Yuting
Zhang
,
Junhua
Kuang
,
Jia
Yu
*,
Yangyang
Dong
,
Jiaran
Li
,
Tianwei
Xue
,
Jing
Wu
,
Junchi
Ma
,
Jinlong
Wan
,
Shiping
Zeng
,
Yong
Sun
,
Yue-Jiao
Zhang
,
Jin-Chao
Dong
,
Li
Peng
*,
Shuliang
Yang
* and
Jian-Feng
Li
*
College of Energy, College of Chemistry and Chemical Engineering, Xiamen University, Xiamen 361102, Fujian, PR China. E-mail: yujia@xmu.edu.cn; li.peng@xmu.edu.cn; ysl@xmu.edu.cn; li@xmu.edu.cn
First published on 12th September 2024
The electrochemical conversion of 5-hydroxymethylfurfural (HMF) into 2,5-furandicarboxylic acid (FDCA), powered by renewable electricity, provides an efficient strategy for upgrading biomass to produce high value-added products. In this work, sulfur-doped hollow metal–organic framework (MOF) nanotubes were synthesized in a scalable and controllable manner by using hydrogen-bonded organic frameworks (HOFs) as the template and precursor. Interestingly, sulfur doping of the catalysts could be simply achieved by replacing nickel nitrate with nickel sulfate at the synthesis stage, and the designed sulfur-doped catalysts demonstrated excellent electrocatalytic activity in the HMFOR and remarkable durability, achieving nearly 100% conversion of HMF and high FDCA faradaic efficiency. Moreover, the HMFOR activity could be further enhanced by additional sulfur doping of the prepared catalyst, while maintaining its unique tubular morphology. The combination of sulfur doping with hollow nanotubes greatly increases the electrochemically active surface area and improves electron transport. In situ Raman and electrochemical impedance spectroscopy disclose that the continuously generated Ni3+ active species act as intermediates to promote HMF oxidation. In addition, this facile fabrication method is more environmentally friendly and significantly expands the scope of preparing non-precious metal-based nanotubes for electrocatalytic applications.
In contrast to the traditional thermochemical catalytic conversion method, which typically requires harsh reaction conditions (temperatures ranging from 50–160 °C and pressures of 1–40 bar oxygen) and precious metal catalysts like Pd and Ru, the electrochemical HMF oxidation reaction (HMFOR) to FDCA presents notable advantages.8,9 These include mild reaction conditions, controllable product distribution, and an environmentally friendly reaction process, rendering it highly appealing for practical applications. The HMFOR exhibits a lower theoretical potential compared to the oxygen evolution reaction (OER).10,11 When coupled with the hydrogen evolution reaction (HER), it could theoretically achieve high overall energy efficiency.12,13 However, catalysts for the HMFOR still suffer from low catalytic activity and unsatisfactory FDCA selectivity, which impedes the practical application of the HMFOR.14,15 Therefore, there is an urgent and highly demanding need to develop highly efficient electrochemical catalysts for HMF conversion in an efficient manner.16,17
Nickel-based electrocatalysts have emerged as promising candidates for the HMFOR due to the abundance of earth's reserves and their relatively high catalytic activity.18,19 Recent studies have demonstrated that the oxidation of HMF relies on the highly active Ni3+ species.20–22 Therefore, constructing catalysts with a higher concentration of high-valent nickel species is key to improving catalytic activity. Additionally, heteroatom doping (such as N, P, B, and S) has been identified as an effective strategy for enhancing the electrocatalytic activity for the HMFOR by adjusting the electronic structure of active sites.23–27 Recent studies have explored the use of sulfur-doped materials,28 including sulfonated carbon-based materials,29 sulfur-functionalized metal oxides, and sulfur-contained metal–organic frameworks (MOFs),30 for the selective oxidation of HMF. For instance, Yan et al. designed an S-modified catalyst by impregnating it in Na2S solution, achieving an FDCA yield of 97.1%.31 Qi et al. reported MOF-supported metal sulfide nanoclusters, showing a 99% yield of FDCA.32 Mu et al. prepared a Co9S8–Ni3S2@N, S, O triple-doped carbon (NSOC) catalyst by adding thiourea as a sulfur source via pyrolysis, achieving a 99.8% yield of FDCA.33 Despite progress in sulfur-doped catalysts for HMF oxidation, significant challenges remain. Firstly, these methods for fabricating S-doped Ni-based catalysts generally require the addition of extra sulfur sources, which complicates the process and raises environmental concerns. Additionally, the ability of a catalyst to adsorb organic molecules is critical for electrocatalytic conversion;34 however, the adsorption and activation capabilities of pure nickel species toward HMF molecules remain suboptimal, severely limiting the catalytic activity for the HMFOR. Furthermore, inadequate electrode stability during continuous operation hinders their practical applications.35 Therefore, the clear objective is to search for an innovative, environmentally friendly method for synthesizing a high-performance and high-stability Ni-based catalyst that could efficiently catalyze the HMFOR while minimizing environmental impacts.
