to the furan ring, HMF can be further transformed into a variety of commodity chemicals through processes, such as oxidation,5 hydrogenation,6–8 hydrogenolysis,9 reductive amination,10,11 open-ring reaction,12 cleavage,13 and esterification.14 Among these upgraded products, 2,5-bis(hydroxymethyl)furan (BHMF), produced by hydrogenating the aldehyde group of HMF, plays a vital role in manufacturing polyurethane foams, polyesters, resins, artificial fibers, and pharmaceuticals.7,15 To date, the industrial production of BHMF still depends on conventional thermocatalytic methods, which demand noble metal catalysts, polluting organic solvents, and harsh operating conditions including high H2 pressures (28–350 bar) and temperatures (403–423 K),15 raising concerns regarding safety, energy consumption, and environmental issues.

Due to its environmentally friendly and carbon-neutral nature, the electrocatalytic hydrogenation of HMF, conducted under ambient conditions with H2O as the proton source, has gained significant attention.9,15–20 The electrocatalytic hydrogenation of HMF towards BHMF can proceed through either a direct electroreduction pathway or an adsorbed hydrogen (H*) route (Scheme 1).21,22 In the direct electroreduction pathway (blue line), the carbonyl group (C[double bond, length as m-dash]O) in HMF acquires one H+/e pair via proton-coupled electron transfer (PCET), leading to the formation of a radical intermediate (C*–OH). The generated C*–OH can then either react with another H+/e pair to produce BHMF or undergo dimerization with another C*–OH radical through C–C coupling (red line in Scheme 1), resulting in the production of bis(hydroxymethyl)hydrofuroin (BHH). In particular, dimerization tends to occur at high HMF concentrations and very negative potential, which favor the over-accumulation of C*–OH and limit the PCET process.17 In contrast, the H* route generates in situ H* on the surface of the electrode via the electrochemical reduction of H+ or H2O (Volmer step). The subsequent hydrogenation of the adjacent adsorbed HMF* with H* follows the Langmuir–Hinshelwood (L–H) mechanism, yielding BHMF (green line in Scheme 1). However, the H2 evolution reaction (HER), which consumes H* via Tafel or Heyrovsky steps, has been considered as the major competing reaction, thereby lowering the faradaic efficiency (FE). This issue is particularly pronounced for platinum-group metals, which possess low hydrogen activation barriers. As a result, the two routes may compete with each other. The preference highly depends on the relative potentials for the formation of H* and C*–OH.


image file: d5ta02755a-s1.tif
Scheme 1 Schematic illustration of the electrocatalytic hydrogenation of HMF towards BHMF through two pathways.

Owing to their exceptional catalytic ability in the electrocatalytic hydrogenation of HMF towards BHMF and relatively low HER activity, Ag-based electrocatalysts have garnered considerable research attention.23–25 Previous studies have demonstrated the outstanding electrocatalytic hydrogenation performance of Ag foil over other metal foil compositions in slightly alkaline electrolytes.26,27 Combined experimental and computational investigations elucidate that the moderate adsorption of HMF on Ag foil leads to the selective hydrogenation of HMF to BHMF.28 Moreover, HMF hydrogenation on Ag-based electrodes is more favorable than the HER, offering a wide potential window and a high FE. To further boost BHMF productivity and selectivity, Ag-based bimetallic catalysts and Ag nanocomposites with synergistic effects have been developed. For instance, AgCu bimetallic electrocatalysts demonstrate superior catalytic performance for the hydrogenation of HMF to BHMF, primarily due to the stronger adsorption of HMF on Cu and a synergistic effect between Ag and Cu.29 Additionally, Li et al. designed a composite catalyst by immobilizing Ag nanoparticles (Ag NPs) on SnO2 sheets, promoting the electrocatalytic HMF hydrogenation to BHMF with a high FE.17 The oxygen vacancies on SnO2 function as electrophilic sites to selectively adsorb the carbonyl group in HMF, which then reacts with the adsorbed H* species generated by Ag NPs to produce BHMF. However, it is highly desired but remains a grand challenge to achieve efficient BHMF formation with high HMF concentrations due to the tendency for dimerization and the severe competition from the HER at various negative potentials.

It is widely recognized that the crystal facet effect strongly modulates the chemisorption of active reactants, hence controlling the reaction pathways and product selectivity/distribution. In 2021, Choi and co-workers investigated the orientation dependence of the electrocatalytic HMF reduction on Ag single-crystal surfaces using density functional theory (DFT) calculations.28 Their findings indicate that adsorbed HMF molecules exhibit the strongest binding affinity to the Ag(110) facet, with decreasing binding strength on the Ag(100) and Ag(111) facets. However, to date, it has been challenging to experimentally modify the arrangement of surface Ag atoms to generate preferentially oriented Ag nanoparticles (PO-Ag NPs) as highly active electrocatalysts for HMF hydrogenation. Herein, we employ a facile in situ electrochemical oxidation–reduction method to rearrange the surface Ag atoms of Ag foil, exposing crystal orientations dominated by the Ag(110) facet and significantly enhancing the performance of electrocatalytic hydrogenation of HMF to BHMF. Multiple characterization techniques and electrochemical measurements have confirmed the successful reconstruction of the Ag foil surface to generate a higher proportion of the Ag(110) facet. The resulting PO-Ag NPs show a high FE and a wide potential window for BHMF production, even at high HMF concentrations. Operando characterization experiments together with DFT calculations reveal that the stronger adsorption of HMF on Ag(110) is the key factor for the enhanced electrocatalytic performance of PO-Ag NPs.

Experimental section

Chemicals and materials

Ag foil (thickness: 0.25 mm, 99.99%) and single-crystal surfaces of Ag(111), Ag(110), and Ag(100) were purchased from Hefei Kejing Materials Technology Co., Ltd. 5-Hydroxymethylfurfural (HMF, 99.5%), boric acid (H3BO3, 99.9%), 2,5-bishydroxymethylfuran (BHMF, >99.5%), potassium hydroxide (KOH, 99.99%), and potassium bicarbonate (KHCO3, 99.5%) were obtained from Shanghai Aladdin Biochemical Technology Co., Ltd and used without any further purification.

