Next Article in Journal
The Effect of CeO2 on the Catalytic Activity and SO2 Resistance of the V2O5-MoO3/TiO2 Catalyst Prepared Using the Ball Milling Method for the NH3-SCR of NO
Previous Article in Journal
Effect of N-Doped Carbon on the Morphology and Oxygen Reduction Reaction (ORR) Activity of a Xerogel-Derived Mn(II)O Electrocatalyst
Previous Article in Special Issue
Low-Temperature NH3-SCR Technology for Industrial Application of Waste Incineration: An Overview of Research Progress
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Fluorine-Doped Graphene Oxide-Modified Graphite Felt Cathode for Hydrogen Peroxide Generation

1
College of Chemistry, Beijing University of Chemical Technology, Beijing 100029, China
2
Sinopec Qingdao Research Institute of Safety Engineering, Qingdao 266000, China
*
Authors to whom correspondence should be addressed.
Catalysts 2024, 14(11), 793; https://doi.org/10.3390/catal14110793
Submission received: 26 September 2024 / Revised: 3 November 2024 / Accepted: 4 November 2024 / Published: 6 November 2024
(This article belongs to the Special Issue Recent Advances in Environment and Energy Catalysis)

Abstract

:
Electrochemical oxygen reduction via the two-electron pathway (2e-ORR) is an emerging method for producing H2O2. It is cleaner, safer, and more convenient compared to the anthraquinone process. Graphite felt is one of the cathode candidates for large-scale cells due to its excellent mechanical properties. However, commercial graphite felt often fails to achieve the desired hydrogen-peroxide yield because of its low catalytic selectivity for the 2e-ORR pathway. Fluorine-doped carbon materials are expected to enhance 2e-ORR selectivity. This is because the electronic structure of carbon atoms adjacent to fluorine atoms may facilitate the production of hydrogen peroxide while hindering its further reduction. In this study, fluorine-doped graphene oxide (FGO) was prepared by the hydrothermal method. Subsequently, graphite felt modified with FGO was fabricated and used as the cathode for H2O2 production. The results indicated that in alkaline media, the graphite felt modified with FGO achieved a catalytic selectivity of 93% and a generation rate of 8.91 mg cm⁻2 h⁻¹. In comparison, commercial graphite felt had a catalytic selectivity of 75% and a generation rate of 2.10 mg cm⁻2 h⁻¹. Moreover, graphite felt modified by FGO also exhibited excellent electrocatalytic performance for H2O2 generation in neutral media. This research provides a fundamental study to promote the application of graphite felt in the environmentally friendly electrocatalytic production of hydrogen peroxide in industries.

Graphical Abstract

1. Introduction

The electrochemical synthesis of H2O2 through the two-electron oxygen reduction reaction (2e-ORR) pathway is an emerging hydrogen-peroxide production method, which can be operated under mild reaction conditions, offering greater cleanliness and safety [1,2]. Moreover, it serves as an effective means to achieve on-site production for the utilization of H2O2 solution [3,4,5]. However, the production efficiency of the electrochemical method is still much lower than that of the state-of-the-art anthraquinone process.
To enhance the production efficiency of H2O2 in a large-scale cell requires attempting to obtain a more concentrated H2O2 solution in a shorter time and with less electricity consumption. In other words, it means improving the production rate and current efficiency. The key to improving current efficiency lies in the selectivity of the catalyst or electrode material. Hence, it is crucial to design and develop active electrocatalysts with selectivity for H2O2. The 2e-ORR pathway for the reaction of O2 molecules on the cathode competes with the thermodynamically favorable 4e-ORR pathway to H2O. OOH* is a common intermediate of both the 2e and 4e pathways. The magnitude of the adsorption energy of OOH* determines the selectivity of H2O2 generation [6]. Therefore, electrocatalysts with appropriate adsorption-free energies of OOH* are expected to improve H2O2 selectivity.
In recent years, carbon materials have attracted much attention as the cathode material for H2O2 production because of their advantages, such as low cost, high storage capacity, and modifiability [7,8,9,10]. Carbon materials, including carbon nanotubes, graphene, mesoporous carbon, activated carbon, and others, have been doped with oxygen functional groups [11,12,13] and heteroatoms, such as N [14,15,16,17], B [18,19,20], S [21,22,23], and others. In these materials, heteroatoms combine with the carbon skeleton, resulting in charge redistribution in the carbon material and modulating the adsorption-free energy of oxygen and the reaction-intermediate OOH* during the reaction process. As a result, the activity of the ORR and the selectivity of H2O2 are improved, and so is the yield of H2O2 in the electrolysis cell.
Moreover, in order to enhance the electrochemical generation of hydrogen peroxide, it is crucial to increase the current density. This requires the use of appropriate substrate materials to facilitate the attainment of the desired current density. Graphite felt (GF) has been widely used as an electrode material due to its large specific surface area, excellent mechanical properties, and good electrical conductivity [24,25]. These characteristics make it an ideal substrate material for cathode electrodes that can withstand high current densities in large-scale applications. Currently, graphite felt is being researched for its potential use as a cathode for the direct production of hydrogen peroxide. However, its selectivity for 2e-ORR is limited, leading to lower production yields [26]. Therefore, modifying graphite felt with high-performance catalysts and using it as the cathode could be one of the approaches to enhance the electrochemical production efficiency of H2O2.
Fluorine atoms doped in carbon materials can enhance the O2 adsorption energy to improve the ORR electrocatalytic activity of carbon materials [27,28,29]. It is found that F doping elongates the O=O bond distance and shortens the C-O distance, thus clearly demonstrating its superiority in O2 adsorption on fluorine-doped carbon materials [30]. In addition, the doped F species with different configurations have a different influence on the selectivity for the ORR pathway. Based on density functional theory (DFT) studies and experimental results, Zhao et al. discovered that the doped F species with different configurations have a different influence on the selectivity for the ORR pathway. That is, the covalent –CFx species (–CF2 and –CF3) can promote H2O2 synthesis by selectively catalyzing the two-electron pathway of oxygen reduction to promote the synthesis of H2O2 [31]. Furthermore, Wang et al. observed that different kinds of F–CNTs had different electrocatalytic properties, which might be related to the contents of –CF2 and –CF3 [32].
In this study, fluorine-doped graphene oxide (FGO) was synthesized as an electrocatalyst by introducing –CF2 groups to graphene oxide (GO) via the hydrothermal method, using GO as the substrate and hydrofluoric acid as the F-source. The synthesized FGO was then used for the preparation of FGO-modified graphite felt cathode material, which exhibited high hydrogen-peroxide selectivity and effective yields in alkaline or neutral media.

