1. Introduction
Graphene is a single-layer, 2D material made of carbon atoms under a crystalline arrangement [
1]; owing to its large surface area and ease of functionalization through heteroatom-doping, this carbon material is extremely valued for application in chemical reactions. In addition, its low density, high conductivity, chemical stability, and resistance to corrosion make graphene an efficient, durable material for catalytic applications [
1,
2,
3].
Graphene comes in a variety of forms, and each form is distinctively characterized by its peculiar structural, electrical, and optical properties. Graphene nanoribbons (GNR) are a special class of graphene which has gained traction among researchers; GNR have been widely applied in catalytic reactions/processes due to their unique physical-chemical properties [
4]. GNR are thin strips of graphene – measuring just a few nanometers wide (commonly described as 1D material), which can exhibit a metallic or semi-conductor characteristic depending on their width and edge structure [
2,
4,
5]. The suitable properties of GNR make them useful for application in electronics, sensing devices, and energy storage [
2,
4]. GNR can be synthesized through a wide range of methods including chemical vapor deposition [
2], patterning with lithography techniques [
6], or through chemical oxidation of carbon nanotubes [
5,
7]. GNR have been successfully applied alone or in combination with other materials in different electrolyte solutions to catalyze several electrochemical reactions including water splitting [
8], CO
2 reduction reaction [
9,
10], and oxygen reduction reaction (ORR) [
7,
11,
12,
13]. Incorporating heteroatoms like nitrogen, sulfur, or oxygen into carbon materials - a mechanism referred to as doping, provides us with considerable benefits, as it leads to the development of carbon materials with suitable properties that are highly efficient for catalytic applications [
14,
15]. Doping alters the electronic structure of the carbon material, and this can effectively boost its ORR catalytic activity in acidic, neutral, and alkaline media. For instance, incorporating nitrogen into graphene leads to the development of pyridinic nitrogen - a material with high electron density, which increases the catalytic activity of graphene in ORR, particularly in alkaline environments [
15]. Similarly, adding sulfur or oxygen to graphene creates defects or dopants, which enhance the catalytic activity and selectivity of the material [
15,
16,
17]. However, the ORR activity and selectivity exhibited by heteroatom-doped carbon catalysts can undergo dramatic variations depending on the pH of the electrolyte [
18], yet these phenomena remain inadequately understood. Certain doped carbon materials showcase heightened activity in either acidic or alkaline conditions, while others demonstrate remarkable performance across a broad pH range [
18,
19,
20,
21,
22]. These differences can be attributed to divergent behaviors of the reactants (i.e., protons or hydroxide ions and their availabilities) and intermediates. Furthermore, heteroatom-doped carbon catalysts may present a multitude of enriched reversible redox couples on their surfaces, serving as conduits for charge transfer to adsorbed oxygen [
19]. The electrochemical activity or inactivity of these redox couples toward the ORR is contingent upon the specific electrolyte employed [
19]. Doping may also help improve the stability and durability of carbon materials, as is the case of nitrogen-doped carbon materials which exhibit greater resistance to oxidative degradation and longer lifetime when applied in catalytic reactions/processes [
23].