Synthesis of PO-Ag NPs

Ag foil (0.5 cm × 0.5 cm) was employed as the working electrode, with a Ag/AgCl (saturated KCl) electrode as the reference electrode and a carbon rod as the counter electrode. Linear sweep voltammetry (LSV) was performed on Ag foil within a potential range from −0.06 to −1.26 V vs. RHE at a scan rate of 5 mV s−1 for 5 cycles in a 0.5 M KHCO3 solution. Subsequently, the potential was set at 8.74 V vs. RHE for 5 seconds to produce a uniform layer of Ag2CO3 on the surface of Ag foil. Finally, PO-Ag NPs were synthesized by conducting LSV within the same potential range (–0.06 to −1.26 V vs. RHE, 5 mV s−1) for 5 cycles, followed by applying a negative potential of −1.76 V vs. RHE for 120 seconds in a fresh 0.5 M KHCO3 solution.

Material characterization studies

Scanning electron microscopy (SEM) and elemental mapping tests were performed using a Thermo Scientific Quattro S instrument equipped with an EDAX ELECT PIUS energy-dispersive X-ray spectroscopy (EDS) system. PO-Ag NPs were detached from Ag foil and further characterized by transmission electron microscopy (TEM) using a FEI-Talos F200S microscope operating at 200 kV. X-ray diffraction (XRD) patterns were recorded on a Bruker D8 Advance diffractometer with Cu Kα radiation. X-ray photoelectron spectroscopy (XPS) measurements were conducted using a Thermo Fisher Scientific K-Alpha spectrometer. The C 1s peak was calibrated to 284.8 eV to ensure accurate alignment of all XPS data.

Operando Raman measurements

Operando Raman tests were conducted using a LabRAM HR Raman microscope (Horiba Jobin Yvon, HR550) equipped with a 532 nm laser for excitation, a Synapse CCD detector, a monochromator with a 1800 grooves per mm grating, and a water immersion objective (Olympus LUMFL, 60×, numerical aperture = 1.10). Each spectrum was acquired over a period of 60 seconds. The electrochemical Raman study was carried out in a custom-made cell, using a Ag/AgCl electrode as the reference electrode and a Pt wire as the counter electrode. The potential was precisely controlled with an Ivium-n-Stat electrochemical workstation.

Quasi in situ electron paramagnetic resonance spectroscopy

Electron paramagnetic resonance (EPR) spectra were acquired using a Bruker EMXplus (298 K, 9.843483 GHz) with a sweep width of 100.0 G and a sweep time of 60 seconds. The EPR measurements were performed in an H-type cell, with a Ag/AgCl electrode functioning as the reference electrode and a carbon rod serving as the counter electrode. After electrolysis at –0.51 V vs. RHE for 3 min, the sample was collected and subjected to EPR tests with 5,5-dimethyl-1-pyrroline-N-oxide (DMPO, 50 mM) as the trapping agent.

Electrochemical measurements

Electrochemical measurements were conducted using a BioLogic SP-150e potentiostat. Both HER and HMF hydrogenation experiments were carried out in an H-type cell (with 15 mL capacity per chamber), separated by a Nafion 211 membrane. In these experiments, Ag foil (0.5 cm × 0.5 cm) and PO-Ag NPs (0.5 cm × 0.5 cm) were used as working electrodes, while a carbon rod worked as the counter electrode and a Ag/AgCl electrode acted as the reference electrode. The electrolyte consisted of a N2-saturated 1 M borate buffer solution (pH = 9.2, 10 mL). All HER and HMF hydrogenation tests were carried out under magnetic stirring and continuous N2 bubbling. Unless otherwise specified, LSV measurements and electrolysis reported in this work were corrected with 85% iR-correction. Operando electrochemical impedance spectroscopy (EIS) tests were conducted at –0.41 V vs. RHE in a frequency range from 0.1 to 100[thin space (1/6-em)]000 Hz. Potential values were converted to the reversible hydrogen electrode (RHE) scale using the equation:
 
E(V vs. RHE) = E(V vs. Ag/AgCl) + 0.059 × pH + 0.197(1)

The overpotentials of the HER and HMF hydrogenation were calculated using the following equations:25

 
Overpotential of HER = 0 − EHER(V vs. RHE)(2)
 
Overpotential of HMF hydrogenation = 0.12 − EHMF(V vs. RHE)(3)
where EHER and EHMF represent the applied potentials for the HER and HMF hydrogenation, respectively.

For rotating disk electrode (RDE) measurements, PO-Ag NPs or Ag foil was fixed onto a glassy carbon electrode with an exposed area of 0.5 cm × 0.5 cm, which served as the working electrode. LSV tests were performed at rotating rates ranging from 1000 rpm to 2000 rpm in the presence of 50 mM HMF. The electron transfer number (n) for HMF hydrogenation was determined from the slope of the Koutecký–Levich plot, defined by the equation:

 
Slope = (0.62nFD2/3ν−1/6C)−1(4)
where F is the Faraday constant (96[thin space (1/6-em)]500 C mol−1), D is the diffusion coefficient of HMF in water (9.169 × 10−6 cm2 s−1), ν is the kinematic viscosity of water (0.01 cm2 s−1), and C is the concentration of HMF (50 mM).24