2. Results and Discussion

2.1. Characterization of FGO and FGO/GF

Many methods were used to identify the existence and the chemical speciation of F in FGO. In the X-ray powder diffraction (XRD) pattern of GO (Figure 1a), the weak broad peak centered at approximately 2θ = 19.8° originates from the carbon skeleton structure. The intensive diffraction peak at 2θ = 9.8° is related to the intercalation of a large number of oxygen functional groups [33], such as –OH, –COOH, and C–O–C. It was found that the peak at 9.8° disappeared and a broad peak at about 24° was observed in the XRD pattern of FGO, indicating that most of the oxygen functional groups in GO had been removed after the hydrothermal treatment with hydrofluoric acid. The peak at about 2θ = 18° was observed in FGO, which may be related to the reduction of oxygen functional groups and the doping of F [34].
The Fourier transform infrared (FTIR) spectra also confirm the introduction of F in FGO compared with that of GO (Figure 1b). The characteristic bands at 1627 cm−1 and 1725 cm−1 in the FTIR spectra of GO are, respectively, assigned to the stretching vibration of C=C and C=O, which originated from the graphene-oxide-like benzene ring structure [35]. The broad bands appeared in the region of 3700–2400 cm−1, and the absorption peak at 1040 cm−1 contributed to O–H stretching and vibration of C–O. For FGO, the characteristic peaks related to the oxygen functional groups weakened or disappeared, and a new peak at 1100 cm−1 could be assigned to the C–F bond connected to a nearby sp2 hybridized carbon atom [36]. Furthermore, the presence of F atoms and their uniform distribution in FGO is also evidenced in EDS mapping images (Figure 1c). The nitrogen adsorption–desorption isotherms of GO and FGO are shown in Figure 1d. The BET surface area of GO and FGO was calculated to be 204.6 m2 g−1 and 237.4 m2 g−1, respectively. The isotherms for GO and FGO exhibited typical type IV curves with type H3 hysteresis loop [37], indicating the existence of micro- and meso-pores. From the inset pore-size distribution in Figure 1d, it can be seen that the pore size of FGO is obviously smaller than that of GO. This may be attributed to the strong noncovalent interactions formed during the hydrothermal treatment of GO sheets under acidic pH conditions. Such interactions lead to the aggregation of FGO, resulting in the formation of many small holes with thick walls [38].
The chemical speciation of F in FGO was further confirmed by X-ray photoelectron spectroscopy (XPS). In the XPS spectrum of FGO (Figure 2a), the F1s peak appeared at 685 eV, and the XPS data showed the presence of carbon, oxygen, and fluorine with atomic percentages of 88.3%, 7.77%, and 3.93%, respectively. This indicates that the F element has been successfully introduced into graphene oxide after hydrothermal treatment with HF. As shown in Figure 2b, the C1s spectra of GO can be fitted with four peaks, centered at about 284.4, 284.8, 286.7, and 288.3 eV, corresponding to the C=C, C–C, C–O, and –C=O groups, respectively [39]. After the HF treatment, the relative intensity of C=C and C–O decreased, and a new peak appeared at 291 eV, which is sure to be related to the C–F bond.
In general, there are three chemical forms of C–F bonds that may be formed in F-doped carbon materials; these are ionic C–F, semi-ionic C–F, and covalent C–F [40,41]. Figure 2c,d show the high-resolution XPS spectra of F 1s and C 1s for FGO, respectively. In the high-resolution F 1s spectra, the prominent peak at approximately 688.6 eV is associated with the –CF2 bond [42,43,44], while the peaks around 685.5 eV and 691.3 eV correspond to the semi-ionic C–F bond and the covalent –CF₃ bond, respectively [45,46,47,48]. Additionally, the broad peak centered at 291 eV in the high-resolution C 1s spectrum provides further evidence that –CF2 is the primary bonding configuration of F [49,50].
Based on the above results, it can be inferred that most of the F atoms in FGO exist in –CF2 bonds. The fluorine atoms have strong electronegativity, which attracts the surrounding electrons and leads to a change in the charge distribution of the neighboring carbon atoms on the graphene-oxide surface.
To clarify the morphological characteristics of the FGO modified GF electrodes, the SEM and EDS mapping images of GF and FGO/GF were analyzed, as shown in Figure 3. The pristine GF surface is smooth and almost defect-free with an interconnected network structure under different magnifications (Figure 3a). In contrast, the GF modified by FGO impregnation (Figure 3b) shows that the dispersed catalyst was locally agglomerated and adhered to the fiber surface of the GF. The results of EDS indicated that after the modification of graphite mats with FGO, the contents of C, O, and F on the surface of the graphite-mat fibers were 89.3 at%, 7.4 at%, and 3.3 at%, respectively (Figure 3c). Due to the large specific surface area provided by the dense nature of the graphite-mat fibers, the catalyst is able to attach to as many fibers as possible to expose the active sites.

2.2. Electrochemical Characterizations

Figure 4a,d show the RRDE polarization curves in O2-saturated 0.1 M KOH solution or 0.1 M Na2SO4 solution with the FGO or GO loaded on the disk electrode and the rotating rate of electrode of 1600 rpm. As shown in Figure 4a, the onset potential (Eonset, at H2O2 current density at 0.1 mA cm−2) of 0.72 V for FGO in 0.1 M KOH was more positive than that of GO (0.59 V), indicating that the ORR catalytic activity of FGO is higher than that of GO. In addition, FGO exhibited a significantly better H2O2 selectivity than GO in the potential range of 0.2–0.7 V (~80–90% for FGO, ~70–80% for GO), in which the transferred electrons number was calculated as 2.13 at 0.70 V. Similarly, the hydrogen-peroxide generation performance of FGO in neutral electrolytes (0.1 M Na2SO4, pH ~ 7) was also significantly increased compared to that of GO, with the measured onset potential positively shifted by about 100 mV (Figure 4d) and the selectivity increased from ~60% for GO to ~80% for FGO (Figure 4f).
These observations suggest that F-atom doping in graphene oxide facilitates the enhancement of electrocatalytic activity and hydrogen-peroxide selectivity for the O2 reduction reaction, which may be attributed to the fact that the –CF2 doping into the carbon skeleton modulates the adsorption-free energies of graphene oxide for O2 and the intermediate OOH* [31].