Herein, we develop a straightforward and green method for synthesizing a nickel, sulfur-codoped carbon nanotubes catalyst (NiSO4@CNTs). This method utilizes HOFs with tubular morphology as the templates and NiSO4 salt as the source of both nickel and sulfur. Hydrogen-bonded organic frameworks (HOFs), composed of organic or metal–organic building blocks linked via intermolecular hydrogen bonding, feature ultrahigh surface areas, tubular morphology and easy coordination with metal ions, making them ideal templates for the preparation of metal catalysts.36–40 For comparison, a catalyst without sulfur, named Ni(NO3)2@CNTs, was also synthesized using only Ni(NO3)2 as the metal salt. Meanwhile, a catalyst with a higher sulfur content, denoted as Ni2.4S0.1@CNTs, was fabricated via introducing additional thiourea as the sulfur source. During the electrocatalytic HMFOR experiments, both NiSO4@CNTs and Ni2.4S0.1@CNTs exhibited higher FDCA yield (>96%) and Faraday efficiency (>99%) at 1.51 V vs. RHE compared to Ni(NO3)2@CNTs (with an FDCA yield ∼71% and a Faraday efficiency of ∼75.5%). Electrochemical tests revealed that sulfur doping promoted the formation of high-valent nickel species, which are highly active species for the HMFOR, enhanced the electron transfer rate, and increased the electrochemically active surface area, thereby accelerating the HMFOR kinetics. Furthermore, in situ electrochemical impedance spectroscopy (EIS) and in situ Raman analysis revealed that the high activity for the HMFOR, as opposed to the competitive OER, was attributed to the combined effects of the rapid formation of NiOOH active sites and the sluggish reaction kinetics of the OER. This work aims to develop a simple and environmentally friendly synthesis method for S-doped Ni-based catalysts by incorporating metal coordinate bonds into a hydrogen-bonded network, while also providing a comprehensive understanding of the role of sulfur doping in enhancing HMFOR performance. These findings could aid in the development of efficient electrocatalysts for the sustainable production of bio-based chemicals and materials.
:
V = 7
:
3) and 15 μL of Nafion, followed by ultrasonication for 1 h at room temperature. Then, 200 μL of catalyst ink was drop-coated onto the hydrophobic carbon paper (geometric area: 1 × 1 cm2). The catalyst loading was about 1 mg cm−2. Linear sweep voltammetry (LSV) measurements were performed at room temperature with a scanning rate of 5 mV s−1 in 0.1 M KOH electrolyte solution. The potential values were converted to potential versus the reversible hydrogen electrode (RHE) using the following equation:| ERHE = EHg/HgO + 0.059 × pH + 0.098 V |
All LSV data were recorded without iR compensation. Electrochemical impedance spectroscopy (EIS) tests were conducted in the frequency range of 0.1 Hz to 100 kHz with 10 mV amplitude at 1.46 V vs. RHE. Double-layer capacitance values (Cdl) were calculated using cyclic voltammetry (CV) curves with scan rates of 20, 40, 60, 80, 100, and 120 mV s−1 in the non-faradaic region. In situ EIS was performed at a potential of 0.15 V ∼0.95 V vs.Hg/HgO with a frequency ranging from 100 kHz to 0.1 Hz at an amplitude of 10 mV.
The conversion of HMF, selectivity and yield of FDCA were calculated using the following equations:
The faradaic efficiency of FDCA formation was calculated using the following equation:
485 C mol−1).
The X-ray diffraction patterns of NiSO4@CNTs, Ni(NO3)2@CNTs, and Ni2.4S0.1@CNTs are shown in Fig. 2a. The diffraction peaks of Ni(NO3)2@CNTs, located at 44.4°, 51.8°, and 76.3°, corresponded to the (111), (200), and (220) planes of metallic Ni (PDF#70-1849), respectively. The diffraction peaks detected in NiSO4@CNTs and Ni2.4S0.1@CNTs at 21.7°, 31.1°, 37.7°, 49.7°, and 55.1° corresponded to the (100), (
10), (111), (210), and (2
1) planes of Ni3S2 (PDF#85-1802), respectively. X-ray photoelectron spectroscopy (XPS) was employed to investigate the elemental compositions and electronic structure of the synthesized catalysts (Fig. 2b–d). The survey spectrum of NiSO4@CNTs revealed the coexistence of Ni, C, N, O, and S elements (Fig. S8†). The high-resolution XPS spectrum of Ni 2p could be deconvoluted into three spin–orbit doublet peaks, with binding energies at 852.6 eV (Ni 2p3/2) and 869.7 eV (Ni 2p1/2) for Ni, 855.4 eV (Ni 2p3/2) and 872.8 eV (Ni 2p1/2) for Ni2+, and 857.4 eV(Ni 2p3/2) and 875.5 eV (Ni 2p1/2) for Ni3+ along with two satellite peaks (denoted as “Sat.”).41,42 Noticeably, a higher content of Ni3+ species was observed in NiSO4@CNTs (22.1%) than that of Ni(NO3)2@CNTs (10.1%). And the Ni3+ content was further improved for Ni2.4S0.1@CNTs (22.3%) with the introduction of higher sulfur content. Moreover, as compared to those in Ni(NO3)2@CNTs, the peaks related to Ni2+ and Ni0 in NiSO4@CNTs and Ni2.4S0.1@CNTs shifted to higher energy. The interaction between Ni and S was further confirmed from the S 2p XPS spectra of NiSO4@CNTs and Ni2.4S0.1@CNTs. The S 2p XPS spectra could be divided into three peaks, such that the peaks at 162.3 eV and 163.7 eV were assigned to the Ni–S bonds.42,43 These results indicated that the presence of S could modify the electronic structure of Ni species, and improve the content of Ni3+ due to the stronger electronic interaction between Ni and S. In the O 1s spectra, three main peaks located at 529.6 eV, 531.5 eV and 533.8 eV are associated with the lattice O2− bonds (Ni–O, OL), hydroxyl groups, and oxygen-species absorbed on the surface (Oads), respectively.44–46 Overall, both XRD and XPS results clearly demonstrated that sulfur doping induced the transformation of Ni species to Ni3S2 with electron transfer between Ni species and S. Moreover, the strong electronic interaction between S and Ni increased the concentration of Ni3+ in Ni2.4S0.1@CNTs, leading to a great enhancement in the subsequent catalytic performance.
![]() | ||
| Fig. 2 (a) XRD patterns of Ni(NO3)2@CNTs, NiSO4@CNTs, and Ni2.4S0.1@CNTs. (b–d) High-resolution XPS spectra of Ni 2p, S 2p, and O 1s for Ni(NO3)2@CNTs, NiSO4@CNTs, and Ni2.4S0.1@CNTs samples. | ||
The electrocatalytic performance of the prepared catalysts was tested in an H-cell with 0.1 M KOH as the electrolyte under ambient conditions. The catalysts were deposited on carbon paper with a loaded mass of 1 mg cm−2. Linear sweep voltammetry (LSV) was conducted on Ni(NO3)2@CNTs, NiSO4@CNTs, and Ni2.4S0.1@CNTs with or without 10 mM HMF to evaluate the relationship between HMF oxidation activities and the catalysts' structure (Fig. 3a). As observed, the current densities significantly increased upon the addition of 10 mM HMF into the electrolyte, indicating a preference for HMF oxidation over the OER on all three catalysts. Compared to Ni(NO3)2@CNTs, NiSO4@CNTs exhibited a higher current density at the same overpotential, indicating that sulfur doping could improve the charge transfer rate. The current density was further enhanced for Ni2.4S0.1@CNTs with higher sulfur content, proving the positive effect of sulfur introduction on catalyst activity (Fig. 3b). The LSV curves of catalysts synthesized with higher thiourea content were also tested (Fig. S9†). Remarkably, when normalized to nickel loading, both Ni2.4S0.1@CNTs and NiSO4@CNTs exhibited superior mass activity toward HMF oxidation compared to Ni(NO3)2@CNTs, while the Ni2.4S0.1@CNTs possessed the best performance among the tested catalysts. The content of Ni in all catalysts were detected via inductively coupled plasma optical emission spectroscopy (ICP-OES; Table S1†). For example, at 1.51 V vs. RHE, the mass activities of Ni2.4S0.1@CNTs (14.3 mA mg−1) and NiSO4@CNTs (10.4 mA mg−1) were 5.5 and 4.0 times higher than that of Ni(NO3)2@CNTs (2.6 mA mg−1), respectively (Fig. 3c).