Product analysis

The liquid samples were initially diluted 25 times or 50 times and then quantitatively analyzed by high-performance liquid chromatography (HPLC, Essentia LC-16) equipped with an ultraviolet photometric detector at 225 nm. HMF and its reductive products were determined using a Shim-pack GIST C18 column (150 mm and 5 μm) maintained at 45 °C. The analysis utilized a gradient elution method: starting with a 15[thin space (1/6-em)]:[thin space (1/6-em)]85 ratio of CH3CN to H2O for 4 minutes, followed by a 2-minute gradient elution that increased the CH3CN ratio to 40% and concluding with a 4-minute gradient elution that reduced the CH3CN ratio back to 15%. In addition, BHH was identified using liquid chromatography mass spectrometry (LC-MS, Agilent 6545B), which was equipped with a 1290 Infinity liquid chromatography system. The flow rate was set at 0.5 mL min−1, and the detection wavelength was 225 nm. The FE, selectivity, and productivity of BHMF were calculated with the following equations:
 
image file: d5ta02755a-t1.tif(5)
 
image file: d5ta02755a-t2.tif(6)
 
image file: d5ta02755a-t3.tif(7)

The FE of BHH was calculated by the following equation, as it is not commercially available.30

 
image file: d5ta02755a-t4.tif(8)

Computational method

DFT calculations were performed using the generalized gradient approximation (GGA) with the Perdew–Burke–Ernzerhof (PBE) functional to calculate the adsorption energy of substrate on the catalyst.31,32 The ionic cores were described using projected augmented wave (PAW) potentials, while the valence electrons were treated with a plane wave basis set, applying a kinetic energy cutoff of 450 eV.33 The convergence criteria for electronic energy and force were set to 10−6 eV and 0.03 eV Å−1, respectively. A supercell was constructed for the face-centered cubic Ag(100), Ag(110), and Ag(111) facets as the support structure. To minimize the interactions between repeating images, a vacuum layer of at least 20 Å was introduced. During geometry optimization, the bottom half layers of the Ag substrate were fixed, while the upper half was allowed to relax. The Brillouin zone was sampled with a 2 × 2 × 1 Monkhorst–Pack k-point mesh. Dispersion interactions were incorporated into the calculations by using the DFT-D3 correlation correction. Adsorption energies (Eads) were calculated using the following equation:
 
Eads = EtotalEAg-subEHMF(9)
where Etotal, EAg-sub, and EHMF represent the total energies of the optimized HMF/Ag-substrate system, the Ag-substrate, and the HMF molecule, respectively.

Results and discussion

Catalyst synthesis and characterization

PO-Ag NPs were synthesized using an in situ electrochemical oxidation–reduction approach, as illustrated in Fig. 1a. Initially, Ag foil was immersed in a 0.5 M KHCO3 solution and subjected to a potential of 8.74 V vs. RHE for 5 seconds. This rapid and intense anodization oxidizes the surface Ag atoms of Ag foil (AD-Ag), forming a layer of monoclinic Ag2CO3. During this process, the crystallographic structure is disrupted, and Ag atoms are selectively removed from the Ag(111) plane due to anisotropic etching and crystallographic strain. In the subsequent reduction step, monoclinic Ag2CO3 transformed back into face-centered cubic (fcc) metallic Ag, accompanied by the accumulation of crystal defects, which results in the preferential exposure of the Ag(110) facet. SEM and XRD were used to characterize Ag foil, AD-Ag, and PO-Ag NPs. Ag foil shows a flat surface (Fig. 1b), and its XRD pattern (Fig. 1e) reveals five peaks corresponding to fcc metallic Ag. Following the electrochemical oxidation of the surface Ag to Ag2CO3, a stacked cube structure emerged (Fig. 1c), along with the formation of interparticle pores, which is confirmed as the monoclinic phase of Ag2CO3 by the XRD pattern (Fig. 1f). Elemental mapping images of AD-Ag (Fig. S1) demonstrate a homogeneous distribution of C, O, and Ag, indicating the formation of a uniform layer of Ag2CO3. As revealed by XRD (Fig. 1g), monoclinic Ag2CO3 is converted back to metallic Ag during the reduction process. Notably, the Ag(110) and Ag(100) facets are not visible in the XRD pattern due to forbidden reflections. However, their existence can be reasonably inferred proportionally to the Ag(220) and Ag(200) peaks, respectively, since these planes are aligned parallel to one another in a fcc structure. Thus, the increased intensity of the Ag(220) peak in PO-Ag NPs implies more exposure of the Ag(110) facet. The resulting PO-Ag NPs exhibit a coral-like nanocolumn morphology (Fig. 1d) with a thickness of approximately 2.5 μm (inset in Fig. 1d), leading to a higher electrochemically active surface area (ECSA) than that of the pristine Ag foil (Fig. S2).
image file: d5ta02755a-f1.tif
Fig. 1 (a) Schematic illustration of the synthesis process for PO-Ag NPs. SEM images of (b) Ag foil, (c) AD-Ag, and (d) PO-Ag NPs (inset: cross-section of PO-Ag NPs). XRD patterns of (e) Ag foil, (f) AD-Ag, and (g) PO-Ag NPs.

The detailed surface structure of PO-Ag NPs was further investigated by high-resolution transmission electron microscopy (HRTEM), XPS, and HCO3 adsorption/desorption experiments. According to the HRTEM images (Fig. 2a), PO-Ag NPs consist of interconnected nanoparticles with an average diameter of ∼60 nm. These interconnected nanoparticles create numerous grain boundaries (Fig. S3), which contain uncoordinated Ag atoms that can facilitate the adsorption of both H2O and HMF molecules, thereby promoting HMF hydrogenation.34 Additionally, clear lattice fringes (Fig. 2a) with interplanar spacings of 0.24, 0.20, and 0.15 nm are observed, corresponding to the (111), (200), and (220) planes of Ag, respectively. Elemental mapping images of PO-Ag NPs (Fig. S4) display a decrease in O content due to the reduction of Ag2CO3 to metallic Ag (Tables S1 and S2). Moreover, the presence of C and O confirms the residual CO32− in PO-Ag NPs.


image file: d5ta02755a-f2.tif
Fig. 2 (a) HRTEM images of PO-Ag NPs. (b) CV curve of PO-Ag NPs in 0.5 M KHCO3 saturated with N2 at a scan rate of 50 mV s−1 (inset: the corresponding facet ratio obtained from HCO3 adsorption/desorption experiments). XPS spectra of (c) Ag 3d and (d) C 1s.