2.3. Electrogeneration of H2O2

The electrochemical hydrogen-peroxide-production performance of FGO/GF was evaluated by using it as cathode in an H-type cell with 1.0 M KOH solution or 0.1 M Na2SO4 solution, and the concentration of H2O2 in the electrolyte after 1200 s of electrolysis was used as the evaluation index. The results in the alkali electrolyte showed that the Faraday efficiency (FE) of FGO/GF was 93% at 0.7 V (Figure 5a), which is significantly higher than that of the commercial GF and GO/GF. Furthermore, the current density during the electrolysis using FGO/GF as a cathode was significantly enhanced to 20.3 mA cm−2 at 0.3 V (Figure 5b). Finally, 371 mg L−1 of H2O2 was yielded in 1.0 M KOH with a Faraday efficiency of 70% and current density of 20.3 mA cm−2 (Figure 5b,c). In the neutral electrolyte, FGO/GF also exhibited good performance: the FE reached ~77% at 0.2 V, and the current density reached ~16 mA cm−2 at 0 V. As a result, 165 mg L−1 of H2O2 was generated using FGO/GF as a cathode, while those levels were 1.28 mg L−1 and 93 mg L−1 using commercial GF and GO/GF as a cathode (Figure 5d–f).
The results of constant potential electrolysis demonstrated that fluorine-doped graphene oxide significantly enhanced the electrocatalytic production of hydrogen peroxide by graphite felt. This improvement was attributed to the excellent oxygen reduction reaction (ORR) activity and two-electron selectivity of FGO.
The H2O2 yield was also evaluated with different FGO loadings (1.2 and 4 mg cm−2) on GF in 1.0 M KOH solution (Figure 6a–c) and 0.1 M Na2SO4 solution (Figure 6d,e). The graphite felts with 1 mg cm−2 FGO loading had the highest hydrogen-peroxide yield in 1.0 M KOH, and the highest H2O2 production rate was calculated to be 8.91 mg·h−1·cm−2. In 0.1 M Na2SO4, the graphite felt with 2 mg cm−2 FGO loading had the highest hydrogen-peroxide yield of 257 mg L−1 at 0.1 V, and the maximum production rate was calculated to be 6.18 mg·h−1·cm−2. This may be due to the fact that the ORR process is limited by the mass-transfer process, where the diffusion of O2 may be relatively slow or the adsorption of OOH* on the catalyst surface may be weak. An appropriate increase in catalyst loading can increase the contact opportunity between O2 and the catalyst and improve the reaction rate.
Compared to the electrocatalytic GF cathodes recently reported in the literature, FGO/GF has a significant advantage in terms of H2O2-generating capacity (Table 1), providing greater current density and yield of hydrogen peroxide.
The stability test was conducted by repeating the constant potential electrolysis test under the optimized conditions in the alkali electrolyte (Figure 7). By replacing the electrolyte and cleaning the electrodes every 20 min at 0.3 V vs. RHE, pH = 13.7, the current density and yield of H2O2 were tested. During the five cycles, the current density was relatively stable, and the yield of hydrogen peroxide decreased slightly but remained within a reasonable range, with the yield of hydrogen peroxide during the last cycle maintained at 91%.

3. Materials and Methods

3.1. Reagents and Chemicals

Graphene oxide (GO) was purchased from Suzhou TanFeng (Suzhou TanFeng Graphene Technology Co., Ltd., Jiangsu, China), graphite felt (GF, G475) was purchased from SCI Material Hub, potassium hydroxide (KOH, AR) and sodium sulfate (Na2SO4, AR) were purchased from Shanghai Aladdin (Shanghai Aladdin Bio-chemistry Technology Co., Ltd., Shanghai, China), and aqueous hydrofluoric acid (HF, 40 Ltd., Beijing, China) and aqueous hydrofluoric acid (HF, 40 wt%) were purchased from Beijing Bailing Wei (Beijing Bailing Wei Technology Co., Ltd., Beijing, China).

3.2. Material Synthesis

3.2.1. Synthesis of FGO

A total of 35 mg of GO and 1 mL of hydrofluoric acid (40%) were added to 35 mL of deionized water and ultrasonicated for 20 min until homogeneous. A total of 30 mL of the mixture was transferred to a 50 mL PTFE-lined reactor and heated to 150 °C for 24 h. After natural cooling to room temperature, the product was filtered through a microporous membrane, washed with distilled water, and dried to obtain FGO.

3.2.2. Preparation of Cathodes

The graphite felts (GFs) were first cut into several 1 cm×2 cm strips, washed with ethanol and deionized water by ultrasonication for 30 min, and then dried at 60 °C for 24 h. The catalyst inks were obtained by adding 1 mg/2 mg/4 mg of GO or FGO to 1 mL/2 mL/4 mL of deionized water and 5 wt% Nafion solution (volume ratio of 0.995:0.005) and ultrasonicated for 1 h. The inks were drop-coated onto the GF fibers, and then dried at room temperature to obtain GO/GF or FGO/GF.

3.3. Characterization of Catalysts and the Composite Cathodes

The X-ray powder diffraction (XRD) patterns were acquired by the Bruker AXS D8 Advance instrument (40 kV, 40 mA), using Cu Kα (λ = 1.5406 A) radiation (Bruker, Billerica, MA, USA). Fourier transform infrared (FTIR) spectra were recorded on a Nicolet iS20 spectrometer (Thermo Nicolet, Waltham, MA, USA). The scanning electron microscopy images were obtained using a field-emission scanning electron microscope (Zeiss supra 55). The N2 adsorption–desorption isotherms were studied on a Micromeritics (Micromeritics, ASAP 2020, Waltham, MA, USA) analyzer to calculate specific surface area and porosities of graphene-based materials using the multi-point Brunauer–Emmett–Teller (BET) technique. X-ray photoelectron spectroscopy (XPS) data were obtained using a Thermo Scientific K-Alpha instrument with an Al Ka X-ray source, and the peak areas were fitted using the software XPSPEAK 4.1 by Lorentz (30%) and Gaussian (https://gaussian.com/) (70%) functions until a fitting criterion of c2 less than 10 was achieved.