The intrinsic activity of different samples in the presence of HMF was further assessed using the Tafel slopes (Fig. 3d). Ni2.4S0.1@CNTs exhibited the lowest Tafel slope (110 mV dec−1) compared to Ni(NO3)2@CNTs (528 mV dec−1) and NiSO4@CNTs (170 mV dec−1), demonstrating its favorable HMF electrooxidation kinetics. The electrochemically active surface area (ECSA) of the three catalysts was estimated by determining the double-layer capacitance (Cdl), calculated by performing cyclic voltammetry (Fig. S10†) in the non-faradaic region. The Cdl of Ni2.4S0.1@CNTs was determined to be 6.61 mF cm−2, which is larger than that of NiSO4@CNTs (1.65 mF cm−2) and Ni(NO3)2@CNTs (1.09 mF cm−2) (Fig. 3e), revealing the largest ECSA of Ni2.4S0.1@CNTs. Clearly, these positive results also revealed that the introduction of sulfur could provide more exposed active sites for the catalytic conversion of HMF to FDCA. Furthermore, electrochemical impedance spectroscopy (EIS) was also conducted (Fig. 3f) to investigate the reaction kinetics of catalysts for HMF oxidation. Nyquist plots were fitted with an equivalent circuit model (Fig. 3f, inset). Compared with Ni(NO3)2@CNTs, NiSO4@CNTs and Ni2.4S0.1@CNTs showed smaller charge-transfer resistance (Rct) indicating the improved charge-transfer behavior at the catalyst interface (Table S2†). All the aforementioned results revealed that the interaction between S and Ni species improved the ECSA, reduced mass transfer resistance, and facilitated charge-transfer kinetics. As a result, these improvements contribute to the enhanced electrocatalytic performance for HMF oxidation.
The presence of both hydroxymethyl and aldehyde groups in HMF allows for oxidation to aldehydes and carboxylic acids. Thus, there are two possible pathways for the HMF oxidation reaction (Fig. 4a). If the aldehyde group in HMF is oxidized initially, it yields 5-hydroxymethyl-2-furanformic acid (HMFCA), whereas if the hydroxymethyl group is oxidized first, 2,5-diformylfuran (DFF) will be produced. Both HMFCA and DFF can undergo further oxidation to generate the FFCA intermediate. Subsequently, the FFCA is further oxidized to produce the desired FDCA. To identify the real reaction path in our catalytic system, the concentration of substrates, intermediates and products changing with time was quantified in 0.1 M KOH with 10 mM HMF at a constant potential of 1.51 V vs. RHE by high-performance liquid chromatography (HPLC). As displayed in Fig. 4b, only two kinds of intermediates, HMFCA and FFCA, were detected by HPLC during the HMF oxidation process, demonstrating that HMF oxidation followed the aldehyde oxidation pathway (path I). This result is consistent with the mechanism reported in the literature. To compare the electrocatalytic HMF oxidation performance of the three catalysts, the faradaic efficiency (FE) and yield of FDCA of Ni(NO3)2@CNTs, NiSO4@CNTs and Ni2.4S0.1@CNTs were tested at the same potential (1.51 V vs. RHE). As shown in Fig. 4c, both NiSO4@CNTs and Ni2.4S0.1@CNTs showed superior FDCA yield (>96%) and higher FE (>99%) at 1.51 V vs. RHE compared to Ni(NO3)2@CNTs (FDCA yield ∼71%, FE ∼75.5%). Moreover, for the optimized Ni2.4S0.1@CNTs, both FE and FDCA yields were over 90% from 1.45 to 1.51 V vs. RHE, indicating a wide range of applied potentials (Fig. S11†). As the reaction progressed (Fig. S12†), the current density gradually decreased to zero due to HMF consumption. It is noteworthy that both NiSO4@CNTs and Ni2.4S0.1@CNTs exhibited excellent electrocatalytic activities for the HMFOR, highlighting the unique contribution of sulfur doping in the HMFOR. The additional sulfur doping using thiourea in Ni2.4S0.1@CNTs further improved the reaction activity and shortened the reaction time. Therefore, Ni2.4S0.1@CNTs was selected as the best catalyst for subsequent stability tests and in situ characterization studies. To evaluate the stability of Ni2.4S0.1@CNTs for the HMFOR, continuous electrolysis experiments were performed (Fig. 4d). The results of the stability tests were presented as time-dependent current density plots (Fig. S13†). The performance remained almost unchanged over five successive cycles, further confirming the excellent stability of Ni2.4S0.1@CNTs for the HMFOR.