To precisely quantify the Ag(110) facet on the surface of PO-Ag NPs, HCO3 adsorption/desorption tests were conducted.35 As shown in the cyclic voltammetry (CV) curve (Fig. 2b), the peaks observed around −0.65 to −0.75 V vs. SHE, −0.45 to −0.55 V vs. SHE, and −0.25 to −0.35 V vs. SHE can be attributed to the electro-adsorption/desorption of HCO3 on Ag(110), Ag(100), and Ag(111) facets, respectively. The ratio of Ag(110), Ag(100), and Ag(111) facets calculated from the integral peak area is 3.7[thin space (1/6-em)]:[thin space (1/6-em)]1.4[thin space (1/6-em)]:[thin space (1/6-em)]1, indicating a predominance of Ag(110) over Ag(100) and Ag(111). In contrast, Ag foil exhibits only one peak around −0.25 to −0.35 V vs. SHE (Fig. S5), confirming the dominance of the Ag(111) facet.

XPS analysis was performed to examine the electronic structures of Ag foil, AD-Ag, and PO-Ag NPs (Fig. S6). The high-resolution XPS spectra of Ag 3d (Fig. 2c) reveal that both Ag foil and PO-Ag NPs are primarily in the metallic state of Ag0, indicating the reduction of AD-Ag to metallic Ag during electroreduction. Notably, the binding energy of PO-Ag NPs shifts to a lower value compared to that of Ag foil, indicating an upshift of the d-band center, which could facilitate the selective adsorption and activation of the C[double bond, length as m-dash]O group in HMF, thereby enhancing the catalytic performance.36 Additionally, the lower binding energy of PO-Ag NPs implies a higher electron density on the surface, which may be ascribed to the charge transfer from the residual CO32− to Ag (Fig. 2d). This increased electron density could lead to stronger adsorption of the C[double bond, length as m-dash]O group in HMF through efficient back donation, thus activating the C[double bond, length as m-dash]O group and improving the catalytic activity.37

Electrocatalytic tests

To evaluate the performance of the electrocatalytic hydrogenation of HMF on PO-Ag NPs, LSV was carried out in a 1 M borate buffer solution (pH = 9.2) containing 50 mM HMF. As shown in Fig. 3a, both Ag foil and PO-Ag NPs demonstrate significantly enhanced electrocatalytic activity for HMF hydrogenation compared to the HER, implying more favorable thermodynamics for HMF hydrogenation. In particular, PO-Ag NPs exhibit the most pronounced shifts in both the onset potential and current density for HMF hydrogenation compared to Ag foil. PO-Ag NPs reach a current density of −32.5 mA cm−2 at −0.45 V vs. RHE, which is 6.5 fold higher than that of Ag foil (Fig. S7). More importantly, the electrocatalytic activities of Ag foil and PO-Ag NPs for the HER are comparable (Fig. 3a and b). The slightly higher current density of PO-Ag NPs right after the onset potential might stem from its larger surface area (Fig. S2). Therefore, while the crystal facet effect has a very limited promoting effect on the HER, it significantly boosts HMF hydrogenation, establishing a wide potential window (>300 mV) between the HER and HMF hydrogenation.
image file: d5ta02755a-f3.tif
Fig. 3 (a) LSV curves for Ag foil and PO-Ag NPs in 1 M borate buffer solution (pH = 9.2) with and without 50 mM HMF saturated with N2 at a scan rate of 5 mV s−1. (b) Tafel slopes for the HER and HMF hydrogenation over Ag foil and PO-Ag NPs. (c) FE and selectivity of BHMF for 50 mM and 100 mM HMF hydrogenation at −0.51 V vs. RHE using Ag foil and PO-Ag NPs as working electrodes. (d) FE and selectivity of BHMF for 50 mM HMF hydrogenation at −0.51 V vs. RHE using single-crystal Ag(111), Ag(110), and Ag(100) as working electrodes.

Subsequently, Tafel plots were utilized to assess the kinetics of the HER and the electrocatalytic HMF hydrogenation (Fig. 3b). The Tafel slopes of Ag foil and PO-Ag NPs for the HER are 116.4 and 112.9 mV dec−1, respectively, which illustrate that the rate-determining step (RDS) is the Volmer step for both electrodes. When 50 mM HMF was introduced into the electrolyte, the Tafel slope of PO-Ag NPs decreased slightly to 103.0 mV dec−1, suggesting more favorable kinetics for the electrocatalytic HMF hydrogenation compared to the HER. Similar Tafel slopes for the HER and HMF hydrogenation on PO-Ag NPs indicate that the introduction of HMF holds a minimal impact on the reaction kinetics; thus, the RDS likely remains as the Volmer step during HMF hydrogenation. In contrast, Ag foil presented a considerably larger Tafel slope (175.5 mV dec−1) upon the addition of HMF, demonstrating sluggish kinetics and a change in the RDS. The distinct kinetic behaviors observed for HMF hydrogenation between Ag foil and PO-Ag NPs presumably imply different reaction mechanisms. Additionally, operando EIS was conducted at –0.41 V vs. RHE in the presence of 50 mM HMF. Fig. S8 presents the Nyquist plots for the HMF hydrogenation over Ag foil and PO-Ag NPs. The semicircular portion of the plots corresponds to the charge transfer resistance (Rct). Compared to Ag foil, PO-Ag NPs exhibit a significantly lower Rct value for HMF hydrogenation, indicating faster charge transport kinetics on PO-Ag NPs.