3.4. Electrochemical Performance Tests

The typical electrochemical study was performed on a CHI760E electrochemical workstation (Chenhua, Shanghai, China) in a standard three-electrode system. For GO and FGO catalysts, the catalytic activity of the catalysts was evaluated using a theoretical collection efficiency of 37% rotating ring-disc electrode (RRDE). A modified glassy carbon electrode (0.2475 cm2) loaded with a GO or FGO catalyst was used as the working electrode, the counter electrode was a platinum sheet (1 cm × 1 cm), and the reference electrode was a saturated calomel electrode (SCE). The catalyst ink was prepared by dispersing 1 mg GO or FGO in 1 mL deionized water and 5 μL 5 wt% Nafion solution. A total of 8 μL of catalyst dispersion was applied dropwise to the glassy carbon electrode at 700 rpm and dried. The electrochemical tests were carried out at ambient temperature and a pressure of (0.1 M KOH) at pH ~ 13, and (0.1 M Na2SO4) at pH ~ 7, respectively. The potentials mentioned in this study have been converted for comparison with the Reversible Hydrogen Electrode (RHE) using Equation (1), which is as follows:
ERHE = E SCE + 0.0592 × pH + 0.2438
For the rotating ring-disc electrode (RRDE), the average number of transferred electrons (n) and the H2O2 selectivity were calculated by the following equation:
n = 4 × I d I d + I r / N
H 2 O 2   ( % ) = 200 × I r / N I d + I r / N
where Ir is the ring electrode current, Id is the disc electrode current, and N is the ring electrode collection efficiency (0.37). The ring current was recorded by keeping the Pt ring at 1.2 V (vs. RHE) throughout the test.

3.5. Electrogeneration of H2O2 and Analytic Methods

The H2O2 yields of GO- and FGO-modified graphite felt cathodes in (1 M KOH) at pH ~ 13.7, and (0.1 M Na2SO4) at pH ~ 7, respectively, were determined by constant potential electrolysis in an H-type electrolysis bath with a Nafion diaphragm, in which a catalyst-loaded graphite felt was used as the working electrode, a Pt plate as the counter electrode, and an SCE connected to the cathode chamber by a salt bridge as the reference electrode. The cathode chamber volume was 8 m L for both. The electrolysis was conducted in a constant potential model with O2 gas bubbling near the cathode for 20 min. The generated H2O2 in the electrolyte after electrolysis was determined by the Ce(SO4)2 method, which, using a UV Vis spectrophotometer (UV1100, Shanghai, China), measured at 400 nm. The current efficiency (FE) for H2O2 production was calculated as the following equation [55]:
F E ( % ) = n F cV I d t × 100 %
where n is the number of electrons transferred by the reduction of oxygen to H2O2, F is Faraday’s constant (96,485 C mol−1), c is the concentration of H2O2 (M), V is the volume (L), I is the current (A), and t is the electrolysis time (s).
The stability test was conducted by repeating the constant potential electrolysis test under the same conditions. The electrolyte was replaced and we cleaned the electrodes every 20 min.

4. Conclusions

In this study, GO was treated with HF under hydrothermal conditions to obtain F-doped GO, and the main chemical speciation of F elements in FGO was confirmed as –CF2 groups. The introduction of –CF2 groups into GO led to a significant improvement in its catalytic selectivity for 2e-ORR. Then, the FGO-modified graphite felt (FGO/GF) was fabricated and its performance in producing H2O2 was evaluated by using it as the cathode in alkaline and neutral electrolytes. The results showed that the H2O2 yield and current efficiency of the FGO/GF cathode were significantly improved compared with those of commercial GF and GO/GF in both alkaline and neutral electrolytes. A total of 371 mg L−1 of H2O2 was generated after electrolysis at 0.3 V vs. RHE in 1.0 M KOH for 1200s with a current density of 20.3 mA cm−2. The present work provides a facile strategy for the preparation of efficient, low-cost, heteroatom co-doped cathode materials for the electrocatalytic production of hydrogen peroxide.

Author Contributions

Conceptualization, Y.C. and J.H.; methodology, J.H.; writing—original draft preparation, J.H.; writing—review and editing, Y.C., J.H., Z.W. and W.X.; visualization, Z.W.; formal analysis, W.X. All authors have read and agreed to the published version of the manuscript.

Funding

This project was supported by the National Natural Science Foundation of China (No. 22075012).

Data Availability Statement

Data are contained within the article.

Conflicts of Interest

The authors declare no conflicts of interest.