Recently, considerable attention has been devoted toward creating nitrogen, oxygen, sulfur, and phosphorus-doped graphene structures targeted at enhancing the electrocatalytic performance of the carbon materials, especially in ORR processes. Regarding the development of ORR electrocatalytic materials, although a plethora of studies have been reported in the literature, the study conducted by Xiang et al. [
24] appears to be outstanding. Xiang et al. [
24] found that N-doped GNR, produced through the oxidative unzipping of CNTs and N-doping with urea, and its abundant edges have a synergistic effect on the ORR process through the 4-electron pathway; these authors also pointed out that pyridinic and graphitic-N are the main contributors when it comes to catalyzing the ORR process. Based on the application of the hard-templating method, Dong et al. [
25] produced porous graphene foam doped with B, N, and P, which effectively enhanced ORR performance in comparison with graphene foam doped with single or dual-elements. Han and Chen [
26] showed that keeping the N proportion more than twice higher than the proportion of P in graphene co-doped with P and N (G-PN
3) improved the catalytic activity in ORR and selectivity under the 4 electron mechanism. Zhao et al. [
27] found that doping ordered mesoporous carbon (OMC) first with P and later with N led to an increase in the graphitic-N ratio and improved the catalytic activity in ORR. Yang et al. [
28] developed and applied edge-rich graphene nanoribbon co-doped with N and S via thermal annealing in ORR which led to significant improvements in catalytic activity. Wang et al. [
29] used a 3-D structured carbon nanotube/GNR co-doped with N and S from the thermal treatment of thiourea which effectively catalyzed a 4-electron ORR process. Kan et al. [
30] synthesized carbon nanospheres doped with N and N and S (NCSs and NSCSs) through melanosome pyrolysis with NSCS; the application of these carbon-doped material resulted in better catalytic activity in ORR aimed at water production. Li et al. [
31] synthesized 3-D reduced graphene oxide co-doped with N and S(NS-3DrGO) through pyrolysis; the application of this material resulted in better ORR activity with the successful transfer of 4 electrons. Zhai et al. [
32] produced graphene oxide doped with S by using DMSO as solvent and S precursor and the application of the solvothermal method; the authors successfully applied the S-RGO in ORR with good results. Yazdi et al. [
33] produced helical CN
x-GNRs co-doped with N and S (CNx/CS
x-GNRs) through annealing, where the application of the material resulted in efficient ORR activity and 4-electron selectivity. The authors of the aforementioned study attributed the efficient ORR activity to the synergistic effects derived from the co-doping of nitrogen/sulfur and to the helical unzipping mechanism which gave rise to graphene nanoribbons with multifaceted structure; according to the authors, the S co-doping mechanism led to an increase in pyridinic-N groups in the GNR structure – these groups constitute the main active sites which helped enhance ORR activity on the edges [
33]. In another related study, through annealing, Wang et al. [
34] synthesized a co-doped N and P 3-D structured GNR/CNT composite through annealing; the application of the synthesized material as ORR catalyst resulted in a satisfactory performance. Tammeveski and co-workers [
35] employed both wet and dry ball-milling methods to produce sulphur and nitrogen co-doped graphene-based catalysts. These catalysts exhibited a significant presence of pyridinic N in both cases, with the catalysts produced by dry ball-milling being more suitable for facilitating a complete 4e
– oxygen reduction. Additionally, they synthesized silicon carbide-derived carbon (SiCDC) doped with nitrogen and phosphorus moieties using a ball-milling method, which improved the 4e
– ORR pathway by incorporating active sites derived from the nitrogen and phosphorus moieties [
36]. Furthermore, they successfully synthesized N,P-doped SiCDC and CNT with hierarchical pore structures via a ball-milling method, leading to further improvements in the 4e
– ORR pathway [
37]. Dey and co-workers [
38] synthesized a triazine-based covalent organic polymer (Trz-COP) metal-free electrocatalyst with dual-active sites. They also employed a polymer-assisted electrophoretic exfoliation method on graphite to produce graphene–polypyrrole (G-PPy) in dilute acidic medium, followed by a high-temperature treatment to incorporate N atoms into the graphene matrix. This resulted in the formation of an N-doped graphene-PPy (NG-PPy) metal-free catalyst [
39]. Additionally, they synthesized a bis(terpyridine) (hexadentate chelating ligand) with Fe, promoting the formation of FeN
x/C active sites (Fe-N/C
(H,P) electrocatalyst) [
40]. These catalysts demonstrated efficient ORR electrocatalysis, predominantly resulting in the production of water.
Based on the above considerations, it is clear that doping carbon materials with heteroatoms, such as nitrogen, sulfur, oxygen, and phosphorus, leads to the development of carbon materials with suitable properties that have great potential for catalytic applications; strangely though, this technique has still not been fully explored in the literature, especially considering the specific benefits that can be derived from the doping mechanism, depending on the heteroatom, medium, and type of reaction involved. Finding an electrocatalyst that can efficiently produce water or hydrogen peroxide through oxygen reduction reaction (ORR) in different electrolyte solutions remains a daunting challenge today. In the present work, we developed and characterized a series of N, S, and P-doped and co-doped carbon catalysts using the same GNR matrix and thoroughly evaluated the impact of doping on the catalytic activity in ORR and selectivity under acidic, neutral, and alkaline conditions using the rotating ring-disk electrode (RRDE) technique. The results obtained from the characterization of the proposed materials showed that the heteroatom doping process did not significantly alter the GNR structure, though it altered the surface chemistry which was caused by the insertion of the heteroatoms and oxygen depletion. Among the dopants investigated in this study, N was the most easily inserted and detected in the carbon matrix. The electrochemical analyses conducted revealed that heteroatom doping or co-doping and residual oxygen levels influenced the physicochemical properties of the catalyst, as well as the ORR activity and selectivity, which varied with changes in the electrolyte solution. The study provides significant insights to readers in the relevant field, including: 1) the matrix (GNR) used for doping demonstrates limited ORR electrocatalytic performance, 2) the soft doping procedure employed does not substantially alter the matrix. Consequently, this enables a thorough investigation of the doping effects on ORR activity and selectivity in various electrolyte solutions.