To quantitatively identify the reductive products, chronoamperometry experiments were performed over Ag foil and PO-Ag NPs at −0.51 V vs. RHE in an H-type cell with a fixed passing charge equivalent to converting half of the HMF to BHMF with a 100% FE. The liquid products and the remaining HMF were analyzed by HPLC (Fig. S9 and S10). Additionally, BHH was identified through LC-MS (Fig. S11). Fig. 3c illustrates that PO-Ag NPs exhibit significantly higher FEs for the hydrogenation of 50 mM and 100 mM HMF (94.0 and 91.2%, respectively) compared to Ag foil (76.8 and 68.4%, respectively). Notably, the selectivity of BHMF on PO-Ag NPs is nearly 1.5 times and 2 times larger than that on Ag foil for 50 mM and 100 mM HMF hydrogenation, respectively. To further demonstrate the facet effect, single-crystal Ag(111), Ag(110), and Ag(100) were employed directly as working electrodes for HMF hydrogenation. As depicted in Fig. S12a, single-crystal Ag(111), Ag(110), and Ag(100) demonstrated comparable overpotentials for the HER. The Tafel slopes of single-crystal Ag(111), Ag(110), and Ag(100) for the HER are 204.7, 210.5, and 222.1 mV dec−1, respectively (Fig. S12b), which illustrate that the RDS is the Volmer step for three single-crystal electrodes, and Ag(111) has the fastest kinetics for the HER. Regarding HMF hydrogenation (Fig S12c and d), Ag(110) displays a lower overpotential and Tafel slope than those of Ag(100) and Ag(111), indicating more favorable thermodynamics and kinetics for HMF hydrogenation on Ag(110). Furthermore, controlled potential electrolysis (Fig. 3d) reveals that the FEs and selectivity of the three single-crystal electrodes follow the order: Ag(110) > Ag(100) > Ag(111). This result confirms that the facet effect plays a crucial role in promoting HMF hydrogenation. Interestingly, even the single-crystal Ag(110) shows lower FE and selectivity for BHMF compared to PO-Ag NPs, indicating that additional factors such as morphology and the presence of residual CO32− may also contribute to the observed performance.

To investigate the potential dependence of electrocatalytic HMF hydrogenation, chronoamperometry experiments were carried out at various applied potentials ranging from −0.41 V to −0.71 V vs. RHE, with a fixed passing charge equivalent to converting half of the HMF to BHMF with a 100% FE. For 50 mM HMF hydrogenation catalyzed by PO-Ag NPs, only two reduction products (BHMF and BHH) were detected under the investigated conditions, whereas multiple species were observed for Ag foil (Fig. S13). As shown in Fig. 4a and S14, PO-Ag NPs exhibit higher FE and selectivity for BHMF compared to Ag foil across the investigated potentials. Importantly, PO-Ag NPs displayed enhanced performance, achieving a near-100% FE for BHMF at −0.46 V vs. RHE. Upon further increasing the applied potentials, the production of BHH decreased and eventually disappeared beyond −0.56 V vs. RHE, leading to a near-unity selectivity for BHMF. Specifically, PO-Ag NPs achieved a productivity of 0.872 mmol cm−2 h−1 at −0.51 V vs. RHE, which was 23 times higher than that of Ag foil (0.037 mmol cm−2 h−1). Although the selectivity of BHMF increased as the applied potential became more negative, a continuous decline in the FE of BHMF was observed due to the competition of the HER at higher potentials. These results demonstrate that PO-Ag NPs not only realize outstanding performance in the electrocatalytic hydrogenation of HMF but also effectively suppress the HER at relatively low potentials. We also performed chronoamperometry tests over PO-Ag NPs with a higher HMF concentration of 100 mM at different potentials (Fig. 4b). At relatively low potentials, an increase in HMF concentration led to a decrease in the selectivity for BHMF owing to the adequate supply of the C*–OH intermediate for dimerization. At high potentials, PO-Ag NPs achieved remarkable selectivity for BHMF compared to those at lower potentials. This can be attributed to the sufficient H* produced on the catalyst surface at higher potentials. As a consequence, the competing HER becomes more severe. Compared to the reported studies (Table S3), PO-Ag NPs achieve superior catalytic activity for electrocatalytic HMF hydrogenation. Moreover, to demonstrate the scalability of this approach, HMF hydrogenation was conducted using a larger electrode (1 cm × 1 cm) with a proportionally increased electrolyte volume (40 mL) at various potentials. Within the operating potential range, both the FE and the selectivity for BHMF slightly increase compared to the smaller electrode (0.5 cm × 0.5 cm) (Fig. S14c). After 10 reaction cycles at −0.51 V vs. RHE, the FE, selectivity, and current density remain essentially unchanged (Fig. S14d and e). These results prove that PO-Ag NPs exhibit both scalability and outstanding durability.


image file: d5ta02755a-f4.tif
Fig. 4 (a) FE and selectivity of BHMF for 50 mM HMF hydrogenation on PO-Ag NPs at various potentials. (b) FE and selectivity of BHMF for 100 mM HMF hydrogenation on PO-Ag NPs at various potentials. (c) Recycling test of PO-Ag NPs for 50 mM HMF hydrogenation at −0.51 V vs. RHE. (d) Conversion of HMF and yield of BHMF during the chronoamperometry experiment conducted at −0.51 V vs. RHE in 1 M borate buffer containing 50 mM HMF.

To investigate the robustness of PO-Ag NPs, 10 consecutive electrolysis cycles were performed at −0.51 V vs. RHE. Even after 10 reaction cycles, the FE, selectivity, and catalytic current density remained essentially unchanged (Fig. 4c and S14f), indicating the excellent durability of PO-Ag NPs. Post-electrolysis characterization experiments further confirmed the robustness of PO-Ag NPs. SEM images and elemental mappings (Fig. S15) indicate no significant changes in the nanostructure of PO-Ag NPs after electrolysis. The XRD pattern (Fig. S16a) proves that the post-electrolysis PO-Ag NPs still maintain a fcc Ag structure. The XPS spectrum of the Ag 3d region for the post-electrolysis PO-Ag NPs shows a similar binding energy to that of the fresh sample (Fig. S16b), suggesting that the electronic structure of PO-Ag NPs remains unchanged after electrolysis. Fig. S17 presents the CV curve and the corresponding facet ratio obtained from HCO3 adsorption/desorption experiments using post-electrolysis PO-Ag NPs as the working electrode. The calculated ratios of Ag(110), Ag(100), and Ag(111) were 55.75%, 31.58%, and 12.67%, respectively. Thus, Ag(110) still remains dominant after the reaction. To further explore the charge-dependence of BHMF production, we monitored the concentrations of BHMF and HMF at −0.51 V vs. RHE. It was calculated that ∼96.5 C of charge would be required to completely convert 50 mM HMF, assuming a 100% FE for BHMF production. As shown in Fig. 4d, after passing 96.5 C of charge, the yield of BHMF reaches approximately 43.8 mM, with 5.8 mM of HMF remaining, resulting in a selectivity of 99.1% for BHMF.