References

  1. Fukuzumi, S.; Yamada, Y.; Karlin, K.D. Hydrogen Peroxide as a Sustainable Energy Carrier: Electrocatalytic Production of Hydrogen Peroxide and the Fuel Cell. Electrochim. Acta 2012, 82, 493–511. [Google Scholar] [CrossRef] [PubMed]
  2. Siahrostami, S.; Villegas, S.J.; Bagherzadeh Mostaghimi, A.H.; Back, S.; Farimani, A.B.; Wang, H.; Persson, K.A.; Montoya, J. A Review on Challenges and Successes in Atomic-Scale Design of Catalysts for Electrochemical Synthesis of Hydrogen Peroxide. ACS Catal. 2020, 10, 7495–7511. [Google Scholar] [CrossRef]
  3. Jiang, Y.; Ni, P.; Chen, C.; Lu, Y.; Yang, P.; Kong, B.; Fisher, A.; Wang, X. Selective Electrochemical H2O2 Production through Two-Electron Oxygen Electrochemistry. Adv. Energy Mater. 2018, 8, 1801909. [Google Scholar] [CrossRef]
  4. Zhou, W.; Meng, X.; Gao, J.; Alshawabkeh, A.N. Hydrogen Peroxide Generation from O2 Electroreduction for Environmental Remediation: A State-of-the-Art Review. Chemosphere 2019, 225, 588–607. [Google Scholar] [CrossRef]
  5. Armaroli, N.; Balzani, V. Towards an Electricity-Powered World. Energy Environ. Sci. 2011, 4, 3193–3222. [Google Scholar] [CrossRef]
  6. Deng, Z.; Choi, S.J.; Li, G.; Wang, X. Advancing H2O2 Electrosynthesis: Enhancing Electrochemical Systems, Unveiling Emerging Applications, and Seizing Opportunities. Chem. Soc. Rev. 2024, 53, 8137–8181. [Google Scholar] [CrossRef]
  7. Zheng, B.; Cai, X.-L.; Zhou, Y.; Xia, X.-H. Pure Pyridinic Nitrogen-Doped Single-Layer Graphene Catalyzes Two-Electron Transfer Process of Oxygen Reduction Reaction. ChemElectroChem 2016, 3, 2036–2042. [Google Scholar] [CrossRef]
  8. Zhang, H.; Li, Y.; Zhao, Y.; Li, G.; Zhang, F. Carbon Black Oxidized by Air Calcination for Enhanced H2O2 Generation and Effective Organics Degradation. ACS Appl. Mater. Interfaces 2019, 11, 27846–27853. [Google Scholar] [CrossRef]
  9. Zhao, H.; Shen, X.; Chen, Y.; Zhang, S.-N.; Gao, P.; Zhen, X.; Li, X.-H.; Zhao, G. A COOH-Terminated Nitrogen-Doped Carbon Aerogel as a Bulk Electrode for Completely Selective Two-Electron Oxygen Reduction to H2O2. Chem. Commun. 2019, 55, 6173–6176. [Google Scholar] [CrossRef]
  10. Chai, G.-L.; Hou, Z.; Ikeda, T.; Terakura, K. Two-Electron Oxygen Reduction on Carbon Materials Catalysts: Mechanisms and Active Sites. J. Phys. Chem. C 2017, 121, 14524–14533. [Google Scholar] [CrossRef]
  11. Palm, I.; Kibena-Põldsepp, E.; Lilloja, J.; Käärik, M.; Kikas, A.; Kisand, V.; Merisalu, M.; Treshchalov, A.; Paiste, P.; Leis, J.; et al. Impact of Ball-Milling of Carbide-Derived Carbons on the Generation of Hydrogen Peroxide via Electroreduction of Oxygen in Alkaline Media. J. Electroanal. Chem. 2020, 878, 114690. [Google Scholar] [CrossRef]
  12. Lu, Z.; Chen, G.; Siahrostami, S.; Chen, Z.; Liu, K.; Xie, J.; Liao, L.; Wu, T.; Lin, D.; Liu, Y.; et al. High-Efficiency Oxygen Reduction to Hydrogen Peroxide Catalysed by Oxidized Carbon Materials. Nat. Catal. 2018, 1, 156–162. [Google Scholar] [CrossRef]
  13. Han, G.-F.; Li, F.; Zou, W.; Karamad, M.; Jeon, J.-P.; Kim, S.-W.; Kim, S.-J.; Bu, Y.; Fu, Z.; Lu, Y.; et al. Building and Identifying Highly Active Oxygenated Groups in Carbon Materials for Oxygen Reduction to H2O2. Nat. Commun. 2020, 11, 2209. [Google Scholar] [CrossRef] [PubMed]
  14. Iglesias, D.; Giuliani, A.; Melchionna, M.; Marchesan, S.; Criado, A.; Nasi, L.; Bevilacqua, M.; Tavagnacco, C.; Vizza, F.; Prato, M.; et al. N-Doped Graphitized Carbon Nanohorns as a Forefront Electrocatalyst in Highly Selective O2 Reduction to H2O2. Chem 2018, 4, 106–123. [Google Scholar] [CrossRef]
  15. Quílez-Bermejo, J.; Melle-Franco, M.; San-Fabián, E.; Morallón, E.; Cazorla-Amorós, D. Towards Understanding the Active Sites for the ORR in N-Doped Carbon Materials through Fine-Tuning of Nitrogen Functionalities: An Experimental and Computational Approach. J. Mater. Chem. A 2019, 7, 24239–24250. [Google Scholar] [CrossRef]
  16. Peng, W.; Liu, J.; Liu, X.; Wang, L.; Yin, L.; Tan, H.; Hou, F.; Liang, J. Facilitating Two-Electron Oxygen Reduction with Pyrrolic Nitrogen Sites for Electrochemical Hydrogen Peroxide Production. Nat. Commun. 2023, 14, 4430. [Google Scholar] [CrossRef]
  17. To, J.W.F.; Ng, J.W.D.; Siahrostami, S.; Koh, A.L.; Lee, Y.; Chen, Z.; Fong, K.D.; Chen, S.; He, J.; Bae, W.-G.; et al. High-Performance Oxygen Reduction and Evolution Carbon Catalysis: From Mechanistic Studies to Device Integration. Nano Res. 2017, 10, 1163–1177. [Google Scholar] [CrossRef]