To study the relationship between potential, structure, and activity, in situ electrochemical impedance spectroscopy (EIS), an electronic technique, was adopted to explore the interfacial evolution of the Ni2.4S0.1@CNTs catalyst. The low-frequency region (10−1–101 Hz) is related to the oxidation of species at the electrode interface, such as the OER and HMFOR, while the high-frequency region (101–105 Hz) corresponds to the oxidation of the electrode catalysts.47–49 As shown in Fig. 5a, the peak in the low-frequency region from 1.60 V vs. RHE suggested that the OER occurred with sluggish kinetics. In contrast, after adding HMF, the peak belonging to the HMFOR could be observed at the potentials of 1.35 and 1.40 V vs. RHE, where no signals were observed for the OER, suggesting that the HMFOR occurred instead of the OER (Fig. 5b). These results unraveled the sluggish OER kinetics and high HMFOR oxidation activity on Ni2.4S0.1@CNTs. To further quantify the relationship between the structural evolution of the catalyst and electrode interface reaction, EIS data (Table S3†) were fitted through the model as shown in Fig. S14.† The Rp which represented the resistance of the electrode catalyst to oxidation and interface oxidation, decreased when the potential reached 1.25 V vs. RHE, indicating changes in the catalyst structure.49,50 The resistance of interface oxidation (Rct) decreased when potential reached 1.35 V vs. RHE (Fig. 5c). These results indicated that the catalyst was first oxidized to real active sites and then participated in the HMFOR.
An in situ Raman device was employed to directly investigate the oxidized species during the HMFOR (Fig. 5d). In a 0.1 M KOH aqueous solution, two distinct vibration peaks at 472 cm−1 and 553 cm−1 assigned to the Ni–O stretching modes of Ni3+OOH were observed,51,52 confirming that the high valence Ni3+OOH species were formed in the OER (Fig. 5e). The intensity of both peaks increased as the potential increased from 1.35 to 1.70 V vs. RHE. After adding HMF, the characteristic peaks corresponding to Ni3+OOH vanished at potentials ranging from 1.30 to 1.55 V vs. RHE (Fig. 5f), suggesting that the in situ generated Ni3+OOH immediately reacted with HMF and was reduced to Ni2+. When the potential reached 1.60 V vs. RHE, the peaks related to Ni3+OOH reappeared, which was due to the faster kinetics of the OER than that of the HMFOR.19,41 Thus, based on the in situ EIS and in situ Raman results, it is reasonable to infer that during the HMFOR, Ni2+ was initially oxidized to Ni3+OOH, which then facilitated the conversion of HMF to FDCA accompanied by the reduction of Ni3+OOH to Ni2+. The continuously generated Ni3+ active species on the surface of the Ni2.4S0.1@CNTs catalyst are responsible for the highly efficient oxidation of HMF to FDCA.
To gain deeper insight into the catalytic process and mechanism, we conducted further analyses and summarized the findings based on the above discussions. For the HMFOR using Ni2.4S0.1@CNTs the catalyst, HMF oxidation to FDCA followed a stepwise pathway: HMF was first oxidized to HMFCA, then to FFCA, and finally to FDCA, as confirmed by HPLC analysis (Fig. 4b). In situ Raman and EIS techniques revealed that the catalyst was initially oxidized to NiOOH, the active form for HMF oxidation. Sulfur doping facilitated the formation of NiOOH and enhanced the adsorption of HMF onto the Ni3S2 surface. The aldehyde group in HMF was selectively oxidized to form HMFCA, which was further oxidized to FFCA by removing hydrogen from the hydroxyl group. Finally, FFCA was oxidized to FDCA at the NiOOH active sites. The desorption of FDCA from the catalyst surface completed the catalytic cycle, which was facilitated by the reduced surface energy imparted by sulfur doping.
Following the continuous HMFOR experiments at constant potentials, we conducted a series of characterization studies to investigate the possible changes in the morphology and structure of the catalyst during the stability testing process. The SEM image (Fig. S15†) revealed that the tubular morphology of the catalysts remained intact, indicating good structural stability. XRD patterns (Fig. S16†) confirmed that the crystallinity and Ni3S2 phase of the catalyst were preserved, suggesting minimal structural degradation. Furthermore, XPS analysis (Fig. S17†) revealed the changes in the chemical composition and oxidation states of the catalysts. The disappearance of Ni0 was observed, indicating a complete transformation of metallic nickel to higher-valent nickel states. The retention of catalytic activity despite the absence of Ni0 suggested that metallic nickel did not play a significant role in the catalytic mechanism for the oxidation of HMF to FDCA. Instead, the higher-valent nickel species were the active centers responsible for the catalytic conversion.
Footnote |
| † Electronic supplementary information (ESI) available. See DOI: https://doi.org/10.1039/d4ta02469a |
| This journal is © The Royal Society of Chemistry 2024 |