Mechanism studies

To elucidate the mechanisms of HMF hydrogenation, the kinetic isotopic effect (KIE) was utilized to gain insights into whether the dissociation of H2O to generate H* is the RDS. LSV measurements for the HER and HMF hydrogenation were conducted on both Ag foil and PO-Ag NPs in H/D isotopically substituted environments. In the case of the HER, a substantial negative shift is observed in the LSV curve for both Ag foil and PO-Ag NPs (Fig. 5a and b), implying more sluggish kinetics for the D2 evolution reaction (DER) compared to the HER. This suggests that the dissociation of H2O to produce H* is indeed the rate-limiting step. For HMF hydrogenation, PO-Ag NPs display a similar KIE to that of the HER, confirming the crucial role of the formation of H* in the reaction mechanism. In sharp contrast, Ag foil shows only a 51 mV cathodic shift at 10 mA cm−2, which is extremely smaller than a typical KIE. This reduced KIE indicates a fundamentally different mechanism for HMF hydrogenation over Ag foil, where the generation of H* is no longer the RDS. Furthermore, Koutecký–Levich analysis (Fig. S18) demonstrates that the electron transfer number (n) of HMF hydrogenation catalyzed by PO-Ag NPs is 1.84, which is close to 2. This result supports a two-electron L–H mechanism, leading to nearly 100% selectivity for BHMF production. For Ag foil, the n value is 1.39, implying a dominant one-electron reaction pathway and accounting for the increased production of BHH.
image file: d5ta02755a-f5.tif
Fig. 5 LSV curves of (a) Ag foil and (b) PO-Ag NPs in pH = 9.2 or pD = 9.2 electrolyte with and without 50 mM HMF. (c) EPR spectra of Ag foil and PO-Ag NPs in pure 1 M borate buffer (top) and in the presence of 50 mM HMF (bottom) at −0.51 V vs. RHE using DMPO as the spin-trapping agent. (d) EPR spectra of Ag foil and PO-Ag NPs in 1.0 M borate buffer containing 50 mM HMF at −0.51 V vs. RHE using DMPO as the spin-trapping agent.

To gain a better understanding of the reaction process, quasi in situ EPR measurements were conducted to detect reaction intermediates using DMPO as the trapping agent, that is, DMPO–H* and DMPO–C*–OH (Fig. S19).16,17,38 As displayed in Fig. 5c (top part), H* intermediates are detected in both Ag foil and PO-Ag NPs systems with similar signal intensities in the pure electrolyte, suggesting that the facet effect has a minimal influence on the generation of H* (Volmer step). This observation aligns with the LSV measurements, which present comparable HER catalytic performance between Ag foil and PO-Ag NPs. Upon the addition of HMF, the intensity of the H* signal decreased significantly for both Ag foil and PO-Ag NPs (Fig. 5c bottom part), likely due to the competition from adsorbed HMF molecules or the consumption of H* for HMF hydrogenation. Notably, the H* signal for PO-Ag NPs is slightly weaker than that for Ag foil, implying stronger adsorption of HMF on PO-Ag NPs or more consumption of H* for HMF hydrogenation, consistent with the superior catalytic performance of PO-Ag NPs. In addition, C*–OH intermediates were observed for both Ag foil and PO-Ag NPs (Fig. 5d). It is worth mentioning that C*–OH can either dimerize via C–C coupling to produce BHH or hydrogenate with H* or a pair of H+/e to yield BHMF. Previous studies have demonstrated that a low concentration of C*–OH on the electrode surface is less favorable for dimerization. Instead, they are more likely to react with H* or another H+/e pair to produce BHMF.39 As shown in Fig. 5d, PO-Ag NPs present a lower concentration of C*–OH compared to that of Ag foil, which contributes to the higher selectivity for BHMF and the lower yield of BHH.

The HMF hydrogenation process was further evaluated using operando Raman spectroscopy in the presence of 50 mM HMF. Fig. 6a and b display operando Raman signals for Ag foil and PO-Ag NPs in a custom-made cell, respectively. At the open-circuit potential (OCP), both Ag foil and PO-Ag NPs exhibited distinct vibration peaks at 1522 cm−1 and 1665 cm−1, corresponding to the C[double bond, length as m-dash]C symmetric stretching vibration in the furan ring and the C[double bond, length as m-dash]O vibration of the adsorbed HMF molecules, respectively. Additionally, peaks at 1397 cm−1 and 1367 cm−1 were assigned to the C–H scissoring vibration of the furan ring.30,40 Upon applying a slight negative potential of −0.06 V vs. RHE, the ν(C[double bond, length as m-dash]O) on the PO-Ag NPs displayed a noticeable red shift to a lower wavenumber, specifically shifting from 1665 cm−1 to 1640 cm−1. Similar red shifts of approximately 20 cm−1 were observed for the ν(C[double bond, length as m-dash]C) and ν(C–H) peaks on the PO-Ag NPs compared to those on Ag foil. These shifts are attributed to the enhanced adsorption of HMF on the PO-Ag NPs under operating conditions. Notably, the intensity of the HMF vibration peaks on the PO-Ag NPs is stronger than that on Ag foil. Moreover, a new peak at 1560 cm−1 appeared, which was assigned to the asymmetric stretching vibration of the C[double bond, length as m-dash]C bond in the furan ring. These findings suggest stronger HMF adsorption on the PO-Ag NPs. Regarding Ag foil, the intensity of the HMF vibration peaks gradually decreased as the potential increases, while a more pronounced C[double bond, length as m-dash]C vibration peak (1559 cm−1) attributed to BHMF emerged at −0.46 V vs. RHE. In stark contrast, a significantly enhanced HMF signal on the PO-Ag NPs was observed, and the BHMF signal appeared at a relatively low potential of −0.26 V vs. RHE. These observations indicate that PO-Ag NPs promote HMF adsorption under the given operating conditions, which has a positive effect on HMF hydrogenation.