  18. Chen, S.; Chen, Z.; Siahrostami, S.; Higgins, D.; Nordlund, D.; Sokaras, D.; Kim, T.R.; Liu, Y.; Yan, X.; Nilsson, E.; et al. Designing Boron Nitride Islands in Carbon Materials for Efficient Electrochemical Synthesis of Hydrogen Peroxide. J. Am. Chem. Soc. 2018, 140, 7851–7859. [Google Scholar] [CrossRef] [PubMed]
  19. Wang, Z.; Sun, Z.; Li, K.; Fan, K.; Tian, T.; Jiang, H.; Jin, H.; Li, A.; Tang, Y.; Sun, Y.; et al. Enhanced Electrocatalytic Performance for H2O2 Generation by Boron-Doped Porous Carbon Hollow Spheres. iScience 2024, 27, 109553. [Google Scholar] [CrossRef]
  20. Fan, M.; Wang, Z.; Sun, K.; Wang, A.; Zhao, Y.; Yuan, Q.; Wang, R.; Raj, J.; Wu, J.; Jiang, J.; et al. N-B-OH Site-Activated Graphene Quantum Dots for Boosting Electrochemical Hydrogen Peroxide Production. Adv. Mater. 2023, 35, 2209086. [Google Scholar] [CrossRef]
  21. Perazzolo, V.; Durante, C.; Pilot, R.; Paduano, A.; Zheng, J.; Rizzi, G.A.; Martucci, A.; Granozzi, G.; Gennaro, A. Nitrogen and Sulfur Doped Mesoporous Carbon as Metal-Free Electrocatalysts for the in Situ Production of Hydrogen Peroxide. Carbon 2015, 95, 949–963. [Google Scholar] [CrossRef]
  22. Xiang, F.; Zhao, X.; Yang, J.; Li, N.; Gong, W.; Liu, Y.; Burguete-Lopez, A.; Li, Y.; Niu, X.; Fratalocchi, A. Enhanced Selectivity in the Electroproduction of H2O2 via F/S Dual-Doping in Metal-Free Nanofibers. Adv. Mater. 2023, 35, 2208533. [Google Scholar] [CrossRef] [PubMed]
  23. Xu, H.; Zhou, S.; Fang, W.; Liu, J.; Lu, M. Confined Mesospace Synthesis of Sulfur-Doped Graphene Quantum Dots for Direct H2O2 Detection. ChemistrySelect 2022, 7, e202202119. [Google Scholar] [CrossRef]
  24. Li, K.; Liu, J.; Li, J.; Wan, Z. Effects of N Mono- and N/P Dual-Doping on H2O2, OH Generation, and MB Electrochemical Degradation Efficiency of Activated Carbon Fiber Electrodes. Chemosphere 2018, 193, 800–810. [Google Scholar] [CrossRef]
  25. Li, D.; Zheng, T.; Liu, Y.; Hou, D.; Yao, K.K.; Zhang, W.; Song, H.; He, H.; Shi, W.; Wang, L.; et al. A Novel Electro-Fenton Process Characterized by Aeration from inside a Graphite Felt Electrode with Enhanced Electrogeneration of H2O2 and Cycle of Fe3+/Fe2+. J. Hazard. Mater. 2020, 396, 122591. [Google Scholar] [CrossRef]
  26. Sun, Y.-M.; Li, C.; Liu, Y.-H. CO2-Activated Graphite Felt as an Effective Substrate to Promote Hydrogen Peroxide Synthesis and Enhance the Electro-Fenton Activity of Graphite/Fe3O4 Composites in Situ Fabricated from Acid Mine Drainage. J. Water Process Eng. 2024, 57, 104690. [Google Scholar] [CrossRef]
  27. Peera, S.G.; Menon, R.S.; Das, S.K.; Alfantazi, A.; Karuppasamy, K.; Liu, C.; Sahu, A.K. Oxygen Reduction Electrochemistry at F Doped Carbons: A Review on the Effect of Highly Polarized C-F Bonding in Catalysis and Stability of Fuel Cell Catalysts. Coord. Chem. Rev. 2024, 500, 215491. [Google Scholar] [CrossRef]
  28. Zhao, J.; Cabrera, C.R.; Xia, Z.; Chen, Z. Single−sided Fluorine–Functionalized Graphene: A Metal–Free Electrocatalyst with High Efficiency for Oxygen Reduction Reaction. Carbon 2016, 104, 56–63. [Google Scholar] [CrossRef]
  29. Chang, Y.; Chen, J.; Jia, J.; Hu, X.; Yang, H.; Jia, M.; Wen, Z. The Fluorine-Doped and Defects Engineered Carbon Nanosheets as Advanced Electrocatalysts for Oxygen Electroreduction. Appl. Catal. B Environ. 2021, 284, 119721. [Google Scholar] [CrossRef]
  30. Sun, X.; Song, P.; Chen, T.; Liu, J.; Xu, W. Fluorine-Doped BP 2000: Highly Efficient Metal-Free Electrocatalysts for Acidic Oxygen Reduction Reaction with Superlow H2O2 Yield. Chem. Commun. 2013, 49, 10296–10298. [Google Scholar] [CrossRef]
  31. Zhao, K.; Su, Y.; Quan, X.; Liu, Y.; Chen, S.; Yu, H. Enhanced H2O2 Production by Selective Electrochemical Reduction of O2 on Fluorine-Doped Hierarchically Porous Carbon. J. Catal. 2018, 357, 118–126. [Google Scholar] [CrossRef]
  32. Wang, W.; Lu, X.; Su, P.; Li, Y.; Cai, J.; Zhang, Q.; Zhou, M.; Arotiba, O. Enhancement of Hydrogen Peroxide Production by Electrochemical Reduction of Oxygen on Carbon Nanotubes Modified with Fluorine. Chemosphere 2020, 259, 127423. [Google Scholar] [CrossRef] [PubMed]
  33. Seehra, M.S.; Narang, V.; Geddam, U.K.; Stefaniak, A.B. Correlation between X-Ray Diffraction and Raman Spectra of 16 Commercial Graphene–Based Materials and Their Resulting Classification. Carbon 2017, 111, 380–385. [Google Scholar] [CrossRef] [PubMed]
  34. Wang, Z.; Wang, J.; Li, Z.; Gong, P.; Liu, X.; Zhang, L.; Ren, J.; Wang, H.; Yang, S. Synthesis of Fluorinated Graphene with Tunable Degree of Fluorination. Carbon 2012, 50, 5403–5410. [Google Scholar] [CrossRef]
  35. Brusko, V.; Khannanov, A.; Rakhmatullin, A.; Dimiev, A.M. Unraveling the Infrared Spectrum of Graphene Oxide. Carbon 2024, 229, 119507. [Google Scholar] [CrossRef]