2. Experimental Part
2.1. Reagents and Instruments
The chemical compounds used for the experiments were as follows: phosphorus pentoxide (P2O5, 99 %), potassium persulfate (K2S2O8, 99 % purity), hydrogen peroxide (H2O2, 30 %), - all were acquired from Vetec (Duque de Caxias, RJ, Brazil). Potassium permanganate (KMnO4, 98 %) was acquired from Nuclear (Diadema, SP, Brazil). Potassium sulfate (K2SO4, 99%) was obtained from Proquimios (Rio de Janeiro, RJ, Brazil). Sodium nitrate (NaNO3), hydrochloric acid (HCl), potassium hydroxide (KOH) and sulfuric acid (H2SO4, 98%) were obtained from Merck (Darmstadt, Germany). Ammonia hydroxide (NH4OH) solution (28 wt% in H2O), hydrazine sulfate (NH2NH2.H2SO4, 99%), and multi-walled carbon nanotubes (≥ 98 % purity, containing 6–8 tube walls) were purchased from Aldrich (Saint Louis, MO, USA). The water used for the preparation of all solutions was obtained from the Gehaka reverse osmose equipment, with resistivity above 18 MΩ cm and temperature of 25 °C.
A three-electrode glass electrochemical cell was used for the experiments; the cell was composed of a carbon paper HCP030N (geometric area of 3.5 cm2) which was used as counter electrode, a reversible hydrogen electrode which was used as reference electrode, and Teflon-embedded glassy carbon (GC) disk/Pt ring rotating electrode which was used as working electrode (the disk and ring had geometric area of 0.196 cm2 and 0.11 cm2 , respectively, with a collection efficiency of N=0.26 – this information was obtained from the manufacturer - Pine Research Instrumentation).
2.2. Electrode Preparation
The RRDE (GC disk/Pt ring) electrode was polished with alumina paste (1 µm) and cleaned by sonication, alternating in ultrapure water (Gehaka, resistivity > 18 MΩ cm), isopropyl alcohol, and 0.1 M HClO4 (Tedia, suprapure quality) for 5 minutes in each solvent. The GC disk was subjected to 12 scanning cycles at a sweep rate of 50 mVs-1 in the potential range of 0.05 - 1.2 V. The Pt ring was also subjected to 300 scanning cycles at 900 mVs-1 in the potential range of 0.05 -1.2 V. Both cycling experiments were performed in 0.1 M HClO4 solution saturated with N2 (acquired from White Martins, 4.0 of purity); subsequently, the GC disk/Pt ring was washed with ultrapure water and dried with N2 flow. A loading of 150 μg cm−2 of undoped or doped GNR on the GC disk surface was obtained by dripping 30 μL of 1.0 mg mL−1 aqueous solution of undoped or doped GNR on the disk surface, which was then dried at room temperature.
2.3. Apparatuses, Measurements, and Material Characterization
To perform hydrodynamic linear potential (HLS) and cyclic voltammetry (CV) scanning analyses, a bipotentiostat AFP2 WaveDriver 20 − galvanostat (Pine Research Instrumentation) was used – the equipment was connected to AFMSRCE speed modulated rotator.
A PGSTAT128N potentiostat-galvanostat (Autolab) equipped with the FRA2.X module was used during the electrochemical impedance spectroscopy (EIS) experiments. EIS measurements were performed at an open circuit potential with average values of 0.72, 0.89 and 0.82, while the working electrode was placed in the presence of 0.5 M H2SO4, 0.1 M K2SO4, and 0.1 M KOH solutions, respectively, in the frequency range of 10 mHz to 100 kHz, with disturbance potential of 10 mV (rms). The ohmic drop resistance, adjusted from a high frequency EIS intercept, was used to correct each HLS curve. The measured ohmic drop resistances were on average 5.2, 36.8, and 38.5 Ω in 0.5 M H2SO4, 0.1 M K2SO4 and 0.1 M KOH solutions, respectively.
The Fourier transform infrared spectroscopy measurements were carried out using Bomen Fourier transform infrared spectrophotometer with a spectral window of 400 to 4000 cm-1. The catalyst samples were produced by grinding the dried catalyst powder (approximately 30 µg) with potassium bromide (approximately 35 mg) in a mortar until a fine and homogeneous powder was obtained. After pressing the powder, a translucent tablet was formed.
The Raman spectra were obtained at room temperature using LabRam HR Evolution micro-Raman spectrometer from Horiba Jobin-Yvon, a solid-state laser operating at 633 nm, a standard grating (600gr/mm), and EMCCD detector (Synapse EM). To avoid overheating and the occurrence of photochemical phenomena, the samples were excited with low intensity laser (ca. 2 mW). The laser was focused on the sample using a 100 objective (Olympus, MPlan N). The spectra were collected over the course of 12 seconds.
Elemental analysis (EA) was performed using a Scientific Flash 2000 CNHS/O Elemental Analyzer Thermo Equipment under cycle operating conditions (run time) of 720 s and oven temperature of 950°C for CHNS determinations and under the cycle (run time) of 400 s and oven temperature of 1060°C for O determination.
The catalyst nanostructures were characterized by TEM using JEM 2100F (JEOL) or Philips CM200, both operating at 200kV. The XPS analyses were performed using the PHI Quantera II surface analysis equipment. The Al Kα line (1486.6 eV) was used as the ionization source, operating at 15 kV and 25 W. After performing the background subtraction, the spectra were deconvoluted using a combination of Lorentzian (30%) and Gaussian Voigt (70%) functions.
Thermogravimetric analyses were conducted using Shimadzu TGA-50 thermogravimetric analyzer, with FID Synthetic Air 5.0 gas flow of 50 mL min-1, in temperatures ranging from ambient to 900 °C and a heating rate of 10 °C min-1; the samples were placed in alumina ceramic crucibles.
2.4. GNR Synthesis and Syntheses of Different Doped GNR Samples
The synthesis of graphene nanoribbons (GNR) derived from the chemical oxidation of multi-walled carbon nanotubes (MWCNTs) was carried out utilizing the method previously described elsewhere [
7]. Briefly, MWCNTs were dispersed in concentrated H
2SO
4 solution containing K
2S
2O
8 and P
2O
5 and then heated (with stirring) at 80 °C for 6 h. The resulting solution was diluted in water at 0 °C and filtered through a 0.22 µm nylon membrane and finally washed with plenty of high purity water. This pre-oxidized material was re-oxidized in a concentrated H
2SO
4 solution containing NaNO
3 and KMNO
4 at 30°C during 2 h, then diluted in high purity water, and then H
2O
2 (30% solution) was added and again diluted with water. After 24 h, this dispersion was centrifuged and washed with HCl and water (10:90) resulting in graphene oxide nanoribbons (GONRs). The GONRs were immersed in hydrazine sulfate and NH
3OH solution under vigorous stirring, followed by heating to 95 °C during 2 h. After reaching room temperature, the dispersion was filtered initially with a 5% NH
3OH solution, washed well with enough high purity water, finally producing the GNRs.
The doping of GNR with N and/or S and/or P was performed using NH
4OH and/or N
2H
6SO
4 and/or K
2S
2O
8 and/or P
2O
5; the doping was carried out in two hydrothermal synthesis steps. The first step was executed as summarized in
Table 1. This step involved mixing 0.02g GNR with defined amounts of the dopant or dopants (NH
4OH and/or N
2H
6SO
4 and/or K
2S
2O
8 and/or P
2O
5) and pouring water in a 30 mL beaker; this was followed by heating and constant stirring for 3 hours, and finally washing/centrifugation several times.
In the second step of synthesis, the product of step 1 was mixed with 30 mL of water or NH
4OH, together with the respective amount of dopant or dopants (
Table 1) and subjected to constant magnetic stirring until complete dispersion was obtained. The resulting solution was heated in an autoclave system at 150°C for 12 h. The final products were washed by centrifuging using ultrapure H
2O and dried at 60°C for 24 h (
Figure 1).