image file: d5ta02755a-f6.tif
Fig. 6 Operando Raman spectra of the electrocatalytic hydrogenation of HMF on (a) Ag foil and (b) PO-Ag NPs in 1.0 M borate buffer containing 50 mM HMF at different potentials. (c) The most stable configurations of HMF adsorbed on the Ag(111), Ag(110), and Ag(100) surfaces.

To elucidate the remarkable HMF hydrogenation performance of PO-Ag NPs, DFT calculations using the PBE functional were conducted to investigate the adsorption of HMF on three Ag facets: (111), (110), and (100). Fig. 6c illustrates the most stable adsorption configuration of HMF on these surfaces. The adsorption energy of HMF via the aldehyde oxygen atom on the three Ag facets follows this order: Ag(110) > Ag(100) > Ag(111), which aligns with the findings from operando Raman analysis, suggesting stronger interactions between HMF and PO-Ag NPs. Additionally, the C[double bond, length as m-dash]O bond length of HMF adsorbed on Ag(110) is 1.251 Å (Table S4), which is longer than those on Ag(100) and Ag(111). This elongation of the C[double bond, length as m-dash]O bond on Ag(110) indicates a weakened C[double bond, length as m-dash]O bond, potentially facilitating the hydrogenation of HMF to BHMF.

Conclusions

In summary, we successfully synthesized preferentially oriented Ag nanoparticles, predominantly featuring the Ag(110) facet, using a simple electrochemical oxidation–reduction method. These PO-Ag NPs enable the selective hydrogenation of HMF to BHMF at high HMF concentrations and large current densities, achieving superior FE and selectivity across a wide potential range. By combining operando Raman spectroscopy and DFT calculations, we found that the stronger interaction between the C[double bond, length as m-dash]O group in HMF and PO-Ag NPs promotes the electrocatalytic hydrogenation of HMF to BHMF. Furthermore, electrochemical measurements and EPR analysis indicate that the hydrogenation of HMF on PO-Ag NPs primarily follows the L–H mechanism. This study elucidates the facet effect of Ag in the electrochemical conversion of HMF and provides fundamental insights for designing efficient Ag-based electrocatalysts.

Data availability

The data supporting the results of this study have been included as part of the ESI, and additional data are available from the corresponding authors upon reasonable request.

Author contributions

D. Q. synthesized the PO-Ag NPs catalyst, conducted the electrochemical measurements, and drafted the original manuscript under the guidance of N. J. and Y. S. N. J. conceptualized the project and designed the experiments. N. J. and Y. S. reviewed and revised the final manuscript. All authors discussed the content and reviewed the manuscript.

Conflicts of interest

There are no conflicts to declare.

Acknowledgements

This work was supported by the National Natural Science Foundation of China (No. 52362028) and Guizhou Provincial Basic Research Program (Natural Science) (No. [2023]042). NJ acknowledges the support from the Natural Science Research Project of the Education Department of Guizhou Province (No. QJJ[2022]001).

References

  1. C. Li, X. Zhao, A. Wang, G. W. Huber and T. Zhang, Chem. Rev., 2015, 115, 11559–11624 CrossRef CAS PubMed.
  2. Y. Jing, Y. Guo, Q. Xia, X. Liu and Y. Wang, Chem, 2019, 5, 2520–2546 CAS.
  3. J. J. Bozell and G. R. Petersen, Green Chem., 2010, 12, 539–554 RSC.
  4. C. O. Tuck, Science, 2012, 338, 604 CAS.
  5. N. Jiang, B. You, R. Boonstra, I. M. T. Rodriguez and Y. Sun, ACS Energy Lett., 2016, 1, 386–390 CrossRef CAS.
  6. S. Li, X. Sun, Z. Yao, X. Zhong, Y. Coo, Y. Liang, Z. Wei, S. Deng, G. Zhuang, X. Li and J. Wang, Adv. Funct. Mater., 2019, 29, 1904780 CrossRef CAS.
  7. X.-L. Liu, W. Zhong, Y.-F. Jin, T.-J. Wang, X. Xiao, P. Chen, Y. Chen and X. Ai, Chin. J. Catal., 2025, 69, 241–248 CrossRef.
  8. Z. Zhao, R. Yu, S. Wang, L. Guo, L. Zhang, M. Chen, J. Chen and H. Zheng, ChemSusChem, 2025, 18, e202401278 CrossRef CAS PubMed.
  9. P. Hauke, T. Merzdorf, M. Klingenhof and P. Strasser, Nat. Commun., 2023, 14, 4708 CrossRef CAS PubMed.
  10. H. Liu, W. Tang, D. Xue, J. Xiao and C. Wang, Catal. Sci. Technol., 2024, 14, 5764–5769 RSC.
  11. M. Zhang, S. Xu, M. Boubeche, D. Decarolis, Y. Huang, B. Liu, E. K. Gibson, X. Li, Y. Wang, H. Luo, C. R. A. Catlow and K. Yan, Green Chem., 2022, 24, 9570–9578 RSC.
  12. J. J. Roylance and K.-S. Choi, Green Chem., 2016, 18, 2956–2960 RSC.
  13. Y.-R. Zhang, B.-X. Wang, L. Qin, Q. Li and Y.-M. Fan, Green Chem., 2019, 21, 1108–1113 RSC.
  14. M. Balakrishnan, E. R. Sacia and A. T. Bell, Green Chem., 2012, 14, 1626–1634 RSC.
  15. M. G. Sendeku, K. Harrath, F. T. Dajan, B. Wu, S. Hussain, N. Gao, X. Zhan, Y. Yang, Z. Wang, C. Chen, W. Liu, F. Wang, H. Duan and X. Sun, Nat. Commun., 2024, 15, 5174 CrossRef CAS PubMed.
  16. K. Ji, M. Xu, S.-M. Xu, Y. Wang, R. Ge, X. Hu, X. Sun and H. Duan, Angew. Chem., Int. Ed., 2022, 61, e202209849 CrossRef CAS PubMed.
  17. X. Guo, H. Fu, J. Yang, L. Luo, H. Zhou, M. Xu, X. Kong, M. Shao, H. Duan and Z. Li, ACS Catal., 2023, 13, 13528–13539 CrossRef CAS.
  18. B. Chen, Y. Feng, S. Ma, W. Xie, G. Yan, Z. Li, J. Sperry, S. Yang, X. Tang, Y. Sun, L. Lin and X. Zeng, J. Energy Chem., 2023, 76, 421–428 CrossRef CAS.
  19. C. Tian, J. Yu, D. Zhou, H. Ze, H. Liu, Y. Chen, R. Xia, P. Ou, W. Ni, K. Xie and E. H. Sargent, Adv. Mater., 2024, 36, 2312778 CrossRef CAS PubMed.
  20. X. Yuan, K. Lee, J. R. Schmidt, K.-S. Choi and J. Am, Chem. Soc., 2023, 145, 20473–20484 CrossRef CAS PubMed.
  21. J. J. Roylance, T. W. Kim and K.-S. Choi, ACS Catal., 2016, 6, 1840–1847 CrossRef CAS.
  22. X.-Q. Pan, X.-Y. Zhang, G.-X. Huang, S.-C. Mei, J.-W. Huang, J.-J. Chen, W.-J. Liu and H.-Q. Yu, Appl. Catal., B, 2023, 323, 122191 CrossRef CAS.
  23. G. S. de Luna, P. H. Ho, A. Sacco, S. Hernandez, J.-J. Velasco-Velez, F. Ospitali, A. Paglianti, S. Albonetti, G. Fornasari and P. Benito, ACS Appl. Mater. Interfaces, 2021, 13, 23675–23688 CrossRef PubMed.
  24. X. H. Chadderdon, D. J. Chadderdon, T. Pfennig, B. H. Shanks and W. Li, Green Chem., 2019, 21, 6210–6219 RSC.
  25. H. Liu, T.-H. Lee, Y. Chen, E. W. Cochran and W. Li, Green Chem., 2021, 23, 5056–5063 RSC.
  26. Y. Kwon, Y. Y. Birdja, S. Raoufmoghaddam and M. T. M. Koper, ChemSusChem, 2015, 8, 1745–1751 CrossRef CAS PubMed.
  27. Y. Kwon, E. de Jong, S. Raoufmoghaddam and M. T. M. Koper, ChemSusChem, 2013, 6, 1659–1667 CrossRef CAS PubMed.
  28. D. K. Lee, S. R. Kubota, A. N. Janes, M. T. Bender, J. Woo, J. R. Schmidt and K.-S. Choi, ChemSusChem, 2021, 14, 4563–4572 CrossRef CAS PubMed.
  29. G. S. de Luna, P. H. Ho, A. Lolli, F. Ospitali, S. Albonetti, G. Fornasari and P. Benito, ChemElectroChem, 2020, 7, 1238–1247 CrossRef.
  30. X. Cao, Y. Ding, D. Chen, W. Ye, W. Yang and L. Sun, J. Am. Chem. Soc., 2024, 146, 25125–25136 CrossRef CAS PubMed.
  31. J. P. Perdew, K. Burke and M. Ernzerhof, Phys. Rev. Lett., 1997, 77, 3865–3868 CrossRef PubMed.
  32. K. Mathew, V. S. C. Kolluru, S. Mula, S. N. Steinmann and R. G. Hennig, J. Chem. Phys., 2019, 151, 234101 CrossRef PubMed.
  33. P. E. Blöchl, Phys. Rev. B: Condens. Matter Mater. Phys., 1994, 50, 17953–17979 CrossRef PubMed.
  34. Y.-C. Chen, Y.-Z. Chen, C.-M. Wang and Y.-K. Hsu, Int. J. Hydrogen Energy, 2023, 48, 3316–3327 CrossRef CAS.
  35. N. Hoshi, M. Kato and Y. Hori, J. Electroanal. Chem., 1997, 440, 283–286 CrossRef CAS.
  36. K. Mariandry, R. Kokate, S. V. Somerville, L. Gloag, S. Cheong, L. R. Carroll, P. V. Kumar, J. J. Gooding and R. D. Tilley, Chem. Mater., 2023, 35, 8636–8644 CrossRef CAS.
  37. Y. Ye, H. Yang, J. Qian, H. Su, K.-J. Lee, T. Cheng, H. Xiao, J. Yano, W. A. Goddard and E. J. Crumlin, Nat. Commun., 2019, 10, 1875 CrossRef PubMed.
  38. S. Han, Y. Shi, C. Wang, C. Liu and B. Zhang, Cell Rep. Phys. Sci., 2021, 2, 100337 CrossRef CAS.
  39. X. H. Chadderdon, D. J. Chadderdon, J. E. Matthiesen, Y. Qiu, J. M. Carraher, J.-P. Tessonnier and W. Li, J. Am. Chem. Soc., 2017, 139, 14120–14128 CrossRef CAS PubMed.
  40. M. Li, T. Zheng, D. Lu, S. Dai, X. Chen, X. Pan, D. Dong, R. Weng, G. Xu and F. Wang, J. Energy Chem., 2023, 84, 101–111 CrossRef CAS.

Footnote

Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d5ta02755a

This journal is © The Royal Society of Chemistry 2025
Click here to see how this site uses Cookies. View our privacy policy here.