  36. Sato, Y.; Itoh, K.; Hagiwara, R.; Fukunaga, T.; Ito, Y. On the So-Called “Semi-Ionic” C–F Bond Character in Fluorine–GIC. Carbon 2004, 42, 3243–3249. [Google Scholar] [CrossRef]
  37. Kumar, N.; Setshedi, K.; Masukume, M.; Ray, S.S. Facile Scalable Synthesis of Graphene Oxide and Reduced Graphene Oxide: Comparative Investigation of Different Reduction Methods. Carbon Lett. 2022, 32, 1031–1046. [Google Scholar] [CrossRef]
  38. Wasalathilake, K.C.; Galpaya, D.G.D.; Ayoko, G.A.; Yan, C. Understanding the Structure-Property Relationships in Hydrothermally Reduced Graphene Oxide Hydrogels. Carbon 2018, 137, 282–290. [Google Scholar] [CrossRef]
  39. Reiche, S.; Blume, R.; Zhao, X.C.; Su, D.; Kunkes, E.; Behrens, M.; Schlögl, R. Reactivity of Mesoporous Carbon against Water—An in-Situ XPS Study. Carbon 2014, 77, 175–183. [Google Scholar] [CrossRef]
  40. Raudsepp, R.; Türk, K.; Zarmehri, E.; Joost, U.; Rauwel, P.; Saar, R.; Mäeorg, U.; Dyck, A.; Bron, M.; Chen, Z.; et al. Boron and Fluorine Co-Doped Graphene/Few-Walled Carbon Nanotube Composite as Highly Active Electrocatalyst for Oxygen Reduction Reaction. ChemNanoMat 2024, 10, e202300546. [Google Scholar] [CrossRef]
  41. Panomsuwan, G.; Saito, N.; Ishizaki, T. Simple One-Step Synthesis of Fluorine-Doped Carbon Nanoparticles as Potential Alternative Metal-Free Electrocatalysts for Oxygen Reduction Reaction. J. Mater. Chem. A 2015, 3, 9972–9981. [Google Scholar] [CrossRef]
  42. Wang, X.; Chen, Y.; Dai, Y.; Wang, Q.; Gao, J.; Huang, J.; Yang, J.; Liu, X. Preparing Highly Fluorinated Multiwall Carbon Nanotube by Direct Heating-Fluorination during the Elimination of Oxygen-Related Groups. J. Phys. Chem. C 2013, 117, 12078–12085. [Google Scholar] [CrossRef]
  43. Gu, Y.; Fu, H.; Huang, Z.; Lin, R.; Wu, Z.; Li, M.; Cui, Y.; Fu, R.; Wang, S. O/F Co-Doped CNTs Promoted Graphite Felt Gas Diffusion Cathode for Highly Efficient and Durable H2O2 Evolution without Aeration. J. Clean. Prod. 2022, 341, 130799. [Google Scholar] [CrossRef]
  44. Shulga, Y.M.; Tien, T.-C.; Huang, C.-C.; Lo, S.-C.; Muradyan, V.E.; Polyakova, N.V.; Ling, Y.-C.; Loutfy, R.O.; Moravsky, A.P. XPS Study of Fluorinated Carbon Multi-Walled Nanotubes. J. Electron. Spectrosc. Relat. Phenom. 2007, 160, 22–28. [Google Scholar] [CrossRef]
  45. Wang, Y.; Lee, W.C.; Manga, K.K.; Ang, P.K.; Lu, J.; Liu, Y.P.; Lim, C.T.; Loh, K.P. Fluorinated Graphene for Promoting Neuro-Induction of Stem Cells. Adv. Mater. 2012, 24, 4285–4290. [Google Scholar] [CrossRef]
  46. Tressaud, A.; Durand, E.; Labrugère, C. Surface Modification of Several Carbon-Based Materials: Comparison between CF4 Rf Plasma and Direct F2-Gas Fluorination Routes. J. Fluor. Chem. 2004, 125, 1639–1648. [Google Scholar] [CrossRef]
  47. Zhao, F.-G.; Zhao, G.; Liu, X.-H.; Ge, C.-W.; Wang, J.-T.; Li, B.-L.; Wang, Q.-G.; Li, W.-S.; Chen, Q.-Y. Fluorinated Graphene: Facile Solution Preparation and Tailorable Properties by Fluorine-Content Tuning. J. Mater. Chem. A 2014, 2, 8782–8789. [Google Scholar] [CrossRef]
  48. Lee, J.-M.; Kim, S.J.; Kim, J.W.; Kang, P.H.; Nho, Y.C.; Lee, Y.-S. A High Resolution XPS Study of Sidewall Functionalized MWCNTs by Fluorination. J. Ind. Eng. Chem. 2009, 15, 66–71. [Google Scholar] [CrossRef]
  49. Wang, X.; Dai, Y.; Gao, J.; Huang, J.; Li, B.; Fan, C.; Yang, J.; Liu, X. High-Yield Production of Highly Fluorinated Graphene by Direct Heating Fluorination of Graphene-Oxide. ACS Appl. Mater. Interfaces 2013, 5, 8294–8299. [Google Scholar] [CrossRef]
  50. Sun, C.; Feng, Y.; Li, Y.; Qin, C.; Zhang, Q.; Feng, W. Solvothermally Exfoliated Fluorographene for High-Performance Lithium Primary Batteries. Nanoscale 2014, 6, 2634–2641. [Google Scholar] [CrossRef]
  51. Zarei, M.; Salari, D.; Niaei, A.; Khataee, A. Peroxi-Coagulation Degradation of C.I. Basic Yellow 2 Based on Carbon-PTFE and Carbon Nanotube-PTFE Electrodes as Cathode. Electrochim. Acta 2009, 54, 6651–6660. [Google Scholar] [CrossRef]
  52. Yu, F.; Tao, L.; Cao, T. High Yield of Hydrogen Peroxide on Modified Graphite Felt Electrode with Nitrogen-Doped Porous Carbon Carbonized by Zeolitic Imidazolate Framework-8 (ZIF-8) Nanocrystals. Environ. Pollut. 2019, 255, 113119. [Google Scholar] [CrossRef] [PubMed]
  53. Yu, F.; Zhou, M.; Yu, X. Cost-Effective Electro-Fenton Using Modified Graphite Felt That Dramatically Enhanced on H2O2 Electro-Generation without External Aeration. Electrochim. Acta 2015, 163, 182–189. [Google Scholar] [CrossRef]
  54. Li, M.; Lan, H.; An, X.; Qin, X.; Zhang, Z.; Li, T. Highly Efficient Electrosynthesis of Hydrogen Peroxide through the Combination of Side Aeration and Vacuum Filtration Modified Graphite Felt. Appl. Catal. B Environ. 2023, 339, 123125. [Google Scholar] [CrossRef]
  55. Jing, L.; Tian, Q.; Wang, W.; Li, X.; Hu, Q.; Yang, H.; He, C. Unveiling Favorable Microenvironment on Porous Doped Carbon Nanosheets for Superior H2O2 Electrosynthesis in Neutral Media. Adv. Energy Mater. 2024, 14, 2304418. [Google Scholar] [CrossRef]
Figure 1. (a) XRD patterns of GO before and after doping; (b) FTIR spectra of FGO and GO; (c) EDS elemental-mapping images of FGO; (d) Nitrogen adsorption–desorption isotherms and pore-size distributions of FGO and GO.
Figure 1. (a) XRD patterns of GO before and after doping; (b) FTIR spectra of FGO and GO; (c) EDS elemental-mapping images of FGO; (d) Nitrogen adsorption–desorption isotherms and pore-size distributions of FGO and GO.
Catalysts 14 00793 g001
Figure 2. (a) The XPS survey spectra of GO, FGO; (b) the deconvoluted C1s spectra of GO; (c) the deconvoluted F1s spectra; and (d) the C1s spectra of FGO.
Figure 2. (a) The XPS survey spectra of GO, FGO; (b) the deconvoluted C1s spectra of GO; (c) the deconvoluted F1s spectra; and (d) the C1s spectra of FGO.
Catalysts 14 00793 g002
Figure 3. SEM images of (a) pristine GF; (b) FGO-modified graphite felt; and (c) EDS elemental mapping images of FGO/GF.
Figure 3. SEM images of (a) pristine GF; (b) FGO-modified graphite felt; and (c) EDS elemental mapping images of FGO/GF.
Catalysts 14 00793 g003
Figure 4. RRDE polarization curves (solid lines) and H2O2 detection currents (dashed lines) at the catalyst-loaded electrodes in 0.1 M KOH (a) and 0.1 M Na2SO4 (d); calculated H2O2 selectivity in 0.1 M KOH (b) and 0.1 M Na2SO4 (e); calculated number of transferred electrons in 0.1 M KOH (c) and 0.1 M Na2SO4 (f).
Figure 4. RRDE polarization curves (solid lines) and H2O2 detection currents (dashed lines) at the catalyst-loaded electrodes in 0.1 M KOH (a) and 0.1 M Na2SO4 (d); calculated H2O2 selectivity in 0.1 M KOH (b) and 0.1 M Na2SO4 (e); calculated number of transferred electrons in 0.1 M KOH (c) and 0.1 M Na2SO4 (f).
Catalysts 14 00793 g004
Figure 5. (a,d) Faraday efficiency of graphite felt in 1.0 M KOH (a) or 0.1 M Na2SO4 (d); (b,e) current density in 1.0 M KOH (b) and 0.1 M Na2SO4 (e); (c,f) hydrogen peroxide yield in 1.0 M KOH (c) and 0.1 M Na2SO4 (f).
Figure 5. (a,d) Faraday efficiency of graphite felt in 1.0 M KOH (a) or 0.1 M Na2SO4 (d); (b,e) current density in 1.0 M KOH (b) and 0.1 M Na2SO4 (e); (c,f) hydrogen peroxide yield in 1.0 M KOH (c) and 0.1 M Na2SO4 (f).
Catalysts 14 00793 g005
Figure 6. (a,d) Faraday efficiency of graphite felts with different FGO loadings in 1.0 M KOH (a) or 0.1 M Na2SO4 (d); (b,e) current densities in 1.0 M KOH (b) and 0.1 M Na2SO4 (e); (c,f) hydrogen peroxide yield in 1.0 M KOH (c) and 0.1 M Na2SO4 (f).
Figure 6. (a,d) Faraday efficiency of graphite felts with different FGO loadings in 1.0 M KOH (a) or 0.1 M Na2SO4 (d); (b,e) current densities in 1.0 M KOH (b) and 0.1 M Na2SO4 (e); (c,f) hydrogen peroxide yield in 1.0 M KOH (c) and 0.1 M Na2SO4 (f).
Catalysts 14 00793 g006
Figure 7. The stability test of FGO/GF over 5 cycles. Conditions: 1.0 M KOH, 0.3V vs. RHE, 1 mg cm−2 catalyst loadings.
Figure 7. The stability test of FGO/GF over 5 cycles. Conditions: 1.0 M KOH, 0.3V vs. RHE, 1 mg cm−2 catalyst loadings.
Catalysts 14 00793 g007
Table 1. Comparison of H2O2 electro-generation with other studies.
Table 1. Comparison of H2O2 electro-generation with other studies.
CathodepHCurrent Density (A m−2)t (min)[H2O2] (mg·h−1·cm−2)Refs
Graphite felt31323000.11[51]
GF-(4)7101200.74[52]
Modified graphite felt750602.5[53]
gc-GF967606.89[54]
FGO/GF7203206.18This work
FGO/GF13.7166208.91This work
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Hu, J.; Wang, Z.; Chen, Y.; Xu, W. Fluorine-Doped Graphene Oxide-Modified Graphite Felt Cathode for Hydrogen Peroxide Generation. Catalysts 2024, 14, 793. https://doi.org/10.3390/catal14110793

AMA Style

Hu J, Wang Z, Chen Y, Xu W. Fluorine-Doped Graphene Oxide-Modified Graphite Felt Cathode for Hydrogen Peroxide Generation. Catalysts. 2024; 14(11):793. https://doi.org/10.3390/catal14110793

Chicago/Turabian Style

Hu, Junling, Zhaohui Wang, Yongmei Chen, and Wei Xu. 2024. "Fluorine-Doped Graphene Oxide-Modified Graphite Felt Cathode for Hydrogen Peroxide Generation" Catalysts 14, no. 11: 793. https://doi.org/10.3390/catal14110793

APA Style

Hu, J., Wang, Z., Chen, Y., & Xu, W. (2024). Fluorine-Doped Graphene Oxide-Modified Graphite Felt Cathode for Hydrogen Peroxide Generation. Catalysts, 14(11), 793. https://doi.org/10.3390/catal14110793

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop