Next Article in Journal
Natural Compounds and Health Benefits of Ganoderma capense
Previous Article in Journal
Pressure Sensitivity of UiO-66 Framework with Encapsulated Spin Probe: A Molecular Dynamics Study
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Metal-Free Graphene-Based Derivatives as Oxygen Reduction Reaction Electrocatalysts in Energy Conversion and Storage Systems: An Overview

Institute of Condensed Matter Chemistry and Technologies for Energy, ICMATE, National Research Council of Italy, CNR, Corso Stati Uniti, 4, 35127 Padua, Italy
Molecules 2025, 30(10), 2248; https://doi.org/10.3390/molecules30102248
Submission received: 7 March 2025 / Revised: 30 April 2025 / Accepted: 2 May 2025 / Published: 21 May 2025
(This article belongs to the Section Materials Chemistry)

Abstract

:
Oxygen reduction reaction (ORR) is one of the most important reactions in electrochemical energy storage and conversion devices. To overcome the slow kinetics, minimize the overpotential, and make this reaction feasible, efficient, and stable, electrocatalysts are needed. Metal-free graphene-based systems are considered promising and cost-effective ORR catalysts with adjustable structures. This review is meant to give a rational overview of the graphene-based metal-free ORR electrocatalysts, illustrating the huge amount of related research developed particularly in the field of fuel cells and metal–air batteries, with particular attention to the synthesis procedures. The novelty of this review is that, beyond general aspects regarding the synthesis and characterization of graphene, above 90% of the various graphene (doped and undoped species, composites)-based ORR electrocatalysts have been reported, which represents an unprecedented thorough collection of both experimental and theoretical studies. Hundreds of references are included in the review; therefore, it can be considered as a vademecum in the field.

Graphical Abstract

1. Introduction

The oxygen reduction reaction (ORR) is one of the processes at the base of the functioning of energy conversion and storage systems such as fuel cells and metal–air batteries, and the key energy conversion reaction involved in the production of hydrogen peroxide.
The electrochemical reduction of molecular oxygen can proceed through two or four-electron transfers in aqueous electrolyte. In alkaline and neutral media, the reaction mechanism can be described as follows [1]:
Four-electron pathway: O2 + 2H2O + 4e- → 4OH- E = 0.401 V
Two-electron pathway: O2 + H2O + 2e- → HO2-+ OH-E = −0.065 V
HO2- + H2O + 2e- → 3OH- E = 0.867 V
2HO2- → 2OH- + O2
In acidic electrolyte, the final products are H2O and H2O2 for the four and the two-electron pathway, respectively.
Four-electron pathway: O2 + 4H+ + 4e- → 2H2O E = +1.229 V
Two-electron pathway: O2 + 2H+ + 2e- → H2O2 E = +0.67 V
H2O2 + 2H+ + 2e- → 2H2O E = +1.77 V
This reaction occurring at the cathode, however, presents a high activation energy related to the strong bond energy (498 kJmol−1) of the O=O bond, so that the improvement of kinetics and reduction in overpotentials require the utilization of electrocatalysts lowering the energy barrier for the bond activation and cleavage.
The electrochemical assessment of the ORR electrocatalyst performance occurs through the determination of key kinetic parameters of ORR, such as onset potential, half-wave potential, kinetic current density, mass and specific activity, Tafel slope, and conductivity using conventional voltammetry techniques, while the Rotating Disk Electrode (RDE) and rotating ring-disk electrode (RRDE) techniques under hydrodynamic conditions analyze the influence of kinetics and the mass-transfer effect on ORR [2]. Additionally, the number of electrons transferred (n) can be preliminarily obtained by the Koutecky–Levich equation shown below:
1 I = 1 I K + 1 0.62 FnAB v 1 / 6 D 2 / 3 ω 1 / 2 C O 2
where I is the reaction current; Ik is the kinetic current; n is the number of exchanged electrons per O2 molecule; F is Faraday constant; A is the electrode area; D is the diffusion coefficient of O2 in the electrolyte; v is the kinematic viscosity of the electrolyte; CO2 is the bulk concentration of O2; and ω is the rotation rate of the electrode [3].
To date, the most commonly used ORR electrocatalysts are platinum-based catalysts consisting of Pt or Pt alloy nanoparticles supported on porous carbon species. However, despite their high activity, platinum-based catalysts present some drawbacks related to the high cost of platinum and limited stability under operating conditions.
Therefore, research has moved to other noble and non-noble metal-based catalysts as well as metal-free species, particularly metal-free carbon materials.
Among the various forms of carbon materials, graphene, a monolayer of graphite containing a honeycomb structure of carbon atoms, represents an emerging material generating huge interest because of its enormous surface area (2630 m2g−1), excellent electrical conductivity (106 Scm−1), good thermal conductivity (∼5000 Wm−1K−1), high charge mobility (200,000 cm2V−1s−1), great mechanical strength (breaking strength of 42 Nm−1 and Young’s modulus of 1.0 TPa), low optical absorbance (2.3%) and density (<1 gcm−3), and unusual flexibility [4].
For electrochemical applications, graphene-based electrodes are expected, theoretically, to react much faster, showing that graphene has faster electron mobility in comparison to all other potential materials [5].
Owing to the unique features of graphene, a great deal of research has been focused on its electrochemical applications as a promising material for electrodes, particularly in fuel cells [6], aqueous metal–air batteries [7], and lithium batteries [8,9,10].
Fuel cells, which produce energy upon the electrochemical oxidation of a fuel and reduction of oxygen, have been developed to respond to the increasing global energy demand facing the increasing scarcity of fossil fuels, and to meet the requirement for environmental sustainability [11].
Among fuel cells, polymer electrolyte membrane fuel cells specifically operating at low temperature (below 120 °C) convert the energy of a fuel such as hydrogen or methanol into electricity with low or zero emissions and high efficiency. They can be classified in terms of the electrolyte in the cell [12] into Proton Exchange Membrane Fuel Cell (PEMFC) (Figure 1A) or Anionic Exchange Membrane Fuel Cell (AEMFC) (Figure 1B). In acidic electrolytes, proton transfer from the anode to the cathode is promoted, while in alkaline electrolytes, hydroxyl ion is transferred from the cathode to the anode, so that the type of components, such as the catalyst layers and the membrane, may differ depending on the electrolyte.
A special type of polymer electrolyte membrane fuel cell, very attractive for its environmental applications, is the microbial fuel cell (MFC), which converts the chemical energy of organic matter such as carbohydrates and organic acids into electricity directly using microorganisms. The wastewater is used as fuel for producing electrical power (Figure 2) [14].
MFCs can, therefore, be applied for wastewater treatment with the double benefit of recovering energy from wastewater as well as reducing the biochemical oxygen demand and excess sludge production.
Besides fuel cells, which are important energy conversion systems, other power supply devices include the aqueous metal batteries, which constitute an emerging energy storage technology for renewable energy. In particular, multivalent ion batteries, such as zinc (Zn), calcium (Ca), aluminum (Al), and magnesium (Mg) ion batteries, are of particular interest, considering the natural abundance of Al, Ca, Mg, and Zn in the Earth’s crust, their low redox potentials, and relevant energy density due to the transfer of several electrons per metallic cation in the electrochemical reactions involving the metals [15].
The basic operating scheme is presented in Figure 3; during the discharge process, the metal releases electrons and metal hydroxide is formed; electrons are supplied to the cathode for the oxygen reduction.
The overall reaction occurring in the battery is as follows:
M + n 4 O 2 + n 2   H 2 O M(OH) n
Hence, on the basis of the reported electrochemical applications where the oxygen reduction reaction is a fundamental process, the goal of this review is to present the development of metal-free graphene as ORR electrocatalysts, comprehending sections describing the synthetic methods, specific characterizations, and the doped and undoped graphene-based specifically as ORR electrocatalysts, ECs. Tables containing ORR parameters such as onset potential (Eonset) and catalyst loading of the electrocatalysts in alkaline, acidic, and neutral electrolytes are also presented. A list of the used abbreviations is provided in the supporting information.

2. Graphene Synthesis Methods

Since the first synthesis of graphene via mechanical exfoliation (repeated peeling) of small mesas of highly oriented pyrolytic graphite (HOPG) [17], the huge interest in such a material has pushed research in finding preparation methods that can produce a low-cost graphene in good yields and in the most effective way in terms of composition, morphology, structure, and properties such as stacked layer, lateral size, defect and impurity contents, surface chemistry, and solubility, as well as electrical and thermal conductivities.
Several techniques have been used, depending on the synthetic strategy, which can be based on a “top-down” or a “bottom-up” approach.
The top-down technique proceeds as follows: (i) isolation of graphene from the stacked parent materials by solid-phase, liquid-phase, or electrochemical exfoliation of pristine graphite and graphite intercalated compounds, and (ii) reduction of graphene oxide (GO) obtained via exfoliation of graphite. The bottom-up approach involves building up graphene from molecular precursors, typically including arc discharge, chemical vapor deposition (CVD), epitaxial growth, on-surface synthesis, laser irradiation, and pyrolysis [18].
In general, top-down methods are more advantageous from the viewpoints of scalability and cost-effectiveness, whereas bottom-up processes are appropriate for producing large-area, high-quality films [19].
Some synthesis routes possessing different scalability and generating graphene with different characteristics are listed in Table 1, including their advantages and disadvantages [20].
Table 1. Summary of techniques for synthesizing graphene.
Table 1. Summary of techniques for synthesizing graphene.
ApproachTechnique and Brief DescriptionAdvantage/Disadvantage
Top-downMicromechanical exfoliation: exfoliation of layers from graphite by Scotch tapeLow output but simple process with high quality
Chemical reduction of graphite derivative: exfoliation of graphite derivative in suspension and chemical reduction Large-scale production with significant defects
Bottom-upChemical vapor deposition: decomposition of hydrocarbon on a metal substrate at high temperatureLarge area graphene, but very poor yield
On-surface synthesis: covalent fusion of molecular building blocks onto a metallic surface Atomic precision in regularly ordered porous graphene sheets, nanographenes, and graphene nanoribbons, but poor scalability
Laser irradiation: conversion of carbon-based precursors induced by laser Fast, low-cost, and energy-saving process, but inhomogeneity and difficult control of morphology
Pyrolysis: thermal conversion of carbon-based precursorsLarge-area monolayer graphene films onto a variety of substrates. Limited yield, energy consumption
Epitaxial growth: evaporation of SiC on Si-wafer at high temperatureMicron-length graphene with few defects, but poor scalability and costly method

2.1. Top-Down Graphene Synthesis

In the top-down approach, the final graphene/heteroatom-doped graphene is obtained starting from a precursor that already includes one-atom-thick layers, such as HOPG and graphene oxide, through both physical and chemical exfoliation processes, and, where necessary, chemical reduction.
Mechanical exfoliation of pristine graphite using Scotch tape produced high-quality graphene layers with good crystalline structure, low defect density, and high electrical conductivity, but at low yields [13].
Since large-scale exfoliation is needed, more effective techniques were developed, including liquid exfoliation based on the use of organic solvents like N-methyl-pyrrolidone (NMP) and dimethylformamide (DMF): in this way, the energy required to exfoliate graphene could be balanced by the solvent–graphene interaction for solvents whose surface energies match that of graphene [21]. It was also observed that further improvement of the yield was obtained if exfoliation of graphite in polar solvents is assisted by other processes, such as a solvothermal process. Qian et al. [22] readily produced monolayer and bilayer graphene by solvothermal-assisted exfoliation and dispersion in acetonitrile (Figure 4), followed by centrifugation in 10–12 wt% yield by using expanded graphite (ExG) as the starting material.
In fact, expanded graphite has the advantage of a larger interlayer spacing than pristine graphite, which weakens the van der Waals interactions and facilitates exfoliation [23]. However, because of several drawbacks in using polar organic solvents (e.g., toxicity, hazard, etc.), research has been advanced to develop water-based processes using surfactants [24] such as sodium dodecylbenzene sulfonate [25], sodium cholate [21], and alkaline lignin [26], or ionic liquids [27] such as 1-butyl-3-methylimidazolium peroxydisulfate [28] and 1-ethyl-3-methylimidazolium tetrafluoroborate [29], as well as butyltrimethylammonium bis(trifluoromethylsulfonyl)imide [30].
Ionic liquids have also been used in the fabrication of graphene nanosheets via electrochemical exfoliation of graphite [31,32]. This process is considered a potentially scalable method for synthesizing graphene in a relatively short time at ambient temperature, generating gram-scale quantities of graphene materials with tunable quality using operating voltages, graphite precursors, and electrolytes [33].
Natural graphite [31,34] and synthetic graphite [32,35] can be used in different forms (powder [36], foil [37], rod [38,39], flake [40], flexible sheet [41], or plat [42,43]) and the electrolytes can be a mixture of water and ionic liquids [44,45], aqueous solution of inorganic salts [46], alkaline electrolyte [47], or mineral acids [47]; under such conditions, graphene is obtained via anodic exfoliation of the precursors. On the other hand, cathodic exfoliation of graphite occurs in organic solvents, such as propylene carbonate [48], dimethyl sulfoxide [49], and N-methyl-2-pyrrolidone [50], containing lithium or alkylammonium salts.
An extension to liquid-phase exfoliation consists of using supercritical fluids (SCFs), which have both gas- and liquid-like physicochemical properties: at the supercritical state, i.e., when temperature and pressure are above the critical points, SCFs have near-zero surface tension, low viscosity, high solvating power, and diffusion coefficients, which are advantageous for penetrating graphite’s interlayer spaces and facilitating exfoliation. Compared to ordinary liquids, the density and solvent strength of the SCF can be tuned by changing pressure and/or temperature.
The choice of SCF is crucial as the interaction between solvent and graphite layers should be sufficient to offset the van der Waals attraction between graphitic layers, thereby making the graphene layers loose [51]; actually, intercalation and exfoliation under SCF conditions imply rapid penetration of solvent molecules through the interlayers of graphite, achieving delamination of graphitic materials.
SCFs comprehend supercritical organic solvents, like DMF, NMP, and alcohols [52], and dimethyl sulfoxide [53]. Tomai et al. [54] prepared nanographene with a diameter of less than 100 nm from platelet carbon nanofibers as an alternative to graphite in supercritical ethanol. More recently, water, whose use would be advantageous due to its low cost, minimal safety concerns, and relatively low environmental impact, has been utilized alone in supercritical conditions as reported by Ibarra et al. [55], who prepared large graphene layers with sizes of around 5 μm from graphite powder. Most graphene exfoliation using supercritical fluids has been carried out using CO2 since first synthesis by Pu et al. 2010 [56,57,58]. Supercritical CO2 is widely used as a nonflammable, nontoxic, environmentally friendly solvent with an easily accessible critical point (TC = 31.1 °C and PC = 73.8 b). However, the non-polar nature of CO2 and the rapid escape of CO2 molecules from the intercalated layers make exfoliation of graphite with ScCO2 alone less effective than exfoliation with organic solvents [51].
Therefore, various enhancements were applied to improve the quality and yield of exfoliated graphene using the scCO2 exfoliation process, including mixing CO2 with liquids [59,60] or using assistant methods such as ultrasonic treatment [61], rotor/stator mixing [62], and ball milling [63].
It is also interesting to report that exfoliation and N doping can be achieved simultaneously using supercritical ammonia (scNH3) [64] (Figure 5).
N-doped graphene consisting of a few layers (<5) with a sizeable lateral size was produced with a 40% yield.
In addition to physical methods, exfoliation of graphite can also be realized chemically to produce graphene oxide [65] characterized by various oxygen moieties (Figure 6), which in turn can be chemically converted to reduced graphene oxide (rGO), with restoration (albeit partially) of the sp2 carbon configuration in the graphene state.
This method is unique and attractive because single-layer graphene can be produced on a large scale and at a relatively low cost [67]. Graphene oxide is therefore an important precursor that can be obtained from graphite using suitable oxidizing agents, which mainly consist of KMnO4 and NaNO3 in H2SO4 solution according to the well-known Hummers’ method developed in 1958 [68].
However, toxic gases such as NO2 and N2O4 are released during this highly efficient procedure, and the residual Na+ and NO3- ions are difficult to remove from the wastewater formed from both the GO synthesis and purification processes.
In 2010, Marcano et al. [69] eliminated NaNO3 and introduced H3PO4 in the oxidizing system since graphene oxide nanoribbons were obtained from multiwalled carbon nanotubes and KMnO4 with more intact graphitic basal planes upon addition of H3PO4 [70].
An oxidation procedure for graphite flakes was developed using a larger amount of KMnO4 and a 9:1 mixture of concentrated H2SO4/H3PO4 (Figure 7). This is called the “improved method”, giving “improved” graphene oxide (IGO); graphene oxide was also prepared using the classical Hummers’ method (HGO) and a modified Hummers’ method (HGO+).
As can be seen in the figure, a very small amount of under-oxidized material is produced with the IGO method, indicative of its increased efficiency with respect to the other methods. Moreover, the reaction is not largely exothermic, and no toxic gas is produced. This IGO-generated product is more oxidized and possesses a more regular structure than the other materials produced using HDO and HGO+.
Besides the use of a greater amount of KMnO4 instead of NaNO3 [71,72], further modification of Hummers’ method was accomplished by introducing a pre-oxidation step before KMnO4 oxidation (in the absence of NaNO3) [73] or by replacing KMnO4 with K2FeO4 while NaNO3 was removed [74,75]. Yu et al. [76] addressed the problems of high consumption of oxidants, intercalating agents, and time-consuming procedure by partly replacing KMnO4 with K2FeO4 and using a smaller amount of concentrated sulfuric acid, according to the scheme in Figure 8, offering new possibility for the production of GO in an economical, eco-friendly, and efficient way.
Such a thorough study on graphene oxide preparation is strictly related to its important role as a starting material in the production of large quantities of rGO, which can also be considered as chemically derived, converted, or modified graphene closely resembling pristine graphene in both structure and properties [77]. Reduced graphene oxide is obtained via direct reduction of GO through both chemical and thermal methods: the removal of the oxygen moieties is a vital topic that determines the properties of the ultimate product [78]. The reduction efficiency largely depends on the methods and processing parameters, resulting in graphene with various properties [77,79,80]. H2S was first used as a reducing agent in 1934 [81]; since then, a large number of reductants (Scheme 1) [82] have been used, among which hydrazine gave rGO with particularly improved electrical and structural properties, resembling pristine graphene to a large extent [83].
In addition to hydrazine [84,85], hydrogen sulphide [86] and metal borohydrides such as NaBH4 [87,88], hydrohalic acids [84,85,89,90,91,92], and various alkaline solutions [93,94,95] are used as reductant agents.
However, to reduce toxic chemicals and complicated manufacturing procedures, researchers introduced green reduction agents comprehending sugars (saccharides) [96,97,98], ascorbic acid (Vitamin C) [96,97,98,99,100], and micro-organisms such as bacteria [101,102], as well as amino acids [103,104,105,106] and plant extracts [107,108,109,110]; ascorbic acid, in particular, has produced rGO with electrical conductivity up to 800 Sm−1, which is almost comparable to that of rGO prepared using hydrazine [111].
Mature alternatives to wet-chemical reduction methods are represented by the solvothermal process as well as the hydrothermal process; they offer the opportunity to modify the physico-chemical properties of the solvent and overwhelmingly address problems of scalability and environmental concerns [112].
The hydrothermal technique has been largely used in the reduction of graphene oxide because it offers the advantages of being a water-soluble reaction, and the ionic product of water is high. Under hydrothermal conditions, the protonation of ·OH functionalities originates dehydration on the edges or basal planes of GO, followed by the π-bonding restoration [113,114,115,116,117,118,119,120].
In 2010, Long et al. [121] prepared nitrogen-doped graphene sheets through hydrothermal reduction of colloidal dispersions of graphite oxide in the presence of hydrazine and ammonia at a pH of 10 via multiple reactions such as ring opening of epoxides induced by hydrazine as well as reorganization of the carbon skeleton with incorporation of nitrogen (Figure 9).
The reduction process can also be achieved by thermal treatment (thermal annealing) in inert or reductive atmospheres at temperatures above 200 °C [122,123,124,125,126]; with this method, oxygen functionalities are removed in the form of water, carbon dioxide, and carbon monoxide.
Thermal annealing is a well-established process for the preparation of heteroatom-doped graphene using GO and suitable molecules as the heteroatom source. Sheng et al. [127] reported a facile and catalyst-free approach for the synthesis of nitrogen-doped graphene by thermal annealing of graphite oxide with a low-cost industrial material, melamine, as a nitrogen source.
The process reported in Scheme 2 consists of the adsorption of melamine into graphite oxide layers at T < 300 °C (1); formation of carbon nitride at T < 600 °C (2), and decomposition into graphene layers at T > 600 °C (3).
Such methods can also be extended to prepare other heteroatom (such as B, P, S, Se, Si and I)-doped graphene nanosheets if appropriate precursors are used, e.g., boron oxide (B2O3), benzyl disulfide, and triphenyl phosphine as the sources of B, S and P elements, respectively [128], diphenyl diselenide [129], triphenyl silane for Si [130] and I2 for I [131].
This methodology has been largely applied to synthesize suitable metal-free graphene-based ORR electrocatalysts. It takes advantage of having different compounds as heteroatom sources and regulating the temperature and atmosphere to conveniently introduce the heteroatom into graphene in order to modulate the electrochemical properties.

2.2. Bottom-Up Graphene Synthesis

The typical “bottom-up” production techniques are arc discharge, CVD on active metals, hydro-/solvothermal synthesis, on-surface synthesis, laser irradiation, and pyrolysis of carbon sources.
In order to obtain high-quality graphene on a large scale under controllable conditions, the arc discharge technique used to vaporize elemental carbon to form nanotubes [132], fullerenes [133], and onion structures [134] has been also applied for synthesizing graphene sheets [135].
In the classic arc discharge set-up illustrated in Figure 10 [136], an arc is generated between two graphite rods upon application of a high voltage, with the anode sprayed up to the atomic state. Formation of carbon structures occurs via the condensation of carbon vapor cooled in a buffer gas.
Separation of the graphite layers into graphene sheets proceeds via high-temperature exfoliation, or the graphite layers can even get broken into small clusters/atoms, which then form graphene sheets via catalytic growth, demonstrating that two mechanisms exist in an arc discharge process simultaneously.
Among the process parameters of arc discharge (arc current/gap voltage, type and ratio of buffer gas, ambient media, cooling conditions, external fields and catalysts) influencing graphene formation in terms of layer numbers, purity, crystallinity, morphology, etc., investigations have been carried out on the buffer gas since arc plasma is enhanced by the ionization of buffer gas.
The use of H2, Ar, CO2, air, NH3, O2, or various mixtures [135,137,138,139,140,141,142,143] as buffer gases could produce graphene sheets of integral structure, high thermal stability, and good dispersibility in organic solvents [142,144].
First studies were conducted using hydrogen (pure or mixed with He, N2 or, Ar) evidencing its key role in the formation of graphene and its etching effect on amorphous carbon; in arc discharge, the dangling carbon bonds are terminated with hydrogen, the rolling of sheets into nanotubes and graphitic polyhedral particles is minimized, and hydrogen also imparts a high temperature on plasma with in situ defect-healing effect [145,146,147].
Carbon dioxide mixed with helium is a better and safer buffer gas. Wu et al. [135] produced tens of grams of high-quality graphene sheets with four to five layers in minutes under optimized conditions consisting of a low voltage (<35 V), high buffer gas pressure (>1270 Torr), high current (about 150 A), and 25–40% (v/v) CO2 in the helium–carbon dioxide mixture. Moreover, the graphene sheets showed excellent solution processability, makingpossible to directly fabricate a variety of films using the organic solutions as precursors.
Arc discharge synthesis was run in liquid media, too, such as liquid nitrogen [148] and water [149,150], which also allows the collection of graphene sheets floating on the water surface.
Other implementations of the arc discharge synthesis in promoting the nucleation, growth, and separation of graphene were obtained by application of external fields [138,151,152,153] or using catalysts such as ZnO, ZnS, Cu [154,155], TiO2 [155], and Ni–Y2O3 powder [152].
Liu et al. [154] combined flake graphite as a carbon source, Cu as a catalyst, and melamine as a nitrogen source and simultaneously synthesized N-doped graphene sheets with nitrogen content as high as 4.92 at% and undoped graphene sheets in the different parts of the apparatus according to two different mechanisms: catalytic growth mechanism for the cathode-part graphene sheets and an exfoliation mechanism for the wall-part graphene sheets.
Since the first arc discharge synthesis of B and N-doped graphene was reported by Panchakarla et al. [156], this technique has also been used to introduce fluorine [157] and silicon [158] as heteroatoms into the graphene lattice.
Another important technique widely used to obtain large-area, single-crystalline graphene is Chemical Vapor Deposition. In a typical CVD process, a hydrocarbon in the gas phase is introduced as a carbon source for graphene growth; hydrogen is usually required to achieve high-quality graphene [159]. Decomposition and deposition usually occur on transition metal foils (Ni or Cu) [160,161,162,163,164,165] that act as catalysts and supports [165]. The carbon solubility of the catalyst is an important factor for controlling the number of graphene layers, enabling different growth mechanisms (carbon adsorption, absorption, or segregation). In the case of Cu, its low carbon solubility, also at high temperatures, allows the growth of single-layer graphene through an adsorption mechanism.
Given the importance of producing large single crystals for the implementation of graphene into practical applications, the nucleation growth, ∼106 nuclei/cm2 in conventional cases, needs to be significantly decreased [162]. Yan et al. [166] conducted an interesting study where, by constructing a controlled chamber pressure CVD, it was possible to grow single crystal monolayer graphene up to ∼4.5 mm2 on commercial Cu foils from methane and hydrogen; they also showed that Cu pretreatments, electrochemical polishing, and high-pressure annealing are critical for suppressing graphene nucleation site density. The growth mechanism for Cu-based graphene was explored by studying the influence of varied growth parameters on graphene domain sizes (Figure 11).
In the proposed mechanism, the active carbon species from the dissociated CH4 are apt to agglomerate into thermodynamically stable (CnHy)s species on the active sites of the Cu surface, ultimately leading to the formation of graphene nuclei. Once the graphene nuclei are formed, most of the active carbon species are captured and consumed for the growth of graphene, and the growth of larger-size graphene domains during the extended growth time will be enabled.
Another strategy for decreasing the nucleation growth is the passivation of Cu surface active sites by utilizing a growth–etching–regrowth strategy or O2 pretreatment prior to growth [167,168,169,170,171,172,173] so that centimeter-sized single-crystalline graphene can be grown [170].
The CVD process has also been revealed to be an effective approach for achieving substitutional doping without affecting the crystalline nature. N-doping was obtained by introducing gaseous NH3 [174,175,176,177,178,179,180] or N2 [181,182,183,184] in the CVD apparatus. Qu et al. [185] prepared large N-doped graphene films, ca. 4 cm2 in size, by flowing a nitrogen-containing reaction gas mixture (NH3:CH4:H2:Ar = 10:50:65:200 standard cubic centimeters per minute) at 1000 °C over a Ni-coated SiO2/Si wafer. These N-doped graphene films were shown for the first time to have ORR electrocatalytic properties, which determined the development of the CVD technique for the preparation of metal-free graphene ORR electrocatalysts.
As an alternative to ammonia and nitrogen gases, liquid N-containing substances, such as acetonitrile [186], pyridine [187,188,189,190,191], DMF [192,193], urea [194,195], various amines [196,197], and solid species such as melamine [198] and triazine [199], were used as a nitrogen source.
Xue et al. [189] demonstrated a rational route based on self-assembly of pyridine molecules on Cu surface at temperatures as low as 300 °C, producing single-layer, single-crystal, tetragonal-shaped, N-doped graphene domain arrays with typical n-type field-effect transistors (FETs) for N-doped graphene both in air and vacuum and mobilities in the range of 53.5–72.9 cm2V−1s−1.
Analogously, B-doped graphene films can be obtained from gaseous boron sources such as diborane [200] and solid sources such as boric acid [194]. In the last case, Wu et al. [194], by easily adjusting the amount of boric acid and polystyrene as boron and carbon sources, respectively, modulated the boron content from 0.7 to 4.3%, obtaining boron-based graphene FET devices with a mobility of 450–650 cm2V−1s−1.
Boric acid was also used to produce N,B-doped graphene. Bepete et al. [184], using a single-step CVD route with methane, boric acid powder, and nitrogen gas, synthesized large-area graphene, incorporating small BN domains in a facile and safe process that avoided the use of boranes and ammonia. XPS studies evidenced that BN species were incorporated in the graphene in small BN domains where the N in the BN is bonded to carbon, forming B–N–C systems.
The CVD technique was also used to incorporate phosphorus atoms [201,202] in the graphene lattice by using solid triphenylphosphine and sulfur [203,204] via simple and efficient CVD methods. In particular, Hassani et al. [204] grew S-doped graphene with high sulfur content (5 at%, determined by XPS) from sulfur powder and acetylene gas at 600 °C. They produced non-porous sulfur-doped graphenes with high surface areas, where sulfur atoms are successfully incorporated into the carbon matrix of samples with covalent bonds, as revealed by Raman and XPS analyses.
In the synthesis of heteroatom-doped graphene, hydro- and solvothermal processes have been revealed to be proper bottom-up, straightforward synthesis methodologies.
N-doped graphene was prepared by Deng et al. [205] by reacting lithium nitride with tetrachloromethane at 250 °C for 10 h in a stainless-steel autoclave in the presence of nitrogen (sample NG−1) or at 360 °C in the presence of cyanuric chloride (sample NG-2), obtaining N-graphene with nitrogen species content in the range of 4.5–16.4%. Preliminary tests of NG−1 and NG-2 as ORR electrocatalysts showed that they are more active than pure graphene and commercial XC-72, while NG-2 exhibits a higher activity than NG−1 because of higher N species content.
It has also been proposed, for graphene synthesis using lithium nitride in tetrachloromethane at 250 °C, a reaction mechanism that assumes that the graphene structure derives from mutual coupling of intermediates originating from transformation of sp3-hybridized carbon in CCl4 to sp2-hybridized carbon, such as dichlorocarbene, free -C=C-, and -C=N- groups. The small domains of sp2-hybridized carbon containing N are obtained through further dechlorination and then grow into N-doped graphene sheets (Scheme 3).
The strategy of using CCl4 as solvent and as carbon source at the same time has also been successfully applied in the solvothermal synthesis of B-doped graphene [206] and dual-doped N,X (X = B, P, or S) graphene [207,208].
The fundamental step is the Wurtz-type reactive coupling reaction of CCl4 with potassium, which involves a bottom-up process for rapid preparation of high-quality pristine graphene [207]. In such a reaction, carbon radicals and potassium chloride are formed, then the carbon radicals form carbon–carbon (C-C) bonds on the surface of potassium particles to give graphene (Figure 12).
Doping of graphene occurred as the above-mentioned doping agents were present in the autoclave.
As an alternative to using alkaline metals, magnesium powder was used for the hydrothermal synthesis of N-doped graphene from hexamethylenetetramine, which was pyrolyzed in water at 500 °C for 20 h [209].
Another synthetic method starting from molecular precursors is the on-surface synthesis of porous graphene, which results in the collection of graphene-related materials with nanopores in the plane [210,211].
Thanks to the fabrication of regular two-dimensional polyphenylene networks with single-atom-wide pores and sub-nanometer periodicity by Bieri et al. [212], it was possible to develop a process based on surface-assisted coupling (generally Ullman reaction) and cyclodehydrogenation of specifically designed molecular building blocks consisting of variously condensed and covalently bonded benzene-ring-based systems [213,214].
This method generally produces graphene nanoribbons that undergo cross-dehydrogenative coupling at high temperature to obtain atomically precise nanometer-sized pores [215].
Jacobse et al. [216] created fully conjugated nanoporous graphene (C-NPG) through a single, mild annealing step following initial polymer formation by depositing methyl and methylene-bearing precursors 1 or 2 (Figure 13) onto a Au(111) substrate and annealing at 400 °C. This new 2D material emerges due to the reactivity of methyl and methylene crosslinking handles toward cross-dehydrogenative coupling at low temperatures.
The on-surface process also allows the formation of low-dimensional carbon-based nanostructures containing periodic nonplanar pores [217].
Qin et al. [218] recently selectively synthesized one-dimensional graphene nanoribbons containing [14]annulene pores on Ag(111) and two-dimensional graphene nanosheets containing [30]annulene pores on Au(111) starting from the same precursor TBDBTP (Figure 14).
They showed that (1) the formation of 1D graphene nanoribbons containing [14]annulene pores on Ag(111) is a thermodynamically controlled pathway due to the reversibility and flexibility of C−Ag−C-linked organometallic intermediate motifs; (2) the formation of 2D graphene nanosheets containing [30]annulene pores on Au(111) is a kinetics-driven pathway where the debromination process is the rate limiting step for the initial C−C coupling reaction.
Another technique that has gained much attention for the construction of 3D porous graphene structures is based on irradiation of certain carbon materials with a laser beam: the obtained laser-induced graphene (LIG) is characterized by high porosity, excellent electrical conductivity, and good mechanical flexibility.
Since the first conversion of a commercial polyimide (PI) film into 3D porous laser-induced graphene (LIG) under ambient conditions using a mid-infrared (MIR) CO2 laser system by Lin et al. [219], there has been significant development in the application of lasers to synthesize carbonaceous materials and in particular LIG as an alternative to conventional thermal treatment. Most types of lasers generate a photothermal effect, i.e., target materials absorb the incident photon energy and convert it into thermal energy [220]. The laser irradiation leads to a sharp rise in the localized temperature (>2500 °C), which can cause the breakage and recombination of weak chemical bonds, such as –COOH and –OH [221], while releasing gaseous products. In the PI network, the C-O, C=O, and N-C bonds are broken, and the aromatic compounds rearrange to form graphitic structures.
PI is commonly used as a precursor because its chemical structure is particularly well suited for LIG formation: it has a high content of aromatic sp2 carbons, which are more likely to form the hexagonal graphene structure than other organic sources that only produce amorphous carbon [222].
By lasing PI under various gas environments, such as O2, Ar, and SF6 under various atmospheres, Li et al. [223] prepared LIG with different surface morphologies and chemical compositions: under ambient or oxidizing atmospheres, LIG with superhydrophilic surface is obtained (contact angle∼0°, Figure 15c–e), while a superhydrophobic surface (contact angle > 150°) can be obtained when inert or reducing gas is applied.
In addition to PI, various materials have been successfully used to prepare LIG, including polysulfones [225], cloth, paper, food [226], wood [227], lignin [228], and phenolic resin [229].
Another synthetic aspect to be taken into account is the laser type (UV, IR, or visible lasers), since the laser properties must match the adsorption properties of the carbon precursor for effective energy transfer and desired outcomes, considering that many carbon precursors exhibit strong light absorption in the UV, visible, and MIR regions. UV lasers primarily drive the reaction through photochemical processes; because of the high energy involved in breaking molecular bonds directly with minimal thermal contribution, a less uniform and more defect-rich graphene structure can be obtained with UV lasers. Conversely, IR lasers rely on photothermal effects, as described above, with the formation of graphene with more uniform structural properties. Visible lasers harness both photothermal and photochemical mechanisms, offering a balance between direct bond cleavage and thermal decomposition, thus affecting the crystallinity and uniformity of the resulting graphene [230].
Moreover, laser processing in a special gas, liquid, or solid environment that contains the target atoms allows the heteroatom-doping of LIG [221]; in particular, in situ doping methods, which offer benefits such as precise control over dopant distribution and enhanced stability, can be performed in two ways. The first consists of lasing a PI film, coated with a thin dopant layer obtained by solvent evaporation of a dopant slurry on the PI film. The second involves the formation of a homogenous solution of the PI precursor solution, typically poly(amic acid) (PAA) with dopants, before laser irradiation, which occurs after thermal curing to create a PI film [231,232,233].
Yang et al. [233] developed a one-step method for selectively generating phosphorus-doped hierarchically porous graphene via 10.6 μm CO2 laser induction (LIPG) with significantly improved electrochemical performance, innovatively using ammonium polyphosphate (APP) as a flame retardant and P supplier as well (Figure 16).
The last bottom-up method is pyrolysis, where decomposition of suitable carbon-based precursors is brought about by high temperatures, which can also be achieved using plasma-based heating methods [234,235].
Pyrolysis is largely used because of the great availability of carbon sources, particularly chemical products such as carbonaceous salts [236,237,238], synthetic polymers [239], and plastic wastes [239].
Zhu et al. [236] developed a novel general synthesis strategy based on the direct pyrolytic conversion of sodium carboxylate in the presence of Na2CO3 to monolayer graphene (Figure 17).
Another important group of carbon precursors is organic species such as sugar [240], organic natural polymers such as lignin [241], and more generally biomass precursors consisting of natural sources such as plants or wood parts, agricultural waste, fruits, vegetables, biopolymers, or even insect parts [242].
Roy et al. [243] produced graphene from nonconventional sources such as tannic acid, alginic acid, and green tea using a controlled pyrolysis technique (Figure 18).
The process of graphitization of the precursors was thoroughly studied, and it was found that the temperature needed is 1100 °C.
Moreover, the pyrolysis technique is also very useful for the synthesis of heteroatom-doped graphene [244]; for instance, starting from a low-cost crystal sugar, Pan et al. [240] developed a simple strategy for fabricating nitrogen-doped graphene (NG) as efficient metal-free ORR electrocatalysts via one-step pyrolysis of sugar in the presence of urea.

3. Characterization of Graphene

Effective characterization techniques play an important role in accurately analyzing the surface morphology, the structural features, composition, and electrochemical behavior of the graphene-based ECs.
In particular, since the electrochemical performance and durability of metal-free graphene-based ECs are strongly dependent on the structure and morphology of the materials, this section will describe techniques based on (1) Electron Microscopy (Transmission Electron Microscopy—TEM; High Resolution Transmission Electron Microscopy—HRTEM; Scanning Electron Microscopy—SEM; and Scanning Tunneling Electron Microscopy—STEM); (2) Scanning Probe Microscopy (Atomic Force Microscopy (AFM) and Scanning Tunneling Microscopy (STM)); (3) Raman Spectroscopy; (4) X-ray Photoelectron Spectroscopy (XPS); (5) X-ray Diffraction (XRD); (6) UV−Vis Spectroscopy; (7) X-ray Absorption Near-Edge Structure (XANES).

3.1. Electron Microscopy (TEM, HRTEM, SEM, and STEM)

The fundamental understanding of the structural characteristics of a nanocomposite is significant for building an optimal nanocatalyst to maximize its electroactivity. SEM has been largely used to unravel the morphology of graphene, and TEM has been used to image nano-sized materials at atomic scale resolutions.
Zhang et al. [245] illustrated accurately via SEM the photochlorination of vertically oriented graphene (VG) grown on carbon cloth (CC), indicated as CC@VG, with chlorination time (Figure 19).
Interesting TEM observations of carbon allotropes [graphene, multi-walled carbon nanotubes (mWCnT) and nanohorns] that are covalently bound to polypyrrole (PPy) (Graphene-g-PPy, mWCnT-PPy, carbon nanohorn-g-PPy) (Figure 20) [246] showed that, for chosen carbons, the polymer coating looks completely different; in particular, in the graphene-based system (Figure 20A), spherical particles of the polypyrrole are evenly distributed on the carbon surface and the size of polymers is in the range 20–80 nm.
However, TEM is characterized by limited resolution (0.2 nm size uncertainty) at low operating voltage, while at high voltage, melting effects of the smallest nanoparticles under the electron beam can occur with damage to the monolayers. Therefore, a new class of TEM (High-Resolution TEM, HRTEM) has been introduced, which, in combination with a monochromator via aberration correction, can provide 1 Å resolution at an acceleration voltage of only 80 kV [247,248]. Using electrons at 80 kV not only minimizes the knock-on damage but also improves detection sensitivity due to the higher scattering power of carbon at a lower acceleration voltage of the electrons [249]. HRTEM, therefore, permits the study of the size and crystalline structure of relatively small metal nanocrystals; it can resolve the in-structural features of graphene (individual carbon atoms, defects, adatoms, etc.) [250] and the number of layers of graphene flakes can be detected from the diffraction of the electron nanobeam [249].
A very nice example of HRTEM potentialities is offered by Gómez-Navarro et al. [247], who were able to provide insight into the exact atomic structure of the rGO layers. At the atomic resolution, an aberration-corrected TEM image of a single-layer reduced-graphene oxide membrane (Figure 21) showed the hexagonal lattice of the well-crystallized graphene sheet, as well as various defects and deformations, e.g., disordered carbon networks, trapped carbonaceous adsorbates, heavy atoms, isolated pentagon−heptagon pairs, and clustered defects evidenced by the different colors in Figure 21.
However, HRTEM may have the drawback of generating defects and edge reconstructions due to the high-energy electrons [245,251].
Investigation of morphology, crystal structure, and composition distribution has also been achieved using STEM [252], which combines the principles of transmission electron microscopy and scanning electron microscopy.
Its primary advantage over conventional TEM is the possibility of using other signals that cannot be spatially correlated in TEM, including secondary electrons, scattered beam electrons, characteristic X-rays, and electron energy loss, while its primary advantage over conventional SEM imaging is the improvement in spatial resolution. In particular, the imaging method in STEM, based on collecting scattered electrons with a dark-field detector both in medium (ADF) and high-angle aberration-corrected annular configuration (HAADF), allows characterization at the atomic level.
In 2011, Huang et al. [253] grew predominately single-layer graphene films on copper foils, and by using atomic-resolution imaging, determined the location and identity of every atom at a grain boundary (Figure 22), overcoming the problem of the five-order-of- magnitude size difference between grains and the atoms at grain boundaries.
It was found that the grain boundary is not straight, with not periodic defects along the boundary; different grains stitch together predominantly through pentagon–heptagon pairs.
In 2008, Blangert et al. [254] evidenced regions including one to a few layers of graphenes via HAADF images in graphene membranes, while in 2012, Zhou et al. [255] analyzed graphene grown by CVD on a Ni film on a silicon wafer, where Si is a common impurity. Using ADF and HAADF images, it was possible to track atoms diffusing in the graphene lattice, providing a full description of their local chemical environment.

3.2. Scanning Probe Microscopy

Imaging of solid surfaces at an atomic resolution is commonly achieved using Atomic Force Microscopy (AFM) and Scanning Tunneling Microscopy (STM), the first based on the strong interatomic forces between atoms of the surface and sharp probe tips at a very short distance, the second exploiting the tunneling effect and monitoring the related current as the conducting tip moves across the surface.
AFM pictures are collected in three different modes: contact, noncontact, and tapping. Amplitude modulation (AM) AFM, the so-called tapping mode, has been predominantly employed for graphene imaging under ambient conditions and measurement of the graphene sheets’ thickness on different substrates [256,257,258,259,260,261].
Pristine graphene thickness is reported to be around 0.34 nm, much smaller than that of graphene oxide (1–1.5 nm), which may depend on various functional groups or adsorbed molecules present on the surface [262].
An interesting study on the structure of GO sheets deposited on a SiO2/Si substrate was carried out by Mkhoyan et al. [263] who could readily identify several GO sheets consisting of mono-, bi-, and trilayers in the AFM image (Figure 23); from the histograms of the AFM-depth profiles, quantitative analysis of flake height (or thickness) was conducted, showing that the thickness of a single sheet is about 1.6 nm and the ratio of thicknesses of the mono-, bi-, and trilayers scale is 1:1.6:2.2 (with actual values being 1.6:2.6:3.6 nm).
Further information on the lattice structure and surface morphology at atomic resolution can be obtained using STM. The possibility of carrying out STM measurements at low and high temperatures makes STM a powerful investigation method for thermal phenomena in real time [264,265].
In 2013, Günther et al. [265] studied the ordering transition of an amorphous carbon layer into graphene at 920 K by monitoring it as a function of time using STM (Figure 24).
Time sequence shows that graphene is not formed homogeneously over the entire layer; a peculiar growth process involves motion in the carbon layer over unexpectedly large distances and in a seemingly directional, non-Brownian manner of topographic holes directly connected with the formation of the Moiré structure.
Another interesting investigation of the thermal evolution of individual graphene oxide sheets to graphene through high temperature (≥1773 K) heat treatment was carried out by Rozada et al. [266], who rationalized the process on the basis of the generation and dynamics of atomic vacancies in the carbon lattice.
It was evidenced via STM that a thorough conversion of graphene oxide/reduced graphene oxide into graphene sheets of high structural quality can be achieved via graphitization at temperatures as low as 2073 K if, prior to annealing, the amount of oxygen groups on the graphene oxide sheets is drastically decreased, e.g., by chemical reduction (Figure 25).
In fact, in this way, the thermal evolution of carbon atoms in CO/CO2 is almost negligible, and no atomic vacancies are created. Secondly, the substrate onto which the sheets must be supported must be inert, such as pristine atomically flat HOPG, in order to prevent interactions with the vacancies that could alter their migration dynamics.

3.3. Raman Spectroscopy

Raman Spectroscopy has found significant applications in the characterization of ordered and disordered crystal structures of carbon materials [267]; in particular, it can be efficiently used to monitor the number and quality of layers, bonding configuration of carbon atoms in graphene sheets, doping level, and confinement in graphene nanostructures. Graphene Raman spectra have some fingerprints that reflect changes in the electronic structure and electron–phonon interactions, allowing unambiguous, high-throughput, nondestructive identification of graphene layers [268,269].
The graphene Raman spectrum (Figure 26) has three characteristic peaks at 1580 cm−1, 1350 cm−1, and 2700 cm−1 [270].
The first one is the G band at 1580 cm−1, associated with the first-order scattering of the E2g mode arising from sp2 carbon atom domains.
The D band at 1350 cm−1 is associated with second-order Raman scattering process consisting of one-elastic and one-inelastic scatterings, and it is the so-called disorder-induced D-band. It is related to the vibrations of sp3 carbon atoms of disordered graphene nanosheets that usually appear in a disordered sample or at the edge of a graphene sample.
Raman data facilitate quantifying defects in graphene, which is a key step toward understanding the limits of its ultimate mobility. In graphene with zero-dimensional point-like defects, the distance between defects, LD, is a measure of the amount of disorder.
The third peak at 2700 cm−1 is the 2D band, also known as G’, and it is associated with a two-phonon double resonance Raman scattering process consisting of two inelastic scatterings. Cançado et al. [272] showed that the position and shape of the second-order 2D feature in the Raman spectra of graphene are sensitive to the number of layers of graphene sheets and can be used to evaluate the number of layers in a multilayer graphene sheet.
The first two Raman peaks, and in particular the D band, have been studied to obtain a deeper insight into the graphene quality.
For this purpose, Cançado et al. [270] bombarded single-layer graphenes prepared on silicon with argon cations in a controlled and reproducible way, obtaining layers with various LD, which were analyzed by means of Raman spectroscopy (Figure 27).
The Raman spectra of graphene change with the energy of the excitation source as well as the ratio between the D and G peak intensities, for a given defect density. The D band moves to higher Raman shifts, and ID decreases as EL increases (Figure 27b), which is a well-known effect in the Raman scattering of sp2 carbons.
In the graphene Raman spectra, analysis of the 2D peak is very significant since its shape, width, and position are affected by the number of graphene layers, reflecting the change in the electron bands via a double resonant Raman process [268].
SLG is characterized by a single sharp 2D band (Figure 28a,b), and moving to a bilayer graphene, a much broader and up-shifted 2D band with respect to graphene appears. This can be fitted with four components, 2D1B, 2D1A, 2D2A, 2D2B (Figure 28c), that can be associated with the splitting of π and π* electronic bands due to the interaction between layers.
The 2D band gets broader as the number of layers increases, until, for graphene with more than five layers, the Raman spectrum resembles that of the bulk graphite.
Raman spectroscopy is also useful in the characterization of heteroatom-doped graphene because the presence of the heteroatom is itself broadly a defect, with an increase in the ID/IG ratio with respect to the undoped species. Molina-Garcia and Rees [273] analyzed single boron-doped (B-Gr), nitrogen-doped (N-Gr), phosphorus-doped (P-Gr), sulfur-doped (S-Gr), and quaternary-doped (BNPS-Gr) graphene (Figure 29).
The ID/IG ratio of all the doped samples is higher than that of the starting material. Moreover, it is lower for B-Gr and N-Gr (ca. 1.05) with respect to P-Gr and S-Gr (ca. 1.35), suggesting that the integrity of the graphene structure is dependent on the size of the doping atoms, rather than the amount of doping species incorporated into the graphene layers. The quaternary-doped catalyst has the highest ID/IG ratio (1.24), which reflects the incorporation of the different heteroatoms.

3.4. X-Ray Photoelectron Spectroscopy

XPS measurements allow the investigation of the surface composition and the oxidation states on the surface of graphene and its derivatives.
XPS survey spectra are collected to determine the surface elemental composition, while high-resolution XPS is used to determine the local bonding environments of C atoms.
In the case of graphene oxide and graphene oxide derivatives, this is accomplished by fitting the C1s spectral envelope to include contributions from different surface oxides (e.g., C-O, C=O, O-C=O), all having peaks at different binding energies.
In 2012, Poh et al. [274] prepared graphene oxide via Staudenmaier, Hofmann, and Hummers methods, which underwent thermal treatment to be converted into graphene, named G-ST, G-HO, and G-HU, respectively. XPS evidenced, in particular, the presence of N in the G-HU (Figure 30c) and showed higher C/O ratios in the reduced species than in the parent species, confirming successful reduction.
Further insight into the types of oxygen-containing groups present in the three thermally reduced graphenes was given by high-resolution XPS spectra of C1s for each sample. From careful fitting of the signal, five different carbon stages were quantitatively differentiated: the sp2-hybridized carbon atoms (284.5 eV), the sp3-hybridized carbon atoms (285.7 eV), the C–O of alcohol/ether groups (286.8 eV), the C=O of carbonyl groups (288.0 eV), and O–C=O of carboxyl acid/ester groups (289.2 eV). The π–π* shake-up signal (290.8 eV) was also found in all three thermally reduced graphenes, which is typical for sp2-hybridized carbon.
In the case of heteroatom-doped graphene, XPS is quite useful in distinguishing heteroatom configurations, which can be correlated, for example, with electrocatalytic properties.
Hu et al. [275] collected XPS spectra of various N-doped samples (NG−1, NG-2, NG-3, NG-4); in particular, it was possible to detect various N configurations from the deconvolution of the N1s peak (Figure 31) and the full width at half maximum (FWHM), the nitride- N (~397.5 eV, FWHM = 0.96 eV), pyridinic- N (~398.5 eV, FWHM = 1.29 eV), nitrile- N (~399.3 eV, FWHM = 1.24 eV), pyrrolic- N (~400.2 eV, FWHM = 1.24 eV), and graphitic- N (~401.1 eV, FWHM = 1.30 eV).

3.5. X-Ray Diffraction

XRD is a useful technique for analyzing crystalline nanomaterials and determining the crystal structure and size of nanoparticles on the basis of the Debye–Scherrer equation.
d = K λ β cos θ  
where K is the Scherrer constant; λ is the wavelength of light used for the diffraction; β is the “full-width at half maximum” of the sharp peaks and θ is the measured angle.
Using XRD, it is possible to distinguish and identify graphite, graphene oxide, and reduced graphene oxide. The XRD pattern of pristine graphite with a graphitic structure exhibits a strong and sharp peak at 2θ at approximately 26°, corresponding to an interlayer spacing of 0.334 nm [276]. The successful oxidation of graphite implies the loss of the typical graphitic structure in graphite because of the introduction of oxygen functionalities onto the carbon basal plane [277]. A new diffraction peak, the characteristic GO peak of the (002) plane, appears at 2θ within a range between 9 and 13° with different associated interlayer spacing. The variation in the interlayer spacing of GO is a result of variations in the degree of oxidation of the graphite, and according to Marcano et al. [69], it is proportional to the degree of oxidation.
XRD has also been used in the study of the reduction of GO to rGO carried out by Shin et al. [87], who treated graphite oxide with different amounts of NaBH4.
In the XRD spectra (Figure 32), it is again evident that the increased interlayer distance between graphite and graphene oxide, attributed to the hydroxyl and epoxy groups between the carbon sheets introduced by the oxidation treatment, can lead to the decrease in van der Waals forces between the graphite sheets in the exfoliated GO.
Upon addition of NaBH4, the interlayer distance is further increased up to 9.72 Å nm, but with 150 mM of NaBH4, the peak of the large interlayer distance disappears completely, and a broad peak near 0.3.73 Å (2θ = 23.98°) becomes visible, which suggests that most of the functional groups were removed.
Other reduced graphene oxide obtained via reduction of GO with hydriodic and acetic acid (RGOHI–AcOH) [278] and with sodium–ammonia solution (RGONa-NH3) [279] present XRD spectra with a distinguishable peak for the reduced GO at θ = 24.57° (interplanar distance = 3.62 Å) and at θ = 25.04° (interplanar distance = 3.56 Å) (Figure 33).

3.6. Ultraviolet–Visible Spectroscopy

UV-Vis spectroscopy is quite useful for identifying graphene and its oxide because of their different UV-Vis absorptions, as shown in Figure 34A, where the reduction of GO with hydrazine was monitored during the reaction [280]. By increasing the duration, a large red shift is observed because of the reduction of GO to rGO.
GO presents two characteristic absorption peaks, one intense at 230 nm related to π-π* transition of the C=C bond, and one weak at 305 nm corresponding to n-π* transition of the C=O bonds of GO [281,282]. A different UV graphene characteristic peak is observed at 268 nm.
Another important feature is that the electronic properties of graphene are strongly linked to its thickness, and the absorption depends linearly on the number of layers: the intensity of the band at 268 nm is almost doubled going from monolayer to bilayer graphene (Figure 34B) [281].
UV-Vis also shows evidence of heteroatom-doping, as in the case of N-doped quantum dots NGqDot: the absorption band at ca. 270 nm (Figure 35) [283] in the absorption spectrum is blue-shifted by ca. 50 nm with respect to that of N-free GqDot of similar size [284].

3.7. X-Ray Absorption Near-Edge Structure

The X-ray absorption near-edge structure (XANES) spectroscopy, also referred to as Near-edge X-ray Absorption Fine Structure (NEXAFS) spectroscopy, is an element-specific spectroscopic technique for probing photoexcitation of electrons from a core level (absorption edge) of an atom to levels within an energy range of ∼50 eV above the absorption edge.
Therefore, being sensitive to the local chemistry of the atom under detection using edges as a “fingerprint” of the electronic structure offers the possibility of investigating the chemical bonding, electronic structure, and interactions of the materials with the advantage of not destroying the sample [285,286].
A NEXAFS study of single-layer graphene was carried out in 2008 by Pacilè et al. [287], who observed a splitting of the π* resonance and a clear signature of the predicted interlayer state. By comparing spectra of single-layer, bilayer, and few-layer graphene samples, it was possible to illustrate the rapid evolution of the electronic states from those of a truly two-dimensional (2D) system towards those of bulk graphite.
Since NEXAFS is particularly powerful at characterizing carbon-related materials by providing information on the degree of bond hybridization in mixed sp2/sp3-bonded carbon, the specific bonding configurations of foreign functional atoms, and the degree of alignment of graphitic crystal structures [287], it is particularly suitable for studying doped graphene [288,289,290,291].
XANES measurements of pristine and N-doped graphene on copper foil at the nitrogen K-edge (Figure 36) carried out by Zhao et al. [178] confirmed the presence of N in the system. N-doping results in sharp peaks at 400.7 and 408 eV in the NEXAFS spectrum, associated with 1s-to-π* and -σ* transitions, respectively, for a single molecular species. The new peak in the spectrum at 400.7 eV is due to graphitic N.
XANES measurements provided unambiguous evidence for the presence of more nitrogen species (pyridinic and amino group) in GO treated (Figure 37) at 300 °C (sample NG 300), 500 °C (sample NG 500), 700 °C (sample NG 700), and 800 °C (sample NG 800), with ammonia used as supporting materials to load Pt nanoparticles [292].
The XANES spectra show four well-resolved resonance peaks, N1, N2, N3, and N4, located at ca 397.4 eV, ca 398.2 eV, ca 399.8 eV, and ca 406 eV, respectively. In contrast, peak N4 is assigned to general transitions from the N 1s core level to C–N σ* states, peak N1 and N3 are attributed to the pyridinic (C=N π*) and graphitic type nitrogen species, respectively, and peak N2 is attributed to amino type species.
It is also interesting to observe the evolution of XANES peaks at C edge with the temperature; on raising the temperature, the intensity of C1 at around 285 eV due to C–C π* (ring) transitions increases, while the intensity of C2 at around 292.4 eV due to C–C σ* (ring) transitions decreases, indicating the increasing dominance of the π-network because of the removal of the oxygenated groups during the treatment.
A detailed analysis of the absorption edges of the dopant elements in doped and co-doped GOqDot was carried out by Favaro et al. in 2015 [293]. In the case of S and S,N-GOqDot, the S L edge spectra show the presence of –SH (S1, 162.8 eV), C–S–C (S2, 164.2 eV), and –SO3 (S3, 166.8 eV) groups. The S,N-GOqDot S L edge spectra are also characterized by a broad peak between 165 and 168 eV due to the concomitant presence of a relatively high amount of oxidized S-based groups and S–N groups whose 1s–π* transition is centered at 166.1 eV. Similarly, in the B K edge spectrum of B-GOqDot, two distinct peaks are present, associated with the 1s–π* (at 189.3 eV) and 1s–σ* (at 191.5 eV) absorption transitions of the B–sp2 C and C–BO2 groups, respectively. In the case of N,B-GOqDot, C–BNO bonds are present as indicated by a further feature at 191.1 eV.

4. Nitrogen-Doped Graphene

Since Gong et al.’s work on vertically aligned nitrogen-containing carbon nanotubes [294] and following analogous investigation on N-doped graphene in 2010 [185] showing its superb ORR electrochemical properties in alkaline medium, intense research has been initiated to study the behavior of various N-doped graphenes as ORR electrocatalysts strictly correlated to the synthetic procedure, morphology, structure, and physicochemical properties of the material.
Nitrogen is considered one of the most effective dopants for enhancing the ORR activity of graphitic carbon materials [185,294,295,296]. The electron-accepting ability of the nitrogen atoms creates a net positive charge via intramolecular charge-transfer on adjacent carbon atoms in the nanocarbon structures to readily attract electrons from the anode to facilitate O2 adsorption and the ORR process [185,297].
Also, the nitrogen-induced charge delocalization changes the chemisorption mode of O2 from the usual end-on adsorption to a side-on adsorption, which effectively weakens the O–O bond, thus facilitating the oxygen reduction process [294]. Additionally, all intermediates involved in the oxygen reduction reaction bind to the carbon atom next to the nitrogen dopant [298].
The nitrogen doping introduces asymmetry in spin density and atomic charge density, making it possible for N-graphene to show high electrocatalytic activities for the ORR, as evidenced by the density functional theory [DFT] analysis of the ORR mechanism by Zhang and Xia [299]. Unlike pristine graphene, the nitrogen-doped graphene has a dramatically lower energy cost for the early catalytic ORR steps [300].
Nitrogen follows carbon in the periodic table, and by replacing C with N in the graphitic framework, the total number of electrons in the system can be tailored depending on the doping [301]. Since N has an atomic radius similar to C, significant lattice mismatches are prevented when N is introduced. N-doping can effectively improve the surface energy and reactivity of graphitic carbon frameworks with minimal damage to their properties [302].
The introduction of N atoms into the matrix of sp2-bonded graphitic frameworks can lead to different configurations of nitrogen states: pyridinic N, pyrrolic N, graphitic [or quaternary] N, N-oxides, and amino group (Figure 38) [303].
Pyridinic N is attached to two carbon atoms in six-membered rings, possesses a lone pair of electrons, and donates one p electron to the π system; also, configurations with hydrogen can be present, as shown in Figure 38B. The pyrrolic N substitutes a carbon atom in five-membered rings and donates two p electrons to the π system. Both types of N occur at the edge of the layer but may also occur inside and are associated with vacancies.
Graphitic N refers to a N atom that bonds to three carbon atoms in graphene basal plane (graphitic/quaternary-N center) or occupies graphene zigzag edge or valley sites (graphitic-N valley), Figure 38B, and it is less impressionable to the protonation reaction because of an unavailable lone pair of electrons around N atom in the carbon plane [304].
In addition, nitrogen can be included in the form of nitrogen oxides or bound to the basal plane as an amino group Figure 38B. The detection and identification of the N configuration can be achieved using XPS via analysis of the nitrogen core level peak N1s and its deconvolution.
There is extensive debate on the contribution of the nitrogen configuration (pyridinic, graphitic, pyrrolic, and amino) to ORR activity: the identity and role of the electrocatalytically active center are still controversial, with different nitrogen configurations often used to justify ORR activity and/or different ORR activity among correlated materials.
Beyond the intrinsic nature of the active sites, which is governed by the chemical compositions and bonding configurations of the doped atoms, there are two more factors to be considered that affect the ORR performance of heteroatom-doped nanostructured carbon materials: (1) the number of the available catalytic centers and the transport properties of reactants and electrolyte, which depend largely on the specific surface area and porous architecture; and (2) the electrical conductivity, which is mainly determined by the defects of carbon materials.
Because of the very wide scenario of cases and related studies, this review presents them using, as the first criterion, the claimed nitrogen configuration boosting the oxygen reduction kinetics.
In 2010, Long et al. [121] prepared N-doped graphene sheets via reduction of graphene oxide with hydrazine, first ascribing the high ORR activity in acidic media in part to the large amount of N pyridinic sites, as supported by previous work [305,306,307].
Many other studies followed, evidencing the relevant contribution of pyridinic N [127,275,304,308,309,310,311,312,313,314,315,316,317,318,319,320,321,322,323,324,325,326,327,328,329,330,331,332,333,334,335,336,337,338,339,340,341,342,343,344,345,346,347,348,349,350,351,352,353,354,355,356].
In 2014, Xing et al. [313] carried out an interesting study to determine the catalytic sites of three nitrogen-doped multilayer graphenes prepared from graphene oxide using different nitrogen sources and doping methods for different nitrogen doping configurations. The chemical composition changes before and after ORR were examined via synchrotron-based X-ray photoelectron spectroscopy analyses of chemisorbed oxygen reduction intermediates.
On the basis of the commonly accepted four-electron associative ORR mechanism of N-doped graphene in alkaline solution, intermediate OHads is formed from Oads with the addition of H2O and one electron: the oxygen reduction intermediate OHads, which should chemically attach to the active sites, was found to remain on the carbon atoms neighboring pyridinic nitrogen after ORR, indicating that the carbon atoms close to pyridinic nitrogen are the main active sites in the different nitrogen doping configurations. In addition, a high amount of OHads attachment after ORR corresponds to a high catalytic efficiency and vice versa.
Synchrotron-based X-ray photoelectron spectroscopy analyses of three nitrogen-doped multilayer graphene samples revealed that the oxygen reduction intermediate OHads, which should chemically attach to the active sites, remains on the carbon atoms neighboring pyridinic nitrogen after ORR. Moreover, the location or microstructure of pyridinic nitrogen can affect its catalytic efficiency.
In 2016, Farzaneh et al. [317] hydrothermally synthesized three-dimensional (3D) mesoporous nitrogen-doped reduced graphene oxides [N-rGO3Dmpo] with different nitrogen contents [1.0–4.7%] containing different N-pyridinic amounts and carried out research on the role of π and lone pair electrons. Pyridinic-N contributes one π electron to the aromatic π system and possesses an electron pair in the plane of the carbon matrix. This electron pair increases the electron–donor property of the catalyst, weakening the O-O bond via bonding between oxygen and nitrogen and/or an adjacent carbon atom, and facilitating the reduction of O2. Pyridinic N at the edge of the carbon plane is also more exposed and thus can be more available for O2 molecules [305,356,357].
It was shown for the first time that there are four different regimes governing ORR on N-doped graphenes depending on the applied potential: at lower potentials, N-content of N-rGO3Dmpo [regardless of their bonding configurations] promotes ORR performance due to activation and strengthening of π-electrons of graphene framework, while pyridinic-N content has a significant role in ORR at higher potentials because of activation of their lone pair electrons. Furthermore, N-rGO3Dmpo, containing 4.1% nitrogen, was tested as a cathode electrocatalyst in the direct methanol fuel cells, which produced superior output power compared to commercial 20 wt% Pt/C.
In 2018, Wang et al. [334] developed a surface modification method to identify the ortho-carbon atom of the pyridinic ring as the reactive site for ORR on N-doped graphene, taking advantage of the molecular probe approach. They first prepared N-doped graphene containing higher amounts of pyridinic N and carried out a targeted modification: the pyridinic ring on N-doped graphene was highly selectively grafted by the acetyl group Ac at the pyridinic N atom, denoted as N_Ac, via an electrophilic reaction and ortho-C atom denoted as C_Ac through electrophilic and via a free radical reaction (Scheme 4).
It was observed that attaching a functional group to the pyridinic N atom of N-doped graphene did not affect the ORR activity, which was mostly preserved, while it was completely suppressed when the ortho-carbon of the pyridinic ring was blocked. DFT calculations revealed that the distinct ORR activity between N_Ac and C_Ac mainly comes from the difference in positive charge density and spin density at the ortho-C atom.
Therefore, it was concluded that the active species of N-doped carbon for ORR in acidic solution is the ortho-C atom of the pyridinic ring, on which O2 can be absorbed and reduced favorably.
It is worth mentioning the efforts that have been made to develop facile methods for the synthesis of pyridinic-N-doped graphene, trying to avoid high temperatures and using simple solution chemistry [333,335,349]. In 2019, Liu et al. [349] converted holey graphene into nitrogen-doped graphene by chemically grafting diamino-benzene derivatives to the ortho-quinone sites on the abundant holey graphene edges through a simple condensation reaction. The ORR performance depends on the crafted N types. The 3,4-diaminopyridine grafted holey graphene [A3-HG] containing pyridinic N possesses the best ORR performance, showing excellent ORR activity comparable to Pt/C in alkaline media, as well as good selectivity and stability (Table 2). The Zn–air batteries assembled with A3-HG electrocatalyst outperformed those assembled using commercial Pt/C electrocatalyst, achieving a high energy density of 495 Whkg−1, much higher than that of the Pt/C [198 Whkg−1].
Another interesting method for controlling pyridinic N in graphene and preventing the formation of other nitrogen species was offered by Zhong et al. [358], who carried out pyridine cycloaddition of graphene, obtaining a graphene derivative with an activity comparable to that of the commercial Pt/C electrocatalyst because of the synergistic effect of pyridine and graphene. Such a derivative is characterized by pyridine groups “external” and covalently attached to graphene, being in a vertical direction with respect to the basal plane. DFT calculations revealed that the ortho-carbon of the “external” nitrogen could be a possible adsorption and catalytic site.
In addition to the rich literature evidencing the role of pyridinic nitrogen related to the modification of the band structure of carbon, raising the density of states near the Fermi level, and lowering the work function [205], many other papers illustrate, on the other hand, the contribution of graphitic nitrogen in the ORR electrochemical process. This is based on the fact that the relative electronegativity of graphitic N atoms reduces the electron density on the adjacent C nuclei, which helps electrons to transfer from the adjacent C to N atoms, and N back-donates electrons to adjacent C pz orbitals [205]. The donation and back-donation processes not only facilitate O2 dissociation on the adjacent C atoms but also help form a strong chemical bond between O and C. Graphitic-N is considered to provide active sites for ORR because its two pz electrons can partially occupy the π* anti-bonding orbital around nitrogen [359]. A theoretical study by Nandhini et al. [360] on the identification of site-dependent activity in alkaline media evidenced a strong correlation between the activity of each C site with the occupancy of pz[π] electrons of the corresponding site. OH adsorption is found to have a linear relationship with pz occupancy; the occupancy of pz[π] electrons of C atoms increased near the dopant site due to the back-donation mechanism. Carbon atoms in the vicinity of the N atom are enriched with different pz[π] electron densities due to the back-donation mechanism. Hence, unequal binding strength is observed for each carbon atom of N-doped graphene, and the C atoms that are far from the N dopant site show similar behavior to the C atoms in the pure graphene surface.
Additional theoretical investigations were carried out to analyze the impact of different nitrogen doping on oxygen adsorption and on the energy barrier of oxygen molecule dissociation [299,361,362,363,364,365,366] and to explore in more detail the chemical reactivity, reaction pathway, electronic transport characteristics, and experimental conditions, such as the use of water as a solvent and the pH [360,367,368,369,370]. Additionally, particular structures, such as quantum dots and nanoribbons, have also been studied [371,372,373,374,375,376].
Scardamaglia et al. [377] used first-principle calculations to support the modification of the N1s line-shape observed by in situ high-resolution synchrotron techniques in the interaction of molecular oxygen with graphene doped via nitrogen plasma. Upon exposing N-graphene/Ir samples in situ to 5 × 10−5 mbar of O2 at a temperature of 200 °C for 30 min, oxygen dissociation and formation of carbon–oxygen single bonds occurred on graphene, along with a band gap opening and a rounding of the Dirac cone. The modification of the N 1s line-shape reveals, in particular, a change in the chemical environment of graphitic nitrogen upon oxidation that is consistent with the adsorption of two oxygen atoms in the epoxy position on the nearest carbon atom neighbors of the graphitic nitrogen, as predicted theoretically. The graphitic nitrogen is involved in the observed mechanism: the adsorbed oxygen molecule is dissociated, and the two O atoms chemisorb via epoxy bonds to the nearest carbon neighbors of the graphitic nitrogen. This is the first direct experimental description of the first step in the oxygen reduction reaction, providing a detailed atomic-level understanding of the role of the nitrogen active sites in the interaction with molecular oxygen.
Besides theoretical and spectroscopic investigations, many experimental studies have proven the effectiveness of graphitic N in promoting the ORR [378,379,380,381,382,383,384,385,386,387,388,389,390,391,392,393,394,395,396,397,398,399,400,401,402,403,404,405,406,407,408,409,410,411]. Geng et al. [378] were the first to prepare nitrogen-doped graphene by heating graphene in the presence of ammonia at different temperatures. The sample treated at 900 °C had the best high ORR activity through a four-electron transfer process in alkaline media, comparable to Pt/C (Table 2). By comparison with the samples heated at 800 °C and 1000 °C, which contain lower relative amounts of quaternary nitrogen than the sample heated at 900 °C, they evidenced that quaternary nitrogen atoms could be considered the most important species for the ORR due to the relationship between activity and quaternary N contents.
An interesting study of the dependence of catalytic activity on the nitrogen configurations was carried out by Lin et al. [384], who prepared N-doped graphene via pyrolysis of graphene oxide/polypyrrole composite. Starting from the N pyrrolic configuration [100%] of the initial composite, by heating above 300 °C, the pyrrolic N percentage decreased, while the pyridinic N percentage increased monotonically, with transformation from pyrrolic N to pyridinic N. However, at T > 600 °C, the pyridinic N percentage started to drop with a simultaneous increase in graphitic N percentage, implying the transformation of pyridinic N to graphitic N. Moreover, the direct transformation of pyrrolic N to graphitic N may occur at temperatures above 500 °C. The sample heated at 900 °C, containing the highest N graphitic percentage among all the samples, gave the best ORR performance in alkaline media.
Analogous thermal behavior of nitrogen configurations was driven and observed by Lu et al. [403] in N-doped graphenes prepared via pyrolysis of graphene oxide. First, they observed that the ORR activity is not proportional to the total nitrogen content, which decreases as the pyrolysis temperature increases up to 1000 °C; an increase in the annealing temperature can enhance the content of graphitic N, and the enriched graphitic sample obtained at 1000 °C had the best catalytic ORR activity and outperformed Pt/C (Table 2). Moreover, it showed appreciable catalytic activity also at acidic pH (Table 3).
DFT calculations evidenced that oxygen can be stably adsorbed on the C site neighboring the N atoms, while H2O can be weakly adsorbed and therefore easily released as an ORR product. More recently, another study by Guo et al. [406] on graphitic-N-rich N-doped graphene was conducted to investigate the effective mechanism of the graphitic nitrogen on the ORR activity of graphene and to better understand the relationship between the N dopant species and catalytic activity, indicating the interest in addressing the synthesis of N-doped graphenes towards tailored N configurations. DFT calculations were first performed; on the basis of electrostatic potential analysesit was showed that the oxygen molecule cannot bind to the pyridinic N site. Secondly, on the basis of the average local ionization energy on the molecular surface, widely used to evaluate electron reactivity, the minimum point can be observed near the graphitic N, while no minimum point can be found near the pyrrolic N [411,412]. Even if both pyrrolic N and graphitic N can bind with oxygen, the electron around graphitic N has large reactivity, so that the reaction occurs on the graphitic N site.
The general synthetic method to get enriched N graphitic graphene is based on pyrolysis operating at high temperature; microwave irradiation was alternatively used [389,401]. It is a simple, facile, and efficient heating process [413,414,415] with several advantages, including fast, volumetric, and uniform heat transfer to the reactants, thereby causing a homogeneous synthesis condition at elevated temperatures.
Wang et al. [389] produced N-doped graphene with different characteristics in a very short time. Through microwave heating, an efficient reductive process occurs, with the O content of graphene reduced to around 5 wt% from 19.2 wt%. The N content increased from 0 to over 5 wt%. The N-doping process via microwave heating required only a few seconds (2–30 s), which is much less time than thermal annealing (30 min) [380], and a considerable advantage compared with previous methods.
Two more N-configurations to be mentioned are the N-pyrrolic and N-amino configurations. In 2011, Zhang and Xia [299] studied the electrocatalytic mechanism of N-doped graphene in an acidic environment using DFT. Two different nitrogen-containing graphene sheets [C45NH20 and C45NH18] were built, containing pyridine and pyrrole species for which different active catalytic sites were identified, having either high positive spin density or high positive atomic charge density. Because of asymmetric spin density and atomic charge density, the N-doped graphenes containing pyridinic or pyrrolic nitrogen are predicted to show appreciable electrocatalytic activities in the ORR.
Unni et al. [416] were the first to ascribe the enhanced catalytic activity towards ORR of the prepared N-doped samples obtained from GO and pyrrole used as nitrogen dopant and reducing agent to the enhanced proportions of the pyrrolic nitrogen along with the mesoporous structure of graphene. Then, Liu et al. [417] proposed, reasonably, that the pyrrolic-N structure is the main contributor to the four-electron transfer mechanism of the ORR of N-rGO samples prepared under microwave irradiation. More recently, the amount of N-pyrrolic has been considered fundamental in driving the selectivity of N-rGO toward the formation of H2O2, involving only two electrons [418,419,420,421]. Li et al. [418] showed that the sample with the highest pyrrolic-N content and abundant in-plane pores had superior activity toward electrochemical H2O2 synthesis, with a selectivity as high as 95% and excellent long-term stability. In particular, the critical role of the pyrrolic-N was elucidated by the variable adsorption profiles of OOH* and O* intermediates on C K edge XANES spectra, as well as the dependent negative shifts of the pyrrolic-N peak on N K edge XANES spectra.
Regarding the amino group its role in the ORR graphene-based electrocatalyst is associated with its electron-donating ability: the functionalization of graphene with amine is supposed to be a promising strategy for improving the electrical conductivity of graphene due to the electron-donating effect of amino groups [422,423,424,425,426]. In particular, amine-functionalized holey graphene [424], prepared via hydrothermal reduction of graphene oxide in the presence of ammonia, followed by etching in the presence of KOH, is shown to have high electrocatalytic activity toward ORR, which is more kinetically facile than when the commercial Johnson Matthey Pt/C 40 wt% is used. This is explained by the fact that the presence of the electron-donating group and a large number of holes in its sheet plate can improve its electrical conductivity and give more active edge N atoms, thereby increasing the electrocatalytic activity of amine-functionalized holey graphene.
Considering functionalization more specifically as a way for molecular doping of graphene [427] in a few papers, other nitrogen-based groups, such as ammonium [428], imino [429], nitro [427], heterocycle [430], and amide groups [246,425,431] have been introduced. The basic concept is that such species can effectively modulate the electron density of graphene and therefore affect the electrochemical properties without altering the graphene basal plane.
There are many other studies where the ORR electrocatalytic properties are favored by the coexistence of various N configurations [432,433,434,435,436,437,438,439,440,441,442,443], in particular, by N-pyrrolic with N-pyridinic configuration [120,409,444,445,446,447,448,449,450,451,452] and by N-graphitic with N-pyridinic configuration, which is the most widespread [176,452,453,454,455,456,457,458,459,460,461,462,463,464,465,466,467,468,469,470,471,472,473,474,475,476,477,478,479,480,481,482,483,484,485,486,487].
In the specific case of the combination of N-pyrrolic and N-pyridinic configurations, Ding et al. [452] developed a particular synthesis method based on the polymerization of aniline and pyrolysis carried out inside layered montmorillonite (MMT), which is used as a quasi-closed flat nanoreactor, since it is open only along the perimeter to enable the entrance of aniline monomer molecules. The flat MMT nanoreactor, less than 1 nm, facilitates the formation of pyridinic and pyrrolic N, characterized by planar sp2 hybridization (Figure 39), while it extensively constrains the formation of quaternary N because of its 3D structure. In this way, N-doped graphene (NG) containing selective N planar configurations could be obtained. The confinement effect of MMT ensures that N is incorporated into the structure and that the graphitization is successful without significant loss of N species.
The ORR activities of the prepared NG decreased as the planar N percentage and the electrical conductivity decreased; the sample with 90.3% N planar showed good catalytic activity toward the ORR in acidic media (Table 3), with a difference in the half-wave potential of 60 mV relative to Pt/C, high selectivity toward the ORR, stability, and a remarkable tolerance for methanol.
On dealing with the combination of graphitic and pyridinic nitrogen, it is interesting to report a study by Lai et al. [453] where the different roles of the pyridinic and graphitic nitrogen configurations were evidenced. Different N-doped graphenes were obtained by starting from different reagents: annealing of GO with ammonia preferentially formed graphitic N and pyridinic N centers, while annealing of polyaniline/rGO and polypyrrole/rGO tended to generate pyridinic and pyrrolic N moieties, respectively. Such graphenes were characterized by different ORR performances; while it was recognized that the dependence of electrocatalytic activity on the graphitic N content affected the limiting current density, the pyridinic N content improved the onset potential for ORR and might convert the ORR reaction mechanism from a 2e-dominated process to a 4e-dominated process. However, the ORR process is not affected by the total N content in the N-doped graphenes.
Analogous observations on the role of graphitic and pyridine N were made by Su et al. [488].
There are also computational studies aimed at evaluating the interplay between graphitic and pyridinic N in the ORR process [474,475]. Yan et al. [475] evidenced the synergic effect of graphitic and pyridinic N on the improvement of the adsorption and activation of O2. Combining graphitic and pyridinic N can induce further redistribution of charge among the carbon materials, which subsequently promotes O2 adsorption and lowers the energy barrier of O2 hydrogenation, demonstrating an improved property for ORR; hence, the synergic effect of graphitic and pyridinic N can strengthen the reduction performance of O2 by reducing the activation energy barrier.
As reported at the beginning of the paragraph, there are other elements beyond the particular N configuration that affect the ORR activity of the graphene-based catalysts, for example, the presence of exposed edges and defects, the conductivity, the morphology, and the microstructure of N-doped graphenes.
Defective graphenes containing nitrogen as dopant with an optimized defect site density are generally characterized by enhanced ORR catalytic activity thanks to the synergistic effect of doping and defects/exposed edges [318,322,337,339,340,351,353,439,440,446,448,489,490,491,492,493,494,495,496,497,498,499,500,501,502,503], while an excess of defects can have detrimental effects on the ORR electrocatalytic behavior because of reduced conductivity [499,504]. Han et al. [499] investigated the influence of in-plane carbon lattice defects on the catalytic activity and selectivity toward ORR, and were able to tune the density of defects, identifying the best and worst conditions for significant ORR activity and selectivity to H2O2 formation. Different graphene and graphene oxide precursors (oxo-G, obtained from mild oxidation of graphite, and GO, obtained from Hummers’ method) underwent an oxidative-etching process by mixing with H2O2 under hydrothermal conditions (Figure 40).
In this way, they could obtain four graphene precursors with different defect densities, which were afterwards hydrothermally treated with ammonium hydroxide to achieve nitrogen doping. Hetero-doping was revealed to improve the ORR performance and the presence of a moderate defect density, as the sample obtained from oxo-G induced selectivity to H2O2 production and the highest electron density below 0.5 V. The sample with higher N content, higher defect density, and higher electrochemically active surface area was obtained from the precursor with GO treated with H2O2, but did not show the highest ORR activity and H2O2 selectivity, which may be due to the fact that an excess of in-plane carbon lattice defects lead to reduced conductivity.
Defect engineering is very important for generating active sites [375], and to avoid disrupting the honeycomb lattice of the graphene basal plane, the scientists have tended to study and work on the edge positions to maximize the number of exposed edge sites [preferably zigzag] relative to catalytically inert basal plane sites [325,348,352,496,505,506,507,508,509].
This can be achieved, in particular, in graphene quantum dots (GqDot): the spherical shape and reduced size of graphene quantum dots offer a greater proportion of edge sites compared to graphene [510]. N-doped graphene quantum dots have shown improved ORR electrocatalytic properties [283,316,319,329,391,410,511,512,513] (Table 3) and have been employed as defective surface modification additives to generate defect-rich graphene/carbon quantum dot composites.
The integration of a defect-rich GqDot in the graphene matrix was achieved hydrothermally by Wang et al. [513]. The product was mixed with melamine and heated at different temperatures to allow N doping by melamine and activate fabrication defects. Thanks to a great density of defects and N doping, the sample obtained at 1000 °C has an excellent electrocatalytic activity for the oxygen reduction reaction, with a nearly four-electron pathway. The application potential of such material was also evaluated. An aqueous zinc/air battery was fabricated by employing it as a cathode catalyst, obtaining a much higher discharge voltage at the current density range from 20 to 100 mAcm−2 compared to Pt/C, leading to a higher power density for zinc/air batteries. The stability during the discharge process is also better than that of Pt/C.
The preservation and enhancement of the conductivity of the sp2 carbon network combined with N-doping is a well-defined issue in graphene-based ORR electrocatalysis [304,332,375,380,497,514,515,516,517]. Chao et al. [515] developed a facile solution-processed strategy to fabricate N-doped graphene consisting of two-step electrochemical exfoliation: in the first step of exfoliation, graphite paper is expanded in a weak exfoliation electrolyte, Na2SO4 aqueous solution, permeated with melamine–formaldehyde resin monomer (MFR), and calcined (Figure 41).
The obtained N-doped graphite was further completely exfoliated in (NH4)2SO4 electrolyte, taking advantage of the much larger number of gas species, i.e., NH3, which can be produced, resulting in a stronger force for exfoliating N-doped graphite paper into thin-layer graphene. The resultant N-doped graphene contains a high level of nitrogen (7.9 at%) and presents outstanding electrical conductivity despite abundant oxygen species, which, however, impart good wettability, facilitating effective ionic transport.
An effective way to ensure efficient mass transfer to the active sites and markedly facilitate the ORR diffusion kinetics is to construct graphene networks with convenient and abundant porosity characterized by a high surface area [518,519,520], which makes several active sites exposed and accessible to the electrolyte [352].
In order to increase the ORR active site density, the enlargement of the volumetric surface of the graphene catalysts [518] is used to achieve porous structures [336,355,464,490,521,522,523,524]. The 2D graphene sheets are converted to hollow spheres [409,444,479,525,526], nanocapsules [448], nanoring-integrated boxes [342], 3D graphene structures with various porosities [342,497], 3D interconnected nanostructures [432], aerogels [106,339,462,469,470,527,528,529], and foams/microfoams [442,460,530,531]. Other 3D structures consist of dendritic carbon structure [532], graphene nanoscrolss [530], crumpled/wrinkled graphene [240,327,330,344,404,461,465,533], holey graphene [399], flakes [466], unstacked double-layer templated graphene [534], fibers [348,535], core-shell structure [337], vertically-aligned graphenes [351], and graphene nanoribbon networks [396,527].
Chen et al. [527] fabricated a highly conductive, ultralight, neat, and versatile nitrogen-doped graphene nanoribbons aerogel (GNnRAe) through a facile hydrothermal method for the first time. Graphene nanoribbons prepared by longitudinally unzipping multi-walled carbon keep the carbon skeleton’s structural integrity as well as present abundant active sites related to straight edges possessing an open bandgap that varies with the ribbon width. The hydrothermal treatment allowed assembly into a highly porous structure with a large surface area and good conductivity. GNnRAe was revealed to be an efficient ORR catalyst in alkaline media (Table 3), with better stability in methanol than Pt/C, as well as in acidic media.
Graphene nanoribbons, a type of one-dimensional carbon material derived from carbon nanotubes (CnT), can serve as an ideal bridge [482,536,537] combining the features of graphene sheets and carbon nanotubes [527]; however, they are still limitedly applied. On the contrary, carbon nanotubes have been more widely used to obtain N-doped 3D graphene–CnT composite materials [441,477,481,538,539,540,541,542,543,544,545,546,547,548,549,550,551]. The introduction of one-dimensional carbon nanotubes into graphene materials has different advantages, such as the formation of a highly conductive three-dimensional network to greatly enhance electron transport through a 3D-interpenetrating conductive network. The incorporation of CnT can also solve the mass transfer issue caused by the stacking and agglomeration in graphene-based catalysts, which also reduces the number of active sites; moreover, the hydrophilic and hydrophobic nature of the dual interfaces can be conveniently tuned [539,542,551].
Qazzazie et al. [544] carried out a thorough investigation of N-doped graphene/carbon nanotube nanocomposites prepared by mixing N-doped graphene and CnT, consisting of a few layers of graphene flakes wrapped with CnT network and separated from each other. Its ORR performance was examined by galvanostatic measurements in realistically approached glucose half-cells at neutral pH. It showed remarkable enhancement due to the improved accessibility of oxygen molecules to active reaction sites and the remove OH-; actually, the incorporation of CnT between 2D graphene sheets opens mesoscale pathways for mass transport, forming a 3D porous structure in the nanocomposite. Efficient 4-electron transfer and improved overall catalytic ORR performance are observed as evident from voltammetric studies. 2D simulation models were carried out to clarify the effects of carbon nanotubes on mass transport and the overall catalytic activity of the electrode: the kinetics features extracted from the simulations show qualitative agreement with the experimentally determined kinetics underlining the role of CnT acting as conductive spacers for improving accessibility of pores and providing more reaction sites for reactants in the nanocomposite.
In order to complete the overview on N-doped graphenes, it is important to report some studies that have strictly addressed the application of N-doped graphenes in fuel cells [552,553], the development of alternative mild and/or massive and/or environmentally friendly synthetic strategies of N-doped graphenes [554,555,556,557,558,559,560,561,562,563,564,565,566,567], and studied the effects of chemical or thermal treatment of N-doped graphenes [568,569,570].
The very last papers report theoretical studies that mostly concern DFT calculations for various heteroatom-doped graphene containing B, S, P, or co-doped species [128,361,571,572,573,574,575,576,577,578,579,580,581,582,583] and two papers on graphene characterized by structural deformation [584,585]. Xie et al. [585] compared the ORR reaction at active sites with different deformations under similar structural and chemical environments. Considering the associative mechanism of ORR, since both the tensile strain and the adsorption of O tend to break the NC* bond, the adsorption strength of the O atom is resonantly enhanced by the tensile strain, while those of OH and OOH remain unchanged. The results suggest that it is possible to improve catalytic performance by tuning the structure of the catalyst to be in selective resonance with the adsorption of a specific intermediate, which provides a new way for designing optimal catalysts.

5. Chalcogen (S, Se)-Doped Graphene

This section illustrates graphene doped with sulfur and selenium, the chalcogen elements belonging to group 16.
The first papers, published in 2012 [454,586], evidenced how the introduction of sulfur can boost the ORR electrocatalytic activity. Unlike nitrogen, boron, and phosphorus, sulfur has an electronegativity (2.58) close to that of carbon (2.55). Therefore, the enhancement of ORR catalytic activity in sulfur-doped graphene can be ascribed to the asymmetric spin density rather than to the change of atomic charge distribution [586,587], taking also into account that the d-orbitals of the embedded sulfur atoms are soft nucleophiles, and the local strain, due to the bigger size of S atoms compared to C atoms, favors the ORR reactivity around these sites [588].
A few theoretical investigations [574,575,576,581,583,589,590,591] have been carried out on the catalytic processes involved in the ORR on S-doped graphene. Zhang et al. [589] conducted a specific study on catalytic mechanisms, proposing four types of sulfur-doping structures: surface S-adsorbed, edge S-substituted, edge SO2-substituted, and sulfur-ring connecting graphene clusters (Figure 42).
The S-doped graphene clusters with sulfur or sulfur oxide at graphene edges show electrocatalytic activity for ORR. Catalytically active sites are distributed at the zigzag edge or the neighboring carbon atoms of doped sulfur oxide atoms, which possess large spin or charge densities. For those systems where the sulfur atoms with the highest charge density are identified as the active catalytic sites, the ORR takes a two-electron transfer pathway, while in the case of the carbon atoms with high spin or charge density as active sites, it follows a four-electron transfer pathway. Moreover, evaluating the reaction energy barriers shows that sulfur-doped graphene is predicted to show ORR catalytic properties comparable to platinum, providing theoretical support for the discoveries made by Yang et al. [586], who prepared S-doped graphene by annealing GO with dibenzyl sulfide. The sample obtained at 1050 °C, characterized by a uniform distribution of S present in the plane and at edge positions, under alkaline conditions has an onset potential close to that of the Pt/C catalyst, the highest current density (at −0.8 V), and a comparable number of electron transfers, but with better stability and tolerance to methanol crossover than Pt/C.
In particular, since the theoretical demonstration that doping of a graphene sheet with S is possible with an average binding energy of 0.30 eV was published [591], there has been great effort in the development of efficient sulfurized metal-free ORR electrocatalysts, because the incorporation of S atoms into graphene requires a higher energy of formation than for the case of N or B atoms, making the synthesis of S-doped graphene more difficult [592].
The sulfur doping process is carried out at high temperature using single source precursors (thiophene [593], sulfonic resin [594], p-methylbenzenesulfonic acid [595], polyphenylensulphide [376]) or combining a C source (graphite/graphene derivatives, CH4, Na2CO3, and fructose) with different sulfur compounds [H2S, CS2, SO2, Na2Sn, benzyldisulfide, phenyldisulfide, triphenyl sulphonium chloride, (NH4)2S2O8, KSCN, Na2SO4, Na2S2O3, H2SO4, 4-benzenedithiol dihydrochloride, sulfur, thiourea, polyphenylensulphide] as S source [128,273,454,531,596,597,598,599,600,601,602,603,604,605,606,607,608,609,610,611].
In particular, synthesis via an ion-exchange/activation combination method using a 732-type sulfonic acid ion exchange resin [594] produced 3D S-doped graphene networks (SG3Dχ) characterized by a high crystallization degree that contained appropriate sulfur doping (the sulfur content is as high as 12.8%), mainly in the form of –C–S–C–, which is beneficial for achieving high oxygen reduction. In addition, the S-doped graphene has an extremely high surface area, which can increase the number of catalytic sites exposed to oxygen molecules. Due to their unique structure and composition, the SG3Dχ exhibited similar electrocatalytic activity but superior stability and methanol tolerance to the commercial Pt/C catalyst for four-electron oxygen reduction in alkaline solutions (Table 4).
Another big challenge to directly generating S-doped graphene is its synthesis from easily available, low-cost, nontoxic inorganic salts, such as Na2CO3 and Na2SO4, which are reduced by magnesium at high temperatures (Figure 43) [602].
This magnesiothermic reaction afforded S-doped graphenes characterized not onlyby high electrocatalytic activity for ORR with a four-electron reaction pathway, but also by much better stability and increased tolerance to MeOH crossover effects than commercial Pt/C (Table 4). The high ORR activity is mainly ascribed to highly graphitized structures, S-related active sites, and hierarchically porous textures.
However, other synthetic routes have also been developed to operate under mild conditions. Except for Ma et al. [612], who recovered S-graphene by continuous charge/discharge cycling of graphene–sulfur composites in Li–S batteries discharging and ball milling graphite with sulfur [502,587], alternative synthesis processes are carried out in solution [610,613,614,615,616,617,618,619], particularly under hydro/solvothermal conditions [120,588,614,619,620,621,622,623].
Chen et al. [620] first developed a low temperature (180 °C), economical, and facile one-pot hydrothermal method for synthesizing S-doped reduced graphene oxide nanosheets, S-rGOnSh, 180, by reacting GO with Na2S, which was, therefore, employed not only as a sulfur source but also to reduce GO. Sulfur was distributed uniformly into the graphene plane and at the edges, mainly in the form of thiophenic S. S-rGOnSh, 180, possessing numerous open edge sites and defects on its surface, showed superior electrocatalytic activity, long-term stability, and high methanol tolerance in alkaline media for ORR (Table 4). The authors speculate that the asymmetric spin density and the increase in edge plane defects play a key role in the observed ORR activity of S-graphene.
S-doped graphene nanosheets were synthesized by plasma-assisted [624] or simple electrochemical exfoliation of graphite using thiosulfate and H2SO4 as the sulfur source and the intercalation compound, respectively [610]. In this way, monolayer graphenes were obtained with fewer oxygen-containing functional groups (e.g., hydroxyl, carboxyl, carbonyl, and phenol groups) than conventional Hummers’ GO and had a less corrugated appearance as well. XPS showed that they contained sulfur substitutional defects, and sulfur doping provided materials with stable structural features. The main S2p peaks at ca. 163.4 eV, 164.6 eV, and 168.3 eV were attributed to the contributions of H-S-C, R-S-C, and C-SOx-C (x = 2, 3, 4), respectively. In alkaline media, the ORR activity is correlated with a nearly four-electron pathway characterized by higher stability and methanol tolerance than a benchmark commercial Pt/C (Table 4).
Functionalization of graphene with sulfur groups can also be achieved under mild reaction conditions. Oligothiophene, sulfonic groups, and thiol groups were covalently bonded to graphene without altering the graphitic basal plane [625,626,627,628].
Jeon et al. [626] prepared large-scale edge-selectively functionalized graphene nanoplatelets via ball milling graphite in the presence of hydrogen, carbon dioxide, sulfur trioxide, or a carbon dioxide/sulfur trioxide mixture. Such edge-selectively functionalized graphene nanoplatelets are highly dispersible in various polar solvents, and it was found that the edge polar nature of the newly prepared graphene nanoplatelets without heteroatom doping into their basal plane played an important role in regulating ORR efficiency. The graphene functionalized with SO3H and SO3H/COOOH showed the best electrocatalytic activity, with ORR performance at high potential even better than Pt/C. The ionic interactions between strong acidic sulfonic acids and strong basic KOH in the electrolyte solution make the sulfonated graphenes more hydrophilic, and sulfonic acids at the edges can be fully ionized in an alkaline medium to form a strong negative charge. As a result, they have a strong affinity for electrolytes and oxygen.
In addition to the graphene functionalized with group 16 elements, the results obtained by Pal et al. [629] must also be mentioned. They coupled rGO-I with seleniothiourea (Figure 44).
The Se–graphene derivative has been evaluated as an ORR electrocatalyst. Since selenium is present in very small amounts (0.16% atom), the system is considered a single-atom site catalyst. Coupling with Se results in a prominent four-electron ORR pathway, similar to Pt/C (Table 4). DFT-based calculations show that Se bridging results in a charge density distribution in graphene, and the Se bridge causes a huge distortion in the lattice distance. Based on Hirshfeld charge analyses, Se is shown to possess a positive charge site, and hence, it is considered the active center in this ORR catalytic process.
To the best of our knowledge, only two papers in the literature [129,586] report information on proper selenium-doped graphene, even though selenium has similar chemical properties to sulfur. For example, its electronegativity (2.55) is close to that of sulfur (2.58), and the bond length and bond angle of C–Se–C in selenophene (C4H4Se) are almost equal to those of C–S–C in thiophene (C4H4S). On the other hand, selenium has a larger atomic size and higher polarizability than sulfur: lone selenium pairs can easily interact with molecules in the surrounding electrolyte.
In 2012, Jin et al. [129] fabricated selenium-doped carbon nanotube/graphene composites by facile thermal annealing of the carbon materials with diphenyl diselenide in argon, characterized by excellent catalytic activity, long-term stability, and a high methanol tolerance (Table 4). Se doping was confirmed by XPS and UV-Vis spectroscopy, showing that Se atoms are successfully introduced into the graphene framework via covalent bonds. The significantly enhanced catalytic activity observed is believed to be due to the increased electron transfer since the introduction of Se allows the restoration of defects at the edge of carbon materials and induces the formation of a π conjugation system.

6. Boron-Doped Graphene

Boron belongs to group 13 and is s immediately before C in the periodic table; boron atoms, possessing an atomic size similar to that of carbon and three valence electrons for binding with carbon atoms, can be introduced into the carbon lattice by substitutional doping. Boron doping into a carbon network gives the host carbon array p-type conductivity. Because of the lower electronegativity of B (2.04) compared to that of C (2.55), B atoms are positively polarized in the carbon lattice and hence able to attract the nucleophilic oxygen molecules, leading to efficient chemisorption; on the other hand, boron sites can also act as electron shuttles for the electron density of the graphite π-electron system going through the pz orbital of boron to the chemisorbed O2 molecule [512,624].
The local positive charge density due to the incorporation of boron atoms into the graphene layer and enhancement of CO tolerance in the oxygen reduction reaction enables effective and stable electrochemical behavior [630]. B-doped graphenes have therefore attracted significant scientific attention.
Several theoretical studies have been carried out on the ORR process catalyzed by B-doped graphenes [128,571,572,574,575,577,578,581,582,631,632,633,634,635].
It must be considered that boron, which belongs to the same period as carbon, maintains the graphene surface basically flat, which helps in calculating the contributions of the electrons of the heteroatom to the delocalized density, even if it also depends on the adopted model [579].
Sheng et al. [636] first suggested that the electron-deficient boron atoms doped in graphene might function as active sites for oxygen adsorption and activation of the O–O bond cleavage. Based on the analysis of charge redistribution, the formation of active catalytic sites is attributed to the localized positive charge and electronic dipole induced by the dopant [574].
The kinetic pathways of oxygen reduction catalyzed by B-doped graphene have been analyzed. Fazio et al. [632] carried out a detailed study of the possible reaction paths for ORR as catalyzed by B-doped graphene through the identification of all the intermediates and transition structures for both dissociative and associative mechanisms, computing also the water contribution.
Among all the intermediate and transition structures that have been characterized along the various reaction coordinates, the most significant species is the end-on dioxygen species, an open-shell intermediate resulting from the adsorption of molecular oxygen on the positively charged B-doped atom. From the free energy diagrams at different electrode potentials derived by using the methodology developed by Nørskov et al. [637], they were able to determine the overpotentials both under acidic and alkaline conditions. In particular, at pH = 14, the overpotential is 0.28 V, which corresponds to an onset potential of 0.05 V, in excellent agreement with the experimentally reported value for ORR on B-doped graphene samples [128].
Despite the theoretical potential of B-doped graphene, from the experimental point of view, the literature on B-doped graphene tested as an ORR electrocatalyst is sparse, mainly because of difficulties at the synthesis level and low doping levels [638].
Synthesis of B-doped graphene has been mainly achieved at temperatures ranging from 400 to 1200 °C with various boron sources (boric oxide, boric acid, tetraphenyl phosphonium tetraphenyl borate, triphenyl borane, and Filmtronic B155 compounds) [385,531,636,639,640,641,642,643,644,645,646,647,648,649].
Zuo et al. [641] prepared a 3D B-doped porous graphene framework with hierarchical pores from frozen GO sediment containing boric acid via quick heating. The morphology of the porous graphene and the doping level can be effectively tuned by controlling the reaction temperature. During the heating, the frozen ice not only maintains the volatilization rate of water but also produces steam, which can introduce pores into the graphene sheet at high temperatures [650]. The nanopores are hence produced because of the strong oxidation ability of steam via the so-called water gas reactions at high temperatures. As for the catalyst, the novel hierarchical porous structure of the B-doped graphene framework can provide edge-rich planes and a large specific surface area with more active sites for ORR. The samples treated at 600 and 800 °C have ORR activity close to that of Pt/C, but with better durability against methanol.
3D B-doped graphene was also obtained using supercritical carbon dioxide, a one-pot and green method [651], where GO and borane adducts were mixed and reacted. The porous network structure of the products is likely the result of exfoliation and expansion of the rGO due to the action of the supercritical carbon dioxide fluid.
The choice of operating under higher pressure allowed synthesis at a lower temperature under hydrothermal conditions [638,652,653,654]. Tam et al. [638] developed a simple and green one-pot synthesis method for novel nanocomposites consisting of a graphene hydrogel (GH) and B-doped graphene quantum dots (B-GQDs), (GH-BGQD), fabricating B-doped graphene quantum dots with a uniform size and a controlled boron concentration from glucose and boric acid (Figure 45).
XPS analysis identified various structures, such as the graphite-like BC3, and oxidized B-C bonds, such as BC2O and BCO2, to which the electrochemical behavior was correlated. Under alkaline conditions, the composite containing the higher percentage of BC3, even if with a lower B-doping level with respect to another sample, showed the most positive onset potential (−0.05 V vs. SCE) and highest limiting current density (4.2 mAcm−2), which are close to those of Pt/C (−0.08 V vs. SCE and 4.6 mAcm−2) (Table 5). Theoretical calculations have evidenced that the adsorption of O2 on the BC3 structure in BGQDs is more favorable than those of the BC2O and BCO2 structures, as deduced from the calculated O2 adsorption energies of the BC3 structure (−0.190 and −0.204 eV) and those of the BC2O and BCO2 structures (−0.162 and −0.133 eV, respectively).
Moreover, the prepared GH-BGQD composites, which are characterized by a unique 3D architecture with high porosity and large specific surface area characterized by abundant catalytic active sites of BGQD as well as enhanced electrolyte mass transport and ion diffusion, exhibited superior trifunctional electrocatalytic activity toward the oxygen reduction reaction, oxygen evolution reaction, and hydrogen evolution reaction with excellent long-term stability and durability: a Zn–air battery and a water electrolysis cell were demonstrated using the GH-BGQD electrodes and were found to deliver a superior performance with a highly stable operation.
B-reduced graphene oxide obtained via hydrothermal reduction of GO [654] was recently integrated with activated carbon and carbon black as part of the MFCs’ cathode catalyst, significantly improving the output voltage and power density, with ameliorants of 75% and 58% over their undoped counterparts, respectively.
As an alternative to the reported hydrothermal reactions, boron-doping can also be achieved electrochemically to produce BGqDot and three-dimensional graphene networks [512,655].

7. Phosphorus-Doped Graphene

Phosphorus belongs to group VA, similar to nitrogen, and possesses the same number of valence electrons; however, its electronegativity (2.19) is lower than that of carbon, which determines a consequent polarized distribution of charge density, a relevant feature for imparting ORR electrochemical properties to the phosphorus-doped graphene. Phosphorus has a relatively large atomic radius with respect to carbon, which makes the synthesis of P-doped graphene quite a challenge; hence, there are very few papers dealing specifically with experimental evaluation of synthesized P-doped graphene as ORR electrocatalysts compared with nitrogen-doped graphene [128,273,531,640,642,656,657,658,659,660,661,662,663,664].
The first papers on P-doped graphene and its electrochemical ORR properties were published in 2013 [656,657]. Zhang et al. [657] prepared P-doped graphene (GP) via a facile, low-cost, and scalable thermal annealing method using GO and triphenylphosphine as carbon and phosphorus sources, respectively. GP has a crystalline structure, and P was detected via EDX and XPS (about 1.3–1.8 at% P), indicating the presence of P-C and P-O bonds. GP exhibits outstanding ORR activity (close to Pt/C) as well as excellent selectivity and stability in alkaline media (Table 6).
A few theoretical studies have been carried out to elucidate the active sites and ORR mechanism in P-doped graphene [572,573,574,575,576,579,581,582,583,631,637,665,666,667]. Del Cueto et al. [572] developed a model to show that ORR via a four-electron transfer mechanism is energetically more favorable than the two-electron transfer mechanism. In addition, the energies calculated for each ORR step showed that the P-doped surface favors ORR better than the N and S-doped surfaces.
While Jiao et al. [128] showed that the active sites of ORR are the adjacent carbon atoms of heteroatoms, Zhang et al. [665] indicated that the P atom and its neighboring C atoms are the most preferred adsorption sites and the active center for the ORR. All of the ORR species adsorb on the P atom strongly, and all of the possible ORR elementary reactions take place within a small region around the dopant. Such divergence arises from the coordination and chemical bonding of P in carbon, which play an important role in ORR catalysis. Yang et al. [667] tried to provide a better understanding by considering the species experimentally, considering the possibility of P forming three to five bonds thanks to the hybridization of d orbitals (Figure 46).
Combining stability, conductivity, ΔG of the rate-determining step, and the number of potential active sites of the P-doped graphene structures, OPC3G was the most effective and stable P-doped graphene for ORR among the four types of P-doped structures—PC3G, OPC3G, PC4G, and OPC4G sheets.

8. Halogen-Doped Graphene

The halogen elements, fluorine (F), chlorine (Cl), bromine (Br), and iodine (I) belong to group 17 of the periodic table. They are all characterized by higher electronegativity than carbon, decreasing from F to I, and, except for fluorine, also have larger atomic sizes. They are also resonance-donating due to the existence of a lone pair of electrons.
Fluorine has strong electronegativity and theoretical calculations showed that by partially fluorinating graphene, the material energy band gap can be changed from 0.8 to 2.9 eV [668] and that a particular F-doped configuration is thermodynamically as good as Pt for ORR [581]; nevertheless, to the best of our knowledge, very few articles in the literature report on the evaluation of the ORR electrocatalytic properties of fluorinated graphene [669,670,671,672,673] mostly because of synthesis problems arising, for example, from safety issues. Kakaei and Balavandi [671] fluorinated graphene-based high-performance metal-free ORR electrocatalysts and optimized the amount of fluorine in rGO by mixing rGO in H2SO4 with NaF. They prepared F-rGO by randomly stacking graphene nanosheets with an interconnected 3D porous structure possessing C-F bonds, as clearly identified via IR. Doping with fluorine imparted better electrochemical activity, suggesting a 4-electron process for the best electrocatalyst in alkaline medium.
Moreover, focusing strictly on halogen-doped graphene, fluorine has been found to determine better ORR performance when used as a co-dopant with other halogens [671] and with just chlorine in halogen-doped graphene [674]. In the last case, F,Cl-rGO, synthesized from Cl3F as a single source precursor, is characterized by an onset potential of ORR close to that of the commercial Pt/C catalysts (40% w/w), but with much better tolerance to methanol crossover. F,Cl-rGO is a better ORR electrocatalyst than Cl-rGO, which contains only chlorine (Table 7).
There are few articles reporting on chlorine-doped graphene as an ORR electrocatalyst [245,671,675,676]. Zhang et al. [245] modified CVD graphene using the photochlorination method by irradiating with a Xenon lamp. Cl2 was generated in situ in a photoreactor containing vertically grown graphene on carbon cloth (Figure 19a).
This method induced the covalent attachment of Cl on the basal plane in a homogeneous way, generating a high degree of structural defects. Chlorinated vertically oriented graphene grown on carbon cloth exhibits improved ORR electrocatalytic activity, giving rise to more positive O2 reduction voltage and larger limited current density than benchmark electrocatalysts such as carbon cloth and pristine vertically oriented graphene grown on carbon cloth (Table 7). Moreover, chlorinated vertically oriented graphene grown on carbon cloth served as a bifunctional air-cathode for the construction of solid-state zinc air batteries, which showed stable discharge potential and high power and exhibited long-lasting rechargeability over 108 cycles at a current density of 2 mAcm−2, superior to noble metal-based zinc air batteries. The authors ascribe the good performance to the synergetic effects between defective sites, exposed edges, and the high electronegativity of Cl dopants.
It is, however, interesting that the high electronegativity of the dopant is not always the key factor for high ORR activity. Br- and I-doped graphenes were found to promote ORR more efficiently than analogous Cl-doped graphene [675,676] (Table 7), and therefore, the origin of ORR activity is thought to be related to the local electronic and bonding configuration around the dopant in the carbon lattice [677].
Jeon et al. [675], who prepared edge-selectively Cl-, Br-, I-doped graphene nanoplatelets (GnPl) via ball milling graphite in the presence of X2 (X =, Cl, Br, I), evidenced that the observed order of ORR activities ClGnPl < BrGrnPl < IGnPl follows the atomic sizes (Cl atomic radius < Br atomic radius < I atomic radius) (Table 7). Hence, the valence electrons of Br and I are more loosely bound than those of Cl, facilitating charge polarization in the BrGnPl and IGnPl electrodes. Unlike Cl, Br and I can form partially ionized bonds of –Br+– and –I+. DFT calculations, performed for various edge configurations, showed that the edges of halogenated graphene have a favorable binding affinity with the O2 molecule, and the O-O bond is weakened because halogenation induces charge transfer, whose efficiency increases as the atomic radius increases.
To the best of our knowledge, apart from this paper and two others [671,676], no specific or further research has been conducted on Br-doped graphene. Among the halogenated graphenes, iodine derivatives are definitely the most studied [131,671,675,676,678,679,680,681,682,683,684].
Yao et al. [131] first utilized intercalated iodine graphene, obtained by annealing graphene oxide and iodine at 500–1100 °C in argon, as an ORR electrocatalyst, prompted by the already known application of iodine in conducting polymers [685,686]. Iodine is an important dopant that improves the electrical conductivity of materials, and graphene has a conjugate structure similar to that of conducting polymers.
As an alternative to molecular iodine [131,678,681,682] KI, NaIO4, HI, and deep eutectic solvents have been used to synthesize I-doped graphene [671,676,679,680,681,683,684,687,688]. Hoang et al. [684] exposed their hybrid to vapors of hydrogen iodide to simultaneously reduce the GO/CDot composite and introduce iodine dopants into the structure, while Marinoiu et al. [679,680,681] conducted solution-based electrophile substitution of KI/NaIO4 and graphene or a nucleophile substitution on graphene oxide using HI, also serving as a reducing agent [679,681,683]. The ORR activity of the iodine-doped graphenes was improved, and better performance of PEMFC was obtained when iodinated graphene was included as a microporous layer set between the catalyst and the gas diffusion layer [679,681,683] or in the cathode of the fuel cell with Pt/C [680].
It is interesting to observe that iodine doping occurred via the formation of iodine intercalated graphene, where I3- and I5- species are present, as detected using IR and XPS analysis. As already observed by Yao et al. [131], the presence of the I3- structure plays a crucial role in the enhancement of the ORR activity of graphene. In particular, a DFT study on polyiodide-doped graphenes by Hoyt et al. [689] showed that on the basis of band structure calculations and molecular orbital theory, the I3 molecule acts as a p-type dopant that shifts the Fermi level 0.46 below the Dirac point without noticeably disrupting graphene’s band structure. The I3 chains acquire charge in π* orbitals that are oriented with lobes perpendicular to graphene. Adsorption energy calculations reveal that I3 acts as an effective catalyst for the oxygen reduction reaction on graphene by stabilizing the rate-limiting OOH intermediate. I3 can serve as the active site for OOH adsorption and hence promote the early stages of oxygen reduction.

9. Multielement-Co-doped Graphene

The introduction of dopants into the carbon lattice has been the focus of many studies aimed at producing highly efficient electrocatalysts for the ORR via the alteration of charge or spin distribution of the sp2 carbon plane. Co-doping is considered an effective way of improving the electrochemical properties of the single heteroatom-doped graphene.
Different heteroatoms in the conjugated carbon backbone can create new non-electron-neutral sites, and except for very few cases [674,690,691,692,693], co-doped graphene always contains nitrogen as one of the heteroatoms.
The co-doped graphenes will be illustrated depending on the second heteroatom (S, Se, B, P, halogens) combined with nitrogen, including a subsection with the tri- and tetra-doped species.

9.1. N,Chalcogen (S,Se)-Co-doped Graphene

To adjust the bonding site of N in carbon nanomaterials, the addition of S as a further dopant helps to tailor the activity of the nanocarbon materials for the ORR through a higher degree of electron delocalization and polarization [586]. While nitrogen polarizes the adjacent carbons, the sulfur electronegativity characteristic and bulk size result in easy polarization and induce charge density/spin density on the surrounding graphene matrix, thus enhancing the ORR activity [694,695]. Furthermore, two lone-pair electrons of sulfur help in tuning the electronic band gap in the graphene framework to achieve a favorable electrochemical ORR. In addition, S-doped system alters the intrinsic hydrophobic nature of graphene to hydrophilic, and thus, it creates a strong affinity toward the adsorption of molecular oxygen, which results in enhanced ORR activity [696]. Remarkably, co-doping increases the density of active catalytic sites and induces a synergistic coupling effect in graphene layers, leading to further enhancement of the electrocatalytic ORR activity [697,698].
This effect has also been explained in terms of DFT calculation, where the ORR reactivity was related to the induced spin density after N,S-doping [596]. Song et al. [699] also showed chemical adsorption of the oxygen molecule on the N,S-co-doped graphene, which is much stronger than on either the mono N- or S-doped graphene.
The first design and preparation of N and S dual-doped mesoporous graphene as a metal-free catalyst for ORR in the literature was reported by Liang et al. [596]. Porous N,S-doped graphene was prepared via a very simple one-step doping process using solid and low-cost precursors (melamine and benzyl sulfide), in which porosity was introduced by employing commercially available colloidal silica. The catalyst showed excellent ORR performance comparable to that of commercial Pt/C as well as full fuel tolerance and much better long-term stability than Pt/C in an alkaline environment. DFT calculations revealed that the synergistic performance enhancement results from the redistribution of spin and charge densities were brought about by the dual doping of S and N atoms, which results in a large number of carbon atom active sites.
Right after the first successful co-doping of nitrogen and sulfur, the co-doping of N and Se of a graphene carbon nanotube composite was reported [700]. Se-doping demonstrated much stronger effects on ORR activity compared to S-doping, but the research on co-doping graphene with nitrogen and selenium did not advance.
On the contrary, a few more papers appeared on N,S-co-doped graphene–CnT composites [701,702].
Villemson et al. [702] proposed a novel and effective one-pot synthetic approach to prepare N,S-doped carbon catalysts by using heat treatment of the GO and multiwalled CnTs mixture, together with o-methylisourea bisulfate. The composite showed superior ORR performance in alkaline media (Table 8), and the electrocatalytic mechanism for the reduction of oxygen was well explained by DFT calculations of graphene sheets according to the XPS data and ORR test results. The Cα and Cβ carbon atoms at the graphitic nitrogen located close to the sulfur atom correspond to the global minimum and are thereby energetically most favorable (−1.14 eV) for oxygen electroreduction.
Recently, Huang et al. [701] reported CnT-N,S-co-doped graphene-based 3D carbon composites (CnT/N,S-graphene3D) with high activity toward a set of important electrochemical reactions and high performance in Zn–air batteries. In the composite, graphene is found to strongly couple with the surface of CnTs, achieving uniform distribution of both components and combining optimal nitrogen configuration (graphitic and pyridinic) and well-developed porosity. The effective integration of 1D CnTs and 2D graphene into 3D conductive frameworks was achieved in a facile and sustainable preparation via an in situ direct pyrolysis in the presence of CnT of a layered biomolecule guanine and guanine sulfate (Scheme 5), which typically exists in the form of a layered structure through intermolecular H-bond interactions.
The strategy of using just one compound as a nitrogen and sulfur source is the most exploited one, thanks to the large availability of reagents [209,608,609,614,623,658,701,702,703,704,705,706,707,708,709,710,711,712,713,714,715,716,717,718,719,720,721,722,723,724,725,726]. The preferred procedure is based mainly on the thermal treatment of suitable precursors [608,658,701,702,703,704,705,706,708,711,715,717,719,720,721]. Pan et al. [713] reported the facile and convenient preparation of N,S-co-doped graphene from low-cost and green raw materials, consisting of the pyrolysis of cysteine in the presence of NaCl, which serves as a structure-directing template to direct the growth of graphene with the formation of curved graphene domains. The product has remarkable ORR activity based on an almost 4-electron transfer process (Table 8). Arunchander et al. [720] carried out simultaneous doping of nitrogen and sulfur using in situ polymerization of 6-N,N-dibutylamine−1,3,5-triazine-2,4-dithiol on a graphene framework and subsequent pyrolysis. The sample treated at 1000 °C (N-S-Gr−1000), exhibiting an enhanced ORR activity dominated by the 4-electron pathway compared to other catalysts (Table 8), was the first N,S-co-doped graphene whose potential as a cathode catalyst was validated in a membrane electrode assembly and in a real AEMFC. A peak power density of 20 mWcm−2, which is ~3.4-fold lower than that of the Pt/C catalyst, was achieved under ambient temperature and pressure, which makes N-S-Gr−1000 a promising alternative nonprecious metal catalyst in AEMFCs.
On the other hand, good electrocatalysts based on N,S-co-doped graphenes can also be obtained under mild reaction conditions [209,609,614,623,707,716,718,723,727]. Wu et al. [722] prepared N,S-co-doped graphene aerogel with a 3D hierarchical porous structure and high specific surface area from the amino acid methionine and GO under hydrothermal conditions and subsequent freeze-drying. Methionine acted as a reductant, as well as an N and S source. The N,S-co-doped graphene aerogel exhibited good mechanical properties, excellent capacitive performance, and ORR performance, with selectivity tending towards a 4-electron process with increasing voltage. Favaro et al. [293] used an electrochemical approach to synthesize doped graphene quantum dots via electrochemical etching of GO with cysteine. It was observed that the reaction selectivity is controlled by the oxidation degree of the materials: as-prepared graphene oxide quantum dots, which have highly oxidized functional groups, follow a two-electron reduction pathway and produce hydrogen peroxide, whereas after reduction treatment with NaBH4, the same materials favor a four-electron reduction of oxygen to water.
N,S-co-doped graphene quantum dots were obtained in the last few years by hydro- or solvothermal treatment [376,637]. In the presence of NH4OH and Na2S as precursors, Fan et al. [619] synthesized N,S-GqDots, which possessed enhanced electrocatalytic activity compared to GqDot, NGqDot, and SGqDot and their composite with GO, even comparable to Pt/C (Table 8). The introduction of S changes the state of N species in N,S-GqDots, creates asymmetrical spin and charge density, and enables the synergetic effect of N and S, which facilitates the electron transfer in the ORR process.
It is interesting to observe that the use of two compounds as separate sources of nitrogen and sulfur opens the opportunity for a one [376,593,596,608,617,618,728,729,730,731,732,733,734] or two-step doping process [598,604,607,735]; however, except for very few examples [617,618,703,730], thermal treatment at high temperature is the common synthesis method. The temperature approach is commonly used to tune the features of the active centers during material fabrication for ORR performance improvement.
Li et al. [735] developed a well-structured process (Figure 47) where crosslinking of GO by the addition of melamine and formaldehyde occurred using a hydrothermal method; after the addition of dibenzylsulfide and pyrolysis, 3D N,S-co-doped graphenes (NS-3DrGO) were obtained.
The sample obtained at 950 °C had the optimal ORR activity compared with the others (Table 8), which may be due to highest amount (74.8 at.%) of the two active nitrogen species (pyridinic N and graphitic N) and the highest amount (79.8 at.%) of active thiophene-S together with the desirable specific surface (391.9 m2g−1) area and multi-porous structure. Furthermore, the N,S-3D rGO catalysts also exhibit superior methanol tolerance and favorable durability.
Moreover, very recently, Zhang et al. [734] were able to tune the configuration between graphitic-N and thiophenic-S dopants in one-step-synthesized graphene nanosheets by selecting suitable S precursors, obtaining a material which performed well in a Zn–air battery as cathode, producing a high-power density of 146 mWcm−2.

9.2. N,B-Co-doped Graphene

Co-doping with nitrogen, which has a higher electronegativity (3.04), and boron, which has a lower electronegativity (2.04) than C (2.55), can create a unique electronic structure with a synergistic coupling effect between the heteroatoms.
The introduction of N and B in graphene systems was predicted to enhance ORR activity [297,342] even before any experimental proof; this was first achieved by Wang et al. [736], who developed a facile low-cost approach for mass production of N,B-co-doped graphene with tunable N-/B-doping levels simply by thermal annealing graphene oxide in the presence of boric acid under ammonia atmosphere. The resultant N,B-co-doped graphene samples showed ORR electrocatalytic activities that were even better than the commercial Pt/C electrocatalyst, and with an electron transfer close to 4. DFT calculations indicated a model with a low bandgap and therefore high conductivity; however, it also had many more carbon atoms with relatively high spin density and charge density compared to pure graphene, thus providing more active sites to catalyze ORR: the theoretical results agreed with the experimental data.
It is interesting to observe theoretical research on the mechanism of ORR catalyzed by N,B-co-doped graphene, analogous to what is reported for the single B-doped graphene [512,572,574,577,578,639,648,737,738,739,740,741,742,743].
The BN systems have been analyzed considering mostly different BN clusters; Sinthika et al. [577] analyzed a co-doped graphene containing a simple BN cluster. They identified a descriptor of N,B-co-doped graphene ORR performance, the variance of the average of pz occupancies of the carbon atoms adjacent to boron, finding a linear correlation with the free energy of OH adsorption. Kattel et al. [737] identified graphitic BN3 motifs as active sites for complete O2 electroreduction in the N,B-co-doped graphene electrocatalyst. Increasing N coordination to B gradually shifts the band gap below the Fermi level, so that graphene with the BN3 defect motif has a small band gap below the Fermi level, showing the increase in electron (donor) concentration states. On the other hand, Tang et al. [740], considering that the BN clusters embedded in graphene involve different sizes and shapes as well as edge termination, discussed the effect of edge termination and the shape of substitutional BN clusters on oxygen dissociation barriers; they found that N-terminated triangular BN (t-BN) cluster doping can reduce the energy barrier more effectively compared to a t-BN with a B edge or quadrangular BN cluster. The enhancement of catalytic activity by triangular BN doping is attributed to the positively charged active site and its large density of unoccupied electronic states around the Fermi level.
However, from the experimental point of view, the presence of BN bond, and in particular the formation of hexagonal boron nitride, is sometimes recognized as detrimental for the ORR performance, with a preference for a structural B-C-N configuration. This is the reason why Zheng et al. [639] developed a two-step co-doping method: GO was annealed with NH3 at an intermediate temperature (e.g., 500 °C), and then B was introduced by pyrolysis of the intermediate material (N-graphene) with H3BO3 at a higher temperature (e.g., 900 °C), enabling the incorporation of heteroatoms at selected sites in the graphene framework to induce a synergistic enhancement of the activity of the N,B graphene. In this way, the formation of inactive by-products was prevented, and the new catalyst showed excellent activity in the ORR and perfect (nearly 100%) selectivity for the four-electron ORR pathway in an alkaline medium (Table 8). The better ORR performance than that of the single boron and nitrogen-doped graphenes was ascribed to the synergistic coupling effect between heteroatoms, which arises from the charge transfer process in which N has the role of an electron-withdrawing group, indirectly activating B, and thus makes the latter an active site to enhance the ORR activity.
The importance of the presence of the B-C-N hetero-ring or the absence of BN bond to achieve the synergistic effect has been recognized by other authors [645,649,742,744,745,746], who accordingly analyzed a dual-step synthesis procedure using two separate N and B sources in order to avoid direct combination of the heteroatoms. In this perspective, Qin et al. [742] succeeded in controlling BN configurations and producing N,B-co-doped porous graphene with ORR catalytic activity comparable to that of Pt/Calso in acidic media (Table 9).
The use of different nitrogen and boron sources is the most preferred way for co-doping materials, mainly because of the larger availability of compounds [177,512,639,644,645,646,648,649,655,739,742,744,745,747,748].
Gong et al. [739] built 3D interpenetrating networks from numerous flexible ribbons and showed the importance of an optimal doping level of boron and nitrogen in graphene nanoribbons to obtain a material with excellent ORR electrocatalytic activity, even better than the commercial Pt-C catalysts (Table 8). A low BN content results in a low number of active sites, while BN content that is too high undermines the conductivity of the material, weakening the charge transport from the electrode to oxygen and decreasing electron transfer.
Even if less exploited, the strategy of using a single source of N and B for doping may have the advantages of energy-saving, less time-consuming, less waste of precursor, and less complicated operating apparatus [208,415,643,746,749,750,751].
Jiang et al. [643] prepared boron and nitrogen-co-doped hollow graphene microspheres using a simple template sacrificing method in which doping of the amino-modified SiO2 nanoparticles supported GO with N and B occurred in the presence of NH3BF3 under hydrothermal conditions. Silica was then removed through HF etching. The as prepared microsphere present high electrocatalytic activity (Table 8), which could be attributed to the synergetic effect of the N,B-co-doped graphitic structure and the specific microspherical hollow morphology, which promotes the exposure of more surface area accessible to electrolytes and decreases the overpotential for the ORR by reducing the contribution from the mass transport limitation.
Other porous structures consisting of graphene aerogels with tunable contents of B and N configurations were obtained by Chen et al. [751] from NH4B5O8 using a hydrothermal method and freeze-drying process. Linear sweep voltammetry results confirmed that the increasing doping contents of pyridinic N and BC3 phases in N,B-doped graphene aerogels can boost ORR activity (Table 8). The proposed co-doped aerogels not only show similar ORR activity to commercial Pt/C, but also superior stability and excellent methanol tolerance; moreover, the performance was appreciable when it was tested as an air cathode for a rechargeable zinc–air battery device.
N,B-doping can also be achieved by functionalizing graphene with L-cysteine [752] or with amino-functionalized boron nitride [753] via cross-linking.

9.3. N,P-Codoped Graphene

Phosphorus like boron, has lower electronegativity than carbon, disrupts the charge uniformity of carbon, creates new charge and structural defect sites, and increases the electron delocalization [754]. On the other hand, like sulfur, it has a relatively larger atomic size than that of C, and consequently produces large defect sites, more edge sites, and high curvature in the carbon lattice [755,756].
This system and the N,P synergistic effect has been studied via DFT calculations [573,663,757] but starting from different models; in one, Gracia-Espino et al. [573] considered substitutional P and substitutional N separate atoms, while in the other one, Xue et al. [757] showed that N and P atoms are inclined to bond with each other, forming N-P clusters in graphene. Gracia-Espino et al. [573] illustrated that the possibility of quantifying and determining the spatial distribution of catalytic sites permitted the construction of reduced overpotential (ηORR) maps; such maps provided an insight into the source of the synergistic effect of N,P-co-doped graphene commonly reported in the literature. Xue et al. [757] evidenced that the catalytic performances of the N-P clusters are sensitive to their geometries, and especially to the N:P ratios.
Experimentally, the existence of P-N moieties and their catalytic effects have been reported by Li et al. [660], who successfully constructed 3D porous carbon nanotube/graphene hybrid foams with embedded N,P-coupled active species by one-pot pyrolysis method (Figure 48) using graphene oxides and carbon nanotubes as building bricks and hexamethylphosphoric triamide (HMPA) as the special source of P and N.
Impressively, the HMPA compound not only facilitated the remarkable increase in the content of both P and N in the carbon frameworks but also offered a high density of the coupled N-P moieties embedded in the graphene surface. Together with the optimized local structures, including open porosities and high electron transportation capacity, this foam exhibits excellent electrocatalytic performance for the ORR in acidic medium. In particular, the ORR onset potential and half-wave potential reached 0.98 and 0.78 V in 0.1 M HClO4 (Table 9).
In another research, Yang et al. [758] prepared N,P-co-doped graphene-based electrocatalysts with high activity also in acidic media (Table 9). They assembled carbon nanotubes and graphene into hybrid nanospheres via a one-step aerosol route, which, after doping with nitrogen and phosphorus, showed higher oxygen reduction reaction activity, better stability, and better methanol tolerance for ORR than the commercial Pt/C catalyst in alkaline solution.
N,P-co-doping was mainly carried out at high temperature, which was necessary to push the large phosphorus atom into the graphene backbone or to form P-C on the edge [660,662,663,758,759,760,761,762,763,764,765,766,767]. There are very few papers dealing with synthesis under mild conditions [208,661,768,769]. Ma et al. [208] reported a safe, facile, and feasible solvothermal method for synthesizing N, X (X = B, P, S) dual-doped graphene by utilizing CCl4 and metal K as precursors and introducing ionic liquids into a Wurtz-type reductive coupling reaction (Scheme 6).
Ionic liquids, which are easily soluble in various solvents [770,771], play a dual role in the fabrication process: on one hand, the ionic liquids containing an organic, nitrogen-containing cation and a bulky inorganic anion can provide two or more doping elements with target materials. On the other hand, the released gas resulting from the decomposition of ionic liquids will act as a porogenic agent, giving rise to some pores in the resultant doped graphene. The improved electrochemical performance of all co-doped graphene samples indicates that dual-doping actually exerts a positive influence on electrochemical activity, which could be regarded as the synergistic effect. All three types of co-doped graphene have good selectivity for a four-electron pathway. Among them, N,P-co-doped graphene exhibited the best catalytic activity in terms of half-wave potential, methanol tolerance, and long-term durability (Table 8). This could be attributed to the change of charge density and high distortion of carbon structures resulting from the combination of the large electronegativity of N and the large covalent radius of P.

9.4. N,Halogen-Co-Doped Graphene

Fluorine and iodine are the only halogens used for co-doping graphene with nitrogen.
While in the case of a co-doped sulfur iodine graphene [690], the halogen is introduced via a well-defined mechanochemical ball milling, co-doping with nitrogen was achieved mainly via thermal treatment of graphene with iodine [678,772].
Hassan et al. [772] performed, for the first time, iodine and nitrogen co-doping of graphene using iodine, poly(aniline) (PANI), and activated graphene as starting materials (Figure 49).
Compared with N-doped graphene (NG), iodine and nitrogen co-doped graphene (NIG) exhibited enhanced surface area, resulting in a direct four-electron reaction pathway, high onset potential, and high current density as well as improved resistance to methanol poisoning and superior catalytic durability for ORR in alkaline medium. The synergistically enhanced ORR performance of NIG was found to be a result of a high strain and size advantage of the larger iodine atom clusters (compared to nitrogen), which facilitate the simultaneous enrichment of anode electrons and O2 and H2O molecule transport at catalytic sites, inducing four-electron transfer in a single step.
Simple reduction of graphene oxide with 4-iodophenylhydrazine produced nitrogen and iodine co-functionalized graphene oxide with enhanced dispersibility in organic solvents and electrochemical ORR activity because of the electronic modifications of the graphene structure [773].
Analogously, graphene sheets functionalized with amino groups and fluorine in an out-plane scenario were obtained by Zhao et al. [774]; optimization of the ratio of NH2 and F groups allowed the achievement of a material with outstanding electrocatalytic properties, even delivering slightly better electrocatalysis than commercial Pt/C catalyst (Table 8).
Such materials were developed as alternatives to the other F,N-co-doped graphenes that have appeared since 2014 [669,670,692,775,776,777,778,779].
Vineesh et al. [669] developed bulk nitrogen and fluorine-doped graphene (FNG) via a simple method. The obtained FNG is characterized by enhanced electrocatalytic efficiency (Table 8) with respect to the N and F individually doped graphene, benchmarked electrocatalyst—Pt/C and other reported doped graphene systems, which is ascribed to the synergistic effect of the heteroatoms. In order to understand the synergistic doping effects and the role of ‘electron spin density’ in determining the catalytic efficiency of graphene, DFT calculations were carried out. According to the DFT calculations, FNG with all three types of nitrogen at close proximity exhibited the most catalytic activity towards ORR. The order of activity towards ORR deduced from the DFT calculations follows the order FNG > FG > NG and is in excellent agreement with the experimental results.
Qiao et al. [776], who prepared reduced graphene oxide co-doped with nitrogen and fluorine in a simple one-step thermal annealing process using a mixture of graphene oxide and ammonium fluoride in an argon atmosphere, evidenced that the additional doping with F considerably enhanced the performance of the catalyst. This could be attributed to fluorine’s high electronegativity and presence in the form of the most active semi-ionic C-F bond, which could have (i) increased the adsorption of oxygen, (ii) activated O-O bond cleavage, and (iii) formed a synergetic effect with the N dopant.
In any case, fluorine plays a determining role in activating the graphene systems not only when co-doped with nitrogen but also with boron. Nitrogen and fluorine-co-doped graphene and boron and fluorine-co-doped graphene prepared via a simple one-pot hydrothermal treatment method exhibit analogous catalytic and definitely better than the respective single-doped species [690].

9.5. Tri- and Tetra-Doped Graphenes

Multiple heteroatom species doping not only introduce more active sites but also generate new properties or create synergistic catalytic effects [780,781], a point that Wang et al. [782] showed in a bit of a satiric way in 2020. To complete the description of the various doped graphenes, this subsection reports graphenes with three or four different heteroatoms, consisting of nitrogen in combination with phosphorus and sulfur [658,783,784,785], with phosphorus and boron [691,786], with sulfur and fluorine [616], and with phosphorus, boron, and sulfur [273,787].
Graphene doped with N,P and S is the most studied tri-doped system and was the first to be reported [658]. P dopant is known to disrupt the charge uniformity of carbon, create new charge and structural defect sites, and increase the electron delocalization. On the other hand, S dopant can induce an enhanced spin density of carbon and polarizability and increase structural defects. Furthermore, these two elements, which have relatively larger atomic sizes than C, produce large defect sites, more edge sites, and high curvature in the carbon lattice [755,756]. Additionally, topographic defects produced under high temperature treatment may be another effective factor for improving onset potential.
Fan et al. [715] developed N,S,P-tri-doped graphenes from N,S-co-doped graphene, prepared by hydrothermal process of GO with 2-aminothiazole as the sulfur and nitrogen source, soaked (in phosphoric acid), and freeze-dried before final carbonization under N2 (Figure 50).
Fan et al. showed that multiple heteroatom species doping is beneficial for improving the ORR catalytic properties (Table 8); with increasing heteroatom species, the onset potentials of graphene positively move, and the electron number approaches the 4e pathway by the synergistic effect of different heteroatom species.

10. Graphene and Reduced Graphene Oxide

In this section, we deal with pristine graphene and reduced graphene oxide, in which the sp2 carbon skeleton may contain oxygen in variable amounts.
Evidence of ORR activity of graphene sheets in alkaline media was first given in 2009 by Tang et al. [788]; later on, ORR activity was also shown in acidic [789] and neutral media [790,791] (Table 10).
It is interesting to observe that in the early years, appreciable electrocatalytic activity in acidic and neutral media was achieved using graphene functionalized with covalently attached anthraquinone derivatives. The π-conjugated system and the quinone group could synergistically interact to enhance the electrocatalytic performance for O2 reduction [791,792,793] (Table 9). The highly conductive rGO provides convenient channels for charge transfer, thus resulting in more favorable electron transfer kinetics, which could also be achieved by linking directly C60 [794].
As an alternative to functionalization, a way for improving the electrochemical properties consisted of enriching graphene with edges and defects in order to perturb the electron configuration and activate the carbon π electrons for effective utilization by O2 [496,506,795,796,797,798,799,800,801,802,803,804,805,806,807]. It was demonstrated that CVD-grown high-quality graphene consisting of almost defect-free multilayer graphene on Ni substrate was rather inactive for oxygen reduction in alkaline medium compared with bare glassy carbon [808].
The role of lattice defects has been well illustrated by Zhang et al. [809] who showed via DFT that a pure graphene cluster could efficiently facilitate the ORR by introducing a point defect like a pentagon carbon ring at the edge or a line defect like a pentagon–pentagon–octagon grain line. These defects interact with the edge structure to generate active sites for ORR catalysis. The four-electron and two-electron ORR can occur on these defective graphenes simultaneously, and these sub-reactions are energetically favorable because the reaction-free energy of the sub-reactions is negative.
Experimentally, it was reported that defects can also be more effective than heteroatom-doping for the ORR [494,810]; it was also possible to carry out a large scale and suitable cost synthesis of holey graphene with controllable size and distribution of pores in the basal plane correlating the defects density, the conductivity, and the ORR performance of the holey graphene [799]. Such holey graphene (Figure 51) was successfully and controllably prepared by carbothermal reaction of immobilized CoOx with graphene in an Ar atmosphere.
The content of edge or defective carbon atom, that is the defect density, could be controlled by regulating the size and the density of holes based on different loading of Co precursor on graphene. The holey graphene with the highest defect density exhibited excellent ORR performance, with the highest ORR activity, extraordinary stability, and good methanol tolerance (Table 11).
On the other hand, because of the role of edge sites in the ORR mechanism [811] in the literature, it is well known that the importance of significantly increasing the edge sites on the basal plane is also crucial in order to overcome the ORR inertness of the basal plane [796]. This can be accomplished by reducing the size of graphene to achieve graphene nanostructures [506,795,804] mainly by mechanically reducing the dimension of graphite. Benson et al. [506] produced few layernanosized graphene with low oxygen content, with lateral dimensions smaller than a few hundred nanometers, via ionic liquid-assisted grinding of high-purity graphite and sequential centrifugation. Defect formation on the crystalline plane of graphene, or chemical reactions due to mechanochemical effects, are avoided, resulting in high-quality material with a plethora of edges. Benson et al. [506] showed that the size and thickness affect the catalytic activity of graphene nanosheets. The enhancement of the ORR performance of nanosized graphene is attributed to the abundance of edge sites accrued from the small lateral size and the efficient electron transfer between the active edge sites and the electrode.
Other graphenes with plenty of edges were obtained via, heat treatment [801], plasma [797,802] and PECVD [803]. San Roman et al. [803] produced a novel graphene-based hybrid nanomaterial, nanowire-templated out-of-plane three-dimensional fuzzy graphene, on which they exerted precise control over the size and density of out-of-plane graphene flakes and edges by varying the temperature of PECVD. In this way, they could get a tunable ORR activity as a function of graphene edge site density. The hierarchical structure allows for many exposed single-layer graphene edges that readily become oxygenated in situ under alkaline ORR conditions. Their combined experiment–theory approach suggests that edge sites are saturated by carbonyl (C=O) or hydroxyl (C-OH) groups under reaction conditions. The zigzag edge sites with high carbonyl group coverage, as well as other edge configurations with a similar local coordination environment, are predicted to enable selective, two-electron ORR.
Carbon materials possessing oxygen species are known to favor the 2e- process in ORR [677], but the 4e- process is also concomitant [791,792,796,798,799,802,803,812,813,814,815,816,817] and very recently Nagappan et al. [817] showed that a near-direct four-electron transfer reaction occurred with undoped graphene samples prepared via thermal annealing of GO. Substantial investigations have demonstrated that the 2e- ORR process is mainly determined by geometrical and electronic structures of carbon nanostructures because the change in geometrical and electronic structures affects the binding strength of the OOH intermediate and the breaking of the O-O bond [469,818,819,820].
Among the parameters directing the ORR reaction, the catalyst loading on the rotating ring-disk electrode has an important impact on activity and selectivity towards ORR in alkaline medium as evidenced by Zhang et al. [814] who investigated four representative catalysts, including graphene, exhibiting different behaviors for ORR with a 4e− pathway, 2e− pathway, or a series ‘2e− + 2e−’ pathway (Table 11).
The results confirmed that the catalyst loading influences the activity and the selectivity of the catalysts. Lower loadings favored the H2O2 electrogeneration, and therefore, for a new catalyst, it is recommended to study the loading effect in order to fully understand its electrochemical properties. The optimum loading range allows the full coverage of the electrode substrate, preventing at the same time the formation of a too-thick catalyst layer in order to maximize the catalyst utilization and to avoid the underestimation of the electrogenerated H2O2.
The presence of particular carbon oxygen groups in the catalyst is important to address the ORR activity as reported by Zhang et. al. [815]; 3D graphene materials were prepared via thermal treatment of GO, showing comparable ORR catalytic activity, better tolerance to methanol crossover effect as well as higher stability than those of commercial Pt/C. It was found, for the first time, that the C=O bonds on 3D graphene display a vital role in catalytic kinetics toward ORR, as subsequently observed also by Vasiliev et al. [624]; the C=O bonds might influence the adsorption type between the oxygen molecule and the catalytically active sites on the graphene surface. The remarkable results of a 3D graphene catalyst heated at 600 °C originate from the synergetic effect between broken C=O bonds on 3D graphene and the unique porous structure of 3D graphene. Actually, the unique porous structure of 3D graphene materials is able to trap oxygen molecules, contributing to a decrease in diffusion resistance while enhancing electrolyte-electrode accessibility for fast mass transport.
Furthermore the adsorption of gaseous oxygen molecules at the liquid–solid phase boundary and the formation of a high-density gas–liquid–solid triple-phase boundary is important for accelerating the ORR as observed in aluminum–air batteries constructed using cathode based on sparked reduced graphene oxide possessing hierarchical pores [821] (Figure 52); moreover, hydrophilically treated graphene was found to be a much better material to enhance the energy production and performance of the dual-chamber MFC system [822].

11. Graphene Composites

As an alternative to heteroatom doping, another way to achieve active ORR electrocatalysts where the basal plane of graphene is preserved is the formation of a composite by combining graphene with carbons (doped and undoped), silicon, boron nitride, carbon nanotubes, fibers, carbon dots, electroactive molecules, and polymers.
In the literature, most composites contain graphene and carbons, and apart from a few articles dealing with composites with activated carbons [823,824], silicon nanosheets [825], coaxial cable-like electrocatalyst based on carbon fibers [826], nanodiamonds [827], fullerene [828], and with phosphorus or boron single-doped carbons [829,830], carbons always contain nitrogen; it is well known that nitrogen doping within a graphitic/turbostratic network of carbon atoms generates active sites for the ORR via C-N bond polarization, which induces a reduced energy barrier towards the ORR on the adjacent carbon atom [831]. About the chemical nature of such nitrogen-doped species, they are mostly reported as generic nitride or simply N-doped carbons, with more attention usually paid to the porosity requirements of the material [831,832,833,834,835,836,837,838,839,840,841,842,843,844,845,846,847,848].
However, excluding the research by Kim et al. [849], who developed specific N-rich mesoporous carbon nitride with C3N5 stoichiometric and studied its graphene hybrids, in the case of graphitic carbon nitride, g-C3N4, we observe a more defined study. Graphitic carbon nitride is a promising catalyst for ORR. It is easily accessible but has a low specific surface area and low conductivity, which strongly restricts the electron transportation during the ORR process and also reduces its electrocatalytic activity. Therefore, the properties most directly related to ORR performance are thought to be manipulated by modulating the electronic structures of g-C3N4 with 2D materials [850]. There has been some effort to immobilize it onto graphene [851,852,853,854,855,856,857,858,859], which has a similar structure and relevant conductivity, since the first composite was reported by Sun et al. in 2010 [851]. Wang et al. [856] prepared monoatomic thick g-C3N4 dots (MTCs) by oxidative exfoliation of bulk g-C3N4 and put in an intimate contact to the basal plane of the graphene sheet to form the monolayer g-C3N4 dots on graphene (MTCG) through hydrothermal self-assembly (Scheme 7).
The face-to-face contact between g-C3N4 and graphene is believed to induce a synergistic coupling interaction in facilitating the ORR and electron transport through the composite with an electrocatalytic activity rivaling that of the commercial Pt/C catalyst (Table 12). Theoretical calculations and RRDE measurements indicate that the composite displayed a higher efficiency for the reduction of OOH− than the g-C3N4 alone, resulting in an enhanced performance in the ORR with an efficient four-electron pathway.
Very recently, Mane et al. [858] developed a rapid, straightforward and cost-effective synthesis technique for preparing a composite of g-C3N4 and few-layered graphene sheets; solar radiation was used to reduce GO, which forms its composite with g-C3N4 in a highly crumpled, porous, and wrinkled 2D/2D morphology with enriched Lewis base sites as active centers. This material catalyzes the ORR activity through a four-electron process, demonstrating high activity, stability, and durability, even in a methanol atmosphere. The superior performance of the catalyst is attributed to its structural characteristics and the relationship between these features and electrocatalytic activity, particularly Lewis bases as active centers.
A few more studies have been published on attempts to utilize single and multi-doped carbon dots in ORR electrocatalysis by constructing composites with graphene [623,829,860,861,862,863].
Anchoring onto rGO sheets N-doped Cdot, S-doped Cdot, and N,S-co-doped Cdot prepared through similar hydrothermal reactions [864], Zhang et al. [623] formed composites which outperformed their rGO counterparts in terms of electrochemical properties in ORR. N,S-co-doping is a more effective strategy for realizing a four-electron transfer pathway in ORR due to the synergistic effects of N,S-co-doping. As a result, optimal N,S-CDot/rGO samples exhibit the most positive half-wave potential, the highest kinetic current density, and the largest electron transfer number. With respect to substrate-constructing, active sites on Cdot surfaces exhibit better ORR activities than those in rGO planes, because the former are located in abundant edges/defects and have more access to oxygen molecules.
There are a few more examples of multi-doped carbon/graphene composites where the beneficial synergistic effect on ORR electrocatalytic performance is evident [834,865,866,867,868,869,870]. Tan et al. [870] reported 2D sandwich-like mesoporous phosphorus- and nitrogen-doped carbon nanosheets (rGO@PN/C) as a metal-free electrocatalyst for ORR (Table 12) by using graphene oxide as the substrate, triblock copolymer F127 micelles as the soft template, and resin and phytic acid as the organic precursor (Scheme 8).
The carbonization of the pre-prepared GO/F127/resin (from m-aminophenol and formaldehyde) precursor creates a 2D sandwich-structured nanoarchitecture composed of an inner layer of reduced graphene oxide and an outer layer of mesoporous nitrogen-doped carbon (rGO@N/C). The products consist of abundant mesopores formed by the assembly of F127/resin micelles, as well as the removal of F127 micelles. After activation of rGO@N/C with phytic acid (PA), phosphorus- and nitrogen-doped carbon (rGO@PN/C) nanosheets with well-maintained bimodal pores are obtained. PA can activate the carbon matrix during heat treatment under high temperatures, leading to increased porosity in carbon materials. Moreover, the incorporation of P determined considerable improvement in catalytic activity and mass transfer for the ORR. The introduction of P generates new active sites for the ORR, and the increased micro- and mesoporosity via activation with PA further facilitates mass transfer during the ORR.
In addition, the rGO@PN/C nanosheets show better durability and high selectivity compared to the commercial Pt/C catalyst.
There has been significant interest in developing 2D sandwich-like carbon nanostructures embedding graphene with ORR electrocatalytic activity [831,832,834,838,841], since the first one was reported and synthesized by Yang et al. [852] via a nanocasting method.
Liu et al. [838], following a rGO-templated 4,40-dicyanobiphenyl self-polymerization process, were able to construct a sandwich-like pyridinic N-enriched covalent triazine-based framework/rGO hybrid, which subsequently affords a sheet-like N-doped and hierarchically porous carbon/rGO composite obtained at 950 °C with dominant pyridinic and graphitic N incorporation, as a highly efficient ORR catalyst (Table 12). Thanks to the synergistic effect between the N-doped carbons and 2D rGO in terms of high surface area, hierarchical pores, controllable incorporation of active N species, favorable mass transfer, and high conductivity, the layered composite exhibits superb ORR activity and ultrahigh stability as well as excellent methanol tolerance.
It is interesting to observe that the composites of graphene with doped carbon are synthesized at high temperatures, resulting in high energy consumption, and several syntheses start from graphene polymer composites. On the other hand, composites of graphene with suitable polymers can themselves be electrocatalysts for ORR, and they can also be synthesized under mild conditions.
In the literature, composites are formed with poly(diallyldimethylammonium chloride) (PDDA) [871,872,873], polyethyleneimine (PEI) [874], amino-substituted sulfonic acid [875], polydopamine [876], adenine [877], but most with conductive polymers such as poly(3,4-ethylenedioxythiophene) (PEDOT) [878] and linear conjugated polymers based on covalently linked pyridine and thiophene [879], Nile blue [880], and PANI [881,882].
It is interesting to underline that the heteroatoms are not in the graphene carbon lattice, but in the polymers.
Barrios Cossio [881] et al. reported unique additive-free bifunctional electrodes composed of graphene oxide, reduced graphene oxide, PANI, and poly(vinyl alcohol) (PVA) on carbon cloths, which were fabricated using an easy and innovative in situ aniline polymerization strategy under mild conditions. The electrostatic interactions at GO/PANI interfaces controlled both the PANI distribution pattern and the electrical conductivity of the resulting nanohybrids, while PVA increased the ion accessibility. The GO/PANI/PVA nanocomposites showed improved conductivity properties, as well as rather uniformly distributed PANI networks, which have been directly associated with the unique synergistic interactions at the PANI/GO interfaces between the N-rich polymeric structure and the functional groups at the GO surfaces. The resulting electrodes were characterized by pH-universal oxygen reduction reaction properties, with an outstanding ORR onset potential value of 0.93 V vs. reversible hydrogen electrode in basic media. Additionally, the nanocomposite showed remarkable stability in acid media for the electroreduction of oxygen, maintaining 98% of the initial current applied after 25,000 s.
The use of polymers to impart better electrochemical properties has also been exploited by Lee et al. [873], who combined PDDA-functionalized hydrophobic multiwalled nanotubes (pMWNTs) with hydrophilic rGO nanosheets having a large surface area for the construction of metal-free carbon electrocatalysts to allow effective oxygen access in both gas and aqueous forms. They obtained 3D interconnected network structures of pMWNTs with interdispersed rGO nanosheets, driven from electrostatic stacking between oppositely charged carbon nanomaterials, which exhibited a remarkably enhanced ORR activity without extreme conditions for synthesis.
While pMWNTs provided the majority of the catalytically active sites due to their hydrophobic nature and the induced partial positive charge on their surface, introduction of rGO into their hybrid materials also improves the utility of both dissolved and gaseous oxygen and facilitates electron transfer during ORR due to its hydrophilic nature.
In this unique architecture, both two- and three-phase reactions in ORR can be maximized with a quasi-four-electron pathway.
Although composites of CnT with graphene have already been illustrated in the previous paragraphs, a few are missing [883,884,885,886,887].
In particular, Cai et al. [884] prepared free-standing, N-doped vertically aligned carbon nanotubes (N-VA-CNTs) supported by graphene foam (GF), whose well-aligned structure not only works as a highly efficient electrocatalyst but also provides appropriate channels/pores for intermediate ions and gas diffusion. Due to their highly efficient catalytic activity and structural merits such as being lightweight and flexible, and possessing high conductivity, N-doped VACNTs/GF as integrated air-cathodes show significantly improved performances and current- and power-densities. A rechargeable Zn–air battery assembled from this N-VA-CNTs/GF hybrid electrode exhibits much higher power density than that of commercial Pt/C electrocatalysts and long-term durability.
To complete the overview, composites of graphene with boron nitride [888] and electroactive molecules such as antraquinone [889], tetracyanoethylene [890], and melanin [891] showed appreciable activity as ORR catalysts. In particular, Wu et al. [891] developed melanin-modified carboxylated graphene as a novel cathode catalyst in a microbial fuel cell with significantly increased bioelectricity production; this research showed the promising application of melanin in enhancing the ORR kinetics and improving microbial electron transfer in bio-electrochemical systems with interesting prospects for future sustainable and environmentally friendly applications.

12. Conclusions

This review offers a very wide overview of the research in the field of the ORR metal-free graphene-based ECs, paying particular attention to the synthetic strategy to improve the specific properties, determining the electroactivity of a species, which is one of the most crucial factors to be considered to fully utilize the structure of graphene and enhance its application value. The synthesis should be developed through an easily scalable protocol to be compatible with large-scale industrial production, and even in greener and more sustainable ways to meet ecological requirements.
Most of both theoretical and experimental studies on ORR metal-free graphene-based ECs (above 600 papers) which show how the electrocatalytic properties and the ORR performance of graphene-based catalysts strongly depend on their composition, including the type and number of doping elements, the materials’ structure and morphology, such as size and shape, as well as electronic effects, have been reported; this should help the scientist in addressing the research towards the design of a suitable electrocatalyst even if its activity cannot be easily predicted.
This review indicates that the first strategy, which can be suggested for a good electrocatalyst, is the introduction of heteroatoms into the carbon structure to bring about modifications to both its chemical and physical properties.
Nitrogen-doping has been largely investigated, resulting in active ECs; however, there are other dopants with suitable physicochemical properties, such as B, P, and S, that may be included in the carbon core. Future development would involve investigating the usage of other heteroatoms, such as selenium, which proved to be promising but was not developed further. It may also be beneficial to start from natural products that contain it, acting as green, single-source precursor. It can also be considered that the creation of desirable active motifs with distinctive coordination structures and electronic states that are beneficial for improved electrochemical performance could be guided by computational explorations such as artificial-intelligence-assisted material screening or DFT predictions with machine learning.
Moreover, it is still a challenge to control the number of heteroatoms, their distributional homogeneity, bonding forms, and other aspects, which are particularly evident in co-doped species. The position and number of heteroatoms in the carbon basal plane are still difficult to accurately adjust as well as the establishment of a clear connection between both the doped structure and catalytic performance deserving further experimental studies, state-of-the-art characterizations, and robust computational simulations to a deeper understanding of the structure, mechanism, and thermodynamics of the catalytic core.
In particular, multi-doping, creating sites simultaneously carrying positive charge and high spin density, is also considered a promising way to achieve the goal of enhancing ORR activity in acidic electrolytes in order to potentially or effectively substitute, in particular, platinum. The metal-free graphene ORR ECs offer the natural advantages of durability in acid due to the absence of metal leaching, metal-ion contamination, and Fenton reaction–related degradation but unfortunately, the best ORR performances are generally obtained under alkaline conditions, even if with appreciable durability.
Beyond the selection of elements, the catalytic performance, which is definitely related to the intrinsic activity, enrichment degree, and accessibility of the active sites, can be improved by the design of graphene structures via morphological and vacancy engineering of catalysts and interface effects.
Catalysts with 3D structures, such as hollow structures, core-shell structures, etc., usually have high mass transfer efficiency. In particular, the reaction kinetics of ORR can be effectively promoted in 3D interconnected layered porous structures and high specific surface areas: additional active sites can be provided and facilitate the adsorption and desorption of intermediates because of efficient charge/material transport. Therefore, porosity is very important, even if what better defines the catalytic activity is the electrochemically active surface area.
Considering vacancy engineering of catalysts and interface effects, it is important to control the vacancy defects and edges, which can induce charge redistribution, change the local electronic structure, but without shattering the conductivity, and effectively improve intrinsic catalytic activity. Interface engineering is also one of the effective strategies for improving catalytic activity. The interface formed between the components in the electrocatalyst can effectively change the electronic structure and atomic configuration of the active site.
It has to be highlighted that most of the existing catalysts were tested in the laboratory, and, therefore, it is necessary to convert laboratory techniques into practical results to achieve high efficiency and stability in practical environments. A lot of effort must therefore be made to fill the gap between “simple” RDE or RRDE measurements and, for example, tests in real MEA.
To conclude, the author must warn that what are largely claimed to be metal-free electrocatalysts often lack proper, substantial elemental analysis. This can be achieved, for example, via inductively coupled plasma, which should evidence the absence of metals; metals derived, for instance, from starting reagents, even in trace amounts, can act as effective ORR electrocatalytic species. The presence of metal impurities is a current issue in catalysis; however, this observation does not negate that graphene is a promising material for ORR electrocatalysis.
Table 2. Comparative analysis of some ORR parameters of the synthesized N-doped graphene-based ECs 1 with best performance in alkaline electrolytes.
Table 2. Comparative analysis of some ORR parameters of the synthesized N-doped graphene-based ECs 1 with best performance in alkaline electrolytes.
Catalyst 2 Catalyst Loading 3 (μgcm−2)Eonset
vs. RHE 4 (V)
Electrolyte
pH
Synthetic Method 5Ref.
N-GQDs_3
N-doped graphene quantum dots
NGqDor283.080.61013electrochemical method[316]
N-rGO
N-doped reduced graphene oxide
N-rGO121.430.62013annealing[561]
N/3D-GNS 850
three-dimensional N-doped graphene nanosheets
NG3DnSh 850
NG3DnSh 850
200
200
0.630
0.870
13
1 6
thermal heating[400]
TDMAC-RGO
Tridodecylmethylammonium chloride (TDMAC)-functionalized reduced graphene oxide
TDMAC-RGO
TDMAC = Tridodecylmethylammonium chloride
50.960.65013solution process[428]
Nr-GO4
N-doped reduced graphene oxide
N-rGO (III)250.65013annealing[516]
GF
N-doped 3D graphene framework
NG3Dχ120.67013hydrothermal process[521]
NG60
nitrogen-doped graphene
NG (II)1187.310.68013sonochemical process[436]
NB-graphene
Nitrobenzene-doped graphene
Nitrobenzene-G151.650.70013solution process[427]
1F-800
N-doped graphene foam
NGFo 8004000.70013carbonization[442]
N-rGO (E)
N-doped reduced graphene oxide
N-rGO (II)150.71013hydrothermal process[570]
NG−1
N-doped graphene
NG(I)400.71013pyrolysis[421]
NGQDs/G
N-doped graphene quantum dots supported by graphene
NGqDot/G283.080.71013hydrothermal process[283]
run 7
N-doped reduced graphene oxide
N-rGO (II)10000.71013microwave irradiation[417]
N-dGOA
N-doped graphene oxide aerogel (GOA)
NGOAe27.750.72013solution process[529]
NG/CCN light
N-doped graphene/carbon-rich C3N4 composite
NG/C3N45000.73013calcination[439]
PyNG-3
Pyridine-nitrogen-doped graphene sheets
Pyridine-Gsh (II)141.540.73813solution process[358]
N-GQDs/G-12
N-doped graphene quantum dots/graphene hybrid
NGqDot/G (I)70.770.74013hydrothermal process[319]
N-Gr
nitrogen-doped graphene
NG
NG
NG
81.53
81.53
81.53
0.740
0.440
0.210
13
7 6
0.3 6
electrochemical process[558]
N-FLG
Nm-FLG8 nitrogen-rich
few-layered graphene
NGfl (III)101.210.74013annealing[418]
NGSH
Nitrogen-doped graphene/carbon nanotube hybrids
NCswnT/NG254.770.74013CVD[541]
NG
N-doped graphene
NG1410.74413pyrolysis[381]
NHG
Nitrogen-doped holey graphene
NGhl250.75013hydrothermal process[490]
N-RGO3
N-doped reduced graphene oxide
N-rGO (III)70.770.75013annealing[310]
G N-max
N-doped graphene
NG (I)1500.75013thermal treatment[473]
HPN-MQGs
high-power N-doped medium-quality graphene sheets
NGSh (III)2000.75013microwave plasma torch [318]
HT-N-RGO
N-doped reduced graphene oxide
N-rGO6000.75013hydrothermal process[472]
N8,3-NGR
Nitrogen-doped graphene nanoribbon
NGnR (IV)141.540.76513pyrolysis[456]
NG
N-doped graphene
NG142.860.76513thermal annealing[380]
HTNG/GCE
N-doped graphene
NG (II)79.620.76913thermal annealing[383]
N-rGO-90
N-doped reduced graphene oxide
N-rGO 9020000.77013solution process[443]
N-G 900
N-doped graphene
NG1166.860.77813annealing[462]
NG10
N-doped graphene
NG20.300.78013thermal annealing[437]
NGnP
N-doped graphene nanoplatelets
NGnPl76.430.78013ball milling[505]
NG-2/GC
N-doped graphene
NG (II)137.620.78013hydrothermal process[504]
N-RGO10
3-D mesoporous N-doped reduced graphene oxide
N-rGO3Dmpo (II)1600.78013annealing[317]
N/G1050
N-doped graphene
NG 10501280.78513annealing[323]
rGO-sp3-rGO
Diaminobutano-rGO (II)51.020.79013solution process[425]
NG-PPy
N-doped graphene
NGpoL5000.79013carbonization[407]
NG
nitrogen-doped graphene
NG2430.79513hydrothermal method[457]
N-rGO900
N-doped reduced graphene oxide
N-rGO141.540.79813thermal annealing[557]
NG-NCNT
Nitrogen-doped graphene/carbon nanotube nanocomposite
NCnT/NG50.960.79913hydrothermal process[538]
NG 900
N-doped graphene
NG 9001410.80013pyrolysis[384]
N-doped MSMG-P
N-doped mechanochemically synthesized multilayer graphene
NGmL15300.80013mechanochemical synthesis[502]
rGO N2H4
N-doped reduced graphene oxide
N-rGO4240.80013hydrothermal process[120]
NMGF
nitrogen-doped mesoporous graphene framework
NGmpoχ255.100.80113CVD[497]
N-GP 4-hr 600 °C
N-doped graphene nanoplatelets
NGnPl 600 (II)4240.80513pyrolysis[446]
N-G
N-doped graphene
NG2000.80813annealing[128]
G-NH3
H2O N-doped graphene
NG (II)204.080.81013annealing[313]
GO/NH3·H2O
oxo-G-derived nitrogen-doped graphene
N-rGO (III)1000.81013hydrothermal treatment[499]
PB-N-rGO
1-pyrenebutyrate functionalized N-doped graphene
1-pyrenebutyrate-rGO1700.81013solution process[320]
 
N5-rGO
N-doped reduced graphene oxide
N-rGO5100.81013hydrothermal process[476]
A-rGO
ammonia-reduced graphene oxide
rGON597.130.81013Solution process[568]
N-pGF
pyridinic-N-doped graphene film
NGFm (II)
NGFm (II)
50
50
0.815
0.211
13
0.3 6
solution process[333]
NG-750
N-doped graphene
NG 75061.220.81513thermal treatment[392]
NG 1
nitrogen-doped reduced graphene oxides
N-rGO (I)1030.81513annealing[395]
NGR−1000
nitrogen-doped graphene
NG 10001400.81513pyrolysis[445]
IRnG-A2
A2 imine-rich nitrogen-doped graphene nanosheets
[(CF3SO2)C6H4N)]-G2000.81513solution process[429]
Graphene:glycine 1:4
N-doped graphene
NG (II)509.550.81513pyrolysis[388]
NHG
nitrogen-doped holey graphene
NGhl100.81913ball milling[491]
T 1000
N-doped graphene-wrapped carbon nanoparticles
NG 1000/NCnP152.870.81913solvothermal process[459]
Lem-rGO (_8)
N-functionalized rGO
N-rGO (I)51.020.82013solution process[533]
NG 1
N-doped graphene
NG1030.82013thermal annealing[398]
N-rGO 800
N-doped reduced graphene oxide
N-rGO (IV)203.110.82013annealing[328]
N-rGO-180
N-doped reduced graphene oxide
N-rGO (IV)492.960.82013microwave treatment[335]
GD1
N-doped graphene
NG (I)509.550.82013solution process[487]
CS@N-G/CNT
carbon spheres@nitrogen-doped graphene/carbon nanotubes hybrid
CSp/NG/CnT5700.82513ultrasonic-assisted process[545]
NG 900
N-doped graphene
NG 9002000.82813calcination[338]
N-RGO
N-doped graphene
N-rGO1900.82913annealing[489]
1000-O2
N-doped graphene
NG (II)1200.83013annealing[494]
NG
N-doped graphene
NG1940.83013microwave plasma[438]
GC900
three-dimensional N-doped graphene
NG3D 9002000.83013pyrolysis[405]
N-rGO–CNT-0.2
N-doped rGO–carbon nanotube composites
N-rGO/NCnT (II)6000.83013annealing[542]
NG
N-doped graphene
NG39.810.83513annealing[127]
N-graphene
N-doped reduced graphene oxide
N-rGO1600.83513shock synthesis of graphene material from CO2[559]
NCDs-NG−12
N-doped reduced graphene oxide
NCDot/NG (II)79.220.84013hydrothermal process[326]
3-NG
N-doped graphene
NG (III)1000.84013pyrolysis[311]
N2-3DrGO
N-doped three-dimensional reduced graphene oxide
N-rGO3D (II)101.910.84013calcination[464]
N-OMMC-G
N-doped graphene natively grown on
hierarchical ordered porous carbon
NG/CNmamop416.670.84013heating and silica etching[522]
N-rGO 800
N-doped reduced graphene oxide
N-rGO 80011000.84013pyrolysis[564]
NDG
N-doped graphene
NG510.84513annealing[435]
CN1000
N-doped graphene
NG 1000
NG 1000
144
144
0.847
0.567
14
7 6
thermal treatment[390]
NDTG
N-doped double-layer templated graphene
NG (I)254.80.84913CVD[534]
NG−1000
nitrogen-doped graphene
NG 1000 (I)97.990.85013annealing[386]
CHs material
N-doped graphene
NGpoL106.160.85013carbonization[449]
GN-CNT 2
N-rich graphene nanoclusters–carbon nanotube composite
NCnT/NGnCl106.160.85013deflagration with NaNO3[477]
PM-NGr/NCNT
N-doped graphene/carbon nanotube
NG/NCnT (I)106.160.85013pyrolysis[548]
NGM
N-doped graphene mesh
NGmh
NGmh
254.78
254.78
0.850
0.540
13
1 6
carbonization[496]
DNGS 480 900
dendritic N-doped graphene spheres
NGddSp (I)
NGddSp (I)
NGddSp (I)
1000
1000
1000
0.770
0.550
0.470
13
0.3 6
7 6
pyrolysis[532]
NG-SCCf
N-doped graphene grown on carbon fibers derived from silk cocoon (SCCf)
NG/NCFb3200.85013thermal treatment[524]
G800
N-doped few-layered graphene sheets
NGfLSh4000.85013carbonization[451]
NVG-30
N-doped vertical graphene nanosheets
NGvSh4700.85013Plasma-enhanced CVD[351]
HNG-900
holey
N-doped graphene
NGhl 9002000.86013thermal heating[399]
NGF
N-doped graphene framework
NGχ2000.86013thermal heating[402]
NGE 1000
nitrogen-doped graphene
NG 10002250.86013heat treatment [416]
N-Gr
N-doped graphene
NG4200.86013thermal treatment[408]
NG-30
N-doped graphene
NG (IV)4850.86813microwave heating[389]
NG/NCNT-BR
N-doped graphene from biuret
NG/NCnT (III)1000.86913pyrolysis[321]
N-MG-800
3D bicontinuous N-doped mesoporous graphene
NG3Dmpo 8001000.87013thermal heating[336]
NGA−150
N-doped graphene aerogels
NGAe (III)1020.87013pyrolysis[469]
N-graphene 900
N-doped graphene
NG (I)1600.87013thermal treatment[378]
ENG
N-doped graphene
NG (II)2000.87013electrochemical exfoliation[515]
CN 900
N-doped graphene
NG 9002800.87013carbonization[556]
N-GNRs-A
N-doped graphene nanoribbons aerogel
NGhl 900800.87513pyrolysis[527]
DG
N-doped graphene annealed at 1150 °C
NG (II)
NG(II)
79.62
79.62
0.880
0.530
13
1 6
annealing[810]
N-GNR
nitrogen-doped graphene nanoribbon
NGnR175.070.88013solution process[325]
N-RGO-PPV(c)-CNTs
three-dimensional N-doped carbon nanotube/reduced graphene oxide composite
NCnT/N-rGO300.570.88013pyrolysis[546]
N-rGO Diethyl ether
N-doped reduced graphene oxide
N-rGO (III)6000.88013annealing [498]
N-GRW
three-dimensional (3D) graphene nanoribbon networks
NG3DnRχ6000.88014pyrolysis[396]
N-aGS-900
N-doped activated graphene/single-wall carbon nanotube hybrid
NG/NCsWnT1770.88513pyrolysis/annealing[547]
NG-C
N-doped graphene
NG (III)2000.88513ultrasound[315]
ENR-GNPs
edge-nitrogen-rich graphene nanoplatelets
NGedrnPl
NGedrnPl
283
283
0.885
0.584
13
0.3 6
ball milling[387]
NHGs
nitrogen-doped hollow graphene microspheres
NGhwmSp100.89013pyrolysis[525]
NMG−1/4
nitrogen-doped mesoporous graphene
NGmpo141.540.89013annealing[526]
N-rGO/PAA-900
N-doped reduced graphene oxide
N-rGO394.890.89013pyrolysis[484]
NOGB-800
N,O-co-doped graphene nanorings-integrated boxes
ONGnBx 800
ONGnBx 800
400
400
0.890
0.710
13
1 6
calcination[342]
NGA
N-doped three-dimensional porous graphene frameworks
NG3Dpoχ101.910.89813calcination[470]
PHNG-800
porous holey nitrogen-doped graphene
NGpohl 800254.780.90013pyrolysis[465]
NG-2
N-doped graphene
NGpoL606.610.90714solvothermal process[205]
NG1000
N-doped graphene
NG 100038.220.90813thermal treatment[454]
N-GQD/rGO
nitrogen-doped graphene quantum dots anchored on N-doped graphene
NGqDot/rGO1000.91013calcination[410]
NG 1000
N-doped graphene
NG 1000
NG 1000
152
152
0.910
0.750
13
1 6
pyrolysis[495]
O,N-graphene
O,N-co-doped 3D graphene hollow sphere
ONGhSp2500.91013annealing[479]
DN-UGNR 3nm 900
N-doped ultranarrow graphene nanoribbons
NGunwnR (I)3000.91013thermal treatment[341]
N-GN
N-doped graphene
NG4000.91013thermal process[433]
GHN-C 900
graphene hydrogel-based nitrogen carbon materials
NC/NG 900
NC/NG 900
NC/NG 1000
450
450
250
0.910
0.640
0.620
13
1 6
13
pyrolysis[440]
N-GPp(few layer)
N-doped graphene (few layers)
NG (I)8000.91013electrochemical exfoliation[343]
Twostep
N-doped graphene
NG (I)19000.91013hydrothermal process[401]
NG 1000
nitrogen-doped graphene
NG 10002830.91513pyrolysis [240]
N-rGO-P
N-doped reduced graphene oxide
N-rGO (II)1000.92013pyrolysis[406]
NGS_1000
porous N-doped graphene layers
NGpoL
NGpoL
204.08
204.08
0.920
0.680
13
0.3 6
pyrolysis[404]
N-hG6
holey N-doped graphene
NGhl2500.92013annealing[352]
NGA950
N-doped graphene after annealing at 950 °C
NG 950283.090.92013pyrolysis[458]
3D-PNG
three-dimensional nanoporous nitrogen-doped graphene
NG3Dnpo4850.92013pyrolysis and etching[461]
NGA
N-doped graphene aerogel
NGAe101.910.93013annealing[472]
NG 800
N-doped graphene
NG 8001800.93013CVD[304]
N-GNRs/G
N-doped graphene nanoribbons
NGnR/NG2550.93013pyrolysis[482]
Ai-HGs
3 3,4-diaminopyridine grafted N-doped holey graphene
3,4diaminopyridine-Ghl305.730.93013solution chemistry[349]
g-C3N4@N-G
Graphitic carbon nitrides supported by N-doped graphene
g-C3N4/NG2000.94013carbonization[514]
NDGs-800
N-dominated doped defective graphene
NGdf2040.94013thermal heating[339]
NGNs-900
N-doped edge-rich graphene nanosphere
NGedrSp254.780.94013pyrolysis[483]
N-G 1000
N-doped graphene
NG (I)
NG (I)
485
485
0.940
0.800
13
1 6
pyrolysis[403]
CrG-900
N-doped crumpled graphene
NGcr 9006000.94013solvothermal process[397]
N-SMCTs@N-rGO
N-doped sub-micron carbon tubes with N-doped
reduced graphene oxide
NCsμT/N-rGO2040.95013carbonization[481]
N-GNR@CNT
N-doped graphene nanoribbons on carbon nanotubes
NGnR/NCnT
NGnR/NCnT
398
398
0.950
0.700
13
0.3 6
annealing[549]
HPGF−1
hierarchical porous N-doped graphene foams
NGhipo-Fo (I)
NGhipo-Fo (I)
485
485
0.950
0.780
13
1 6
calcination and silica etching[460]
N-rGO
N-doped reduced graphene oxide
N-rGO1700.95214annealing[176]
NA-3DGFs 600
N-doped densely arranged sharp edges graphene fibers
NG3DFb (II)1500.96013thermal treatment[348]
NSAOrGO
N-doped macroporous carbon materials
NSC/rGON (I)159.240.96013pyrolysis[478]
NGS4 1000
nitrogen-doped graphene nanosheets
Nnsha 10004070.96013annealing[350]
N-GN15
N-doped graphene
NG (I)
NG (I)
102
102
0.980
0.850
13
1 6
annealing[393]
NPG 1-05
N-doped graphene nanoporous graphene
NGpo (I)3500.98013pyrolysis[330]
NHGNs
nitrogen-doped graphene sheets
NGhlnCs394.150.9813hydrothermal process[448]
NGTB-900
high-density pyridinic-N-doped graphene−
nanotube complexes with hierarchical networks
NCnT/NG 900
NCnT/NG 900
600
600
1.01
0.780
13
16
annealing[550]
1 For proper comparison, this table contains only data from samples whose catalyst loading can be defined. 2 The catalyst name in italic is the catalyst name reported in the paper with the best Eonset, reported in the third column and below the explanation reported in the paper; in bold, the name is set using the terms reported in the legend following mainly the indications reported in [892]. 3 Catalyst loading, when not reported in the paper, has been calculated from data provided by the authors. 4 Eonset was generally extracted from the LSV curves analyzed using the WebPlot program. 5 The synthesis method is not fully described but usually refers to the last treatment. 6 For a direct comparison, values at different pH have also been reported.
Table 3. Comparative analysis of some ORR parameters of the synthesized N-doped graphene-based ECs 1 with best performance in acidic electrolytes.
Table 3. Comparative analysis of some ORR parameters of the synthesized N-doped graphene-based ECs 1 with best performance in acidic electrolytes.
Catalyst 2 Catalyst Loading 3 (μgcm−2)Eonset
vs. RHE 4 (V)
Electrolyte
pH
Synthetic Method 5Ref.
N-Gr
N-doped graphene
NG
NG
NG
81.53
81.53
81.53
0.210
0.740
0.440
0.3
13 6
7 6
electrochemical process[558]
N-pGF
pyridinic-N-doped graphene film
NGFm (II)50
50
0.211
0.815
0.3
13 6
solution process[333]
N-GNS 800
N-doped graphene
NGnSh2000.3100.3pyrolysis[434]
N-GrapheneNG1000.4201annealing[334]
NG
N-doped graphene
NG1060.4700.3annealing[345]
DG
N-doped graphene
NG (II)
NG (II)
79.62
79.62
0.530
0.880
1
13 6
annealing[810]
NGM
N-doped graphene mesh
NGmh
NGmh
254.78
254.78
0.540
0.850
1
13 6
carbonization[496]
DNGS 480 900
dendritic N-doped graphene spheres
NGddSp/Csp 900 (I)
NGddSp/Csp 900 (I)
NGddSp/Csp 900 (I)
1000
1000
1000
0.550
0.770
0.470
0.3
13 6
7 6
pyrolysis[532]
nitrogen-doped graphene foamNGFo5800.5701solvothermal process[492]
ENR-GNPs
edge-nitrogen-rich graphene nanoplatelets
NGedrnPl
NGedrnPl
283
283
0.584
0.885
0.3
13 6
ball milling[387]
Nr-GO4
N-doped reduced graphene oxide
N-rGO (III)
N-rGO (III)
25
25
0.580
0.660
5
1
annealing
annealing
[516]
N-MG-801
3D bicontinuous N-doped mesoporous graphene
NG3Dmpo 8004000.5900.3thermal heating[336]
N-HPC/RGO−1
N-doped hierarchical porous carbon reduced graphene oxide
NChipo/NrGO101.910.6250.3pyrolysis[463]
GHN-C 900
graphene hydrogel-based nitrogen/carbon materials
NC/NG 900
NC/NG 900
450
450
250
0.640
0.910
0.620
1
13 6
13 6
pyrolysis[440]
N-Gr
N-doped graphene
NG707.710.6701pyrolysis[640]
HS-MW 1200
nanoporous N-doped reduced graphene oxide
NrGOnpo (V)3250.6801annealing[503]
N-GNR@CNT
N-doped graphene nanoribbons on carbon nanotubes
NGnR/NCnT
NGnR/NCnT
398
398
0.700
0.950
0.3
13 6
annealing[549]
NOGB-800
N,O-co-doped graphene nanorings-integrated boxes
ONGnBx 800
ONGnBx 800
400
400
0.710
0.890
1
13 6
calcination[342]
GN 1000
N-doped graphene
NG 1000
NG 1000
152
152
0.750
0.910
1
13 6
pyrolysis[495]
GCA 4
N-doped graphene–carbon nanotube self-assembly
NCnT/NG (IV)707.710.7601thermal heating[540]
HPGF−1
hierarchical porous N-doped graphene foams
NGhipoFo (I)
NGhipoFo (I)
485
485
0.780
0.950
1
13 6
calcination and silica etching[460]
NGTB-900
high-density pyridinic-N-doped graphene−
nanotube complexes with hierarchical networks
NCnT/NG 900
NCnT/NG 900
600
600
0.780
1.010
1
13 6
annealing[550]
PNGF
porous N-doped graphene foam
NGpoFo4850.7901pyrolysis and etching[518]
N-G 1000
N-doped graphene
NG (I)
NG (I)
485
485
0.800
0.940
1
13 6
pyrolysis[403]
NGCA
N-doped graphene/CNT self-assembly
NG/CNmamop7140.8101pyrolysis[539]
CNG-3
3D interconnected carbon nitride (CNx) tetrapods wrapped with nitrogen-doped graphene
NG/CNxtPod (II)2550.8201annealing[432]
NG@MMT
N-doped graphene
NG6000.8301pyrolysis[452]
N-GN15
N-doped graphene
NG (I)
NG (I)
102
102
0.850
0.980
1
13 6
annealing [393]
N/3D-GNS 850
three-dimensional N-doped graphene nanosheets
NG3DnSh 850
NG3DnSh 850
200
200
0.630
0.870
0.3
13 6
thermal heating[400]
NGS_1000
porous N-doped graphene layers
NGpoL204.08
204.08
0.680
0.920
0.3
13 6
pyrolysis[404]
1 For a proper comparison, this table contains only data from samples whose catalyst loading can be defined. 2 The catalyst name in italic is the catalyst name reported in the paper with the best Eonset, reported in the third column and below the explanation reported in the paper; in bold is the name set using the terms reported in the legend, following mostly the indications reported in [892]. 3 Catalyst loading, when not reported in the paper, has been calculated from data provided by the authors. 4 Eonset was generally extracted from the LSV curves analyzed via the WebPlot program. 5 The synthesis method is not fully described but it usually refers to the last treatment. 6 For a direct comparison, values at different pH have also been reported.
Table 4. Comparative analysis of ORR parameters of the synthesized S-doped graphene-based ECs 1 with best performance in alkaline electrolytes.
Table 4. Comparative analysis of ORR parameters of the synthesized S-doped graphene-based ECs 1 with best performance in alkaline electrolytes.
Catalyst 2 Catalyst Loading 3
(μgcm−2)
Eonset
vs. RHE 4 (V)
Electrolyte
pH
Synthetic Method 5Ref.
S-GNF4
N-doped graphene nanoflakes
SGnFk (I)1000.61013thermal plasma[605]
c-FLGS
sulfur-containing few-layer graphene
SGfL(II)400.68013plasma-assisted electrochemical exfoliation[624]
SG 900
S-doped graphene
SG 90038.220.72813thermal treatment[454]
S-rGO
S-doped reduced graphene oxide
S-rGO
S-rGO
255
255
0.740
0.170
13
1 5
solution process[615]
SPG
S-doped porous holey graphene frameworks
SGpohoχ1250.74513pyrolysis[595]
SGnP
edge-selectively sulfurized graphene nanoplatelet
GSnPl76.430.75013ball milling[587]
S-RGO−180
sulfur-doped reduced graphene oxide
S-rGO 180500.76013hydrothermal process[620]
rGO H2SO4
S-doped reduced graphene oxide
S-rGO4240.76013hydrothermal process[120]
rGO_Se
selenium-doped reduced graphene oxide
Se-rGO169.850.77013solution process[629]
SDGN(10)
S-doped graphene nanosheets
SGnSh (IV)476.190.79013electrochemical exfoliation[610]
SG-700
S-doped graphene
SG 7003200.81013thermal reduction [602]
S-doped MSMG-C
mechanochemically synthesized multilayer graphene
SGmL15300.81013mechanochemical synthesis [502]
S-GNs1000-CB
sulfur-doped graphene nanosheets/carbon black composite
SGnSh/Cb 1000203.820.82013annealing[599]
rGO3TP
2,2′:5′,2″-terthiophene-rGO
3TP-rGO (I)
3TP= 2,2′:5′,2-terthiophene
50.960.82513solution process[625]
S-GQDs/CNPs
hybrid S-doped graphene quantum-dot-decorated carbon nanoparticles
SGqDot/CnP318.470.84013pyrolysis[607]
S-G
S-doped graphene
SG2000.86013annealing[128]
Se-CNTs-graphene-900
selenium-doped carbon nanotube/graphene networks
CSenT 900/GSe141.540.88313annealing[129]
3D S-GNs
3D sulfur-doped graphene networks
SG3Dχ2000.92013thermal annealing[594]
1 For a proper comparison, this table contains only data from samples whose catalyst loading can be defined. 2 The catalyst name in italic is the catalyst name reported in the paper with the best Eonset, reported in the third column and below the explanation reported in the paper; in bold is the name set using the terms reported in the legend, following mostly the indications reported in [892]. 3 Catalyst loading, when not reported in the paper, has been calculated from data provided by the authors. 4 Eonset was generally extracted from the LSV curves analyzed via the WebPlot program. 5 The synthesis method is not fully described but usually refers to the last treatment.
Table 5. Comparative analysis of some ORR parameters of the synthesized B-doped graphene ECs 1 with best performance in alkaline electrolytes.
Table 5. Comparative analysis of some ORR parameters of the synthesized B-doped graphene ECs 1 with best performance in alkaline electrolytes.
Catalyst 2 Catalyst Loading 3 (μgcm−2)Eonset
vs. RHE 4 (V)
Electrolyte
pH
Synthetic Method 5Ref.
BG
B-doped graphene
GB39.810.79013thermal annealing[636]
PN-B2G
polynitrogen N8- (PN) deposited on boron-doped graphene
N8-/GB (II)
N8- = polynitrogen
141.50.82813hydrothermal process[653]
B-G
B-doped graphene
GB2000.80813annealing[128]
B2-3DrGO
B-doped 3D-reduced graphene oxide
B-rGO3D (I)101.910.92513supercritical fluid process[651]
GH-BGQD2
B-doped graphene quantum dots anchored on a graphene hydrogel
GBqDot/GHy7961.78
7961.78
0.940
0.860
13
0.3
hydrothermal process[638]
1 For a proper comparison, this table contains only data from samples whose catalyst loading can be defined. 2 The catalyst name in italic is the catalyst name reported in the paper with the best Eonset, reported in the third column and below the explanation reported in the paper; in bold is the name set using the terms reported in the legend, following mostly the indications reported in [892]. 3 Catalyst loading, when not reported in the paper, has been calculated from data provided by the authors. 4 Eonset was generally extracted from the LSV curves analyzed via the WebPlot program. 5 The synthesis method is not fully described but usually refers to the last treatment.
Table 6. Comparative analysis of some ORR parameters of the synthesized P-doped graphene-based ECs 1 with best performance in alkaline electrolytes.
Table 6. Comparative analysis of some ORR parameters of the synthesized P-doped graphene-based ECs 1 with best performance in alkaline electrolytes.
Catalyst 2 Catalyst Loading 3 (μgcm−2)Eonset
vs. RHE 4 (V)
Electrolyte
pH
Synthetic Method 5Ref.
PG
P-doped graphene
GP2550.76013thermal annealing [273]
PG
P-doped graphene
GP2550.76013calcination[869]
P-G
P-doped Graphene
GP2000.77213annealing[128]
PG
Graphene and Carbon black composite
Cb/GP50.960.82813thermal annealing [657]
P-TRG
Metal-free phosphorus-doped graphene nanosheets
GPnSh141.50.92013thermal annealing [656]
1 For a proper comparison, this table contains only data from samples whose catalyst loading can be defined. 2 The catalyst name in italic is the catalyst name reported in the paper with the best Eonset, reported in the third column and below the explanation reported in the paper; in bold is the name set using the terms reported in the legend, following mostly the indications reported in [892]. 3 Catalyst loading, when not reported in the paper, has been calculated from data provided by the authors. 4 Eonset was generally extracted from the LSV curves analyzed via the WebPlot program. 5 The synthesis method is not fully described but usually refers to the last treatment.
Table 7. Comparative analysis of some ORR parameters of the synthesized halogen-doped graphene-based ECs 1 with best performance in alkaline electrolytes.
Table 7. Comparative analysis of some ORR parameters of the synthesized halogen-doped graphene-based ECs 1 with best performance in alkaline electrolytes.
Catalyst 2 Catalyst Loading 3 (μgcm−2)Eonset
vs. RHE 4 (V)
Electrolyte
pH
Synthetic Method 5Ref.
ClFG
Fluoro and chloro-doped graphene
FClG283.090.76013pyrolysis[674]
HIGnP
high-quality iodine-doped graphene nanoplatelets
IGnPl101.910.77013annealing[682]
IGnP
edge-selectively-iodinated graphene nanoplatelets
BrGnP
edge-selectively-chlorinated graphene nanoplatelets
ClGnP
edge-selectively-iodinated graphene nanoplatelets
IGnPl


BrGnPl


ClGnPl
76.43


76.43


76.43
0.812


0.806


0.768
13


13


13
ball milling


ball milling


ball milling
[675]


[675]


[675]
FIIRGO
Fluorinated reduced graphene oxide
F-rGO (II)8000.83013solution process[672]
I-graphene-900
Iodine-doped graphene
IG 90028.310.85513annealing[131]
HI-RGO/CDs
Iodine-doped reduced graphene oxide/carbon dot composite
ICDot/I-rGO101.910.86013solution process[684]
I/rGO
Iodine-doped reduced graphene oxide
I-rGO661.380.87013microwave heating[688]
FG−1100
Fluorine-doped graphene
FG 11001000.94013thermal treatment[673]
1 For a proper comparison, this table contains only data from samples whose catalyst loading can be defined. 2 The catalyst name in italic is the catalyst name reported in the paper with the best Eonset, reported in the third column and below the explanation reported in the paper; in bold is the name set using the terms reported in the legend, following mostly the indications reported in [892]. 3 Catalyst loading, when not reported in the paper, has been calculated from data provided by the authors. 4 Eonset was generally extracted from the LSV curves analyzed via the WebPlot program. 5 The synthesis method is not fully described but usually refers to the last treatment.
Table 8. Comparative analysis of some ORR parameters of the synthesized multidoped graphene-based ECs 1 with best performance in alkaline electrolytes.
Table 8. Comparative analysis of some ORR parameters of the synthesized multidoped graphene-based ECs 1 with best performance in alkaline electrolytes.
Catalyst 2 Catalyst Loading 3 (μgcm−2)Eonset
vs. RHE 4 (V)
Electrolyte
pH
Synthetic Method 5Ref.
N–F-rGO
N and F-doped reduced graphene oxide
F,N-rGO5000.73013hydrothermal process[692]
NPHG 8
N,P-enriched hierarchically porous graphene
NGPhipo1270.73513carbonization[766]
SN-rGO 900
N and S-doped reduced graphene oxide
N,S-rGO 900139,7950.74013pyrolysis[725]
N/S co-doped graphene
Nitrogen and sulfur co-doped graphene
NSG194.110.75013annealing[709]
B&N-rGO
N- and B-doped reduced graphene oxide
N,B-rGO324.970.77013pyrolysis[649]
CNPS -900
N-, P-, S-tridoped graphene
NSGP (III)668.790.78013annealing[783]
0.05-S, N-rGO
N,S co-doped graphene
N,S-rGO (II)800.790
0.290
13
1 6
hydrothermal process[727]
L-Cy-rGO
heteroatom (N, O, and S)-based
L-cysteine-functionalized rGO
LCy-rGO169.730.79013microwave treatment[752]
BNG
boron and nitrogen co-doped graphene
NGB101.910.79513annealing[744]
N,S-CD/rGO
nitrogen and sulfur-doped carbon dot/graphene composite
NSCDot/rGO353.860.80013solvo-/hydrothermal process[623]
GC-NLS
graphene/CnT composite doped sequentially with both nitrogen and sulfur
NSG/CSNnT (I)407.430.80013pyrolysis[598]
NSP-G1636,94
00
N,P,S-doped graphene
NSGP 1000283.090.82013hydrothermal process[715]
S-N-rGO, 180
nitrogen and phosphorus co-doped reduced graphene oxide
N,S-rGO (I)636.940.82013solution process[730]
N-P-rGO
nitrogen and phosphorus-co-doped graphene nanoribbons/CNTs composite
N,P-rGO (I)191.390.83013solution process[661]
nGR-NF
graphene co-doped with nitrogen and fluorine
FNG4200.83013mechanochemical treatment[777]
B,N-graphene
B- and N-co-doped graphene
NGB282.690.83513annealing[639]
NSG cv
S and N-doped graphene
NSG (I)1280.84013freeze drying[618]
NS-3DrGO 950
nitrogen and sulfur dual-doped three-dimensional reduced graphene oxide
N,S-rGO3D 950203.820.84013pyrolysis[735]
BNG 1000
boron and nitrogen-dual-self-doped graphene sheets
NBGSh 1000607.150.84013pyrolysis[750]
NPG1000
N,P-co-doped graphene
NGP 1000800.84513pyrolysis[760]
N,S-GQDs/rGO
defect-rich nitrogen and sulfur-co-doped
graphene quantum dots
NSGqDot/rGO50.960.85013hydrothermal process[619]
BCN 4
B, N-co-doped graphene
NGB (III)1000.85013pyrolysis[646]
F,N,S-rGO
fluorine, nitrogen, and sulfur-doped graphene
F,N,S-rGO
F,N,S-rGO
255
255
0.850
0.640
13
1 6
pyrolysis[616]
carbonaceous rGO/E. coli
nitrogen and phosphorus-doped graphene–bacteria composite
NCP/N,P-rGO509.550.8513annealing[867]
N-S/Gr1000
N- and S-doped graphene
NSG 1000600
400
200
100
0.850
0.850
0.830
0.740
13annealing[720]
SNGL-20
S, N co-doped, few-layered graphene oxides
NSGOfL3060.85513CVD[593]
PN-rGO
N and P-doped reduced graphene oxide
N,P-rGO5000.85813thermal annealing [761]
N–S-GAs 900
nitrogen and sulfur dual-doped three-dimensional (3D) graphene aerogels
NSG3DAe (I)4250.85913pyrolysis[714]
NBGHS
boron and nitrogen co-doped hollow graphene microspheres
NGBhwmSp1300.86013calcination and etching[643]
NPS-G1
nitrogen, Sulfur, and Phosphorus co-doped graphene
NSGP2550.86013pyrolysis[784]
NS-G
nitrogen, sulfur-doped graphene nanoribbon
GSN2830.86013pyrolysis[713]
BNr-GO(S)
N- and B-doped reduced graphene oxide
N,B-rGOedr4080.86013freeze drying[746]
N,S-GNRs-2s
N and S-doped graphene nanoribbons
NSGnR (III)161.620.87013Solution process[732]
N, S-GCNT
N-doped graphene
Sulfur, nitrogen co-doped nanocomposite of graphene and carbon nanotube
NSCnT/NSG2300.87013annealing[608]
HRBNG
highly reduced boron and nitrogen co-doped graphene
N,B-rGO203.820.87013ball milling[435]
NPG
nitrogen and phosphorus-doped graphene
NGP3200.87013solvothermal process[208]
C/oMUS
nitrogen- and sulfur-co-doped carbon
NSCnT/NSG4000.87013pyrolysis[702]
BN-rGO
N and B-doped reduced graphene oxide
N,B-rGO1019.110.87013annealing[453]
CDs/M-rGO
N,S, and P-doped carbon decorated reduced graphene oxide
NSCPDot/N,S,P-rGO101.920.87913hydrothermal process[862]
BN-GAs-2
boron and nitrogen co-doped graphene aerogels
NGBAe (II)1400.88013freeze drying[751]
NPCGF-7:3
carbon nanotubes (C) loaded on graphene microfolds
N,P-CnT/NGP3000.88013annealing[764]
inG
Iodine/nitrogen co-doped graphene
NIG509.550.88013annealing[678]
N, P-GDs/N-3DG
nitrogen and phosphorus co-doped graphene dots supported on nitrogen-doped three-dimensional graphene
NGPDot/GN3D8000.88013annealing[663]
N,P-CGHNs
nitrogen and phosphorous-doped CNTs and graphene hybrid nanospheres
(NPCnT/NGP)nSp
(NPCnT/NGP)nSp
300
600
0.890
0.830
13
1 6
annealing[758]
GSP
pyrolyzed PG with SO3*Py
NSGpo (II)3000.89013pyrolysis [721]
Gl300G-900
N,S,B, P-doped graphene
NSGPB (I)3000.89013pyrolysis[786]
NFG
nitrogen-doped fluorinated graphene
FNG2830.89813annealing[669]
BNPGA−1000−15
boron, nitrogen, phosphorus ternary doped graphene aerogels
NGPBAe 1000 (I)6000.90013thermal treatment[691]
N,B-GA-900
nitrogen and boron co-doped graphene aerogel
NGBAe 9002120.90913pyrolysis[645]
NSGCB
carbon black incorporated nitrogen and sulfur co-doped graphene
NSG/NSCb306.120.91013pyrolysis[708]
PNGF_ADP(op)
N and P co-doped graphene framework
NGPχ (IV)100.570.91013calcination[768]
BN-Gas
N and B-three-dimensional graphene aerogels
NGB3DAe1600.91013thermal reduction[748]
N-F/rGO
nitrogen and fluorine-doped reduced graphene oxide
F,N-rGO5000.91013thermal annealing[776]
N/S-G-7d
nitrogen/sulfur doped graphene
NSG (III)242.640.91113annealing[719]
NH2-G-F5
amino (NH2)-/fluoro (F)-cofunctionalized graphene
NH2-FG (II)2360.91513solution process[774]
NF–MG3
nitrogen and fluorine-dual-doped mesoporous graphene
FNGmpo (III)4000.92013pyrolysis[670]
NB-3DGN
N and B-doped Three-dimensional graphene networks
NGB3Dχ160.92813electrochemical doping[655]
tdgc-900
three-dimensional (3D DOPED graphene-based carbon (interconnected porous foam)
NCP 900/NGP3DpoFo1530.93013pyrolysis[726]
CNx/CSx-GNRs 1000 2.5h
nitrogen/sulfur co-doped graphene nanoribbons
NSGhenR 1000
NSGhenR 1000
81.53
81.53
0.930
0.500
13
0.3 6
thermal annealing[604]
NSG@CNT-2
N/S co-doped CNT-graphene hybrids
NSG/CnT (II)1800.93013pyrolysis [701]
NSG
nitrogen and sulfur dual-doped graphene
NSG2300.93013pyrolysis[609]
NP + NPY/PG
nitrogen and phosphorus co-doped graphene
NGP6000.93013pyrolysis[769]
NS/G-AA
N,S co-doped graphene nanosheets
NGnSh (II)6000.93013pyrolysis[734]
BN-GQD/G-30
boron- and nitrogen-doped graphene quantum
NGBqDot/G (II)81.530.93513CVD[747]
NSG
three-dimensional hierarchical porous nitrogen and sulfur co-dopedgraphene nanosheets
NSG3DhiponSh
NSG3DhiponSh
102
102
0.935
0.685
13
0.3 6
pyrolysis[606]
N-S-G 1000
nitrogen and sulfur co-doped graphene
NSG 10005660.93913solution polymerization[705]
N,S-PGN 700 rGO
nitrogen/sulfur co-doped porous graphene networks
NSGpoχ 700 (III)3000.94013annealing[729]
NS graphene (1:1)
N and S-doped graphene
NSG (IV)6000.94013pyrolysis[733]
BNG etched
boron, nitrogen-doped graphene
NGB (I)80.950.94513annealing[745]
N,P-GCNS
nitrogen and phosphorus dual-doped graphene/carbon nanosheets
NCP/NGPnSh1410.95013electrochemical preparation[759]
P–N-GFs-HMPA
high performancemetal-free P,N-doped carbon catalysts
NCPnT/NGPFo (I)
NCPnT/NGPFo (I)
360
600
0.950
0.880
13
1 6
solution process[660]
FN3SG
N,F, and S tri-doped porous graphene
FNSGpo
FNSGpo
200
200
0.960
0.710
13
0.3 6
pyrolysis[778]
B,N-PG-O−15
boron and
nitrogen-doped porous graphene
N,B-rGOpo (I)255
255
0.960
0.760
13
0.3 6
annealing[742]
GNF-H/N-F
heteroatom (N,F)-doped various graphitic carbon nanofibers
FNCFb/FNGnSh(III)5000.98013pyrolysis[779]
1 For a proper comparison, this table contains only data from samples whose catalyst loading can be defined. 2 The catalyst name in italic is the catalyst name reported in the paper with the best Eonset, reported in the third column and below the explanation reported in the paper; in bold is the name set using the terms reported in the legend, following mostly the indications reported in [892]. 3 Catalyst loading, when not reported in the paper, has been calculated from data provided by the authors. 4 Eonset was generally extracted from the LSV curves analyzed via the WebPlot program. 5 The synthesis method is not fully described but usually refers to the last treatment. 6 For a direct comparison, values at different pH have also been reported.
Table 9. Comparative analysis of some ORR parameters of the synthesized graphene, composites, and variously doped graphene-based ECs 1 with best performance in acidic electrolytes.
Table 9. Comparative analysis of some ORR parameters of the synthesized graphene, composites, and variously doped graphene-based ECs 1 with best performance in acidic electrolytes.
Catalyst 2 Catalyst Loading 3 (μgcm−2)Eonset
vs. RHE 4 (V)
Electrolyte
pH
Synthetic Method 5Ref.
EG-AQ 1
electrochemical grafted anthraquinone functionalized graphene sheets
AQ-GSh (I)
AQ = anthraquinone C14H8O2
102.04
102.04
0.140
0.729
1
13 6
electrochemical grafting and functionalization[793]
GNR1
graphene nanoribbons
GnR (I)
GnR (I)
GnR (I)
80
80
80
0.170
0.450
0.760
1
7 6
13 6
solution process[447]
S-rGO
S-doped reduced graphene oxide
S-rGO
S-rGO
255
255
0.170
0.740
1
13 5
solution process[615]
GP-BM
graphene prepared via ball milling
G310.700.2691ball milling[795]
0.05-S, N-rGO
N,S co-doped graphene
N,S-rGO (II)
N,S-rGO (II)
800.290
0.790
1
13 6
hydrothermal process[727]
S-rGO
S-doped reduced graphene oxide
S-rGO2550.3001 6reflux solution[615]
P(1,5-DAAQ)/RGO
Poly(1,5-
diaminoanthraquinone)/reduced graphene oxide nanohybrid
P(1,5-DAAQ)/rGO
P(1,5-DAAQ) =
Poly(1,5-
Diaminoanthraquinone) = (C4H7O2N2)n
2000.3270.3potentiodynamic deposition [792]
CNx/CSx-GNRs 1000 2.5h
nitrogen/sulfur co-doped graphene nanoribbons
NSGhenR 1000
NSGhenR 1000
81.53
81.53
0.500
0.930
0.3
13 6
thermal annealing[604]
rGO/PANI/PVAPANI/PVA/rGO
PVA = polyvinyl alcohol (C2H4O)n
PANI =
POYLANILINE [C6H4NH]n
90.6100.3chemical reduction in solution[881]
F,N,S-rGO
fluorine nitrogen and sulfurl-doped graphene
F,N,S-rGO
F,N,S-rGO
255
255
0.640
0.850
1
13 6
pyrolysis[616]
NSG
three-dimensional hierarchical porous nitrogen and sulfur-codoped graphene nanosheets
NSG3DhiponSh NSG3DhiponSh102
102
0.685
0.935
0.3
13 6
pyrolysis[606]
FN3SG
N, F, S tri-doped porous graphene
FNSGpo
FNSGpo
200
200
0.710
0.960
0.3
13 6
Reflux solution[778]
NPC/G
N, P dual-doped carbon/graphene
G/NCP (I)2550.7301calcination[869]
B,N-PG-O-15
Boron and
nitrogen-doped porous graphene
N,B-rGOpo (I)
N,B-rGOpo (I)
255
255
0.760
0.960
0.3
13 6
annealing[742]
N-HC@G-900
ultrathin N-doped holey carbon/graphene hybrid
CNhl 900/G
CNhl 900/G
250
250
0.770
0.940
0.3
13 6
calcination[842]
PNGr
phosphorous and nitrogen-doped graphene
NGP707.710.7701pyrolysis[640]
NSG@CNT-3
hybrid N,S-doped CNT-graphene nanocomposites
NSG/CnT (III)3000.8201pyrolysis[701]
N,P-CGHNs
nitrogen and phosphorous doped CNTs and graphene hybrid nanospheres
(NCPnT/NGP)nSp
(NCPnT/NGP)nSp
600
300
0.830
0.890
1
13 6
annealing[758]
GH-BGQD2
B-doped graphene quantum dots anchored on a graphene hydrogel
GBqDot/GHy (II)
GBqDot/GHy (II)
7961.78
7961.78
0.860
0.940
0.3
13 6
pyrolysis[638]
P–N-GFs-HMPA
high performancemetal-free P,N-doped carbon catalysts
NCPnT/NGPFo (I)
NCPnT/NGPFo (I)
600
0.360
0.880
0.950
1
13 6
solution process[660]
1 For a proper comparison, this table contains only data from samples whose catalyst loading can be defined. 2 The catalyst name in italic is the catalyst name reported in the paper with the best Eonset, reported in the third column and below the explanation reported in the paper; in bold is the name set using the terms reported in the legend, following mostly the indications reported in [892]. 3 Catalyst loading, when not reported in the paper, has been calculated from data provided by the authors. 4 Eonset was generally extracted from the LSV curves analyzed via the WebPlot program. 5 The synthesis method is not fully described but usually refers to the last treatment. 6 For a direct comparison, values at different pH have also been reported.
Table 10. Comparative analysis of some ORR parameters of the synthesized graphene, composites, and variously heteroatom-doped graphene-based ECs 1 with best performance in neutral electrolytes.
Table 10. Comparative analysis of some ORR parameters of the synthesized graphene, composites, and variously heteroatom-doped graphene-based ECs 1 with best performance in neutral electrolytes.
Catalyst 2 Catalyst Loading 3 (μgcm−2)Eonset
vs. RHE 4 (V)
Electrolyte
pH
Synthetic Method 5Ref.
AQS/RGO
anthraquinone monosulfonate sodium/reduced graphene oxide nanocomposite
AQS/rGO
AQS = C14H7NaO5S
5.670.4407solution process[889]
N/S-G
nitrogen and sulfur dual-doped graphene
NSG141.540.4417pyrolysis[731]
RGSs
reduced graphene oxide sheets amino-functionalized graphene
rGOSh2000.4607electrochemical reduction[790]
DNGS 480 900
dendritic N-doped graphene spheres
NGddSp 900
NGddSp 900
NGddSp 900
1000
1000
1000
0.470
0.550
0.770
7
0.3 6
13 6
pyrolysis[532]
CN 1000
N-doped graphene
NG 1000
NG 1000
144
144
0.567
0.847
7,4
146
thermal treatment [390]
GNR1
graphene nanoribbons
GnR
GnR
GnR
80
80
80
0.450
0.760
0.170
7
13 6
1 6
solution process[447]
NG/CB 10
porous N-doped graphene/carbon black composite
NG/Cb (IV)1000.7807,4pyrolysis[344]
5rG@NHCS
reduced GO-modified ultra-thin nitrogen-doped hollow carbon sphere composites
rGO/CNhwnSp (II)
rGO/CNhwnSp (II)
1400
1400
0.851
0.882
7
13 6
pyrolysis[845]
I-NG
N-doped graphene with implantation of nitrogen active sites via C3N4
C3N4/NG283.090.9907sonochemical process[312]
1 For a proper comparison, this table contains only data from samples whose catalyst loading can be defined. 2 The catalyst name in italic is the catalyst name reported in the paper with the best Eonset, reported in the third column and below the explanation reported in the paper; in bold is the name set using the terms reported in the legend, following mostly the indications reported in [892]. 3 Catalyst loading, when not reported in the paper, has been calculated from data provided by the authors. 4 Eonset was generally extracted from the LSV curves analyzed via the WebPlot program. 5 The synthesis method is not fully described but usually refers to the last treatment. 6 For a direct comparison, values at different pH have also been reported.
Table 11. Comparative analysis of some ORR parameters of the synthesized graphene-based ECs 1 with best performance in alkaline electrolytes.
Table 11. Comparative analysis of some ORR parameters of the synthesized graphene-based ECs 1 with best performance in alkaline electrolytes.
Catalyst 2 Catalyst Loading 3 (μgcm−2)Eonset
vs. RHE 4 (V)
Electrolyte
pH
Synthetic Method 5Ref.
LIG-O
oxidized laser-induced graphene
G (I)150.980.690 13plasma treatment[802]
Gdots
graphene quantum dts
GqDot100.900.53913electrochemical process[775]
GO 1000
edge-rich non-doped graphene
Gedr 800400.64013pyrolysis[801]
H-rGO-sp
heat-treated reduced graphene oxide-based hollow spheres
rGOhwSp141.540.71013solution process[444]
EG-AQ 2
electrochemical-grafted anthraquinone functionalized graphene sheets
AQ-GSh (II)
AQ = anthraquinone
102.04
102.04
0.729
0.140
13
1 6
electrochemical grafting and functionalization[793]
GrapheneG90.8
453.9
0.740
0.770
13
13
thermal treatment[814]
G
graphene
G121.460.74013pyrolysis[816]
graphene–C60 hybridC60-G500.75013freeze drying[794]
10K
graphene nanosheets
GnSh (III)2830.75813solution process[506]
rGO pH 10
reduced graphene oxide nanoscrolls
rGOnSc600.76013freeze casting process[530]
GNR1
graphene nanoribbon
GnR (I)
GnR (I)
GnR (I)
80
80
80
0.760
0.170
0.450
13
1 6
7 6
solution process[447]
GC/Graphene1G(I)1000.78013solution process[813]
DMG−10
doping-free, multi-defect graphene
Gdf (I)253.440.81013hydrothermal process[805]
DTG
3D double-layer templated graphene
G3D2550.82013CVD[798]
MC-G700
mechanochemically synthesized multilayer graphene
GmL (II)15300.82014thermal treatment[806]
H-rGO
hydrazine-reduced graphene oxide
rGO151.650.84013solution process[500]
P-G
plasma-treated graphene
G(I)151.650.84013Plasma treatment[797]
GP-BM
graphene prepared via ball milling
G310.700.86714ball milling[795]
8Co/G 700-HCl
holey graphene
G 700 (V)152.80.87013chemical etching[799]
F-graphene
graphene fragments
GFg (I)283.090.88013grinding[800]
3D-PGM−1150
3D interconnected hierarchical porous graphene mesh
G3Dhipomh 11501000.93013self-assembly and chemical etching[807]
1 For a proper comparison, this table contains only data from samples whose catalyst loading can be defined. 2 The catalyst name in italic is the catalyst name reported in the paper with the best Eonset, reported in the third column and below the explanation reported in the paper; in bold is the name set using the terms reported in the legend, following mostly the indications reported in [892]. 3 Catalyst loading, when not reported in the paper, has been calculated from data provided by the authors. 4 Eonset was generally extracted from the LSV curves analyzed via the WebPlot program. 5 The synthesis method is not fully described but usually refers to the last treatment. 6 For a direct comparison, values at different pH have also been reported.
Table 12. Comparative analysis of some ORR parameters of the synthesized graphene composite-based ECs 1 with best performance in alkaline electrolytes.
Table 12. Comparative analysis of some ORR parameters of the synthesized graphene composite-based ECs 1 with best performance in alkaline electrolytes.
Catalyst 2 Catalyst Loading 3 (μgcm−2)Eonset
vs. RHE 4 (V)
Electrolyte
pH 5
Synthetic Method 6Ref.
C50
vertical graphene sheets on separated papillary granules formed
nanocrystalline diamond films
Gv/CDdFm (III)800.67013plasma etching[827]
Si–GQD1.0 NCs
hybrid silicon nanosheets (NSs)–graphene quantum dot nanocomposites
SinSh/GqDot10400.67013solution process[825]
GOBN5 700
reduced graphene oxide/boron nitride composite annealed at 700
BNhg/rGO42.260.68013annealing[888]
NCGO0.05 1000
nitrogen-doped carbon graphene composite
CNposhell/rGOcore 10001270.70013thermal treatment[831]
C3N4@rGO
graphitic carbon nitride on reduced graphene oxide
C3N4/rGO8490.72013pyrolysis[859]
N-DC/G
three-dimensional honeycomb-like nitrogen-doped carbon nanosheet/graphene
NC3DhcnSh/G99.80.73513carbonization[839]
CNx/G-600
carbon-nitrogen/graphene composite
NC 600/rGO509.550.76013pyrolysis[832]
GH-GQD 90
graphene
quantum dots/graphene hydrogel composites
GqDot/GHy (II)1000.78013hydrothermal process[804]
NCNTs/G
three-dimensional (3D) nitrogen-doped carbon nanotubes/graphene
NCnT/G3D10000.78013pyrolysis[866]
MCN−1_G3
triazole-based mesoporous C3N5
a graphene hybrid
C3N5mpo/G (I)13.8760.78013calcination and etching[849]
PEDOT/rGO/H2SO4
PEDOT:PSS/Graphene Composite
PEDOT + PSS/rGO (I)
PEDOT = Poly(3,4-ethylenedioxythiophene)
PSS = Polystyrene
1200.78513heating in solution[878]
 
g-C3N4/rGO
ultrathin g-C3N4 nanosheets and reduced graphene oxide hybrid
g-C3N4uth/rGO (I)3000.79513photoreduction[855]
G-PMF1−1
graphene-based composite sheets
NC/G/NC (II)800.79913annealing[833]
N-CDs/G
nitrogen-doped carbon dots
decorated on the graphene surface
NCtDot/G50.950.82013hydrothermal process [860]
hp-GSGCN_2x
hierarchically porous graphene sheets/graphitic carbon nitride intercalated composites
GSh/g-C3N4 (I)283.080.82513thermal annealing[854]
HMCG Cys
heteroatom (N and S) doped mesoporous carbon/graphene
NSCmpo/G (I)2000.83813nanocasting approach[865]
0.47%
boron-doped carbon catalysts on graphene
BC/G (I)2230.84013liquid-phase electrodeposition[830]
A/graphene
adenine functionalized graphene
A/G
A = adenine
2500.84013ultrasound[877]
N-VA-CNTs/GF
free-standing vertically aligned nitrogen-doped
carbon nanotube supported by graphene foam
NCvalnT/GFo143.310.85013PECVD[884]
P-T/rGO
covalently linked
pyridine and thiophene molecule (P-T) polymer composite with reduced graphene oxide
(C45H74N4B2F2S)n/rGO521.130.85013sonication[879]
GMC-s
graphene-based microporous carbons
SC/G/SCL6000.85013pyrolysis[834]
FPCFs@rGO
fullerene-derived porous carbon fibers (FPCFs)/reduced graphene oxide composite
C60poFb/rGO (II)2500.87013calcination[828]
NS-CD@gf_a900
carbon dot-embedded porous graphene
NSCDot 900/rGO2800.87013annealing[863]
G-CN 800
graphene-based carbon nitride nanosheets
CNx 800/G70,70.87813pyrolysis[852]
5rG@NHCS
reduced GO (rG) modified ultra-thin nitrogen-doped hollow carbon sphere (NHCS) composites
rGO/NCuthwnSp (II)
rGO/NCuthwnSp (II)
1400
1400
0.882
0.851
13
76
pyrolysis[845]
N-APC-Gr-900
N-doped hybrid carbons
NC 900/G4000.89013carbonization[848]
mNC/rGO
mesoporous nitrogen-doped carbon/reduced graphene oxide
NCmpo/rGO2400.90013annealing[843]
NPC/G
N, P dual-doped carbon/graphene
G/NCP (I)2550.90013calcination[869]
rGO@PN\C-2
phosphorus- and
nitrogen-doped carbon nanosheets
(NCP/rGO/NPC)nSh (II)2000.91013carbonization[870]
MTCG
monoatomic-thick g-C3N4dots@
graphene
C3N4mtDot/G800.91513hydrothermal process[856]
N-PC@G-0.02
N-doped porous carbon@graphene composites
NCpo/G/NCpo (I)407.640.92013pyrolysis[838]
N-MCS@rGO
N-doped mesoporous carbon spheres loaded on reduced graphene
oxide nanosheets
NCmpoSp/rGOnSh7200.92013carbonization[847]
rGO/PANI/PVA
reduced GO polyvinyl alcohol
polyaniline nanocomposites
PANI/PVA/rGO
PVA = polyvinyl alcohol
PANI =
polyaniline
90.93013.7chemical reduction in solution[881]
PANRGO700
porous carbon nanoballs on graphene composite
NCnBl 700/rGO2000.93013pyrolysis [846]
N-HC@G-900
ultrathin N-doped holey carbon/graphene hybrid
CNhl900/G
CNhl900/G
250
250
0.940
0.770
13
0.3 6
calcination[842]
N-MC/rGO-800
graphene-supported nitrogen-doped mesoporous carbons
NCmpo/rGO/NCmpo 9002040.95013pyrolysis[840]
NPCS-850
2D nitrogen-doped porous carbon sheet
NC2DpoSh 850/G2500.97013pyrolysis[841]
G-GCN
graphitic carbon nitride immobilized on graphene
g-C3N4/G714.290.97913solution chemistry[851]
1 For a proper comparison, this table contains only data for samples whose catalyst loading can be defined. 2 The catalyst name in italic is the catalyst name reported in the paper with the best Eonset, reported in the third column and below the explanation reported in the paper; in bold is the name set using the terms reported in the legend, following mostly the indications reported in [892]. 3 Catalyst loading, when not reported in the paper, has been calculated from data provided by the authors. 4 Eonset was generally extracted from the LSV curves analyzed via the WebPlot program. 5 The synthesis method is not fully described but it usually refers to the last treatment. 6 For a direct comparison, values at different pH have also been reported.

Supplementary Materials

The following supporting information can be downloaded at https://www.mdpi.com/article/10.3390/molecules30102248/s1.

Funding

This research received no external funding.

Conflicts of Interest

The author declares no conflict of interest.

References

  1. Wang, M.; Wang, S.Y.; Yang, H.Q.; Ku, W.; Yang, S.C.; Liu, Z.N.; Lu, G.L. Carbon-Based Electrocatalysts Derived from Biomass for Oxygen Reduction Reaction: A Minireview. Front. Chem. 2020, 8, 5. [Google Scholar] [CrossRef] [PubMed]
  2. Bhuvanendran, N.; Ravichandran, S.; Xu, Q.; Maiyalagan, T.; Su, H. A Quick Guide to the Assessment of Key Electrochemical Performance Indicators for the Oxygen Reduction Reaction: A Comprehensive Review. Int. J. Hydrogen Energy 2022, 47, 7113–7138. [Google Scholar] [CrossRef]
  3. Zhu, Y.S.; Zhang, B.S. NanoCarbon-Based Metal-Free and Non-Precious Metal Bifunctional Electrocatalysts for Oxygen Reduction and Oxygen Evolution Reactions. J. Energy Chem. 2021, 58, 610–628. [Google Scholar] [CrossRef]
  4. Dai, L.M. Functionalization of Graphene for Efficient Energy Conversion and Storage. Acc. Chem. Res. 2013, 46, 31–42. [Google Scholar] [CrossRef]
  5. Lv, W.; Li, Z.J.; Deng, Y.Q.; Yang, Q.H.; Kang, F.Y. Graphene-Based Materials for Electrochemical Energy Storage Devices: Opportunities and Challenges. Energy Storage Mater. 2016, 2, 107–138. [Google Scholar] [CrossRef]
  6. Yang, L.J.; Shui, J.L.; Du, L.; Shao, Y.Y.; Liu, J.; Dai, L.M.; Hu, Z. Carbon-Based Metal-Free ORR Electrocatalysts for Fuel Cells: Past, Present, and Future. Adv. Mater. 2019, 31, 20. [Google Scholar] [CrossRef]
  7. Deng, J.; Fang, S.Y.; Fang, Y.; Hao, Q.Q.; Wang, L.; Hu, Y.H. Multiple Roles of Graphene in Electrocatalysts for Metal-Air Batteries. Catal. Today 2023, 409, 2–22. [Google Scholar] [CrossRef]
  8. Zhang, L.; Wang, S.; Wang, Q.; Shao, H.Y.; Jin, Z. Dendritic Solid Polymer Electrolytes: A New Paradigm for High-Performance Lithium-Based Batteries. Adv. Mater. 2023, 35, 35. [Google Scholar] [CrossRef]
  9. Wang, S.; Zhang, L.; Zeng, Q.H.; Guan, J.Z.; Gao, H.Q.; Zhang, L.Y.; Zhong, J.; Lai, W.Y.; Wang, Q. Designing Polymer Electrolytes via Ring-Opening Polymerization for Advanced Lithium Batteries. Adv. Energy Mater. 2024, 14, 37. [Google Scholar] [CrossRef]
  10. Xiao, S.J.; Ren, L.T.; Liu, W.; Zhang, L.; Wang, Q. High-Voltage Polymer Electrolytes: Challenges and Progress. Energy Storage Mater. 2023, 63, 35. [Google Scholar] [CrossRef]
  11. Qasem, N.A.A.; Abdulrahman, G.A.Q. A Recent Comprehensive Review of Fuel Cells: History, Types, and Applications. Int. J. Energy Res. 2024, 2024, 36. [Google Scholar] [CrossRef]
  12. Zhou, X.J.; Qiao, J.L.; Yang, L.; Zhang, J.J. A Review of Graphene-Based Nanostructural Materials for Both Catalyst Supports and Metal-Free Catalysts in PEM Fuel Cell Oxygen Reduction Reactions. Adv. Energy. Mater. 2014, 4, 25. [Google Scholar] [CrossRef]
  13. Li, F.J.; Chan, S.H.; Tu, Z.K. Recent Development of Anion Exchange Membrane Fuel Cells and Performance Optimization Strategies: A Review. Chem. Rec. 2024, 24, 31. [Google Scholar] [CrossRef] [PubMed]
  14. ElMekawy, A.; Hegab, H.M.; Losic, D.; Saint, C.P.; Pant, D. Applications of Graphene in Microbial Fuel Cells: The Gap between Promise and Reality. Renew. Sustain. Energy Rev. 2017, 72, 1389–1403. [Google Scholar] [CrossRef]
  15. Lu, J.H.; Jaumaux, P.; Wang, T.Y.; Wang, C.Y.; Wang, G.X. Recent Progress in Quasi-Solid and Solid Polymer Electrolytes for Multivalent Metal-Ion Batteries. J. Mater. Chem. A 2021, 9, 24175–24194. [Google Scholar] [CrossRef]
  16. Liu, Q.F.; Pan, Z.F.; Wang, E.D.; An, L.; Sun, G.Q. Aqueous Metal-Air Batteries: Fundamentals and Applications. Energy Storage Mater. 2020, 27, 478–505. [Google Scholar] [CrossRef]
  17. Novoselov, K.S.; Geim, A.K.; Morozov, S.V.; Jiang, D.; Zhang, Y.; Dubonos, S.V.; Grigorieva, I.V.; Firsov, A.A. Electric Field Effect in Atomically Thin Carbon Films. Science 2004, 306, 666–669. [Google Scholar] [CrossRef]
  18. Yang, Y.K.; Han, C.P.; Jiang, B.B.; Iocozzia, J.; He, C.G.; Shi, D.; Jiang, T.; Lin, Z.Q. Graphene-Based Materials with Tailored Nanostructures for Energy Conversion and Storage. Mater. Sci. Eng. R Rep. 2016, 102, 1–72. [Google Scholar] [CrossRef]
  19. Chang, D.W.; Baek, J.B. Eco-Friendly Synthesis of Graphene Nanoplatelets. J. Mater. Chem. A 2016, 4, 15281–15293. [Google Scholar] [CrossRef]
  20. Mitra, S.; Banerjee, S.; Datta, A.; Chakravorty, D. A Brief Review on Graphene/Inorganic Nanostructure Composites: Materials for the Future. Indian J. Phys. 2016, 90, 1019–1032. [Google Scholar] [CrossRef]
  21. Coleman, J.N. Liquid Exfoliation of Defect-Free Graphene. Acc. Chem. Res. 2013, 46, 14–22. [Google Scholar] [CrossRef] [PubMed]
  22. Qian, W.; Hao, R.; Hou, Y.L.; Tian, Y.; Shen, C.M.; Gao, H.J.; Liang, X.L. Solvothermal-Assisted Exfoliation Process to Produce Graphene with High Yield and High Quality. Nano Res. 2009, 2, 706–712. [Google Scholar] [CrossRef]
  23. Hossain, M.M.; Hahn, J.R.; Ku, B.C. Synthesis of Highly Dispersed and Conductive Graphene Sheets by Exfoliation of Preheated Graphite in a Sealed Bath and Its Applications to Polyimide Nanocomposites. Bull. Korean Chem. Soc. 2014, 35, 2049–2056. [Google Scholar] [CrossRef]
  24. Wang, S.; Yi, M.; Shen, Z.G. The Effect of Surfactants and Their Concentration on the Liquid Exfoliation of Graphene. RSC Adv. 2016, 6, 56705–56710. [Google Scholar] [CrossRef]
  25. Lotya, M.; Hernandez, Y.; King, P.J.; Smith, R.J.; Nicolosi, V.; Karlsson, L.S.; Blighe, F.M.; De, S.; Wang, Z.M.; McGovern, I.T.; et al. Liquid Phase Production of Graphene by Exfoliation of Graphite in Surfactant/Water Solutions. J. Am. Chem. Soc. 2009, 131, 3611–3620. [Google Scholar] [CrossRef]
  26. Badri, M.A.S.; Salleh, M.M.; Noor, N.F.M.; Abd Rahman, M.Y.; Umar, A.A. Green Synthesis of Few-Layered Graphene from Aqueous Processed Graphite Exfoliation for Graphene Thin Film Preparation. Mater. Chem. Phys. 2017, 193, 212–219. [Google Scholar] [CrossRef]
  27. Chaban, V.V.; Fileti, E.E.; Prezhdo, O.V. Exfoliation of Graphene in Ionic Liquids: Pyridinium versus Pyrrolidinium. J. Phys. Chem. C 2017, 121, 911–917. [Google Scholar] [CrossRef]
  28. Yu, Y.Y.; Han, Z.; Zhang, Y.B.; Dong, B.; Kong, A.G.; Shan, Y.K. Synthesis of High-Quality Graphene Sheets in Task-Specific Ionic Liquids and Their Photocatalytic Performance. N. J. Chem. 2016, 40, 3147–3154. [Google Scholar] [CrossRef]
  29. Chaban, V.V.; Fileti, E.E. Graphene Exfoliation in Ionic Liquids: Unified Methodology. RSC Adv. 2015, 5, 81229–81234. [Google Scholar] [CrossRef]
  30. Najafabadi, A.T.; Gyenge, E. High-Yield Graphene Production by Electrochemical Exfoliation of Graphite: Novel Ionic Liquid (IL)-Acetonitrile Electrolyte with Low IL Content. Carbon 2014, 71, 58–69. [Google Scholar] [CrossRef]
  31. Liu, N.; Luo, F.; Wu, H.; Liu, Y.; Zhang, C.; Chen, J. One-Step Ionic-Liquid-Assisted Electrochemical Synthesis of Ionic-Liquid-Functionalized Graphene Sheets Directly from Graphite. Adv. Funct. Mater. 2008, 18, 1518–1525. [Google Scholar] [CrossRef]
  32. Sahoo, S.K.; Mallik, A. Fundamentals of Fascinating Graphene Nanosheets: A Comprehensive Study. Nano 2019, 14, 1930003. [Google Scholar] [CrossRef]
  33. Yang, S.; Lohe, M.R.; Müllen, K.; Feng, X. New-Generation Graphene from Electrochemical Approaches: Production and Applications. Adv. Mater. 2016, 28, 6213–6221. [Google Scholar] [CrossRef] [PubMed]
  34. Hsieh, C.T.; Hsueh, J.H. Electrochemical Exfoliation of Graphene Sheets from a Natural Graphite Flask in the Presence of Sulfate Ions at Different Temperatures. RSC Adv. 2016, 6, 64826–64831. [Google Scholar] [CrossRef]
  35. Gee, C.M.; Tseng, C.C.; Wu, F.Y.; Chang, H.P.; Li, L.J.; Hsieh, Y.P.; Lin, C.T.; Chen, J.C. Flexible Transparent Electrodes Made of Electrochemically Exfoliated Graphene Sheets from Low-Cost Graphite Pieces. Displays 2013, 34, 315–319. [Google Scholar] [CrossRef]
  36. Munuera, J.M.; Paredes, J.I.; Villar-Rodil, S.; Ayán-Varela, M.; Pagán, A.; Aznar-Cervantes, S.D.; Cenis, J.L.; Martínez-Alonso, A.; Tascón, J.M.D. High Quality, Low Oxygen Content and Biocompatible Graphene Nanosheets Obtained by Anodic Exfoliation of Different Graphite Types. Carbon 2015, 94, 729–739. [Google Scholar] [CrossRef]
  37. Huang, X.; Li, S.; Qi, Z.; Zhang, W.; Ye, W.; Fang, Y. Low Defect Concentration Few-Layer Graphene Using a Two-Step Electrochemical Exfoliation. Nanotechnology 2015, 26, 105602. [Google Scholar] [CrossRef]
  38. Liu, J.; Poh, C.K.; Zhan, D.; Lai, L.; Lim, S.H.; Wang, L.; Liu, X.; Gopal Sahoo, N.; Li, C.; Shen, Z.; et al. Improved Synthesis of Graphene Flakes from the Multiple Electrochemical Exfoliation of Graphite Rod. Nano Energy 2013, 2, 377–386. [Google Scholar] [CrossRef]
  39. Punith Kumar, M.K.; Nidhi, M.; Srivastava, C. Electrochemical Exfoliation of Graphite to Produce Graphene Using Tetrasodium Pyrophosphate. RSC Adv. 2015, 5, 24846–24852. [Google Scholar] [CrossRef]
  40. Hofmann, M.; Chiang, W.Y.; Nguyn, T.D.; Hsieh, Y.P. Controlling the Properties of Graphene Produced by Electrochemical Exfoliation. Nanotechnology 2015, 26, 335607. [Google Scholar] [CrossRef]
  41. Wei, D.; Grande, L.; Chundi, V.; White, R.; Bower, C.; Andrew, P.; Ryhänen, T. Graphene from Electrochemical Exfoliation and Its Direct Applications in Enhanced Energy Storage Devices. Chem. Commun. 2012, 48, 1239–1241. [Google Scholar] [CrossRef] [PubMed]
  42. Sima, M.; Enculescu, I.; Sima, A. Preparation of Graphene and Its Application in Dye-Sensitized Solar Cells. Optoelectron. Adv. Mater. Rapid Commun. 2011, 5, 414–418. [Google Scholar]
  43. Tang, J.; Chen, S.; Yuan, Y.; Cai, X.; Zhou, S. In Situ Formation of Graphene Layers on Graphite Surfaces for Efficient Anodes of Microbial Fuel Cells. Biosens. Bioelectron. 2015, 71, 387–395. [Google Scholar] [CrossRef] [PubMed]
  44. Lu, J.; Yang, J.X.; Wang, J.; Lim, A.; Wang, S.; Loh, K.P. One-Pot Synthesis of Fluorescent Carbon Nanoribbons, Nanoparticles, and Graphene by the Exfoliation of Graphite in Ionic Liquids. ACS Nano 2009, 3, 2367–2375. [Google Scholar] [CrossRef] [PubMed]
  45. Xu, J.; Shi, Z.; Zhang, X.; Haarberg, G.M. Mechanism of Graphene Formation by Graphite Electro-Exfoliation in Ionic Liquids-Water Mixtures. Mater. Res. Express 2015, 1, 045606. [Google Scholar] [CrossRef]
  46. Parvez, K.; Wu, Z.S.; Li, R.; Liu, X.; Graf, R.; Feng, X.; Müllen, K. Exfoliation of Graphite into Graphene in Aqueous Solutions of Inorganic Salts. J. Am. Chem. Soc. 2014, 136, 6083–6091. [Google Scholar] [CrossRef]
  47. Su, C.Y.; Lu, A.Y.; Xu, Y.; Chen, F.R.; Khlobystov, A.N.; Li, L.J. High-Quality Thin Graphene Films from Fast Electrochemical Exfoliation. ACS Nano 2011, 5, 2332–2339. [Google Scholar] [CrossRef]
  48. Wang, J.; Manga, K.K.; Bao, Q.; Loh, K.P. High-Yield Synthesis of Few-Layer Graphene Flakes through Electrochemical Expansion of Graphite in Propylene Carbonate Electrolyte. J. Am. Chem. Soc. 2011, 133, 8888–8891. [Google Scholar] [CrossRef]
  49. Zhou, M.; Tang, J.; Cheng, Q.; Xu, G.; Cui, P.; Qin, L.C. Few-Layer Graphene Obtained by Electrochemical Exfoliation of Graphite Cathode. Chem. Phys. Lett. 2013, 572, 61–65. [Google Scholar] [CrossRef]
  50. Cooper, A.J.; Wilson, N.R.; Kinloch, I.A.; Dryfe, R.A.W. Single Stage Electrochemical Exfoliation Method for the Production of Few-Layer Graphene via Intercalation of Tetraalkylammonium Cations. Carbon 2014, 66, 340–350. [Google Scholar] [CrossRef]
  51. Pang, Y.X.; Yew, M.; Yan, Y.X.; Khine, P.; Filbert, A.; Manickam, S.; Foo, D.C.Y.; Sharmin, N.; Lester, E.; Wu, T.; et al. Application of Supercritical Fluid in the Synthesis of Graphene Materials: A Review. J. Nanopart. Res. 2021, 23, 204. [Google Scholar] [CrossRef]
  52. Rangappa, D.; Sone, K.; Wang, M.S.; Gautam, U.K.; Golberg, D.; Itoh, H.; Ichihara, M.; Honma, I. Rapid and Direct Conversion of Graphite Crystals into High-Yielding, Good-Quality Graphene by Supercritical Fluid Exfoliation. Chem.-Eur. J. 2010, 16, 6488–6494. [Google Scholar] [CrossRef] [PubMed]
  53. Balaji, S.S.; Raj, J.A.; Karnan, M.; Sathish, M. Supercritical Fluid Assisted Synthesis of S-Doped Graphene and Its Symmetric Supercapacitor Performance Evaluation Using Different Electrolytes. Synth. Met. 2019, 255, 116111. [Google Scholar] [CrossRef]
  54. Tomai, T.; Kawaguchi, Y.; Honma, I. Nanographene Production from Platelet Carbon Nanofiber by Supercritical Fluid Exfoliation. Appl. Phys. Lett. 2012, 100, 233110. [Google Scholar] [CrossRef]
  55. Ibarra, R.M.; Goto, M.; García-Serna, J.; Montes, S.M.G. Graphene Exfoliation with Supercritical Fluids. Carbon Lett. 2021, 31, 99–105. [Google Scholar] [CrossRef]
  56. Pu, N.W.; Wang, C.A.; Sung, Y.; Liu, Y.M.; Ger, M.D. Production of Few-Layer Graphene by Supercritical CO2 Exfoliation of Graphite. Mater. Lett. 2009, 63, 1987–1989. [Google Scholar] [CrossRef]
  57. Sim, H.S.; Kim, T.A.; Lee, K.H.; Park, M. Preparation of Graphene Nanosheets through Repeated Supercritical Carbon Dioxide Process. Mater. Lett. 2012, 89, 343–346. [Google Scholar] [CrossRef]
  58. Wang, W.C.; Wang, Y.; Gao, Y.H.; Zhao, Y.P. Control of Number of Graphene Layers Using Ultrasound in Supercritical CO2 and Their Application in Lithium-Ion Batteries. J. Supercrit. Fluids 2014, 85, 95–101. [Google Scholar] [CrossRef]
  59. Xu, Q.Q.; Zhao, W.; Zhi, J.T.; Yin, J.Z. Exfoliation of Graphite in CO2 Expanded Organic Solvents Combined with Low Speed Shear Mixing. Carbon 2018, 135, 180–186. [Google Scholar] [CrossRef]
  60. Maraschin, T.G.; Gonçalves, R.V.; de Vargas, M.C.; Correa, R.; Basso, N.R.S.; Galland, G.B.; Cassel, E. Few-Layer Graphene Production through Graphite Exfoliation in Pressurized CO2 Assisted by Natural Surfactant. FlatChem 2024, 45, 12. [Google Scholar] [CrossRef]
  61. Gao, Y.H.; Shi, W.; Wang, W.C.; Wang, Y.; Zhao, Y.P.; Lei, Z.H.; Miao, R.R. Ultrasonic-Assisted Production of Graphene with High Yield in Supercritical CO2 and Its High Electrical Conductivity Film. Ind. Eng. Chem. Res. 2014, 53, 2839–2845. [Google Scholar] [CrossRef]
  62. Gai, Y.Z.; Wang, W.C.; Xiao, D.; Tan, H.J.; Lin, M.Y.; Zhao, Y.P. Exfoliation of Graphite into Graphene by a Rotor-Stator in Supercritical CO2: Experiment and Simulation. Ind. Eng. Chem. Res. 2018, 57, 8220–8229. [Google Scholar] [CrossRef]
  63. Tan, H.J.; Navik, R.; Liu, Z.Y.; Xiang, Q.X.; Zhao, Y.P. Scalable Massive Production of Defect-Free Few-Layer Graphene by Ball-Milling in Series with Shearing Exfoliation in Supercritical CO2. J. Supercrit. Fluids 2022, 181, 9. [Google Scholar] [CrossRef]
  64. Sasikala, S.P.; Huang, K.; Giroire, B.; Prabhakaran, P.; Henry, L.; Penicaud, A.; Poulin, P.; Aymonier, C. Simultaneous Graphite Exfoliation and N Doping in Supercritical Ammonia. ACS Appl. Mater. Interfaces 2016, 8, 30964–30971. [Google Scholar] [CrossRef]
  65. Chen, J.; Yao, B.; Li, C.; Shi, G. An Improved Hummers Method for Eco-Friendly Synthesis of Graphene Oxide. Carbon 2013, 64, 225–229. [Google Scholar] [CrossRef]
  66. Gao, W.; Alemany, L.B.; Ci, L.; Ajayan, P.M. New Insights into the Structure and Reduction of Graphite Oxide. Nat. Chem. 2009, 1, 403–408. [Google Scholar] [CrossRef]
  67. Bai, H.; Li, C.; Shi, G. Functional Composite Materials Based on Chemically Converted Graphene. Adv. Mater. 2011, 23, 1089–1115. [Google Scholar] [CrossRef]
  68. Hummers, W.S., Jr.; Offeman, R.E. Preparation of Graphitic Oxide. J. Am. Chem. Soc. 1958, 80, 1339. [Google Scholar] [CrossRef]
  69. Marcano, D.C.; Kosynkin, D.V.; Berlin, J.M.; Sinitskii, A.; Sun, Z.; Slesarev, A.; Alemany, L.B.; Lu, W.; Tour, J.M. Improved Synthesis of Graphene Oxide. ACS Nano 2010, 4, 4806–4814. [Google Scholar] [CrossRef]
  70. Higginbotham, A.L.; Kosynkin, D.V.; Sinitskii, A.; Sun, Z.; Tour, J.M. Lower-Defect Graphene Oxide Nanoribbons from Multiwalled Carbon Nanotubes. ACS Nano 2010, 4, 2059–2069. [Google Scholar] [CrossRef]
  71. Liou, Y.J.; Tsai, B.D.; Huang, W.J. An Economic Route to Mass Production of Graphene Oxide Solution for Preparing Graphene Oxide Papers. Mater. Sci. Eng. B 2015, 193, 37–40. [Google Scholar] [CrossRef]
  72. Hu, Y.; Song, S.; Lopez-Valdivieso, A. Effects of Oxidation on the Defect of Reduced Graphene Oxides in Graphene Preparation. J. Colloid Interface Sci. 2015, 450, 68–73. [Google Scholar] [CrossRef]
  73. Sun, J.; Yang, N.; Sun, Z.; Zeng, M.; Fu, L.; Hu, C.; Hu, S. Fully Converting Graphite into Graphene Oxide Hydrogels by Preoxidation with Impure Manganese Dioxide. ACS Appl. Mater. Interfaces 2015, 7, 21356–21363. [Google Scholar] [CrossRef]
  74. Peng, L.; Xu, Z.; Liu, Z.; Wei, Y.; Sun, H.; Li, Z.; Zhao, X.; Gao, C. An Iron-Based Green Approach to 1-h Production of Single-Layer Graphene Oxide. Nat. Commun. 2015, 6, 5716. [Google Scholar] [CrossRef]
  75. Yu, C.; Wang, C.F.; Chen, S. Facile Access to Graphene Oxide from Ferro-Induced Oxidation. Sci. Rep. 2016, 6, 17071. [Google Scholar] [CrossRef]
  76. Yu, H.; Zhang, B.; Bulin, C.; Li, R.; Xing, R. High-Efficient Synthesis of Graphene Oxide Based on Improved Hummers Method. Sci. Rep. 2016, 6, 36143. [Google Scholar] [CrossRef]
  77. Eda, G.; Chhowalla, M. Chemically Derived Graphene Oxide: Towards Large-Area Thin-Film Electronics and Optoelectronics. Adv. Mater. 2010, 22, 2392–2415. [Google Scholar] [CrossRef]
  78. Si, Y.; Samulski, E.T. Synthesis of Water Soluble Graphene. Nano Lett. 2008, 8, 1679–1682. [Google Scholar] [CrossRef]
  79. Wan, X.; Huang, Y.; Chen, Y. Focusing on Energy and Optoelectronic Applications: A Journey for Graphene and Graphene Oxide at Large Scale. Acc. Chem. Res. 2012, 45, 598–607. [Google Scholar] [CrossRef]
  80. Chua, C.K.; Pumera, M. Chemical Reduction of Graphene Oxide: A Synthetic Chemistry Viewpoint. Chem. Soc. Rev. 2014, 43, 291–312. [Google Scholar] [CrossRef]
  81. Hofmann, U.; Frenzel, A. Die Reduktion von Graphitoxyd mit Schwefelwasserstoff. Kolloid Z. 1934, 68, 149–151. [Google Scholar] [CrossRef]
  82. De Silva, K.K.H.; Huang, H.H.; Joshi, R.K.; Yoshimura, M. Chemical Reduction of Graphene Oxide Using Green Reductants. Carbon 2017, 119, 190–199. [Google Scholar] [CrossRef]
  83. Stankovich, S.; Dikin, D.A.; Piner, R.D.; Kohlhaas, K.A.; Kleinhammes, A.; Jia, Y.; Wu, Y.; Nguyen, S.T.; Ruoff, R.S. Synthesis of Graphene-Based Nanosheets via Chemical Reduction of Exfoliated Graphite Oxide. Carbon 2007, 45, 1558–1565. [Google Scholar] [CrossRef]
  84. Xu, Y.; Sheng, K.; Li, C.; Shi, G. Highly Conductive Chemically Converted Graphene Prepared from Mildly Oxidized Graphene Oxide. J. Mater. Chem. 2011, 21, 7376–7380. [Google Scholar] [CrossRef]
  85. Park, S.; An, J.; Potts, J.R.; Velamakanni, A.; Murali, S.; Ruoff, R.S. Hydrazine-Reduction of Graphite- and Graphene Oxide. Carbon 2011, 49, 3019–3023. [Google Scholar] [CrossRef]
  86. Zhang, C.; Lv, W.; Zhang, W.; Zheng, X.; Wu, M.B.; Wei, W.; Tao, Y.; Li, Z.; Yang, Q.H. Reduction of Graphene Oxide by Hydrogen Sulfide: A Promising Strategy for Pollutant Control and as an Electrode for Li-s Batteries. Adv. Energy Mater. 2014, 4, 1301565. [Google Scholar] [CrossRef]
  87. Shin, H.J.; Kim, K.K.; Benayad, A.; Yoon, S.M.; Park, H.K.; Jung, I.S.; Jin, M.H.; Jeong, H.K.; Kim, J.M.; Choi, J.Y.; et al. Efficient Reduction of Graphite Oxide by Sodium Borohydride and Its Effect on Electrical Conductance. Adv. Funct. Mater. 2009, 19, 1987–1992. [Google Scholar] [CrossRef]
  88. Chua, C.K.; Pumera, M. Reduction of Graphene Oxide with Substituted Borohydrides. J. Mater. Chem. A Mater. 2013, 1, 1892–1898. [Google Scholar] [CrossRef]
  89. Chen, Y.; Zhang, X.; Zhang, D.; Yu, P.; Ma, Y. High Performance Supercapacitors Based on Reduced Graphene Oxide in Aqueous and Ionic Liquid Electrolytes. Carbon 2011, 49, 573–580. [Google Scholar] [CrossRef]
  90. Cui, P.; Lee, J.; Hwang, E.; Lee, H. One-Pot Reduction of Graphene Oxide at Subzero Temperatures. Chem. Commun. 2011, 47, 12370–12372. [Google Scholar] [CrossRef]
  91. Pei, S.; Zhao, J.; Du, J.; Ren, W.; Cheng, H.M. Direct Reduction of Graphene Oxide Films into Highly Conductive and Flexible Graphene Films by Hydrohalic Acids. Carbon 2010, 48, 4466–4474. [Google Scholar] [CrossRef]
  92. Zhao, J.; Pei, S.; Ren, W.; Gao, L.; Cheng, H.M. Efficient Preparation of Large-Area Graphene Oxide Sheets for Transparent Conductive Films. ACS Nano 2010, 4, 5245–5252. [Google Scholar] [CrossRef] [PubMed]
  93. Fan, X.; Peng, W.; Li, Y.; Li, X.; Wang, S.; Zhang, G.; Zhang, F. Deoxygenation of Exfoliated Graphite Oxide Under Alkaline Conditions: A Green Route to Graphene Preparation. Adv. Mater. 2008, 20, 4490–4493. [Google Scholar] [CrossRef]
  94. Chen, W.; Yan, L.; Bangal, P.R. Chemical Reduction of Graphene Oxide to Graphene by Sulfur-Containing Compounds. J. Phys. Chem. C 2010, 114, 19885–19890. [Google Scholar] [CrossRef]
  95. Zhou, T.; Chen, F.; Liu, K.; Deng, H.; Zhang, Q.; Feng, J.; Fu, Q. A Simple and Efficient Method to Prepare Graphene by Reduction of Graphite Oxide with Sodium Hydrosulfite. Nanotechnology 2011, 22, 045704. [Google Scholar] [CrossRef]
  96. Zhu, C.; Guo, S.; Fang, Y.; Dong, S. Reducing Sugar: New Functional Molecules for the Green Synthesis of Graphene Nanosheets. ACS Nano 2010, 4, 2429–2437. [Google Scholar] [CrossRef]
  97. Kim, Y.K.; Kim, M.H.; Min, D.H. Biocompatible Reduced Graphene Oxide Prepared by Using Dextran as a Multifunctional Reducing Agent. Chem. Commun. 2011, 47, 3195–3197. [Google Scholar] [CrossRef]
  98. Xu, C.; Shi, X.; Ji, A.; Shi, L.; Zhou, C.; Cui, Y. Fabrication and Characteristics of Reduced Graphene Oxide Produced with Different Green Reductants. PLoS ONE 2015, 10, e0144842. [Google Scholar] [CrossRef]
  99. Fernández-Merino, M.J.; Guardia, L.; Paredes, J.I.; Villar-Rodil, S.; Solís-Fernández, P.; Martínez-Alonso, A.; Tascón, J.M.D. Vitamin C Is an Ideal Substitute for Hydrazine in the Reduction of Graphene Oxide Suspensions. J. Phys. Chem. C 2010, 114, 6426–6432. [Google Scholar] [CrossRef]
  100. Zhang, J.; Yang, H.; Shen, G.; Cheng, P.; Zhang, J.; Guo, S. Reduction of Graphene Oxide Vial-Ascorbic Acid. Chem. Commun. 2010, 46, 1112–1114. [Google Scholar] [CrossRef]
  101. Akhavan, O.; Ghaderi, E. Escherichia coli Bacteria Reduce Graphene Oxide to Bactericidal Graphene in a Self-Limiting Manner. Carbon 2012, 50, 1853–1860. [Google Scholar] [CrossRef]
  102. Salas, E.C.; Sun, Z.; Lüttge, A.; Tour, J.M. Reduction of Graphene Oxide via Bacterial Respiration. ACS Nano 2010, 4, 4852–4856. [Google Scholar] [CrossRef] [PubMed]
  103. Chen, D.; Li, L.; Guo, L. An Environment-Friendly Preparation of Reduced Graphene Oxide Nanosheets via Amino Acid. Nanotechnology 2011, 22, 325601. [Google Scholar] [CrossRef]
  104. Bose, S.; Kuila, T.; Mishra, A.K.; Kim, N.H.; Lee, J.H. Dual Role of Glycine as a Chemical Functionalizer and a Reducing Agent in the Preparation of Graphene: An Environmentally Friendly Method. J. Mater. Chem. 2012, 22, 9696–9703. [Google Scholar] [CrossRef]
  105. Ma, J.; Wang, X.; Wu, T.; Liu, Y.; Guo, Y.; Li, R.; Sun, X.; Wu, F.; Li, C.; Gao, J.; et al. Reduction of Graphene Oxide with L-Lysine to Prepare Reduced Graphene Oxide Stabilized with Polysaccharide Polyelectrolyte. J. Mater. Chem. A Mater. 2013, 1, 2192–2201. [Google Scholar] [CrossRef]
  106. Yang, L.; Wang, T.; Wu, D. Porous Nitrogen-Doped Reduced Graphene Oxide Gels as Efficient Supercapacitor Electrodes and Oxygen Reduction Reaction Electrocatalysts. Chin. J. Chem. 2020, 38, 1123–1131. [Google Scholar] [CrossRef]
  107. Wang, Y.; Shi, Z.; Yin, J. Facile Synthesis of Soluble Graphene via a Green Reduction of Graphene Oxide in Tea Solution and Its Biocomposites. ACS Appl. Mater. Interfaces 2011, 3, 1127–1133. [Google Scholar] [CrossRef]
  108. Lei, Y.; Tang, Z.; Liao, R.; Guo, B. Hydrolysable Tannin as Environmentally Friendly Reducer and Stabilizer for Graphene Oxide. Green Chem. 2011, 13, 1655–1658. [Google Scholar] [CrossRef]
  109. Tavakoli, F.; Salavati-Niasari, M.; Badiei, A.; Mohandes, F. Green Synthesis and Characterization of Graphene Nanosheets. Mater. Res. Bull. 2015, 63, 51–57. [Google Scholar] [CrossRef]
  110. Zikalala, N.E.; Azizi, S.; Mpeta, L.S.; Ahmed, R.; Dube, A.; Mketo, N.; Zinatizadeh, A.A.; Mokrani, T.; Maaza, M.M. Green Synthesis of Reduced Graphene Oxide Using Persea americana Mill. Extract: Characterization, Oxygen Reduction Reaction and Antibacterial Application. Diam. Relat. Mater. 2024, 149, 13. [Google Scholar] [CrossRef]
  111. Park, S.; Ruoff, R.S. Chemical Methods for the Production of Graphenes. Nat. Nanotechnol. 2009, 4, 217–224. [Google Scholar] [CrossRef] [PubMed]
  112. Sasikala, S.P.; Poulin, P.; Aymonier, C. Advances in Subcritical Hydro-/Solvothermal Processing of Graphene Materials. Adv. Mater. 2017, 29, 1605473. [Google Scholar] [CrossRef] [PubMed]
  113. Bosch-Navarro, C.; Coronado, E.; Martí-Gastaldo, C.; Sánchez-Royo, J.F.; Gómez, M.G. Influence of the PH on the Synthesis of Reduced Graphene Oxide Under Hydrothermal Conditions. Nanoscale 2012, 4, 3977–3982. [Google Scholar] [CrossRef]
  114. Ding, J.N.; Liu, Y.B.; Yuan, N.Y.; Ding, G.Q.; Fan, Y.; Yu, C.T. The Influence of Temperature, Time and Concentration on the Dispersion of Reduced Graphene Oxide Prepared by Hydrothermal Reduction. Diam. Relat. Mater. 2012, 21, 11–15. [Google Scholar] [CrossRef]
  115. Liu, Z.; Duan, X.; Qian, G.; Zhou, X.; Yuan, W. Eco-Friendly One-Pot Synthesis of Highly Dispersible Functionalized Graphene Nanosheets with Free Amino Groups. Nanotechnology 2013, 24, 045609. [Google Scholar] [CrossRef]
  116. Hong, Z.; Jiao, X.; Gao, Q.; Zhang, Y.; He, C.; Yang, J.; Liu, Y. Tuning the Band Gap of Stable and Dispersible Graphene Aqueous Solution via Hydrothermal Reduction Method. Integr. Ferroelectr. 2013, 145, 115–121. [Google Scholar] [CrossRef]
  117. Díez, N.; ͆liwak, A.; Gryglewicz, S.; Grzyb, B.; Gryglewicz, G. Enhanced Reduction of Graphene Oxide by High-Pressure Hydrothermal Treatment. RSC Adv. 2015, 5, 81831–81837. [Google Scholar] [CrossRef]
  118. Johra, F.T.; Jung, W.G. Hydrothermally Reduced Graphene Oxide as a Supercapacitor. Appl. Surf. Sci. 2015, 357, 1911–1914. [Google Scholar] [CrossRef]
  119. Mei, X.; Meng, X.; Wu, F. Hydrothermal Method for the Production of Reduced Graphene Oxide. Phys. E Low Dimens. Syst. Nanostruct. 2015, 68, 81–86. [Google Scholar] [CrossRef]
  120. Hayes, W.I.; Joseph, P.; Mughal, M.Z.; Papakonstantinou, P. Production of Reduced Graphene Oxide via Hydrothermal Reduction in an Aqueous Sulphuric Acid Suspension and Its Electrochemical Behaviour. J. Solid State Electrochem. 2015, 19, 361–380. [Google Scholar] [CrossRef]
  121. Long, D.; Li, W.; Ling, L.; Miyawaki, J.; Mochida, I.; Yoon, S.H. Preparation of Nitrogen-Doped Graphene Sheets by a Combined Chemical and Hydrothermal Reduction of Graphene Oxide. Langmuir 2010, 26, 16096–16102. [Google Scholar] [CrossRef] [PubMed]
  122. Wang, X.; Zhi, L.; Müllen, K. Transparent, Conductive Graphene Electrodes for Dye-Sensitized Solar Cells. Nano Lett. 2008, 8, 323–327. [Google Scholar] [CrossRef]
  123. Becerril, H.A.; Mao, J.; Liu, Z.; Stoltenberg, R.M.; Bao, Z.; Chen, Y. Evaluation of Solution-Processed Reduced Graphene Oxide Films as Transparent Conductors. ACS Nano 2008, 2, 463–470. [Google Scholar] [CrossRef]
  124. Acik, M.; Lee, G.; Mattevi, C.; Chhowalla, M.; Cho, K.; Chabal, Y.J. Unusual Infrared-Absorption Mechanism in Thermally Reduced Graphene Oxide. Nat. Mater. 2010, 9, 840–845. [Google Scholar] [CrossRef] [PubMed]
  125. Qiu, Y.; Guo, F.; Hurt, R.; Külaots, I. Explosive Thermal Reduction of Graphene Oxide-Based Materials: Mechanism and Safety Implications. Carbon 2014, 72, 215–223. [Google Scholar] [CrossRef]
  126. Zhang, H.B.; Wang, J.W.; Yan, Q.; Zheng, W.G.; Chen, C.; Yu, Z.Z. Vacuum-Assisted Synthesis of Graphene from Thermal Exfoliation and Reduction of Graphite Oxide. J. Mater. Chem. 2011, 21, 5392–5397. [Google Scholar] [CrossRef]
  127. Sheng, Z.H.; Shao, L.; Chen, J.J.; Bao, W.J.; Wang, F.B.; Xia, X.H. Catalyst-Free Synthesis of Nitrogen-Doped Graphene via Thermal Annealing Graphite Oxide with Melamine and Its Excellent Electrocatalysis. ACS Nano 2011, 5, 4350–4358. [Google Scholar] [CrossRef]
  128. Jiao, Y.; Zheng, Y.; Jaroniec, M.; Qiao, S.Z. Origin of the Electrocatalytic Oxygen Reduction Activity of Graphene-Based Catalysts: A Roadnnap to Achieve the Best Performance. J. Am. Chem. Soc. 2014, 136, 4394–4403. [Google Scholar] [CrossRef]
  129. Jin, Z.P.; Nie, H.G.; Yang, Z.; Zhang, J.; Liu, Z.; Xu, X.J.; Huang, S.M. Metal-Free Selenium Doped Carbon Nanotube/Graphene Networks as a Synergistically Improved Cathode Catalyst for Oxygen Reduction Reaction. Nanoscale 2012, 4, 6455–6460. [Google Scholar] [CrossRef]
  130. Wang, Z.; Chen, Y.; Li, P.; He, J.; Zhang, W.; Guo, Z.; Li, Y.; Dong, M. Synthesis of Silicon-Doped Reduced Graphene Oxide and Its Applications in Dye-Sensitive Solar Cells and Supercapacitors. RSC Adv. 2016, 6, 15080–15086. [Google Scholar] [CrossRef]
  131. Yao, Z.; Nie, H.G.; Yang, Z.; Zhou, X.M.; Liu, Z.; Huang, S.M. Catalyst-Free Synthesis of Iodine-Doped Graphene via a Facile Thermal Annealing Process and Its Use for Electrocatalytic Oxygen Reduction in an Alkaline Medium. Chem. Commun. 2012, 48, 1027–1029. [Google Scholar] [CrossRef] [PubMed]
  132. Arora, N.; Sharma, N.N. Arc Discharge Synthesis of Carbon Nanotubes: Comprehensive Review. Diam. Relat. Mater. 2014, 50, 135–150. [Google Scholar] [CrossRef]
  133. Dallas, P.; Meysami, S.S.; Grobert, N.; Porfyrakis, K. Classification of Carbon Nanostructure Families Occurring in a Chemically Activated Arc Discharge Reaction. RSC Adv. 2016, 6, 24912–24920. [Google Scholar] [CrossRef]
  134. Zheng, Y.; Zhu, P. Carbon Nano-Onions: Large-Scale Preparation, Functionalization and Their Application as Anode Material for Rechargeable Lithium Ion Batteries. RSC Adv. 2016, 6, 92285–92298. [Google Scholar] [CrossRef]
  135. Wu, Y.; Wang, B.; Ma, Y.; Huang, Y.; Li, N.; Zhang, F.; Chen, Y. Efficient and Large-Scale Synthesis of Few-Layered Graphene Using an Arc-Discharge Method and Conductivity Studies of the Resulting Films. Nano Res. 2010, 3, 661–669. [Google Scholar] [CrossRef]
  136. Veena, C.; Singh, B.P.; Mathur, R.B. Carbon Nanotubes and Their Composites. In Syntheses and Applications of Carbon Nanotubes and Their Composites; Suzuki, S., Ed.; Intechopen: Rijeka, Croatia, 2013; pp. 193–222. [Google Scholar]
  137. Kumar, R.; Singh, R.K.; Dubey, P.K.; Kumar, P.; Tiwari, R.S.; Oh, I.K. Pressure-Dependent Synthesis of High-Quality Few-Layer Graphene by Plasma-Enhanced Arc Discharge and Their Thermal Stability. J. Nanoparticle Res. 2013, 15, 1847. [Google Scholar] [CrossRef]
  138. Karmakar, S.; Nawale, A.B.; Lalla, N.P.; Sathe, V.G.; Kolekar, S.K.; Mathe, V.L.; Das, A.K.; Bhoraskar, S.V. Gas Phase Condensation of Few-Layer Graphene with Rotational Stacking Faults in an Electric-Arc. Carbon 2013, 55, 209–220. [Google Scholar] [CrossRef]
  139. Wang, Z.; Li, N.; Shi, Z.; Gu, Z. Low-Cost and Large-Scale Synthesis of Graphene Nanosheets by Arc Discharge in Air. Nanotechnology 2010, 21, 175602. [Google Scholar] [CrossRef]
  140. Li, N.; Wang, Z.; Zhao, K.; Shi, Z.; Gu, Z.; Xu, S. Large Scale Synthesis of N-Doped Multi-Layered Graphene Sheets by Simple Arc-Discharge Method. Carbon 2010, 48, 255–259. [Google Scholar] [CrossRef]
  141. Qin, B.; Zhang, T.; Chen, H.; Ma, Y. The Growth Mechanism of Few-Layer Graphene in the Arc Discharge Process. Carbon 2016, 102, 494–498. [Google Scholar] [CrossRef]
  142. Shen, B.; Ding, J.; Yan, X.; Feng, W.; Li, J.; Xue, Q. Influence of Different Buffer Gases on Synthesis of Few-Layered Graphene by Arc Discharge Method. Appl. Surf. Sci. 2012, 258, 4523–4531. [Google Scholar] [CrossRef]
  143. Chen, Y.; Zhao, H.; Sheng, L.; Yu, L.; An, K.; Xu, J.; Ando, Y.; Zhao, X. Mass-Production of Highly-Crystalline Few-Layer Graphene Sheets by Arc Discharge in Various H 2-Inert Gas Mixtures. Chem. Phys. Lett. 2012, 538, 72–76. [Google Scholar] [CrossRef]
  144. Wu, X.; Liu, Y.; Yang, H.; Shi, Z. Large-Scale Synthesis of High-Quality Graphene Sheets by an Improved Alternating Current Arc-Discharge Method. RSC Adv. 2016, 6, 93119–93124. [Google Scholar] [CrossRef]
  145. Subrahmanyam, K.S.; Panchakarla, L.S.; Govindaraj, A.; Rao, C.N.R. Simple Method of Preparing Graphene Flakes by an Arc-Discharge Method. J. Phys. Chem. C 2009, 113, 4257–4259. [Google Scholar] [CrossRef]
  146. Wu, Z.S.; Ren, W.; Gao, L.; Zhao, J.; Chen, Z.; Liu, B.; Tang, D.; Yu, B.; Jiang, C.; Cheng, H.M. Synthesis of Graphene Sheets with High Electrical Conductivity and Good Thermal Stability by Hydrogen Arc Discharge Exfoliation. ACS Nano 2009, 3, 411–417. [Google Scholar] [CrossRef]
  147. Li, B.; Song, X.; Zhang, P. Raman-Assessed Structural Evolution of as-Deposited Few-Layer Graphene by He/H2 Arc Discharge during Rapid-Cooling Thinning Treatment. Carbon 2014, 66, 426–435. [Google Scholar] [CrossRef]
  148. Compagnini, G.; Sinatra, M.; Russo, P.; Messina, G.C.; Puglisi, O.; Scalese, S. Deposition of Few Layer Graphene Nanowalls at the Electrodes during Electric Field-Assisted Laser Ablation of Carbon in Water. Carbon 2012, 50, 2362–2365. [Google Scholar] [CrossRef]
  149. Kim, S.; Song, Y.; Wright, J.; Heller, M.J. Graphene Bi- and Trilayers Produced by a Novel Aqueous Arc Discharge Process. Carbon 2016, 102, 339–345. [Google Scholar] [CrossRef]
  150. Li, Y.; Chen, Q.; Xu, K.; Kaneko, T.; Hatakeyama, R. Synthesis of Graphene Nanosheets from Petroleum Asphalt by Pulsed Arc Discharge in Water. Chem. Eng. J. 2013, 215–216, 45–49. [Google Scholar] [CrossRef]
  151. Seo, D.H.; Rider, A.E.; Kumar, S.; Randeniya, L.K.; Ostrikov, K. Vertical Graphene Gas- and Bio-Sensors via Catalyst-Free, Reactive Plasma Reforming of Natural Honey. Carbon 2013, 60, 221–228. [Google Scholar] [CrossRef]
  152. Yu, H.; Wang, L.; Li, J.; Jia, D. To Promote the Nucleation and Growth of Graphene in Arc Discharge Process by Magnetic Field and H2. Mater. Lett. 2015, 159, 43–46. [Google Scholar] [CrossRef]
  153. Roslan, M.S.; Chaudary, K.T.; Haider, Z.; Zin, A.F.M.; Ali, J. Effect of Magnetic Field on Carbon Nanotubes and Graphene Structure Synthesized at Low Pressure via Arc Discharge Process. In Proceedings of the International Conference on Plasma Science and Applications (ICPSA 2016), Langkawi, Malaysia, 28–30 November 2016; Chaudhary, K.T., Ali, J., Rizvi, S.Z.H., Poznanski, R.R., Eds.; American Institute of Physics Inc.: College Park, MD, USA, 2017; Volume 1824. [Google Scholar]
  154. Huang, L.; Wu, B.; Chen, J.; Xue, Y.; Geng, D.; Guo, Y.; Yu, G.; Liu, Y. Gram-Scale Synthesis of Graphene Sheets by a Catalytic Arc-Discharge Method. Small 2013, 9, 1330–1335. [Google Scholar] [CrossRef] [PubMed]
  155. Poorali, M.S.; Bagheri-Mohagheghi, M.M. Synthesis and Physical Properties of Multi-Layered Graphene Sheets by Arc-Discharge Method with TiO2 and ZnO Catalytic. J. Mater. Sci. Mater. Electron. 2017, 28, 6186–6193. [Google Scholar] [CrossRef]
  156. Panchakarla, L.S.; Subrahmanyam, K.S.; Saha, S.K.; Govindaraj, A.; Krishnamurthy, H.R.; Waghmare, U.V.; Rao, C.N.R. Synthesis, Structure, and Properties of Boron- and Nitrogen-Doped Graphene. Adv. Mater. 2009, 21, 4726–4730. [Google Scholar] [CrossRef]
  157. Shen, B.; Chen, J.; Yan, X.; Xue, Q. Synthesis of Fluorine-Doped Multi-Layered Graphene Sheets by Arc-Discharge. RSC Adv. 2012, 2, 6761–6764. [Google Scholar] [CrossRef]
  158. Shah, S.A.; Cui, L.; Lin, K.; Xue, T.; Guo, Q.; Li, L.; Zhang, L.; Zhang, F.; Hu, F.; Wang, X.; et al. Preparation of Novel Silicon/Nitrogen-Doped Graphene Composite Nanosheets by DC Arc Discharge. RSC Adv. 2015, 5, 29230–29237. [Google Scholar] [CrossRef]
  159. Qing, F.; Jia, R.; Li, B.W.; Liu, C.; Li, C.; Peng, B.; Deng, L.; Zhang, W.; Li, Y.; Ruoff, R.S.; et al. Graphene Growth with “no” Feedstock. 2D Mater. 2017, 4, 025089. [Google Scholar] [CrossRef]
  160. Kim, K.S.; Zhao, Y.; Jang, H.; Lee, S.Y.; Kim, J.M.; Kim, K.S.; Ahn, J.H.; Kim, P.; Choi, J.Y.; Hong, B.H. Large-Scale Pattern Growth of Graphene Films for Stretchable Transparent Electrodes. Nature 2009, 457, 706–710. [Google Scholar] [CrossRef]
  161. Iwasaki, T.; Park, H.J.; Konuma, M.; Lee, D.S.; Smet, J.H.; Starke, U. Long-Range Ordered Single-Crystal Graphene on High-Quality Heteroepitaxial Ni Thin Films Grown on MgO(111). Nano Lett. 2011, 11, 79–84. [Google Scholar] [CrossRef]
  162. Lin, X.; Liu, P.; Wei, Y.; Li, Q.; Wang, J.; Wu, Y.; Feng, C.; Zhang, L.; Fan, S.; Jiang, K. Development of an Ultra-Thin Film Comprised of a Graphene Membrane and Carbon Nanotube Vein Support. Nat. Commun. 2013, 4, 2920. [Google Scholar] [CrossRef]
  163. Li, X.; Cai, W.; An, J.; Kim, S.; Nah, J.; Yang, D.; Piner, R.; Velamakanni, A.; Jung, I.; Tutuc, E.; et al. Large-Area Synthesis of High-Quality and Uniform Graphene Films on Copper Foils. Science 2009, 324, 1312–1314. [Google Scholar] [CrossRef] [PubMed]
  164. Ding, D.; Solís-Fernández, P.; Yunus, R.M.; Hibino, H.; Ago, H. Behavior and Role of Superficial Oxygen in Cu for the Growth of Large Single-Crystalline Graphene. Appl. Surf. Sci. 2017, 408, 142–149. [Google Scholar] [CrossRef]
  165. Marchena, M.; Song, Z.; Senaratne, W.; Li, C.; Liu, X.; Baker, D.; Ferrer, J.C.; Mazumder, P.; Soni, K.; Lee, R.; et al. Direct Growth of 2D and 3D Graphene Nano-Structures over Large Glass Substrates by Tuning a Sacrificial Culate Layer. 2D Mater. 2017, 4, 025088. [Google Scholar] [CrossRef]
  166. Yan, Z.; Lin, J.; Peng, Z.; Sun, Z.; Zhu, Y.; Li, L.; Xiang, C.; Samuel, E.L.; Kittrell, C.; Tour, J.M. Toward the Synthesis of Wafer-Scale Single-Crystal Graphene on Copper Foils. ACS Nano 2012, 6, 9110–9117. [Google Scholar] [CrossRef]
  167. Ma, T.; Ren, W.; Liu, Z.; Huang, L.; Ma, L.P.; Ma, X.; Zhang, Z.; Peng, L.M.; Cheng, H.M. Repeated Growth-Etching-Regrowth for Large-Area Defect-Free Single-Crystal Graphene by Chemical Vapor Deposition. ACS Nano 2014, 8, 12806–12813. [Google Scholar] [CrossRef]
  168. Zhou, H.; Yu, W.J.; Liu, L.; Cheng, R.; Chen, Y.; Huang, X.; Liu, Y.; Wang, Y.; Huang, Y.; Duan, X. Chemical Vapour Deposition Growth of Large Single Crystals of Monolayer and Bilayer Graphene. Nat. Commun. 2013, 4, 2096. [Google Scholar] [CrossRef]
  169. Gan, L.; Luo, Z. Turning off Hydrogen to Realize Seeded Growth of Subcentimeter Single-Crystal Graphene Grains on Copper. ACS Nano 2013, 7, 9480–9488. [Google Scholar] [CrossRef]
  170. Hao, Y.; Bharathi, M.S.; Wang, L.; Liu, Y.; Chen, H.; Nie, S.; Wang, X.; Chou, H.; Tan, C.; Fallahazad, B.; et al. The Role of Surface Oxygen in the Growth of Large Single-Crystal Graphene on Copper. Science 2013, 342, 720–723. [Google Scholar] [CrossRef]
  171. Li, J.; Wang, X.Y.; Liu, X.R.; Jin, Z.; Wang, D.; Wan, L.J. Facile Growth of Centimeter-Sized Single-Crystal Graphene on Copper Foil at Atmospheric Pressure. J. Mater. Chem. C Mater. 2015, 3, 3530–3535. [Google Scholar] [CrossRef]
  172. Gottardi, S.; Müller, K.; Bignardi, L.; Moreno-López, J.C.; Pham, T.A.; Ivashenko, O.; Yablonskikh, M.; Barinov, A.; Björk, J.; Rudolf, P.; et al. Comparing Graphene Growth on Cu(111) versus Oxidized Cu(111). Nano Lett. 2015, 15, 917–922. [Google Scholar] [CrossRef]
  173. Magnuson, C.W.; Kong, X.; Ji, H.; Tan, C.; Li, H.; Piner, R.; Ventrice, C.A., Jr.; Ruoff, R.S. Copper Oxide as a “Self-Cleaning” Substrate for Graphene Growth. J. Mater. Res. 2014, 29, 403–409. [Google Scholar] [CrossRef]
  174. Guo, W.; Jing, F.; Xiao, J.; Zhou, C.; Lin, Y.; Wang, S. Oxidative-Etching-Assisted Synthesis of Centimeter-Sized Single-Crystalline Graphene. Adv. Mater. 2016, 28, 3152–3158. [Google Scholar] [CrossRef] [PubMed]
  175. Li, X.S.; Zhu, Y.W.; Cai, W.W.; Borysiak, M.; Han, B.Y.; Chen, D.; Piner, R.D.; Colombo, L.; Ruoff, R.S. Transfer of Large-Area Graphene Films for High-Performance Transparent Conductive Electrodes. Nano Lett. 2009, 9, 4359–4363. [Google Scholar] [CrossRef]
  176. Lu, Z.J.; Bao, S.J.; Gou, Y.T.; Cai, C.J.; Ji, C.C.; Xu, M.W.; Song, J.; Wang, R.Y. Nitrogen-Doped Reduced-Graphene Oxide as an Efficient Metal-Free Electrocatalyst for Oxygen Reduction in Fuel Cells. RSC Adv. 2013, 3, 3990–3995. [Google Scholar] [CrossRef]
  177. Xue, Y.H.; Yu, D.S.; Dai, L.M.; Wang, R.G.; Li, D.Q.; Roy, A.; Lu, F.; Chen, H.; Liu, Y.; Qu, J. Three-Dimensional B,N-Doped Graphene Foam as a Metal-Free Catalyst for Oxygen Reduction Reaction. Phys. Chem. Chem. Phys. 2013, 15, 12220–12226. [Google Scholar] [CrossRef]
  178. Zhao, L.Y.; He, R.; Rim, K.T.; Schiros, T.; Kim, K.S.; Zhou, H.; Gutierrez, C.; Chockalingam, S.P.; Arguello, C.J.; Palova, L.; et al. Visualizing Individual Nitrogen Dopants in Monolayer Graphene. Science 2011, 333, 999–1003. [Google Scholar] [CrossRef]
  179. Sui, Y.P.; Zhu, B.; Zhang, H.R.; Shu, H.B.; Chen, Z.Y.; Zhang, Y.H.; Zhang, Y.Q.; Wang, B.; Tang, C.M.; Xie, X.M.; et al. Temperature-Dependent Nitrogen Configuration of N-Doped Graphene by Chemical Vapor Deposition. Carbon 2015, 81, 814–820. [Google Scholar] [CrossRef]
  180. Li, J.Y.; Ren, Z.Y.; Zhou, Y.X.; Wu, X.J.; Xu, X.L.; Qi, M.; Li, W.L.; Bai, J.T.; Wang, L. Scalable Synthesis of Pyrrolic N-Doped Graphene by Atmospheric Pressure Chemical Vapor Deposition and Its Terahertz Response. Carbon 2013, 62, 330–336. [Google Scholar] [CrossRef]
  181. Komissarov, I.V.; Kovalchuk, N.G.; Labunov, V.A.; Girel, K.V.; Korolik, O.V.; Tivanov, M.S.; Lazauskas, A.; Andrulevicius, M.; Tamulevicius, T.; Grigaliunas, V.; et al. Nitrogen-Doped Twisted Graphene Grown on Copper by Atmospheric Pressure CVD from a Decane Precursor. Beilstein J. Nanotechnol. 2017, 8, 145–158. [Google Scholar] [CrossRef]
  182. Terasawa, T.; Saiki, K. Synthesis of Nitrogen-Doped Graphene by Plasma-Enhanced Chemical Vapor Deposition. Jpn. J. Appl. Phys. 2012, 51, 4. [Google Scholar] [CrossRef]
  183. Wang, C.D.; Zhou, Y.A.; He, L.F.; Ng, T.W.; Hong, G.; Wu, Q.H.; Gao, F.; Lee, C.S.; Zhang, W.J. In Situ Nitrogen-Doped Graphene Grown from Polydimethylsiloxane by Plasma Enhanced Chemical Vapor Deposition. Nanoscale 2013, 5, 600–605. [Google Scholar] [CrossRef] [PubMed]
  184. Bepete, G.; Voiry, D.; Chhowalla, M.; Chiguvare, Z.; Coville, N.J. Incorporation of Small BN Domains in Graphene during CVD Using Methane, Boric Acid and Nitrogen Gas. Nanoscale 2013, 5, 6552–6557. [Google Scholar] [CrossRef]
  185. Qu, L.T.; Liu, Y.; Baek, J.B.; Dai, L.M. Nitrogen-Doped Graphene as Efficient Metal-Free Electrocatalyst for Oxygen Reduction in Fuel Cells. ACS Nano 2010, 4, 1321–1326. [Google Scholar] [CrossRef]
  186. Reddy, A.L.M.; Srivastava, A.; Gowda, S.R.; Gullapalli, H.; Dubey, M.; Ajayan, P.M. Synthesis of Nitrogen-Doped Graphene Films for Lithium Battery Application. ACS Nano 2010, 4, 6337–6342. [Google Scholar] [CrossRef]
  187. Imamura, G.; Saiki, K. Synthesis of Nitrogen-Doped Graphene on Pt(111) by Chemical Vapor Deposition. J. Phys. Chem. C 2011, 115, 10000–10005. [Google Scholar] [CrossRef]
  188. Jin, Z.; Yao, J.; Kittrell, C.; Tour, J.M. Large-Scale Growth and Characterizations of Nitrogen-Doped Monolayer Graphene Sheets. ACS Nano 2011, 5, 4112–4117. [Google Scholar] [CrossRef]
  189. Xue, Y.Z.; Wu, B.; Jiang, L.; Guo, Y.L.; Huang, L.P.; Chen, J.Y.; Tan, J.H.; Geng, D.C.; Luo, B.R.; Hu, W.P.; et al. Low Temperature Growth of Highly Nitrogen-Doped Single Crystal Graphene Arrays by Chemical Vapor Deposition. J. Am. Chem. Soc. 2012, 134, 11060–11063. [Google Scholar] [CrossRef]
  190. Zabet-Khosousi, A.; Zhao, L.Y.; Palova, L.; Hybertsen, M.S.; Reichman, D.R.; Pasupathy, A.N.; Flynn, G.W. Segregation of Sublattice Domains in Nitrogen-Doped Graphene. J. Am. Chem. Soc. 2014, 136, 1391–1397. [Google Scholar] [CrossRef]
  191. Capasso, A.; Dikonimos, T.; Sarto, F.; Tamburrano, A.; De Bellis, G.; Sarto, M.S.; Faggio, G.; Malara, A.; Messina, G.; Lisi, N. Nitrogen-Doped Graphene Films from Chemical Vapor Deposition of Pyridine: Influence of Process Parameters on the Electrical and Optical Properties. Beilstein J. Nanotechnol. 2015, 6, 2028–2038. [Google Scholar] [CrossRef]
  192. Gao, H.; Song, L.; Guo, W.H.; Huang, L.; Yang, D.Z.; Wang, F.C.; Zuo, Y.L.; Fan, X.L.; Liu, Z.; Gao, W.; et al. A Simple Method to Synthesize Continuous Large Area Nitrogen-Doped Graphene. Carbon 2012, 50, 4476–4482. [Google Scholar] [CrossRef]
  193. Nang, L.V.; Duy, N.V.; Hoa, N.D.; Hieu, N.V. Nitrogen-Doped Graphene Synthesized from a Single Liquid Precursor for a Field Effect Transistor. J. Electron. Mater. 2016, 45, 839–845. [Google Scholar] [CrossRef]
  194. Wu, T.R.; Shen, H.L.; Sun, L.; Cheng, B.; Liu, B.; Shen, J.C. Nitrogen and Boron Doped Monolayer Graphene by Chemical Vapor Deposition Using Polystyrene, Urea and Boric Acid. New J. Chem. 2012, 36, 1385–1391. [Google Scholar] [CrossRef]
  195. Zhang, C.K.; Lin, W.Y.; Zhao, Z.J.; Zhuang, P.P.; Zhan, L.J.; Zhou, Y.H.; Cai, W.W. CVD Synthesis of Nitrogen-Doped Graphene Using Urea. Sci. China Phys. Mech. Astron. 2015, 58, 6. [Google Scholar] [CrossRef]
  196. Bao, J.F.; Kishi, N.; Soga, T. Synthesis of Nitrogen-Doped Graphene by the Thermal Chemical Vapor Deposition Method from a Single Liquid Precursor. Mater. Lett. 2014, 117, 199–203. [Google Scholar] [CrossRef]
  197. Feng, X.M.; Zhang, Y.; Zhou, J.H.; Li, Y.; Chen, S.F.; Zhang, L.; Ma, Y.W.; Wang, L.H.; Yan, X.H. Three-Dimensional Nitrogen-Doped Graphene as an Ultrasensitive Electrochemical Sensor for the Detection of Dopamine. Nanoscale 2015, 7, 2427–2432. [Google Scholar] [CrossRef]
  198. Shinde, S.M.; Kano, E.; Kalita, G.; Takeguchi, M.; Hashimoto, A.; Tanemura, M. Grain Structures of Nitrogen-Doped Graphene Synthesized by Solid Source-Based Chemical Vapor Deposition. Carbon 2016, 96, 448–453. [Google Scholar] [CrossRef]
  199. Liu, Y.; Dai, D.; Jiang, N. Synthesis of the Nitrogen-doped CVD Graphene through Triazine. J. Inorg. Mater. 2017, 32, 517–522. [Google Scholar] [CrossRef]
  200. Cattelan, M.; Agnoli, S.; Favaro, M.; Garoli, D.; Romanato, F.; Meneghetti, M.; Barinov, A.; Dudin, P.; Granozzi, G. Microscopic View on a Chemical Vapor Deposition Route to Boron-Doped Graphene Nanostructures. Chem. Mater. 2013, 25, 1490–1495. [Google Scholar] [CrossRef]
  201. Mastrapa, G.C.; da Costa, M.; Larrude, D.G.; Freire, F.L. Synthesis and Characterization of Graphene Layers Prepared by Low-Pressure Chemical Vapor Deposition Using Triphenylphosphine as Precursor. Mater. Chem. Phys. 2015, 166, 37–41. [Google Scholar] [CrossRef]
  202. Li, X.G.; Qiu, Y.F.; Hu, P.A. Three Dimensional P-Doped Graphene Synthesized by Eco-Friendly Chemical Vapor Deposition for Oxygen Reduction Reactions. J. Nanosci. Nanotechnol. 2016, 16, 6216–6222. [Google Scholar] [CrossRef]
  203. Choi, H.; Jo, H.H.; Hwang, S.; Jeon, M.; Kim, J.H. Synthesis of Sulfur-Doped Graphene by Using Near-Infrared Chemical-Vapor Deposition. J. Korean Phys. Soc. 2016, 68, 1257–1261. [Google Scholar] [CrossRef]
  204. Hassani, F.; Tavakol, H.; Keshavarzipour, F.; Javaheri, A. A Simple Synthesis of Sulfur-Doped Graphene Using Sulfur Powder by Chemical Vapor Deposition. RSC Adv. 2016, 6, 27158–27163. [Google Scholar] [CrossRef]
  205. Deng, D.H.; Pan, X.L.; Yu, L.A.; Cui, Y.; Jiang, Y.P.; Qi, J.; Li, W.X.; Fu, Q.A.; Ma, X.C.; Xue, Q.K.; et al. Toward N-Doped Graphene via Solvothermal Synthesis. Chem. Mater. 2011, 23, 1188–1193. [Google Scholar] [CrossRef]
  206. Lu, X.J.; Wu, J.J.; Lin, T.Q.; Wan, D.Y.; Huang, F.Q.; Xie, X.M.; Jiang, M.H. Low-Temperature Rapid Synthesis of High-Quality Pristine or Boron-Doped Graphene via Wurtz-Type Reductive Coupling Reaction. J. Mater. Chem. 2011, 21, 10685–10689. [Google Scholar] [CrossRef]
  207. Jung, S.M.; Lee, E.K.; Choi, M.; Shin, D.; Jeon, I.Y.; Seo, J.M.; Jeong, H.Y.; Park, N.; Oh, J.H.; Baek, J.B. Direct Solvothermal Synthesis of B/N-Doped Graphene. Angew. Chem. Int. Ed. 2014, 53, 2398–2401. [Google Scholar] [CrossRef]
  208. Ma, R.G.; Xia, B.Y.; Zhou, Y.; Li, P.X.; Chen, Y.F.; Liu, Q.; Wang, J.C. Ionic Liquid-Assisted Synthesis of Dual-Doped Graphene as Efficient Electrocatalysts for Oxygen Reduction. Carbon 2016, 102, 58–65. [Google Scholar] [CrossRef]
  209. Xing, Z.; Ju, Z.C.; Zhao, Y.L.; Wan, J.L.; Zhu, Y.B.; Qiang, Y.H.; Qian, Y.T. One-Pot Hydrothermal Synthesis of Nitrogen-Doped Graphene as High-Performance Anode Materials for Lithium Ion Batteries. Sci. Rep. 2016, 6, 10. [Google Scholar] [CrossRef]
  210. Sinitskii, A. A Recipe for Nanoporous Graphene Nanoporous Graphene Created from Molecular Precursors Shows Promise for Electronic Applications. Science 2018, 360, 154–155. [Google Scholar] [CrossRef]
  211. Moreno, C.; Vilas-Varela, M.; Kretz, B.; Garcia-Lekue, A.; Costache, M.V.; Paradinas, M.; Panighel, M.; Ceballos, G.; Valenzuela, S.O.; Peña, D.; et al. Bottom-up Synthesis of Multifunctional Nanoporous Graphene. Science 2018, 360, 199–203. [Google Scholar] [CrossRef]
  212. Bieri, M.; Treier, M.; Cai, J.M.; Aït-Mansour, K.; Ruffieux, P.; Gröning, O.; Gröning, P.; Kastler, M.; Rieger, R.; Feng, X.L.; et al. Porous Graphenes: Two-Dimensional Polymer Synthesis with Atomic Precision. Chem. Commun. 2009, 45, 6919–6921. [Google Scholar] [CrossRef]
  213. Khatun, S.; Samanta, S.; Sahoo, S.; Mukherjee, I.; Maity, S.; Pradhan, A. Bottom-Up Porous Graphene Synthesis and Its Applications. Chem. Eur. J. 2024, 30, 14. [Google Scholar] [CrossRef]
  214. Talirz, L.; Ruffieux, P.; Fasel, R. On-Surface Synthesis of Atomically Precise Graphene Nanoribbons. Adv. Mater. 2016, 28, 6222–6231. [Google Scholar] [CrossRef]
  215. Gu, Y.W.; Qiu, Z.J.; Müllen, K. Nanographenes and Graphene Nanoribbons as Multitalents of Present and Future Materials Science. J. Am. Chem. Soc. 2022, 144, 11499–11524. [Google Scholar] [CrossRef]
  216. Jacobse, P.H.; McCurdy, R.D.; Jiang, J.W.; Rizzo, D.J.; Veber, G.; Butler, P.; Zuzak, R.; Louie, S.G.; Fischer, F.R.; Crommie, M.F. Bottom-up Assembly of Nanoporous Graphene with Emergent Electronic States. J. Am. Chem. Soc. 2020, 142, 13507–13514. [Google Scholar] [CrossRef]
  217. Ajayakumar, M.R.; Di Giovannantonio, M.; Pignedoli, C.A.; Yang, L.; Ruffieux, P.; Ma, J.; Fasel, R.; Feng, X.L. On-Surface Synthesis of Porous Graphene Nanoribbons Containing Nonplanar 14 Annulene Pores. J. Polym. Sci. 2022, 60, 1912–1917. [Google Scholar] [CrossRef]
  218. Qin, T.C.; Guo, D.Z.; Xiong, J.J.; Li, X.Y.; Hu, L.; Yang, W.S.; Chen, Z.J.; Wu, Y.L.; Ding, H.H.; Hu, J.; et al. Synthesis of a Porous 14 Annulene Graphene Nanoribbon and a Porous 30 Annulene Graphene Nanosheet on Metal Surfaces. Angew. Chem. Int. Ed. 2023, 62, 11. [Google Scholar] [CrossRef]
  219. Lin, J.; Peng, Z.W.; Liu, Y.Y.; Ruiz-Zepeda, F.; Ye, R.Q.; Samuel, E.L.G.; Yacaman, M.J.; Yakobson, B.I.; Tour, J.M. Laser-Induced Porous Graphene Films from Commercial Polymers. Nat. Commun. 2014, 5, 8. [Google Scholar] [CrossRef]
  220. Le, T.S.D.; Phan, H.P.; Kwon, S.; Park, S.; Jung, Y.; Min, J.; Chun, B.J.; Yoon, H.; Ko, S.H.; Kim, S.W.; et al. Recent Advances in Laser-Induced Graphene: Mechanism, Fabrication, Properties, and Applications in Flexible Electronics. Adv. Funct. Mater. 2022, 32, 39. [Google Scholar] [CrossRef]
  221. Alhajji, E.; Zhang, F.; Alshareef, H.N. Status and Prospects of Laser-Induced Graphene for Battery Applications. Energy Technol. 2021, 9, 15. [Google Scholar] [CrossRef]
  222. Avinash, K.; Patolsky, F. Laser-Induced Graphene Structures: From Synthesis and Applications to Future Prospects. Mater. Today 2023, 70, 104–136. [Google Scholar] [CrossRef]
  223. Li, Y.L.; Luong, D.X.; Zhang, J.B.; Tarkunde, Y.R.; Kittrell, C.; Sargunaraj, F.; Ji, Y.S.; Arnusch, C.J.; Tour, J.M. Laser-Induced Graphene in Controlled Atmospheres: From Superhydrophilic to Superhydrophobic Surfaces. Adv. Mater. 2017, 29, 8. [Google Scholar] [CrossRef]
  224. Ye, R.; James, D.K.; Tour, J.M. Laser-Induced Graphene. Acc. Chem. Res. 2018, 51, 1609–1620. [Google Scholar] [CrossRef]
  225. Singh, S.P.; Li, Y.L.; Zhang, J.B.; Tour, J.M.; Arnusch, C.J. Sulfur-Doped Laser-Induced Porous Graphene Derived from Polysulfone-Class Polymers and Membranes. ACS Nano 2018, 12, 289–297. [Google Scholar] [CrossRef]
  226. Chyan, Y.; Ye, R.Q.; Li, Y.L.; Singh, S.P.; Arnusch, C.J.; Tour, J.M. Laser-Induced Graphene by Multiple Lasing: Toward Electronics on Cloth, Paper, and Food. ACS Nano 2018, 12, 2176–2183. [Google Scholar] [CrossRef]
  227. Ye, R.Q.; Chyan, Y.; Zhang, J.B.; Li, Y.L.; Han, X.; Kittrell, C.; Tour, J.M. Laser-Induced Graphene Formation on Wood. Adv. Mater. 2017, 29, 7. [Google Scholar] [CrossRef]
  228. Zhang, H.W.; Li, Q.W.; Hammond, K.D.; He, X.Q.; Lin, J.; Wan, C.X. Probing Laser-Induced Structural Transformation of Lignin into Few-Layer Graphene. Green Chem. 2024, 26, 5921–5932. [Google Scholar] [CrossRef]
  229. Zhang, Z.C.; Song, M.M.; Hao, J.X.; Wu, K.B.; Li, C.Y.; Hu, C.G. Visible Light Laser-Induced Graphene from Phenolic Resin: A New Approach for Directly Writing Graphene-Based Electrochemical Devices on Various Substrates. Carbon 2018, 127, 287–296. [Google Scholar] [CrossRef]
  230. Movaghgharnezhad, S.; Kang, P.L.Y. Laser-Induced Graphene: Synthesis Advances, Structural Tailoring, Enhanced Properties, and Sensing Applications. J. Mater. Chem. C 2024, 12, 6718–6742. [Google Scholar] [CrossRef]
  231. Peng, Z.W.; Ye, R.Q.; Mann, J.A.; Zakhidov, D.; Li, Y.L.; Smalley, P.R.; Lin, J.; Tour, J.M. Flexible Boron-Doped Laser-Induced Graphene Microsupercapacitors. ACS Nano 2015, 9, 5868–5875. [Google Scholar] [CrossRef]
  232. Han, S.; Liu, C.; Li, N.A.; Zhang, S.D.; Song, Y.P.; Chen, L.Q.; Xi, M.; Yu, X.L.; Wang, W.B.; Kong, M.G.; et al. One-Step Fabrication of Nitrogen-Doped Laser-Induced Graphene Derived from Melamine/Polyimide for Enhanced Flexible Supercapacitors. CrystEngComm 2022, 24, 1866–1876. [Google Scholar] [CrossRef]
  233. Yang, W.W.; Liu, Y.; Li, Q.S.; Wei, J.; Li, X.L.; Zhang, Y.; Liu, J.P. In Situ Formation of Phosphorus-Doped Porous Graphene via Laser Induction. RSC Adv. 2020, 10, 23953–23958. [Google Scholar] [CrossRef] [PubMed]
  234. Wang, F.; Wang, F.F.; Hong, R.Y.; Lv, X.S.; Zheng, Y.; Chen, H.Y. High-Purity Few-Layer Graphene from Plasma Pyrolysis of Methane as Conductive Additive for LiFePO4 Lithium Ion Battery. J. Mater. Res. Technol. 2020, 9, 10004–10015. [Google Scholar] [CrossRef]
  235. Tatarova, E.; Dias, A.; Henriques, J.; do Rego, A.M.B.; Ferraria, A.M.; Abrashev, M.V.; Luhrs, C.C.; Phillips, J.; Dias, F.M.; Ferreira, C.M. Microwave Plasmas Applied for the Synthesis of Free Standing Graphene Sheets. J. Phys. D Appl. Phys. 2014, 47, 11. [Google Scholar] [CrossRef]
  236. Zhu, Y.Q.; Cao, T.; Cao, C.B.; Ma, X.L.; Xu, X.Y.; Li, Y.D. A General Synthetic Strategy to Monolayer Graphene. Nano Res. 2018, 11, 3088–3095. [Google Scholar] [CrossRef]
  237. Wu, Y.J.; Yuan, Y.F.; Shuang, W.; Wang, L.G.; Yang, L.; Bai, Z.Y.; Lu, J. Reducing Carbonaceous Salts for Facile Fabrication of Monolayer Graphene. Small Methods 2023, 7, 9. [Google Scholar] [CrossRef]
  238. Speyer, L.; Fontana, S.; Cahen, S.; Hérold, C. Simple Production of High-Quality Graphene Foams by Pyrolysis of Sodium Ethoxide. Mater. Chem. Phys. 2018, 219, 57–66. [Google Scholar] [CrossRef]
  239. Yan, Y.X.; Nashath, F.Z.; Chen, S.R.; Manickam, S.; Lim, S.S.; Zhao, H.T.; Lester, E.; Wu, T.; Pang, C.H. Synthesis of Graphene: Potential Carbon Precursors and Approaches. Nanotechnol. Rev. 2020, 9, 1284–1314. [Google Scholar] [CrossRef]
  240. Pan, F.P.; Jin, J.T.; Fu, X.G.; Liu, Q.; Zhang, J.Y. Advanced Oxygen Reduction Electrocatalyst Based on Nitrogen-Doped Graphene Derived from Edible Sugar and Urea. ACS Appl. Mater. Interfaces 2013, 5, 11108–11114. [Google Scholar] [CrossRef]
  241. Zhao, Y.; Wen, M.Y.; He, C.H.; Liu, C.L.; Li, Z.R.; Liu, Y. Preparation of Graphene by Catalytic Pyrolysis of Lignin and Its Electrochemical Properties. Mater. Lett. 2020, 274, 3. [Google Scholar] [CrossRef]
  242. Prekodravac, J.R.; Kepic, D.P.; Colmenares, J.C.; Giannakoudakis, D.A.; Jovanovic, S.P. A Comprehensive Review on Selected Graphene Synthesis Methods: From Electrochemical Exfoliation through Rapid Thermal Annealing towards Biomass Pyrolysis. J. Mater. Chem. C 2021, 9, 6722–6748. [Google Scholar] [CrossRef]
  243. Roy, A.; Kar, S.; Ghosal, R.; Naskar, K.; Bhowmick, A.K. Facile Synthesis and Characterization of Few-Layer Multifunctional Graphene from Sustainable Precursors by Controlled Pyrolysis, Understanding of the Graphitization Pathway, and Its Potential Application in Polymer Nanocomposites. ACS Omega 2021, 6, 1809–1822. [Google Scholar] [CrossRef] [PubMed]
  244. Gao, K.; Sun, S.; Zhang, B. Recent Advancements in the Development of Graphene-Based Materials for Catalytic Applications. ChemCatChem 2024, 16, 13. [Google Scholar] [CrossRef]
  245. Zhang, B.X.; Zhang, E.H.; Wang, S.Y.; Zhang, Y.Y.; Ma, Z.; Qiu, Y.F. Bifunctional Oxygen Electrocatalyst Derived from Photochlorinated Graphene for Rechargeable Solid-State Zn-Air Battery. J. Colloid. Interface Sci. 2019, 543, 84–95. [Google Scholar] [CrossRef]
  246. Radtke, M.; Ignaszak, A. Grafting of the Carbon Allotropes and Polypyrrole via a Kevlar-Type Organic Linker: The Correlation of Carbon Structure/Morphology with Electrochemistry of the Composite Electrode. Mater. Renew. Sustain. Energy 2017, 6, 1. [Google Scholar] [CrossRef]
  247. Gomez-Navarro, C.; Meyer, J.C.; Sundaram, R.S.; Chuvilin, A.; Kurasch, S.; Burghard, M.; Kern, K.; Kaiser, U. Atomic Structure of Reduced Graphene Oxide. Nano Lett. 2010, 10, 1144–1148. [Google Scholar] [CrossRef]
  248. Girit, C.O.; Meyer, J.C.; Erni, R.; Rossell, M.D.; Kisielowski, C.; Yang, L.; Park, C.H.; Crommie, M.F.; Cohen, M.L.; Louie, S.G.; et al. Graphene at the Edge: Stability and Dynamics. Science 2009, 323, 1705–1708. [Google Scholar] [CrossRef]
  249. Jinschek, J.R.; Yucelen, E.; Calderon, H.A.; Freitag, B. Quantitative Atomic 3D Imaging of Single/Double Sheet Graphene Structure. Carbon 2011, 49, 556–562. [Google Scholar] [CrossRef]
  250. Vicarelli, L.; Heerema, S.J.; Dekker, C.; Zandbergen, H.W. Controlling Defects in Graphene for Optimizing the Electrical Properties of Graphene Nanodevices. ACS Nano 2015, 9, 3428–3435. [Google Scholar] [CrossRef]
  251. Jian, Y.H.; Ding, F.; Yakobson, B.I.; Lu, P.; Qi, L.; Li, J. In Situ Observation of Graphene Sublimation and Multi-Layer Edge Reconstructions. Proc. Natl. Acad. Sci. USA 2009, 106, 10103–10108. [Google Scholar] [CrossRef]
  252. Bachmatiuk, A.; Zhao, J.; Gorantla, S.M.; Martinez, I.G.G.; Wiedermann, J.; Lee, C.; Eckert, J.; Rummeli, M.H. Low Voltage Transmission Electron Microscopy of Graphene. Small 2015, 11, 515–542. [Google Scholar] [CrossRef]
  253. Huang, P.Y.; Ruiz-Vargas, C.S.; Van Der Zande, A.M.; Whitney, W.S.; Levendorf, M.P.; Kevek, J.W.; Garg, S.; Alden, J.S.; Hustedt, C.J.; Zhu, Y.; et al. Grains and Grain Boundaries in Single-Layer Graphene Atomic Patchwork Quilts. Nature 2011, 469, 389–392. [Google Scholar] [CrossRef] [PubMed]
  254. Bangert, U.; Eberlein, T.; Nair, R.R.; Jones, R.; Gass, M.; Bleloch, A.L.; Novoselov, K.S.; Geim, A.; Briddon, P.R. STEM Plasmon Spectroscopy of Free Standing Graphene. Phys. Status Solidi A Appl. Mater. Sci. 2008, 205, 2265–2269. [Google Scholar] [CrossRef]
  255. Zhou, W.; Oxley, M.P.; Lupini, A.R.; Krivanek, O.L.; Pennycook, S.J.; Idrobo, J.C. Single Atom Microscopy. Microsc. Microanal. 2012, 18, 1342–1354. [Google Scholar] [CrossRef] [PubMed]
  256. Eda, G.; Chhowalla, M. Graphene-Based Composite Thin Films for Electronics. Nano Lett. 2009, 9, 814–818. [Google Scholar] [CrossRef]
  257. Paredes, J.I.; Villar-Rodil, S.; Solís-Fernández, P.; Martínez-Alonso, A.; Tascón, J.M.D. Atomic Force and Scanning Tunneling Microscopy Imaging of Graphene Nanosheets Derived from Graphite Oxide. Langmuir 2009, 25, 5957–5968. [Google Scholar] [CrossRef]
  258. Giannazzo, F.; Sonde, S.; Raineri, V.; Patanè, G.; Compagnini, G.; Aliotta, F.; Ponterio, R.; Rimini, E. Optical, Morphological and Spectroscopic Characterization of Graphene on SiO2. Phys. Status Solidi (C) 2010, 7, 1251–1255. [Google Scholar] [CrossRef]
  259. Nemes-Incze, P.; Osváth, Z.; Kamarás, K.; Biró, L.P. Anomalies in Thickness Measurements of Graphene and Few Layer Graphite Crystals by Tapping Mode Atomic Force Microscopy. Carbon 2008, 46, 1435–1442. [Google Scholar] [CrossRef]
  260. Shen, Z.; Li, J.; Yi, M.; Zhang, X.; Ma, S. Preparation of Graphene by Jet Cavitation. Nanotechnology 2011, 22, 365306. [Google Scholar] [CrossRef]
  261. Jang, Y.R.; Kim, K.Y.; Yoo, K.H. Accurate Measurement of Thickness of Large-Area Graphene Layers by Neutron Reflectometry. J. Mater. Sci. 2016, 51, 10059–10065. [Google Scholar] [CrossRef]
  262. Liu, M.; Zhang, R.; Chen, W. Graphene-Supported Nanoelectrocatalysts for Fuel Cells: Synthesis, Properties, and Applications. Chem. Rev. 2014, 114, 5117–5160. [Google Scholar] [CrossRef]
  263. Mkhoyan, K.A.; Contryman, A.W.; Silcox, J.; Stewart, D.A.; Eda, G.; Mattevi, C.; Miller, S.; Chhowalla, M. Atomic and Electronic Structure of Graphene-Oxide. Nano Lett. 2009, 9, 1058–1063. [Google Scholar] [CrossRef] [PubMed]
  264. Hu, T.; Bao, H.; Liu, S.; Liu, X.; Ma, D.; Ma, F.; Xu, K. Near-Free-Standing Epitaxial Graphene on Rough SiC Substrate by Flash Annealing at High Temperature. Carbon 2017, 120, 219–225. [Google Scholar] [CrossRef]
  265. Günther, S.; Dänhardt, S.; Ehrensperger, M.; Zeller, P.; Schmitt, S.; Wintterlin, J. High-Temperature Scanning Tunneling Microscopy Study of the Ordering Transition of an Amorphous Carbon Layer into Graphene on Ruthenium(0001). ACS Nano 2013, 7, 154–164. [Google Scholar] [CrossRef]
  266. Rozada, R.; Paredes, J.I.; López, M.J.; Villar-Rodil, S.; Cabria, I.; Alonso, J.A.; Martínez-Alonso, A.; Tascón, J.M.D. From Graphene Oxide to Pristine Graphene: Revealing the Inner Workings of the Full Structural Restoration. Nanoscale 2015, 7, 2374–2390. [Google Scholar] [CrossRef]
  267. Ferrari, A.C. Raman Spectroscopy of Graphene and Graphite: Disorder, Electron-Phonon Coupling, Doping and Nonadiabatic Effects. Solid State Commun. 2007, 143, 47–57. [Google Scholar] [CrossRef]
  268. Ferrari, A.C.; Meyer, J.C.; Scardaci, V.; Casiraghi, C.; Lazzeri, M.; Mauri, F.; Piscanec, S.; Jiang, D.; Novoselov, K.S.; Roth, S.; et al. Raman Spectrum of Graphene and Graphene Layers. Phys. Rev. Lett. 2006, 97. [Google Scholar] [CrossRef]
  269. Malard, L.M.; Pimenta, M.A.; Dresselhaus, G.; Dresselhaus, M.S. Raman Spectroscopy in Graphene. Phys. Rep. 2009, 473, 51–87. [Google Scholar] [CrossRef]
  270. Cançado, L.G.; Jorio, A.; Ferreira, E.H.M.; Stavale, F.; Achete, C.A.; Capaz, R.B.; Moutinho, M.V.O.; Lombardo, A.; Kulmala, T.S.; Ferrari, A.C. Quantifying Defects in Graphene via Raman Spectroscopy at Different Excitation Energies. Nano Lett. 2011, 11, 3190–3196. [Google Scholar] [CrossRef]
  271. Lucchese, M.M.; Stavale, F.; Ferreira, E.H.M.; Vilani, C.; Moutinho, M.V.O.; Capaz, R.B.; Achete, C.A.; Jorio, A. Quantifying Ion-Induced Defects and Raman Relaxation Length in Graphene. Carbon 2010, 48, 1592–1597. [Google Scholar] [CrossRef]
  272. Cançado, L.G.; Reina, A.; Kong, J.; Dresselhaus, M.S. Geometrical Approach for the Study of G′ Band in the Raman Spectrum of Monolayer Graphene, Bilayer Graphene, and Bulk Graphite. Phys. Rev. B Condens. Matter Mater. Phys. 2008, 77, 245408. [Google Scholar] [CrossRef]
  273. Molina-Garcia, M.A.; Rees, N.V. “Metal-Free” Electrocatalysis: Quaternary-Doped Graphene and the Alkaline Oxygen Reduction Reaction. Appl. Catal. A Gen. 2018, 553, 107–116. [Google Scholar] [CrossRef]
  274. Poh, H.L.; Šaněk, F.; Ambrosi, A.; Zhao, G.; Sofer, Z.; Pumera, M. Graphenes Prepared by Staudenmaier, Hofmann and Hummers Methods with Consequent Thermal Exfoliation Exhibit Very Different Electrochemical Properties. Nanoscale 2012, 4, 3515–3522. [Google Scholar] [CrossRef] [PubMed]
  275. Hu, B.S.; He, X.Y.; Wu, R.; Jin, Y.; Bian, Y.W.; Chen, S.G.; Wei, Z.D. Communication-Controlling Highly Dominated N Configuration in N-Doped Graphene as Oxygen Reduction Catalyst. J. Electrochem. Soc. 2017, 164, F256–F258. [Google Scholar] [CrossRef]
  276. Guo, H.L.; Wang, X.F.; Qian, Q.Y.; Wang, F.B.; Xia, X.H. A Green Approach to the Synthesis of Graphene Nanosheets. ACS Nano 2009, 3, 2653–2659. [Google Scholar] [CrossRef]
  277. Gao, Z.; Yang, W.; Wang, J.; Yan, H.; Yao, Y.; Ma, J.; Wang, B.; Zhang, M.; Liu, L. Electrochemical Synthesis of Layer-by-Layer Reduced Graphene Oxide Sheets/Polyaniline Nanofibers Composite and Its Electrochemical Performance. Electrochim. Acta 2013, 91, 185–194. [Google Scholar] [CrossRef]
  278. Moon, I.K.; Lee, J.; Ruoff, R.S.; Lee, H. Reduced Graphene Oxide by Chemical Graphitization. Nat. Commun. 2010, 1, 73. [Google Scholar] [CrossRef]
  279. Feng, H.; Cheng, R.; Zhao, X.; Duan, X.; Li, J. A Low-Temperature Method to Produce Highly Reduced Graphene Oxide. Nat. Commun. 2013, 4, 1539. [Google Scholar] [CrossRef]
  280. Das, P.; Ibrahim, S.; Chakraborty, K.; Ghosh, S.; Pal, T. Stepwise Reduction of Graphene Oxide and Studies on Defect-Controlled Physical Properties. Sci. Rep. 2024, 14, 10. [Google Scholar] [CrossRef]
  281. Sun, Z.; Yan, Z.; Yao, J.; Beitler, E.; Zhu, Y.; Tour, J.M. Growth of Graphene from Solid Carbon Sources. Nature 2010, 468, 549–552. [Google Scholar] [CrossRef]
  282. Kumar, P.V.; Bardhan, N.M.; Tongay, S.; Wu, J.Q.; Belcher, A.M.; Grossman, J.C. Scalable Enhancement of Graphene Oxide Properties by Thermally Driven Phase Transformation. Nat. Chem. 2014, 6, 151–158. [Google Scholar] [CrossRef]
  283. Li, Y.; Zhao, Y.; Cheng, H.H.; Hu, Y.; Shi, G.Q.; Dai, L.M.; Qu, L.T. Nitrogen-Doped Graphene Quantum Dots with Oxygen-Rich Functional Groups. J. Am. Chem. Soc. 2012, 134, 15–18. [Google Scholar] [CrossRef] [PubMed]
  284. Li, Y.; Hu, Y.; Zhao, Y.; Shi, G.; Deng, L.; Hou, Y.; Qu, L. An Electrochemical Avenue to Green-Luminescent Graphene Quantum Dots as Potential Electron-Acceptors for Photovoltaics. Adv. Mater. 2011, 23, 776–780. [Google Scholar] [CrossRef] [PubMed]
  285. Liang, Y.Y.; Li, Y.G.; Wang, H.L.; Dai, H.J. Strongly Coupled Lnorganic/NanoCarbon Hybrid Materials for Advanced Electrocatalysis. J. Am. Chem. Soc. 2013, 135, 2013–2036. [Google Scholar] [CrossRef]
  286. Liang, X.; Zhong, J.; Shi, Y.; Guo, J.; Huang, G.; Hong, C.; Zhao, Y. Hydrothermal Synthesis of Highly Nitrogen-Doped Few-Layer Graphene via Solid-Gas Reaction. Mater. Res. Bull. 2015, 61, 252–258. [Google Scholar] [CrossRef]
  287. Pacilé, D.; Papagno, M.; Rodríguez, A.F.; Grioni, M.; Papagno, L.; Girit, C.; Meyer, J.C.; Begtrup, G.E.; Zettl, A. Near-Edge X-Ray Absorption Fine-Structure Investigation of Graphene. Phys. Rev. Lett. 2008, 101. [Google Scholar] [CrossRef]
  288. Liang, X.; Pan, D.; Lao, M.; Liang, S.; Huang, D.; Zhou, W.; Guo, J. Structural Evolution of Fluorinated Graphene upon Molten-Alkali Treatment Probed by X-Ray Absorption near-Edge Structure Spectroscopy. Appl. Surf. Sci. 2017, 404, 1–6. [Google Scholar] [CrossRef]
  289. Zitolo, A.; Goellner, V.; Armel, V.; Sougrati, M.T.; Mineva, T.; Stievano, L.; Fonda, E.; Jaouen, F. Identification of Catalytic Sites for Oxygen Reduction in Iron- and Nitrogen-Doped Graphene Materials. Nat. Mater. 2015, 14, 937–942. [Google Scholar] [CrossRef]
  290. Soin, N.; Ray, S.C.; Sarma, S.; Mazumder, D.; Sharma, S.; Wang, Y.F.; Pong, W.F.; Roy, S.S.; Strydom, A.M. Tuning the Electronic and Magnetic Properties of Nitrogen-Functionalized Few-Layered Graphene Nanoflakes. J. Phys. Chem. C 2017, 121, 14073–14082. [Google Scholar] [CrossRef]
  291. Ray, S.C.; Pong, W.F.; Papakonstantinou, P. Electronic Structure and Field Emission Properties of Nitrogen Doped Graphene Nano-Flakes (GNFs:N) and Carbon Nanotubes (CNTs:N). Appl. Surf. Sci. 2016, 380, 301–304. [Google Scholar] [CrossRef]
  292. Zhang, L.S.; Liang, X.Q.; Song, W.G.; Wu, Z.Y. Identification of the Nitrogen Species on N-Doped Graphene Layers and Pt/NG Composite Catalyst for Direct Methanol Fuel Cell. Phys. Chem. Chem. Phys. 2010, 12, 12055–12059. [Google Scholar] [CrossRef]
  293. Favaro, M.; Carraro, F.; Cattelan, M.; Colazzo, L.; Durante, C.; Sambi, M.; Gennaro, A.; Agnoli, S.; Granozzi, G. Multiple Doping of Graphene Oxide Foams and Quantum Dots: New Switchable Systems for Oxygen Reduction and Water Remediation. J. Mater. Chem. A Mater. 2015, 3, 14334–14347. [Google Scholar] [CrossRef]
  294. Gong, K.; Du, F.; Xia, Z.; Durstock, M.; Dai, L. Nitrogen-Doped Carbon Nanotube Arrays with High Electrocatalytic Activity for Oxygen Reduction. Science 2009, 323, 760–764. [Google Scholar] [CrossRef] [PubMed]
  295. Song, S.; Xue, Y.; Feng, L.; Elbatal, H.; Wang, P.; Moorefiel, C.N.; Newkome, G.R.; Dai, L. Reversible Self-Assembly of Terpyridine-Functionalized Graphene Oxide for Energy Conversion. Angew. Chem. Int. Ed. 2014, 53, 1415–1419. [Google Scholar] [CrossRef]
  296. Zhang, L.; Niu, J.; Dai, L.; Xia, Z. Effect of Microstructure of Nitrogen-Doped Graphene on Oxygen Reduction Activity in Fuel Cells. Langmuir 2012, 28, 7542–7550. [Google Scholar] [CrossRef]
  297. Feng, Y.X.; Li, F.F.; Hu, Z.P.; Luo, X.G.; Zhang, L.X.; Zhou, X.F.; Wang, H.T.; Xu, J.J.; Wang, E.G. Tuning the Catalytic Property of Nitrogen-Doped Graphene for Cathode Oxygen Reduction Reaction. Phys. Rev. B 2012, 85, 155454. [Google Scholar] [CrossRef]
  298. Studt, F. The Oxygen Reduction Reaction on Nitrogen-Doped Graphene. Catal. Lett. 2013, 143, 58–60. [Google Scholar] [CrossRef]
  299. Zhang, L.P.; Xia, Z.H. Mechanisms of Oxygen Reduction Reaction on Nitrogen-Doped Graphene for Fuel Cells. J. Phys. Chem. C 2011, 115, 11170–11176. [Google Scholar] [CrossRef]
  300. Boukhvalov, D.W.; Son, Y.W. Oxygen Reduction Reactions on Pure and Nitrogen-Doped Graphene: A First-Principles Modeling. Nanoscale 2012, 4, 417–420. [Google Scholar] [CrossRef]
  301. Jadhav, P.; Joshi, G.M. Recent Trends in Nitrogen Doped Polymer Composites: A Review. J. Polym. Res. 2021, 28, 16. [Google Scholar] [CrossRef]
  302. Park, S.; Srivastava, D.; Cho, K. Generalized Chemical Reactivity of Curved Surfaces: Carbon Nanotubes. Nano Lett. 2003, 3, 1273–1277. [Google Scholar] [CrossRef]
  303. Inagaki, M.; Toyoda, M.; Soneda, Y.; Morishita, T. Nitrogen-Doped Carbon Materials. Carbon 2018, 132, 104–140. [Google Scholar] [CrossRef]
  304. Wu, J.J.; Ma, L.L.; Yadav, R.M.; Yang, Y.C.; Zhang, X.; Vajtai, R.; Lou, J.; Ajayan, P.M. Nitrogen-Doped Graphene with Pyridinic Dominance as a Highly Active and Stable Electrocatalyst for Oxygen Reduction. ACS Appl. Mater. Interfaces 2015, 7, 14763–14769. [Google Scholar] [CrossRef]
  305. Matter, P.H.; Zhang, L.; Ozkan, U.S. The Role of Nanostructure in Nitrogen-Containing Carbon Catalysts for the Oxygen Reduction Reaction. J. Catal. 2006, 239, 83–96. [Google Scholar] [CrossRef]
  306. Montoya, A.; Mondragón, F.; Truong, T.N. Kinetics of Nitric Oxide Desorption from Carbonaceous Surfaces. Fuel Process. Technol. 2002, 77–78, 453–458. [Google Scholar] [CrossRef]
  307. Subramanian, N.P.; Li, X.; Nallathambi, V.; Kumaraguru, S.P.; Colon-Mercado, H.; Wu, G.; Lee, J.W.; Popov, B.N. Nitrogen-Modified Carbon-Based Catalysts for Oxygen Reduction Reaction in Polymer Electrolyte Membrane Fuel Cells. J. Power Sources 2009, 188, 38–44. [Google Scholar] [CrossRef]
  308. Luo, Z.Q.; Lim, S.H.; Tian, Z.Q.; Shang, J.Z.; Lai, L.F.; MacDonald, B.; Fu, C.; Shen, Z.X.; Yu, T.; Lin, J.Y. Pyridinic N Doped Graphene: Synthesis, Electronic Structure, and Electrocatalytic Property. J. Mater. Chem. 2011, 21, 8038–8044. [Google Scholar] [CrossRef]
  309. Yasuda, S.; Yu, L.; Kim, J.; Murakoshi, K. Selective Nitrogen Doping in Graphene for Oxygen Reduction Reactions. Chem. Commun. 2013, 49, 9627–9629. [Google Scholar] [CrossRef]
  310. Park, M.; Lee, T.; Kim, B.S. Covalent Functionalization Based Heteroatom Doped Graphene Nanosheet as a Metal-Free Electrocatalyst for Oxygen Reduction Reaction. Nanoscale 2013, 5, 12255–12260. [Google Scholar] [CrossRef]
  311. Vikkisk, M.; Kruusenberg, I.; Joost, U.; Shulga, E.; Kink, I.; Tammeveski, K. Electrocatalytic Oxygen Reduction on Nitrogen-Doped Graphene in Alkaline Media. Appl. Catal. B Environ. 2014, 147, 369–376. [Google Scholar] [CrossRef]
  312. Feng, L.Y.; Yang, L.Q.; Huang, Z.J.; Luo, J.Y.; Li, M.; Wang, D.B.; Chen, Y.G. Enhancing Electrocatalytic Oxygen Reduction on Nitrogen-Doped Graphene by Active Sites Implantation. Sci. Rep. 2013, 3, 3306. [Google Scholar] [CrossRef]
  313. Xing, T.; Zheng, Y.; Li, L.H.; Cowie, B.C.C.; Gunzelmann, D.; Qiao, S.Z.; Huang, S.; Chen, Y. Observation of Active Sites for Oxygen Reduction Reaction on Nitrogen-Doped Multilayer Graphene. ACS Nano 2014, 8, 6856–6862. [Google Scholar] [CrossRef] [PubMed]
  314. Zhao, H.; Hui, K.S.; Hui, K.N. Synthesis of Nitrogen-Doped Multilayer Graphene from Milk Powder with Melamine and Their Application to Fuel Cells. Carbon 2014, 76, 1–9. [Google Scholar] [CrossRef]
  315. Liao, Y.L.; Gao, Y.; Zhu, S.M.; Zheng, J.S.; Chen, Z.X.; Yin, C.; Lou, X.H.; Zhang, D. Facile Fabrication of N-Doped Graphene as Efficient Electrocatalyst for Oxygen Reduction Reaction. ACS Appl. Mater. Interfaces. 2015, 7, 19619–19625. [Google Scholar] [CrossRef] [PubMed]
  316. Shinde, D.B.; Dhavale, V.M.; Kurungot, S.; Pillai, V.K. Electrochemical Preparation of Nitrogen-Doped Graphene Quantum Dots and Their Size-Dependent Electrocatalytic Activity for Oxygen Reduction. Bull. Mater. Sci. 2015, 38, 435–442. [Google Scholar] [CrossRef]
  317. Farzaneh, A.; Saghatoleslami, N.; Goharshadi, E.K.; Gharibi, H.; Ahmadzadeh, H. 3-D Mesoporous Nitrogen-Doped Reduced Graphene Oxide as an Efficient Metal-Free Electrocatalyst for Oxygen Reduction Reaction in Alkaline Fuel Cells: Role of Pi and Lone Pair Electrons. Electrochim. Acta 2016, 222, 608–618. [Google Scholar] [CrossRef]
  318. Chi, Y.W.; Hu, C.C.; Huang, K.P.; Shen, H.H.; Muniyandi, R. Manipulation of Defect Density and Nitrogen Doping on Few-Layer Graphene Sheets Using the Plasma Methodology for Electrochemical Applications. Electrochim. Acta 2016, 221, 144–153. [Google Scholar] [CrossRef]
  319. Fan, M.M.; Zhu, C.L.; Yang, J.Z.; Sun, D.P. Facile Self-Assembly N-Doped Graphene Quantum Dots/Graphene for Oxygen Reduction Reaction. Electrochim. Acta 2016, 216, 102–109. [Google Scholar] [CrossRef]
  320. Mohapatra, B.D.; Mantry, S.P.; Behera, N.; Behera, B.; Rath, S.; Varadwaj, K.S.K. Stimulation of Electrocatalytic Oxygen Reduction Activity on Nitrogen Doped Graphene through Noncovalent Molecular Functionalisation. Chem. Commun. 2016, 52, 10385–10388. [Google Scholar] [CrossRef]
  321. Ratso, S.; Kruusenberg, I.; Joost, U.; Saar, R.; Tammeveski, K. Enhanced Oxygen Reduction Reaction Activity of Nitrogen-Doped Graphene/Multi-Walled Carbon Nanotube Catalysts in Alkaline Media. Int. J. Hydrogen Energy 2016, 41, 22510–22519. [Google Scholar] [CrossRef]
  322. Sun, J.G.; Wang, L.; Song, R.R.; Yanga, S.B. Enhancing Pyridinic Nitrogen Level in Graphene to Promote Electrocatalytic Activity for Oxygen Reduction Reaction. Nanotechnology 2016, 27, 055404. [Google Scholar] [CrossRef]
  323. Bera, B.; Chakraborty, A.; Kar, T.; Leuaa, P.; Neergat, M. Density of States, Carrier Concentration, and Flat Band Potential Derived from Electrochemical Impedance Measurements of N-Doped Carbon and Their Influence on Electrocatalysis of Oxygen Reduction Reaction. J. Phys. Chem. C 2017, 121, 20850–20856. [Google Scholar] [CrossRef]
  324. Gao, X.C.; Wang, L.W.; Ma, J.Z.; Wang, Y.Q.; Zhang, J.T. Facile Preparation of Nitrogen-Doped Graphene as an Efficient Oxygen Reduction Electrocatalyst. Inorg. Chem. Front. 2017, 4, 1582–1590. [Google Scholar] [CrossRef]
  325. Kundu, S.; Malik, B.; Pattanayak, D.K.; Pillai, V.K. Effect of Dimensionality and Doping in Quasi—“One-Dimensional (1D)” Nitrogen-Doped Graphene Nanoribbons on the Oxygen Reduction Reaction. ACS Appl. Mater. Interfaces 2017, 9, 38409–38418. [Google Scholar] [CrossRef] [PubMed]
  326. Liu, H.; Zhao, Q.S.; Liu, J.Y.; Ma, X.; Rao, Y.; Shao, X.D.; Li, Z.T.; Wu, W.T.; Ning, H.; Wu, M.B. Synergistically Enhanced Activity of Nitrogen-Doped Carbon Dots/Graphene Composites for Oxygen Reduction Reaction. Appl. Surf. Sci. 2017, 423, 909–916. [Google Scholar] [CrossRef]
  327. Lu, X.W.; Li, Z.F.; Yin, X.Y.; Wang, S.W.; Liu, Y.R.; Wang, Y.X. Controllable Synthesis of Three-Dimensional Nitrogen-Doped Graphene as a High Performance Electrocatalyst for Oxygen Reduction Reaction. Int. J. Hydrogen Energy 2017, 42, 17504–17513. [Google Scholar] [CrossRef]
  328. Miao, H.; Li, S.H.; Wang, Z.H.; Sun, S.S.; Kuang, M.; Liu, Z.P.; Yuan, J.L. Enhancing the Pyridinic N Content of Nitrogen-Doped Graphene and Improving Its Catalytic Activity for Oxygen Reduction Reaction. Int. J. Hydrogen Energy 2017, 42, 28298–28308. [Google Scholar] [CrossRef]
  329. Sun, L.; Luo, Y.; Li, M.; Hu, G.H.; Xu, Y.J.; Tang, T.; Wen, J.F.; Li, X.Y.; Wang, L. Role of Pyridinic-N for Nitrogen-Doped Graphene Quantum Dots in Oxygen Reaction Reduction. J. Colloid. Interface Sci. 2017, 508, 154–158. [Google Scholar] [CrossRef]
  330. Yadegari, A.; Samiee, L.; Tasharrofi, S.; Tajik, S.; Rashidi, A.; Shoghi, F.; Rasoulianboroujeni, M.; Tahriri, M.; Rowley-Neale, S.J.; Banks, C.E. Nitrogen Doped Nanoporous Graphene: An Efficient Metal-Free Electrocatalyst for the Oxygen Reduction Reaction. RSC Adv. 2017, 7, 55555–55566. [Google Scholar] [CrossRef]
  331. Hibino, T.; Kobayashi, K.; Heo, P. Oxygen Reduction Reaction over Nitrogen-Doped Graphene Oxide Cathodes in Acid and Alkaline Fuel Cells at Intermediate Temperatures. Electrochim. Acta 2013, 112, 82–89. [Google Scholar] [CrossRef]
  332. Ito, Y.; Qiu, H.J.; Fujita, T.; Tanabe, Y.; Tanigaki, K.; Chen, M.W. Bicontinuous Nanoporous N-Doped Graphene for the Oxygen Reduction Reaction. Adv. Mater. 2014, 26, 4145–4150. [Google Scholar] [CrossRef]
  333. Bang, G.S.; Shim, G.W.; Shin, G.H.; Jung, D.Y.; Park, H.; Hong, W.G.; Choi, J.; Lee, J.; Choi, S.Y. Pyridinic-N-Doped Graphene Paper from Perforated Graphene Oxide for Efficient Oxygen Reduction. ACS Omega 2018, 3, 5522–5530. [Google Scholar] [CrossRef] [PubMed]
  334. Wang, T.; Chen, Z.X.; Chen, Y.G.; Yang, L.J.; Yang, X.D.; Ye, J.Y.; Xia, H.P.; Zhou, Z.Y.; Sun, S.G. Identifying the Active Site of N-Doped Graphene for Oxygen Reduction by Selective Chemical Modification. ACS Energy Lett. 2018, 3, 986–991. [Google Scholar] [CrossRef]
  335. Sudhakar, S.; Jaiswal, K.K.; Ramaswamy, A.P. The Role of Microwave Irradiation Temperature on Nitrogen Doping in Metal-Free Graphene Catalysts for an Efficient Oxygen Reduction Reaction in an Alkaline Condition. ChemistrySelect 2018, 3, 8962–8972. [Google Scholar] [CrossRef]
  336. Han, J.H.; Huang, G.; Wang, Z.L.; Lu, Z.; Du, J.; Kashani, H.; Chen, M.W. Low-Temperature Carbide-Mediated Growth of Bicontinuous Nitrogen-Doped Mesoporous Graphene as an Efficient Oxygen Reduction Electrocatalyst. Adv. Mater. 2018, 30, 1803588. [Google Scholar] [CrossRef]
  337. Hang, C.; Zhang, J.; Zhu, J.W.; Li, W.Q.; Kou, Z.K.; Huang, Y.H. In Situ Exfoliating and Generating Active Sites on Graphene Nanosheets Strongly Coupled with Carbon Fiber toward Self-Standing Bifunctional Cathode for Rechargeable Zn-Air Batteries. Adv. Energy Mater. 2018, 8, 1703539. [Google Scholar] [CrossRef]
  338. She, Y.Y.; Chen, J.F.; Zhang, C.X.; Lu, Z.G.; Ni, M.; Sit, P.H.L.; Leung, M.K.H. Nitrogen-Doped Graphene Derived from Ionic Liquid as Metal-Free Catalyst for Oxygen Reduction Reaction and Its Mechanisms. Appl. Energy 2018, 225, 513–521. [Google Scholar] [CrossRef]
  339. Wang, Q.C.; Ji, Y.J.; Lei, Y.P.; Wang, Y.B.; Wang, Y.D.; Li, Y.Y.; Wang, S.Y. Pyridinic-N-Dominated Doped Defective Graphene as a Superior Oxygen Electrocatalyst for Ultrahigh-Energy-Density Zn-Air Batteries. ACS Energy Lett. 2018, 3, 1183–1191. [Google Scholar] [CrossRef]
  340. Xiang, Q.; Liu, Y.P.; Zou, X.F.; Hu, B.B.; Qiang, Y.J.; Yu, D.M.; Yin, W.; Chen, C.G. Hydrothermal Synthesis of a New Kind of N-Doped Graphene Gel-like Hybrid As an Enhanced ORR Electrocatalyst. ACS Appl. Mater. Interfaces 2018, 10, 10842–10850. [Google Scholar] [CrossRef]
  341. Zhang, J.; Sun, Y.M.; Zhu, J.W.; Gao, Z.H.; Li, S.Z.; Mu, S.C.; Huang, Y.H. Ultranarrow Graphene Nanoribbons toward Oxygen Reduction and Evolution Reactions. Adv. Sci. 2018, 5, 1801375. [Google Scholar] [CrossRef]
  342. Hu, Q.; Li, G.M.; Li, G.D.; Liu, X.F.; Zhu, B.; Chai, X.Y.; Zhang, Q.L.; Liu, J.; He, C. Trifunctional Electrocatalysis on Dual-Doped Graphene Nanorings-Integrated Boxes for Efficient Water Splitting and Zn-Air Batteries. Adv. Energy Mater. 2019, 9, 1803867. [Google Scholar] [CrossRef]
  343. Komba, N.; Wei, Q.L.; Zhang, G.X.; Rosei, F.; Sun, S.H. Controlled Synthesis of Graphene via Electrochemical Route and Its Use as Efficient Metal-Free Catalyst for Oxygen Reduction. Appl. Catal. B Environ. 2019, 243, 373–380. [Google Scholar] [CrossRef]
  344. Liu, Y.; Liu, Z.M.; Liu, H.; Liao, M.L. Novel Porous Nitrogen Doped Graphene/Carbon Black Composites as Efficient Oxygen Reduction Reaction Electrocatalyst for Power Generation in Microbial Fuel Cell. Nanomaterials 2019, 9, 836. [Google Scholar] [CrossRef] [PubMed]
  345. Peediyakkal, H.P.; Yu, L.; Munakata, H.; Kanamura, K. Highly Durable Non-Platinum Catalyst for Protic Ionic Liquid Based Intermediate Temperature PEFCs. Electrochemistry 2019, 87, 35–46. [Google Scholar] [CrossRef]
  346. Ramirez-Barria, C.S.; Fernandes, D.M.; Freire, C.; Villaro-Abalos, E.; Guerrero-Ruiz, A.; Rodriguez-Ramos, I. Upgrading the Properties of Reduced Graphene Oxide and Nitrogen-Doped Reduced Graphene Oxide Produced by Thermal Reduction toward Efficient ORR Electrocatalysts. Nanomaterials 2019, 9, 1761. [Google Scholar] [CrossRef]
  347. Yeddala, M.; Gorle, D.B.; Kulandainathan, M.A.; Ragupathy, P.; Pillai, V.K. Solid-State Thermal Exfoliation of Graphite Nano-Fibers to Edge-Nitrogenized Graphene Nanosheets for Oxygen Reduction Reaction. J. Colloid. Interface Sci. 2019, 545, 71–81. [Google Scholar] [CrossRef]
  348. Zeng, J.; Mu, Y.B.; Ji, X.X.; Lin, Z.J.; Lin, Y.H.; Ma, Y.H.; Zhang, Z.X.; Wang, S.G.; Ren, Z.H.; Yu, J. N,O-Codoped 3D Graphene Fibers with Densely Arranged Sharp Edges as Highly Efficient Electrocatalyst for Oxygen Reduction Reaction. J. Mater. Sci. 2019, 54, 14495–14503. [Google Scholar] [CrossRef]
  349. Liu, X.; Jiang, L.X.; Zhu, Z.J.; Chen, S.; Dou, Y.H.; Liu, P.R.; Wang, Y.; Yin, H.J.; Tang, Z.Y.; Zhao, H.J. Wet-Chemistry Grafted Active Pyridinic Nitrogen Sites on Holey Graphene Edges as High Performance ORR Electrocatalyst for Zn-Air Batteries. Mater. Today Energy 2019, 11, 24–29. [Google Scholar] [CrossRef]
  350. Romero, A.; Lavín-López, M.P.; de la Osa, A.R.; Ordoñez, S.; de Lucas-Consuegra, A.; Valverde, J.L.; Patón, A. Different Strategies to Simultaneously N-Doping and Reduce Graphene Oxide for Electrocatalytic Applications. J. Electroanal. Chem. 2020, 857, 113695. [Google Scholar] [CrossRef]
  351. Wu, Z.H.; Zhang, Y.S.; Li, L.; Zhao, Y.G.; Shen, Y.L.; Wang, S.B.; Shao, G.S. Nitrogen-Doped Vertical Graphene Nanosheets by High-Flux Plasma Enhanced Chemical Vapor Deposition as Efficient Oxygen Reduction Catalysts for Zn-Air Batteries. J. Mater. Chem. A Mater. 2020, 8, 23248–23256. [Google Scholar] [CrossRef]
  352. Bian, Y.R.; Wang, H.; Hu, J.T.; Liu, B.W.; Liu, D.; Dai, L.M. Nitrogen-Rich Holey Graphene for Efficient Oxygen Reduction Reaction. Carbon 2020, 162, 66–73. [Google Scholar] [CrossRef]
  353. Li, D.N.; Chen, H.B.; Zhang, Y.Y.; Yang, J.Z.; Yuan, H.R.; Chen, Y. Upcycling Biomass Tar into Highly Porous, Defective and Pyridinic-n-Enriched Graphene Nanohybrid as Efficient Bifunctional Catalyst for Zn-Air Battery. Electrochim. Acta 2020, 364, 137319. [Google Scholar] [CrossRef]
  354. Manzhos, R.A.; Baskakov, S.A.; Kabachkov, E.N.; Korepanov, V.I.; Dremova, N.N.; Baskakova, Y.V.; Krivenko, A.G.; Shulga, Y.M.; Gutsev, G.L. Reduced Graphene Oxide Aerogel inside Melamine Sponge as an Electrocatalyst for the Oxygen Reduction Reaction. Materials 2021, 14, 322. [Google Scholar] [CrossRef]
  355. Xi, Z.Y.; Han, J.H.; Jin, Z.Y.; Hu, K.L.; Qiu, H.J.; Ito, Y. All-Solid-State Mg-Air Battery Enhanced with Free-Standing N-Doped 3D Nanoporous Graphene. Small 2023, 20, 2308045. [Google Scholar] [CrossRef]
  356. Chen, Z.; Higgins, D.; Tao, H.S.; Hsu, R.S.; Chen, Z.W. Highly Active Nitrogen-Doped Carbon Nanotubes for Oxygen Reduction Reaction in Fuel Cell Applications. J. Phys. Chem. C 2009, 113, 21008–21013. [Google Scholar] [CrossRef]
  357. Rao, C.V.; Cabrera, C.R.; Ishikawa, Y. In Search of the Active Site in Nitrogen-Doped Carbon Nanotube Electrodes for the Oxygen Reduction Reaction. J. Phys. Chem. Lett. 2010, 1, 2622–2627. [Google Scholar] [CrossRef]
  358. Zhong, X.; Yu, H.Y.; Zhuang, G.L.; Li, Q.; Wang, X.D.; Zhu, Y.S.; Liu, L.; Li, X.N.; Dong, M.D.; Wang, J.G. Pyridyne Cycloaddition of Graphene: “External” Active Sites for Oxygen Reduction Reaction. J. Mater. Chem. A Mater. 2014, 2, 897–901. [Google Scholar] [CrossRef]
  359. Jhi, S.H.; Louie, S.G.; Cohen, M.L. Electronic Properties of Oxidized Carbon Nanotubes. Phys. Rev. Lett. 2000, 85, 1710–1713. [Google Scholar] [CrossRef]
  360. Nandhini, S.; Rajkamal, A.; Saha, B.; Thapa, R. First-Principles Identification of Site Dependent Activity of Graphene Based Electrocatalyst. Mol. Catal. 2017, 432, 242–249. [Google Scholar] [CrossRef]
  361. Huang, S.F.; Terakura, K.; Ozaki, T.; Ikeda, T.; Boero, M.; Oshima, M.; Ozaki, J.; Miyata, S. First-Principles Calculation of the Electronic Properties of Graphene Clusters Doped with Nitrogen and Boron: Analysis of Catalytic Activity for the Oxygen Reduction Reaction. Phys. Rev. B 2009, 80, 235410. [Google Scholar] [CrossRef]
  362. Ni, S.; Li, Z.Y.; Yang, J.L. Oxygen Molecule Dissociation on Carbon Nanostructures with Different Types of Nitrogen Doping. Nanoscale 2012, 4, 1184–1189. [Google Scholar] [CrossRef]
  363. Saidi, W.A. Oxygen Reduction Electrocatalysis Using N-Doped Graphene Quantum-Dots. J. Phys. Chem. Lett. 2013, 4, 4160–4165. [Google Scholar] [CrossRef]
  364. Ikeda, T.; Hou, Z.F.; Chai, G.L.; Terakura, K. Possible Oxygen Reduction Reactions for Graphene Edges from First Principles. J. Phys. Chem. C 2014, 118, 17616–17625. [Google Scholar] [CrossRef]
  365. Radovic, L.R.; Salgado-Casanova, A.J.A.; Mora-Vilches, C. V On the Active Sites for the Oxygen Reduction Reaction Catalyzed by Graphene-Based Materials. Carbon 2020, 156, 389–398. [Google Scholar] [CrossRef]
  366. Chen, M.F.; Chao, T.H.; Shen, M.H.; Lu, Q.; Cheng, M.J. Evaluating Potential Catalytic Active Sites on Nitrogen-Doped Graphene for the Oxygen Reduction Reaction: An Approach Based on Constant Electrode Potential Density Functional Theory Calculations. J. Phys. Chem. C 2020, 124, 25675–25685. [Google Scholar] [CrossRef]
  367. Yu, L.; Pan, X.L.; Cao, X.M.; Hu, P.; Bao, X.H. Oxygen Reduction Reaction Mechanism on Nitrogen-Doped Graphene: A Density Functional Theory Study. J. Catal. 2011, 282, 183–190. [Google Scholar] [CrossRef]
  368. Reda, M.; Hansen, H.A.; Vegge, T. DFT Study of Stabilization Effects on N-Doped Graphene for ORR Catalysis. Catal. Today 2018, 312, 118–125. [Google Scholar] [CrossRef]
  369. Matsuyama, H.; Akaishi, A.; Nakamura, J. Effect of Water on the Manifestation of the Reaction Selectivity of Nitrogen-Doped Graphene Nanoclusters toward Oxygen Reduction Reaction. ACS Omega 2019, 4, 3832–3838. [Google Scholar] [CrossRef]
  370. Ma, J.; Gong, L.; Shen, Y.; Sun, D.; Liu, B.; Zhang, J.; Liu, D.; Zhang, L.; Xia, Z. Detrimental Effects and Prevention of Acidic Electrolytes on Oxygen Reduction Reaction Catalytic Performance of Heteroatom-Doped Graphene Catalysts. Front. Mater. 2019, 6, 294. [Google Scholar] [CrossRef]
  371. Zhang, P.; Hu, Q.; Yang, X.J.; Hou, X.L.; Mi, J.L.; Liu, L.; Dong, M.D. Size Effect of Oxygen Reduction Reaction on Nitrogen-Doped Graphene Quantum Dots. RSC Adv. 2018, 8, 531–536. [Google Scholar] [CrossRef]
  372. Wang, X.; Hou, Z.; Ikeda, T.; Huang, S.F.; Terakura, K.; Boero, M.; Oshima, M.; Kakimoto, M.A.; Miyata, S. Selective Nitrogen Doping in Graphene: Enhanced Catalytic Activity for the Oxygen Reduction Reaction. Phys. Rev. B Condens. Matter Mater. Phys. 2011, 84, 245434. [Google Scholar] [CrossRef]
  373. Kim, H.; Lee, K.; Woo, S.I.; Jung, Y. On the Mechanism of Enhanced Oxygen Reduction Reaction in Nitrogen-Doped Graphene Nanoribbons. Phys. Chem. Chem. Phys. 2011, 13, 17505–17510. [Google Scholar] [CrossRef] [PubMed]
  374. Hou, X.L.; Hu, Q.; Zhang, P.; Mi, J.L. Oxygen Reduction Reaction on Nitrogen-Doped Graphene Nanoribbons: A Density Functional Theory Study. Chem. Phys. Lett. 2016, 663, 123–127. [Google Scholar] [CrossRef]
  375. Ekspong, J.; Boulanger, N.; Gracia-Espino, E. Surface Activation of Graphene Nanoribbons for Oxygen Reduction Reaction by Nitrogen Doping and Defect Engineering: An Ab Initio Study. Carbon 2018, 137, 349–357. [Google Scholar] [CrossRef]
  376. Goswami, M.; Mandal, S.; Pillai, V.K. Effect of Hetero-Atom Doping on the Electrocatalytic Properties of Graphene Quantum Dots for Oxygen Reduction Reaction. Sci. Rep. 2023, 13, 10. [Google Scholar] [CrossRef]
  377. Scardamaglia, M.; Susi, T.; Struzzi, C.; Snyders, R.; Di Santo, G.; Petaccia, L.; Bittencourt, C. Spectroscopic Observation of Oxygen Dissociation on Nitrogen-Doped Graphene. Sci. Rep. 2017, 7, 7960. [Google Scholar] [CrossRef] [PubMed]
  378. Geng, D.S.; Chen, Y.; Chen, Y.G.; Li, Y.L.; Li, R.Y.; Sun, X.L.; Ye, S.Y.; Knights, S. High Oxygen-Reduction Activity and Durability of Nitrogen-Doped Graphene. Energy Environ. Sci. 2011, 4, 760–764. [Google Scholar] [CrossRef]
  379. Huan, T.N.; Khai, T.V.; Kang, Y.; Shim, K.B.; Chung, H. Enhancement of Quaternary Nitrogen Doping of Graphene Oxide via Chemical Reduction Prior to Thermal Annealing and an Investigation of Its Electrochemical Properties. J. Mater. Chem. 2012, 22, 14756–14762. [Google Scholar] [CrossRef]
  380. Lin, Z.Y.; Song, M.K.; Ding, Y.; Liu, Y.; Liu, M.L.; Wong, C.P. Facile Preparation of Nitrogen-Doped Graphene as a Metal-Free Catalyst for Oxygen Reduction Reaction. Phys. Chem. Chem. Phys. 2012, 14, 3381–3387. [Google Scholar] [CrossRef]
  381. Lin, Z.Y.; Waller, G.; Liu, Y.; Liu, M.L.; Wong, C.P. Facile Synthesis of Nitrogen-Doped Graphene via Pyrolysis of Graphene Oxide and Urea, and Its Electrocatalytic Activity toward the Oxygen-Reduction Reaction. Adv. Energy Mater. 2012, 2, 884–888. [Google Scholar] [CrossRef]
  382. Lin, Z.Y.; Waller, G.H.; Liu, Y.; Liu, M.L.; Wong, C.P. Simple Preparation of Nanoporous Few-Layer Nitrogen-Doped Graphene for Use as an Efficient Electrocatalyst for Oxygen Reduction and Oxygen Evolution Reactions. Carbon 2013, 53, 130–136. [Google Scholar] [CrossRef]
  383. Zheng, B.; Wang, J.; Wang, F.B.; Xia, X.H. Synthesis of Nitrogen Doped Graphene with High Electrocatalytic Activity toward Oxygen Reduction Reaction. Electrochem. Commun. 2013, 28, 24–26. [Google Scholar] [CrossRef]
  384. Lin, Z.Y.; Waller, G.H.; Liu, Y.; Liu, M.L.; Wong, C.P. 3D Nitrogen-Doped Graphene Prepared by Pyrolysis of Graphene Oxide with Polypyrrole for Electrocatalysis of Oxygen Reduction Reaction. Nano Energy 2013, 2, 241–248. [Google Scholar] [CrossRef]
  385. Xu, X.; Yuan, T.; Zhou, Y.K.; Li, Y.W.; Lu, J.M.; Tian, X.H.; Wang, D.L.; Wang, J. Facile Synthesis of Boron and Nitrogen-Doped Graphene as Efficient Electrocatalyst for the Oxygen Reduction Reaction in Alkaline Media. Int. J. Hydrogen Energy 2014, 39, 16043–16052. [Google Scholar] [CrossRef]
  386. Cong, H.P.; Wang, P.; Gong, M.; Yu, S.H. Facile Synthesis of Mesoporous Nitrogen-Doped Graphene: An Efficient Methanol-Tolerant Cathodic Catalyst for Oxygen Reduction Reaction. Nano Energy 2014, 3, 55–63. [Google Scholar] [CrossRef]
  387. Fu, X.G.; Jin, J.T.; Liu, Y.R.; Wei, Z.Y.; Pan, F.P.; Zhang, J.Y. Efficient Oxygen Reduction Electrocatalyst Based on Edge-Nitrogen-Rich Graphene Nanoplatelets: Toward a Large-Scale Synthesis. ACS Appl. Mater. Interfaces 2014, 6, 3930–3936. [Google Scholar] [CrossRef]
  388. Ouyang, W.P.; Zeng, D.R.; Yu, X.; Xie, F.Y.; Zhang, W.H.; Chen, J.; Yan, J.; Xie, F.J.; Wang, L.; Meng, H.; et al. Exploring the Active Sites of Nitrogen-Doped Graphene as Catalysts for the Oxygen Reduction Reaction. Int. J. Hydrogen Energy 2014, 39, 15996–16005. [Google Scholar] [CrossRef]
  389. Wang, Z.W.; Li, B.; Xin, Y.C.; Liu, J.G.; Yao, Y.F.; Zou, Z.G. Rapid Synthesis of Nitrogen-Doped Graphene by Microwave Heating for Oxygen Reduction Reactions in Alkaline Electrolyte. Chin. J. Catal. 2014, 35, 509–513. [Google Scholar] [CrossRef]
  390. Qazzazie, D.; Beckert, M.; Mulhaupt, R.; Yurchenko, O.; Urban, G. A Nitrogen-Doped Graphene Electrocatalyst for Selective Oxygen Reduction in Presence of Glucose and D-Gluconic Acid in PH-Neutral Media. Electrochim. Acta 2015, 186, 579–590. [Google Scholar] [CrossRef]
  391. Liu, Y.S.; Li, J.; Li, W.Z.; Li, Y.M.; Zhan, F.Q.; Tang, H.; Chen, Q.Y. Exploring the Nitrogen Species of Nitrogen Doped Graphene as Electrocatalysts for Oxygen Reduction Reaction in Al-Air Batteries. Int. J. Hydrogen Energy 2016, 41, 10354–10365. [Google Scholar] [CrossRef]
  392. Haque, E.; Sarkar, S.; Hassan, M.; Hossain, M.S.; Minett, A.I.; Dou, S.X.; Gomes, V.G. Tuning Graphene for Energy and Environmental Applications: Oxygen Reduction Reaction and Greenhouse Gas Mitigation. J. Power Sources 2016, 328, 472–481. [Google Scholar] [CrossRef]
  393. Bayram, E.; Yilmaz, G.; Mukerjee, S. A Solution-Based Procedure for Synthesis of Nitrogen Doped Graphene as an Efficient Electrocatalyst for Oxygen Reduction Reactions in Acidic and Alkaline Electrolytes. Appl. Catal. B Environ. 2016, 192, 26–34. [Google Scholar] [CrossRef]
  394. Cui, H.J.; Yu, H.M.; Zheng, J.F.; Wang, Z.J.; Zhu, Y.Y.; Jia, S.P.; Jia, J.; Zhu, Z.P. N-Doped Graphene Frameworks with Superhigh Surface Area: Excellent Electrocatalytic Performance for Oxygen Reduction. Nanoscale 2016, 8, 2795–2803. [Google Scholar] [CrossRef] [PubMed]
  395. Soo, L.T.; Loh, K.S.; Mohamad, A.B.; Daud, W.R.W.; Wong, W.Y. Effect of Nitrogen Precursors on the Electrochemical Performance of Nitrogen-Doped Reduced Graphene Oxide towards Oxygen Reduction Reaction. J. Alloys Compd. 2016, 677, 112–120. [Google Scholar] [CrossRef]
  396. Yang, H.B.; Miao, J.W.; Hung, S.F.; Chen, J.Z.; Tao, H.B.; Wang, X.Z.; Zhang, L.P.; Chen, R.; Gao, J.J.; Chen, H.M.; et al. Identification of Catalytic Sites for Oxygen Reduction and Oxygen Evolution in N-Doped Graphene Materials: Development of Highly Efficient Metal-Free Bifunctional Electrocatalyst. Sci. Adv. 2016, 2, 1501122. [Google Scholar] [CrossRef]
  397. Zhang, G.X.; Jin, X.Y.; Li, H.Y.; Wang, L.; Hu, C.J.; Sun, X.M. N-Doped Crumpled Graphene: Bottom-up Synthesis and Its Superior Oxygen Reduction Performance. Sci. China Mater. 2016, 59, 337–347. [Google Scholar] [CrossRef]
  398. Soo, L.T.; Loh, K.S.; Mohamad, A.; Daud, W.R.W. The Effect of Varying N/C Ratios of Nitrogen Precursors during Non-Metal Graphene Catalyst Synthesis. Int. J. Hydrogen Energy 2017, 42, 9069–9076. [Google Scholar] [CrossRef]
  399. Cui, H.J.; Jiao, M.G.; Chen, Y.N.; Guo, Y.B.; Yang, L.P.; Xie, Z.J.; Zhou, Z.; Guo, S.J. Molten-Salt-Assisted Synthesis of 3D Holey N-Doped Graphene as Bifunctional Electrocatalysts for Rechargeable Zn-Air Batteries. Small Methods 2018, 2, 1800144. [Google Scholar] [CrossRef]
  400. Kabir, S.; Artyushkova, K.; Serov, A.; Atanassov, P. Role of Nitrogen Moieties in N-Doped 3D-Graphene Nanosheets for Oxygen Electroreduction in Acidic and Alkaline Media. ACS Appl. Mater. Interfaces 2018, 10, 11623–11632. [Google Scholar] [CrossRef]
  401. Rahsepar, M.; Nobakht, M.R.; Kim, H.; Pakshir, M. Facile Enhancement of the Active Catalytic Sites of N-Doped Graphene as a High Performance Metal-Free Electrocatalyst for Oxygen Reduction Reaction. Appl. Surf. Sci. 2018, 447, 182–190. [Google Scholar] [CrossRef]
  402. Shi, J.L.; Tang, C.; Huang, J.Q.; Zhu, W.C.; Zhang, Q. Effective Exposure of Nitrogen Heteroatoms in 3D Porous Graphene Framework for Oxygen Reduction Reaction and Lithium-Sulfur Batteries. J. Energy Chem. 2018, 27, 167–175. [Google Scholar] [CrossRef]
  403. Lu, X.Y.; Wang, D.; Ge, L.P.; Xiao, L.H.; Zhang, H.Y.; Liu, L.L.; Zhang, J.Q.; An, M.Z.; Yang, P.X. Enriched Graphitic N in Nitrogen-Doped Graphene as a Superior Metal-Free Electrocatalyst for the Oxygen Reduction Reaction. N. J. Chem. 2018, 42, 19665–19670. [Google Scholar] [CrossRef]
  404. Chen, X.F.; Liang, Y.; Wan, L.; Xie, Z.L.; Easton, C.D.; Bourgeois, L.; Wang, Z.Y.; Bao, Q.L.; Zhu, Y.G.; Tao, S.W.; et al. Construction of Porous N-Doped Graphene Layer for Efficient Oxygen Reduction Reaction. Chem. Eng. Sci. 2019, 194, 36–44. [Google Scholar] [CrossRef]
  405. Maouche, C.; Zhou, Y.Z.; Li, B.; Cheng, C.; Wu, Y.Y.; Li, J.H.; Gao, S.; Yang, J. Thermal Treated Three-Dimensional N-Doped Graphene as Efficient Metal Free-Catalyst for Oxygen Reduction Reaction. J. Electroanal. Chem. 2019, 853, 113536. [Google Scholar] [CrossRef]
  406. Guo, J.H.; Zhang, S.L.; Zheng, M.X.; Tang, J.; Liu, L.; Chen, J.M.; Wang, X.C. Graphitic-N-Rich N-Doped Graphene as a High Performance Catalyst for Oxygen Reduction Reaction in Alkaline Solution. Int. J. Hydrogen Energy 2020, 45, 32402–32412. [Google Scholar] [CrossRef]
  407. Mohmad, G.; Sarkar, S.; Biswas, A.; Roy, K.; Dey, R.S. Polymer-Assisted Electrophoretic Synthesis of N-Doped Graphene-Polypyrrole Demonstrating Oxygen Reduction with Excellent Methanol Crossover Impact and Durability. Chem. A Eur. J. 2020, 26, 12664–12673. [Google Scholar] [CrossRef]
  408. Posudievsky, O.Y.; Kondratyuk, A.S.; Kozarenko, O.A.; Cherepanov, V.V.; Karbivskiy, V.L.; Koshechko, V.G.; Pokhodenko, V.D. Boosting Graphene Electrocatalytic Efficiency in Oxygen Reduction Reaction by Mechanochemically Induced Low-Temperature Nitrogen Doping. Electrochim. Acta 2021, 399, 9. [Google Scholar] [CrossRef]
  409. Yan, W.; Wang, L.; Chen, C.; Zhang, D.; Li, A.J.; Yao, Z.; Shi, L.Y. Polystyrene Microspheres-Templated Nitrogen-Doped Graphene Hollow Spheres as Metal-Free Catalyst for Oxygen Reduction Reaction. Electrochim. Acta 2016, 188, 230–239. [Google Scholar] [CrossRef]
  410. Zhang, B.; Xiao, C.H.; Xiang, Y.; Dong, B.T.; Ding, S.J.; Tang, Y.H. Nitrogen-Doped Graphene Quantum Dots Anchored on Thermally Reduced Graphene Oxide as an Electrocatalyst for the Oxygen Reduction Reaction. ChemElectroChem 2016, 3, 864–870. [Google Scholar] [CrossRef]
  411. Sjoberg, P.; Murray, J.S.; Brinck, T.; Politzer, P. Average Local Ionization Energies on the Molecular-Surfaces of Aromatic Systems as Guides to Chemical-Reactivity. Can. J. Chem. Rev. Can. Chim. 1990, 68, 1440–1443. [Google Scholar] [CrossRef]
  412. Lu, T.; Chen, F. Quantitative Analysis of Molecular Surface Based on Improved Marching Tetrahedra Algorithm. J. Mol. Graph. Model 2012, 38, 314–323. [Google Scholar] [CrossRef]
  413. Rahsepar, M.; Pakshir, M.; Kim, H. Synthesis of Multiwall Carbon Nanotubes with a High Loading of Pt Bya Microwave-Assisted Impregnation Method for Use in the Oxygenreduction Reaction. Electrochim. Acta 2013, 108, 769–775. [Google Scholar] [CrossRef]
  414. Rahsepar, M.; Kim, H. Microwave-Assisted Synthesis and Characterization of Bimetallic PtRu Alloy Nanoparticles Supported on Carbon Nanotubes. J. Alloys Compd. 2015, 649, 1323–1328. [Google Scholar] [CrossRef]
  415. Kim, I.T.; Song, M.J.; Kim, Y.B.; Shin, M.W. Microwave-Hydrothermal Synthesis of Boron/Nitrogen Co-Doped Graphene as an Efficient Metal-Free Electrocatalyst for Oxygen Reduction Reaction. Int. J. Hydrogen Energy 2016, 41, 22026–22033. [Google Scholar] [CrossRef]
  416. Unni, S.M.; Devulapally, S.; Karjule, N.; Kurungot, S. Graphene Enriched with Pyrrolic Coordination of the Doped Nitrogen as an Efficient Metal-Free Electrocatalyst for Oxygen Reduction. J. Mater. Chem 2012, 22, 23506–23513. [Google Scholar] [CrossRef]
  417. Liu, C.L.; Hu, C.C.; Wu, S.H.; Wu, T.H. Electron Transfer Number Control of the Oxygen Reduction Reaction on Nitrogen-Doped Reduced-Graphene Oxides Using Experimental Design Strategies. J. Electrochem. Soc. 2013, 160, H547–H552. [Google Scholar] [CrossRef]
  418. Li, L.Q.; Tang, C.; Zheng, Y.; Xia, B.Q.; Zhou, X.L.; Xu, H.L.; Qiao, S.Z. Tailoring Selectivity of Electrochemical Hydrogen Peroxide Generation by Tunable Pyrrolic-Nitrogen-Carbon. Adv. Energy Mater. 2020, 10, 2000789. [Google Scholar] [CrossRef]
  419. Zhang, Y.; Melchionna, M.; Medved, M.; Blonski, P.; Stekly, T.; Bakandritsos, A.; Kment, S.; Zboril, R.; Otyepka, M.; Fornasiero, P.; et al. Enhanced On-Site Hydrogen Peroxide Electrosynthesis by a Selectively Carboxylated N-Doped Graphene Catalyst. ChemCatChem 2021, 13, 4372–4383. [Google Scholar] [CrossRef]
  420. Koh, K.H.; Kim, Y.J.; Mostaghim, A.H.B.; Siahrostami, S.; Han, T.H.; Chen, Z. Elaborating Nitrogen and Oxygen Dopants Configurations within Graphene Electrocatalysts for Two-Electron Oxygen Reduction. ACS Mater. Lett. 2022, 4, 320–328. [Google Scholar] [CrossRef]
  421. Peng, Y.Y.; Bian, Z.Y.; Zhang, W.H.; Wang, H. Identifying the Key N Species for Electrocatalytic Oxygen Reduction Reaction on N-Doped Graphene. Nano Res. 2023, 16, 6642–6651. [Google Scholar] [CrossRef]
  422. Naveen, M.H.; Noh, H.B.; Al Hossain, M.S.; Kim, J.H.; Shim, Y.B. Facile Potentiostatic Preparation of Functionalized Polyterthiophene-Anchored Graphene Oxide as a Metal-Free Electrocatalyst for the Oxygen Reduction Reaction. J. Mater. Chem. A Mater. 2015, 3, 5426–5433. [Google Scholar] [CrossRef]
  423. Navaee, A.; Salimi, A. Efficient Amine Functionalization of Graphene Oxide through the Bucherer Reaction: An Extraordinary Metal-Free Electrocatalyst for the Oxygen Reduction Reaction. RSC Adv. 2015, 5, 59874–59880. [Google Scholar] [CrossRef]
  424. Jiang, Z.Q.; Jiang, Z.J.; Tian, X.N.; Chen, W.H. Amine-Functionalized Holey Graphene as a Highly Active Metal-Free Catalyst for the Oxygen Reduction Reaction. J. Mater. Chem. A Mater. 2014, 2, 441–450. [Google Scholar] [CrossRef]
  425. Ahmed, M.S.; Kim, Y.B. 3D Graphene Preparation via Covalent Amide Functionalization for Efficient Metal-Free Electrocatalysis in Oxygen Reduction. Sci. Rep. 2017, 7, 43279. [Google Scholar] [CrossRef]
  426. Zhang, C.Z.; Hao, R.; Liao, H.B.; Hou, Y.L. Synthesis of Amino-Functionalized Graphene as Metal-Free Catalyst and Exploration of the Roles of Various Nitrogen States in Oxygen Reduction Reaction. Nano Energy 2013, 2, 88–97. [Google Scholar] [CrossRef]
  427. Dou, S.; Shen, A.L.; Tao, L.; Wang, S.Y. Molecular Doping of Graphene as Metal-Free Electrocatalyst for Oxygen Reduction Reaction. Chem. Commun. 2014, 50, 10672–10675. [Google Scholar] [CrossRef]
  428. Ahmed, M.S.; Jeon, S. New Functionalized Graphene Sheets for Enhanced Oxygen Reduction as Metal-Free Cathode Electrocatalysts. J. Power Sources 2012, 218, 168–173. [Google Scholar] [CrossRef]
  429. Lee, M.S.; Whang, D.R.; Choi, H.J.; Yang, M.H.; Kim, B.G.; Baek, J.B.; Chang, D.W. A Facile Approach to Tailoring Electrocatalytic Activities of Imine-Rich Nitrogen-Doped Graphene for Oxygen Reduction Reaction. Carbon 2017, 122, 515–523. [Google Scholar] [CrossRef]
  430. Zuo, Z.C.; Li, W.; Manthiram, A. N-Heterocycles Tethered Graphene as Efficient Metal-Free Catalysts for an Oxygen Reduction Reaction in Fuel Cells. J. Mater. Chem. A Mater. 2013, 1, 10166–10172. [Google Scholar] [CrossRef]
  431. Ahmed, M.S.; Kim, Y.B. Amide-Functionalized Graphene with 1,4-Diaminobutane as Efficient Metal-Free and Porous Electrocatalyst for Oxygen Reduction. Carbon 2017, 111, 577–586. [Google Scholar] [CrossRef]
  432. Unni, S.M.; Illathvalappil, R.; Gangadharan, P.K.; Bhange, S.N.; Kurungot, S. Layer-Separated Distribution of Nitrogen Doped Graphene by Wrapping on Carbon Nitride Tetrapods for Enhanced Oxygen Reduction Reactions in Acidic Medium. Chem. Commun. 2014, 50, 13769–13772. [Google Scholar] [CrossRef]
  433. Vazquez-Arenas, J.; Galano, A.; Lee, D.U.; Higgins, D.; Guevara-Garcia, A.; Chen, Z. Theoretical and Experimental Studies of Highly Active Graphene Nanosheets to Determine Catalytic Nitrogen Sites Responsible for the Oxygen Reduction Reaction in Alkaline Media. J. Mater. Chem. A Mater. 2016, 4, 976–990. [Google Scholar] [CrossRef]
  434. Kabir, S.; Artyushkova, K.; Serov, A.; Kiefer, B.; Atanassov, P. Binding Energy Shifts for Nitrogen-Containing Graphene-Based Electrocatalysts—Experiments and DFT Calculations. Surf. Interface Anal. 2016, 48, 293–300. [Google Scholar] [CrossRef]
  435. Tian, L.L.; Yang, J.; Weng, M.Y.; Tan, R.; Zheng, J.X.; Chen, H.B.; Zhuang, Q.C.; Dai, L.M.; Pan, F. Fast Diffusion of O-2 on Nitrogen-Doped Graphene to Enhance Oxygen Reduction and Its Application for High-Rate Zn-Air Batteries. ACS Appl. Mater. Interfaces 2017, 9, 7125–7130. [Google Scholar] [CrossRef] [PubMed]
  436. Alonso-Lemus, I.L.; Figueroa-Torres, M.Z.; Garcia-Hernandez, A.B.; Escobar-Morales, B.; Rodriguez-Varela, F.J.; Fuentes, A.F.; Lardizabal-Gutierrez, D.; Quintana-Owen, P. Low-Cost Sonochemical Synthesis of Nitrogen-Doped Graphene Metal-Free Electrocatalyst for the Oxygen Reduction Reaction in Alkaline Media. Int. J. Hydrogen Energy 2017, 42, 30330–30338. [Google Scholar] [CrossRef]
  437. Faisal, S.N.; Haque, E.; Noorbehesht, N.; Zhang, W.M.; Harris, A.T.; Church, T.L.; Minett, A.I. Pyridinic and Graphitic Nitrogen-Rich Graphene for High-Performance Supercapacitors and Metal-Free Bifunctional Electrocatalysts for ORR and OER. RSC Adv. 2017, 7, 17950–17958. [Google Scholar] [CrossRef]
  438. Chanda, D.; Dobrota, A.S.; Hnat, J.; Sofer, Z.; Pasti, I.A.; Skorodumova, N.V.; Paidar, M.; Bouzek, K. Investigation of Electrocatalytic Activity on a N-Doped Reduced Graphene Oxide Surface for the Oxygen Reduction Reaction in an Alkaline Medium. Int. J. Hydrogen Energy 2018, 43, 12129–12139. [Google Scholar] [CrossRef]
  439. Yang, Y.M.; Hu, B.S.; Zhao, W.B.; Yang, Q.; Yang, F.; Ren, J.C.; Li, X.G.; Jin, Y.; Fang, L.; Pan, Q.J. Bridging N-Doped Graphene and Carbon Rich C3N4 Layers for Photopromoted Multi-Functional Electrocatalysts. Electrochim. Acta 2019, 317, 25–33. [Google Scholar] [CrossRef]
  440. Lei, G.Y.; Ma, J.W.; Zhao, M.Y.; Wu, S.; He, H.W.; Qi, H.; Peng, W.C.; Fan, X.B.; Zhang, G.L.; Zhang, F.B.; et al. Nitrogen-Carbon Materials Base on Pyrolytic Graphene Hydrogel for Oxygen Reduction. J. Colloid. Interface Sci. 2021, 602, 274–281. [Google Scholar] [CrossRef]
  441. Marbaniang, P.; Kapse, S.; Ingavale, S.; Thapa, R.; Kakade, B. Nitrogen Doping Derived Bridging of Graphene and Carbon Nanotube Composite for Oxygen Electroreduction. Int. J. Energy Res. 2021, 45, 21293–21306. [Google Scholar] [CrossRef]
  442. Skorupska, M.; Ilnicka, A.; Lukaszewicz, J.P. Modified Graphene Foam as a High-Performance Catalyst for Oxygen Reduction Reaction. RSC Adv. 2023, 13, 25437–25442. [Google Scholar] [CrossRef]
  443. Roy, D.; Sarkar, S.; Bhattacharjee, K.; Panigrahi, K.; Das, B.K.; Sardar, K.; Chattopadhyay, K.K. Site Speci Fic Nitrogen Incorporation in Reduced Graphene Oxide Using Imidazole as a Novel Reducing Agent for Ef Ficient Oxygen Reduction Reaction and Improved Supercapacitive Performance. Carbon 2020, 166, 361–373. [Google Scholar] [CrossRef]
  444. Wu, L.F.; Feng, H.B.; Liu, M.J.; Zhang, K.X.; Li, J.H. Graphene-Based Hollow Spheres as Efficient Electrocatalysts for Oxygen Reduction. Nanoscale 2013, 5, 10839–10843. [Google Scholar] [CrossRef] [PubMed]
  445. Bo, X.J.; Han, C.; Zhang, Y.F.; Guo, L.P. Confined Nanospace Synthesis of Less Aggregated and Porous Nitrogen-Doped Graphene as Metal-Free Electrocatalysts for Oxygen Reduction Reaction in Alkaline Solution. ACS Appl. Mater. Interfaces 2014, 6, 3023–3030. [Google Scholar] [CrossRef] [PubMed]
  446. Hayes, W.I.; Lubarsky, G.; Li, M.X.; Papakonstantinou, P. Mechanical Exfoliation of Graphite in 1-Butyl-3-Methylimidazolium Hexafluorophosphate (BMIM-PF6) Providing Graphene Nanoplatelets That Exhibit Enhanced Electrocatalysis. J. Power Sources 2014, 271, 312–325. [Google Scholar] [CrossRef]
  447. de Lima, F.; Maia, G. Oxidized/Reduced Graphene Nanoribbons Facilitate Charge Transfer to the Fe(CN)(6)(3-)/Fe(CN)(6)(4-) Redox Couple and towards Oxygen Reduction. Nanoscale 2015, 7, 6193–6207. [Google Scholar] [CrossRef]
  448. Zhang, Z.Y.; Cao, T.F.; Liu, S.S.; Duan, X.M.; Liu, L.M.; Wang, S.; Liu, Y.Q. Substrate-Induced Synthesis of Nitrogen-Doped Holey Graphene Nanocapsules for Advanced Metal-Free Bifunctional Electrocatalysts. Part. Part. Syst. Charact. 2017, 34. [Google Scholar] [CrossRef]
  449. Rybarczyk, M.K.; Gontarek, E.; Lieder, M.; Titirici, M.M. Salt Melt Synthesis of Curved Nitrogen-Doped Carbon Nanostructures: ORR Kinetics Boost. Appl. Surf. Sci. 2018, 435, 543–551. [Google Scholar] [CrossRef]
  450. Yang, W.L.; Zhou, M.H.; Liang, L. Highly Efficient In-Situ Metal-Free Electrochemical Advanced Oxidation Process Using Graphite Felt Modified with N-Doped Graphene. Chem. Eng. J. 2018, 338, 700–708. [Google Scholar] [CrossRef]
  451. Ilnicka, A.; Skorupska, M.; Romanowski, P.; Kamedulski, P.; Lukaszewicz, J.P. Improving the Performance of Zn-Air Batteries with N-Doped Electroexfoliated Graphene. Materials 2020, 13, 2115. [Google Scholar] [CrossRef]
  452. Ding, W.; Wei, Z.D.; Chen, S.G.; Qi, X.Q.; Yang, T.; Hu, J.S.; Wang, D.; Wan, L.J.; Alvi, S.F.; Li, L. Space-Confinement-Induced Synthesis of Pyridinic- and Pyrrolic-Nitrogen-Doped Graphene for the Catalysis of Oxygen Reduction. Angew. Chem. Int. Ed. 2013, 52, 11755–11759. [Google Scholar] [CrossRef]
  453. Lai, L.F.; Potts, J.R.; Zhan, D.; Wang, L.; Poh, C.K.; Tang, C.H.; Gong, H.; Shen, Z.X.; Jianyi, L.Y.; Ruoff, R.S. Exploration of the Active Center Structure of Nitrogen-Doped Graphene-Based Catalysts for Oxygen Reduction Reaction. Energy Environ. Sci. 2012, 5, 7936–7942. [Google Scholar] [CrossRef]
  454. Yang, S.B.; Zhi, L.J.; Tang, K.; Feng, X.L.; Maier, J.; Mullen, K. Efficient Synthesis of Heteroatom (N or S)-Doped Graphene Based on Ultrathin Graphene Oxide-Porous Silica Sheets for Oxygen Reduction Reactions. Adv. Funct. Mater. 2012, 22, 3634–3640. [Google Scholar] [CrossRef]
  455. Zhang, Y.W.; Ge, J.; Wang, L.; Wang, D.H.; Ding, F.; Tao, X.M.; Chen, W. Manageable N-Doped Graphene for High Performance Oxygen Reduction Reaction. Sci. Rep. 2013, 3, 2771. [Google Scholar] [CrossRef]
  456. Liu, M.K.; Song, Y.F.; He, S.X.; Tjiu, W.W.; Pan, J.S.; Xia, Y.Y.; Liu, T.X. Nitrogen-Doped Graphene Nanoribbons as Efficient Metal-Free Electrocatalysts for Oxygen Reduction. ACS Appl. Mater. Interfaces 2014, 6, 4214–4222. [Google Scholar] [CrossRef]
  457. Wang, L.; Yin, F.X.; Yao, C.X. N-Doped Graphene as a Bifunctional Electrocatalyst for Oxygen Reduction and Oxygen Evolution Reactions in an Alkaline Electrolyte. Int. J. Hydrogen Energy 2014, 39, 15913–15919. [Google Scholar] [CrossRef]
  458. Li, J.J.; Zhang, Y.M.; Zhang, X.H.; Han, J.C.; Wang, Y.; Gu, L.; Zhang, Z.H.; Wang, X.J.; Jian, J.K.; Xu, P.; et al. Direct Transformation from Graphitic C3N4 to Nitrogen-Doped Graphene: An Efficient Metal-Free Electrocatalyst for Oxygen Reduction Reaction. ACS Appl. Mater. Interfaces 2015, 7, 19626–19634. [Google Scholar] [CrossRef]
  459. Ma, R.G.; Zhou, Y.; Li, P.X.; Chen, Y.F.; Wang, J.C.; Liu, Q. Self-Assembly of Nitrogen-Doped Graphene-Wrapped Carbon Nanoparticles as an Efficient Electrocatalyst for Oxygen Reduction Reaction. Electrochim. Acta 2016, 216, 347–354. [Google Scholar] [CrossRef]
  460. Zhou, X.J.; Tang, S.; Yin, Y.; Sun, S.H.; Qiao, J.L. Hierarchical Porous N-Doped Graphene Foams with Superior Oxygen Reduction Reactivity for Polymer Electrolyte Membrane Fuel Cells. Appl. Energy 2016, 175, 459–467. [Google Scholar] [CrossRef]
  461. Tang, S.; Zhou, X.J.; Xu, N.N.; Bai, Z.Y.; Qiao, J.L.; Zhang, J.J. Template-Free Synthesis of Three-Dimensional Nanoporous N-Doped Graphene for High Performance Fuel Cell Oxygen Reduction Reaction in Alkaline Media. Appl. Energy 2016, 175, 405–413. [Google Scholar] [CrossRef]
  462. Hou, Z.S.; Jin, Y.Q.; Xi, X.; Huang, T.; Wu, D.Q.; Xu, P.M.; Liu, R.L. Hierarchically Porous Nitrogen-Doped Graphene Aerogels as Efficient Metal-Free Oxygen Reduction Catalysts. J. Colloid. Interface Sci. 2017, 488, 317–321. [Google Scholar] [CrossRef]
  463. Kong, D.W.; Yuan, W.J.; Li, C.; Song, J.M.; Xie, A.J.; Shen, Y.H. Synergistic Effect of Nitrogen-Doped Hierarchical Porous Carbon/Graphene with Enhanced Catalytic Performance for Oxygen Reduction Reaction. Appl. Surf. Sci. 2017, 393, 144–150. [Google Scholar] [CrossRef]
  464. Li, Y.; Yang, J.; Zhao, N.; Huang, J.P.; Zhou, Y.Z.; Xu, K. Facile Fabrication of N-Doped Three-Dimensional Reduced Graphene Oxide as a Superior Electrocatalyst for Oxygen Reduction Reaction. Appl. Catal. A Gen. 2017, 534, 30–39. [Google Scholar] [CrossRef]
  465. Qin, L.; Ding, R.M.; Wang, H.X.; Wu, J.H.; Wang, C.H.; Zhang, C.H.; Xu, Y.; Wang, L.; Lv, B.L. Facile Synthesis of Porous Nitrogen-Doped Holey Graphene as an Efficient Metal-Free Catalyst for the Oxygen Reduction Reaction. Nano Res. 2017, 10, 305–319. [Google Scholar] [CrossRef]
  466. Hassani, S.S.; Samiee, L.; Ghasemy, E.; Rashidi, A.; Ganjali, M.R.; Tasharrofi, S. Porous Nitrogen-Doped Graphene Prepared through Pyrolysis of Ammonium Acetate as an Efficient ORR Nanocatalyst. Int. J. Hydrogen Energy 2018, 43, 15941–15951. [Google Scholar] [CrossRef]
  467. Ma, R.G.; Xing, R.H.; Lin, G.X.; Zhou, Y.; Liu, Q.; Yang, M.H.; Hu, C.; Yan, K.; Wang, J.C. Graphene-Wrapped Nitrogen-Doped Hollow Carbon Spheres for High-Activity Oxygen Electroreduction. Mater. Chem. Front. 2018, 2, 1489–1497. [Google Scholar] [CrossRef]
  468. Sibul, R.; Kibena-Poldsepp, E.; Ratso, S.; Kook, M.; Kaarik, M.; Merisalu, M.; Paiste, P.; Leis, J.; Sammelselg, V.; Tammeveski, K. Nitrogen-Doped Carbon-Based Electrocatalysts Synthesised by Ball-Milling. Electrochem. Commun. 2018, 93, 39–43. [Google Scholar] [CrossRef]
  469. Xue, Q.; Ding, Y.; Xue, Y.Y.; Li, F.; Chen, P.; Chen, Y. 3D Nitrogen-Doped Graphene Aerogels as Efficient Electrocatalyst for the Oxygen Reduction Reaction. Carbon 2018, 139, 137–144. [Google Scholar] [CrossRef]
  470. Zhao, L.; Sui, X.L.; Li, J.Z.; Zhang, J.J.; Zhang, L.M.; Huang, G.S.; Wang, Z.B. Supramolecular Assembly Promoted Synthesis of Three-Dimensional Nitrogen Doped Graphene Frameworks as Efficient Electrocatalyst for Oxygen Reduction Reaction and Methanol Electrooxidation. Appl. Catal. B Environ. 2018, 231, 224–233. [Google Scholar] [CrossRef]
  471. Song, M.; Zhao, J.; Meng, Y.; Riekehr, L.; Hou, P.X.; Grennberg, H.; Zhang, Z.B. Nitrogen-Doped Reduced Graphene Oxide Hydrogel Achieved via a One-Step Hydrothermal Process. ChemNanoMat 2019, 5, 1144–1151. [Google Scholar] [CrossRef]
  472. Xie, B.B.; Zhang, Y.; Zhang, R.J. Pure Nitrogen-Doped Graphene Aerogel with Rich Micropores Yields High ORR Performance. Mater. Sci. Eng. B-adv. Func. Solid-State Mater. 2019, 242, 1–5. [Google Scholar] [CrossRef]
  473. Cardoso, E.S.F.; Fortunato, G.V.; Palm, I.; Kibena-Poldsepp, E.; Greco, A.S.; Junior, J.L.R.; Kikas, A.; Merisalu, M.; Kisand, V.; Sammelselg, V.; et al. Effects of N and O Groups for Oxygen Reduction Reaction on One- and Two-Dimensional Carbonaceous Materials. Electrochim. Acta 2020, 344, 136052. [Google Scholar] [CrossRef]
  474. Behan, J.A.; Mates-Torres, E.; Stamatin, S.N.; Dominguez, C.; Iannaci, A.; Fleischer, K.; Hoque, M.K.; Perova, T.S.; Garcia-Melchor, M.; Colavita, P.E. Untangling Cooperative Effects of Pyridinic and Graphitic Nitrogen Sites at Metal-Free N-Doped Carbon Electrocatalysts for the Oxygen Reduction Reaction. Small 2019, 15, 1902081. [Google Scholar] [CrossRef]
  475. Yan, P.; Liu, J.; Yuan, S.D.; Liu, Y.J.; Cen, W.L.; Chen, Y.Q. The Promotion Effects of Graphitic and Pyridinic N Combinational Doping on Graphene for ORR. Appl. Surf. Sci. 2018, 445, 398–403. [Google Scholar] [CrossRef]
  476. Benavides, R.; Gallardo, C.; Fernandez, S.; De-Casas, E.; Morales-Acosta, D. Influence of Doping Level on the Electrocatalytic Properties for Oxygen Reduction Reaction of N-Doped Reduced Graphene Oxide. Int. J. Hydrogen Energy 2021, 46, 26040–26052. [Google Scholar] [CrossRef]
  477. Gong, L.; Sun, J.; Li, X.D.; Huang, B.; Yang, G.C.; Liu, Y.S. One-Step and Controllable Synthesis of Active N-Rich Graphene Nanoclusters-CNT Composite via an Ultrafast Deflagration Reaction for Oxygen Reduction Electrocatalysis. J. Mater. Sci. 2021, 56, 6349–6360. [Google Scholar] [CrossRef]
  478. Karunagaran, R.; Tran, D.; Tung, T.T.; Shearer, C.; Losic, D. A Unique Synthesis of Macroporous N-Doped Carbon Composite Catalyst for Oxygen Reduction Reaction. Nanomaterials 2021, 11, 43. [Google Scholar] [CrossRef]
  479. Song, R.L.; Cao, X.T.; Xu, J.; Zhou, X.S.; Wang, X.; Yuan, N.Y.; Ding, J.N. O,N-Codoped 3D Graphene Hollow Sphere Derived from Metal-Organic Frameworks as Oxygen Reduction Reaction Electrocatalysts for Zn-Air Batteries. Nanoscale 2021, 13, 6174–6183. [Google Scholar] [CrossRef]
  480. Song, R.L.; Cao, X.T.; Zhou, X.S.; Yuan, N.Y. N-Doped Graphene Supported on N-RGO Nanosheets as Metal-Free Oxygen Reduction Reaction Electrocatalysts for Zn-Air Batteries. N. J. Chem. 2021, 45, 21716–21724. [Google Scholar] [CrossRef]
  481. Zhao, J.; Li, Q.Q.; Zhang, Q.C.; Liu, R. Carbon Tube-Graphene Heterostructure with Different N-Doping Configurations Induces an Electrochemically Active-Active Interface for Efficient Oxygen Electrocatalysis. Chem. Eng. J. 2022, 431, 9. [Google Scholar] [CrossRef]
  482. Zhang, Y.M.; Zhang, H.X.; Sha, W.B.; Song, Y.H.; Liu, P.Z.; Liu, R.; Hou, Y.; Wei, H.; Xu, B.S.; Cao, T.F.; et al. N-Doped Graphene Nanoribbons Intertwined on 3D Graphene Skeleton as Superior Metal-Free Electrocatalyst for Oxygen Reduction. Colloids Surf. A Physicochem. Eng. Asp. 2022, 652, 10. [Google Scholar] [CrossRef]
  483. Han, Y.J.; Shen, Y.Q.; Song, Y.H.; Zhang, H.X.; Liu, P.Z.; Guo, J.J. Edge-Rich Graphene Nanospheres with Ultra-High Nitrogen Loading Metal-Free Electrocatalysts for Boosted Oxygen Reduction. ChemElectroChem 2022, 9, 8. [Google Scholar] [CrossRef]
  484. Eledath, A.N.; Poulose, A.E.; Muthukrishnan, A. O-Functionalization of N-Doped Reduced Graphene Oxide for Topological Defect-Driven Oxygen Reduction. ACS Appl. Nano Mater. 2022, 5, 10528–10536. [Google Scholar] [CrossRef]
  485. Dogan, M.Z.; Gökcen, D.; Bayram, C. Effects of Heteroatom Doping and Physicochemical Character on the Electrochemical Properties of Graphene Sheets. ChemNanoMat 2023, 9, 202300393. [Google Scholar] [CrossRef]
  486. Wang, S.; Song, W.N.; Hu, M.J.; Li, F.C.; Song, Y.P.; Ju, R.; Xu, K.D.; Zhou, H.T. Green Synthesis of N-Doped Graphene Nanosheets with Multi-Wrinkled Textural Properties for Boosting Oxygen Reduction Reaction in Glucose Fuel Cell. Int. J. Hydrogen Energy 2024, 72, 226–236. [Google Scholar] [CrossRef]
  487. Carrillo-Rodriguez, J.C.; Alonso-Lemus, I.L.; Siller-Ceniceros, A.A.; Martinez, E.; Piza-Ruiz, P.; Vargas-Gutierrez, G.; Rodriguez-Varela, F.J. Easy Synthesis of N-Doped Graphene by Milling Exfoliation with Electrocatalytic Activity towards the Oxygen Reduction Reaction (ORR). Int. J. Hydrogen Energy 2017, 42, 30383–30388. [Google Scholar] [CrossRef]
  488. Su, P.; Zhou, M.H.; Lu, X.Y.; Yang, W.L.; Ren, G.B.; Cai, J.J. Electrochemical Catalytic Mechanism of N-Doped Graphene for Enhanced H2O2 Yield and in-Situ Degradation of Organic Pollutant. Appl. Catal. B Environ. 2019, 245, 583–595. [Google Scholar] [CrossRef]
  489. Lu, Z.J.; Xu, M.W.; Bao, S.J.; Tan, K.; Chai, H.; Cai, C.J.; Ji, C.C.; Zhang, Q. Facile Preparation of Nitrogen-Doped Reduced Graphene Oxide as a Metal-Free Catalyst for Oxygen Reduction Reaction. J. Mater. Sci. 2013, 48, 8101–8107. [Google Scholar] [CrossRef]
  490. Yu, D.S.; Wei, L.; Jiang, W.C.; Wang, H.; Sun, B.; Zhang, Q.; Goh, K.L.; Si, R.M.; Chen, Y. Nitrogen Doped Holey Graphene as an Efficient Metal-Free Multifunctional Electrochemical Catalyst for Hydrazine Oxidation and Oxygen Reduction. Nanoscale 2013, 5, 3457–3464. [Google Scholar] [CrossRef]
  491. Jiang, Z.J.; Jiang, Z.Q.; Chen, W.H. The Role of Holes in Improving the Performance of Nitrogen-Doped Holey Graphene as an Active Electrode Material for Supercapacitor and Oxygen Reduction Reaction. J. Power Sources 2014, 251, 55–65. [Google Scholar] [CrossRef]
  492. Liu, J.F.; Takeshi, D.; Orejon, D.; Sasaki, K.; Lyth, S.M. Defective Nitrogen-Doped Graphene Foam: A Metal-Free, Non-Precious Electrocatalyst for the Oxygen Reduction Reaction in Acid. J. Electrochem. Soc. 2014, 161, F544–F550. [Google Scholar] [CrossRef]
  493. Zhao, Z.K.; Dai, Y.T.; Ge, G.F.; Mao, Q.; Rong, Z.M.; Wang, G.R. A Facile Approach to Fabricate an N-Doped Mesoporous Graphene/Nanodiamond Hybrid Nanocomposite with Synergistically Enhanced Catalysis. ChemCatChem 2015, 7, 1070–1077. [Google Scholar] [CrossRef]
  494. Zhao, H.Y.; Sun, C.H.; Jin, Z.; Wang, D.W.; Yan, X.C.; Chen, Z.G.; Zhu, G.S.; Yao, X.D. Carbon for the Oxygen Reduction Reaction: A Defect Mechanism. J. Mater. Chem. A Mater. 2015, 3, 11736–11739. [Google Scholar] [CrossRef]
  495. Bai, X.; Shi, Y.; Guo, J.; Gao, L.; Wang, K.; Du, Y.; Ma, T. Catalytic Activities Enhanced by Abundant Structural Defects and Balanced N Distribution of N-Doped Graphene in Oxygen Reduction Reaction. J. Power Sources 2016, 306, 85–91. [Google Scholar] [CrossRef]
  496. Tang, C.; Wang, H.F.; Chen, X.; Li, B.Q.; Hou, T.Z.; Zhang, B.S.; Zhang, Q.; Titirici, M.M.; Wei, F. Topological Defects in Metal-Free NanoCarbon for Oxygen Electrocatalysis. Adv. Mater. 2016, 28, 6845–6851. [Google Scholar] [CrossRef]
  497. Wang, H.F.; Tang, C.; Zhang, Q. Template Growth of Nitrogen-Doped Mesoporous Graphene on Metal Oxides and Its Use as a Metal-Free Bifunctional Electrocatalyst for Oxygen Reduction and Evolution Reactions. Catal. Today 2018, 301, 25–31. [Google Scholar] [CrossRef]
  498. Dumont, J.H.; Martinez, U.; Artyushkova, K.; Purdy, G.M.; Dattelbaum, A.M.; Zelenay, P.; Mohite, A.; Atanassov, P.; Gupta, G. Nitrogen-Doped Graphene Oxide Electrocatalysts for the Oxygen Reduction Reaction. ACS Appl. Nano Mater. 2019, 2, 1675–1682. [Google Scholar] [CrossRef]
  499. Han, L.; Sun, Y.Y.; Li, S.; Cheng, C.; Halbig, C.E.; Feicht, P.; Hubner, J.L.; Strasser, P.; Eigler, S. In-Plane Carbon Lattice-Defect Regulating Electrochemical Oxygen Reduction to Hydrogen Peroxide Production over Nitrogen-Doped Graphene. ACS Catal. 2019, 9, 1283–1288. [Google Scholar] [CrossRef]
  500. Rivera-Gavidia, L.M.; de la Puente, I.F.; Hernandez-Rodriguez, M.A.; Celorrio, V.; Sebastian, D.; Lazaro, M.J.; Pastor, E.; Garcia, G. Bi-Functional Carbon-Based Catalysts for Unitized Regenerative Fuel Cells. J. Catal. 2020, 387, 138–144. [Google Scholar] [CrossRef]
  501. Liu, S.; Zhang, Y.C.; Ge, B.H.; Zheng, F.C.; Zhang, N.; Zuo, M.; Yang, Y.; Chen, Q.W. Constructing Graphitic-Nitrogen-Bonded Pentagons in Interlayer-Expanded Graphene Matrix toward Carbon-Based Electrocatalysts for Acidic Oxygen Reduction Reaction. Adv. Mater. 2021, 33, 11. [Google Scholar] [CrossRef]
  502. Yuasa, M.; Tanaka, M.; Shimizu, M.; Yoshida, M. Mechanochemical Synthesis of Nitrogen-Doped and Sulfur-Doped Multilayer Graphene for Use in Bifunctional Oxygen Electrodes. J. Electrochem. Soc. 2022, 169, 10. [Google Scholar] [CrossRef]
  503. Satoh, K.; Odawara, T.; Yamazaki, Y.; Murakami, T.; Ono, H.; Iwamura, S.; Mukai, S.R.; Ogino, I. Benefits of Using Rapid Microwave Heating in the Synthesis of Metal-Free Carbon Electrocatalysts. Ind. Eng. Chem. Res. 2024, 63, 4825–4837. [Google Scholar] [CrossRef]
  504. Wu, J.J.; Zhang, D.; Wang, Y.; Hou, B.R. Electrocatalytic Activity of Nitrogen-Doped Graphene Synthesized via a One-Pot Hydrothermal Process towards Oxygen Reduction Reaction. J. Power Sources 2013, 227, 185–190. [Google Scholar] [CrossRef]
  505. Jeon, I.Y.; Choi, H.J.; Ju, M.J.; Choi, I.T.; Lim, K.; Ko, J.; Kim, H.K.; Kim, J.C.; Lee, J.J.; Shin, D.; et al. Direct Nitrogen Fixation at the Edges of Graphene Nanoplatelets as Efficient Electrocatalysts for Energy Conversion. Sci. Rep. 2013, 3, 2260. [Google Scholar] [CrossRef]
  506. Benson, J.; Xu, Q.; Wang, P.; Shen, Y.T.; Sun, L.T.; Wang, T.Y.; Li, M.X.; Papakonstantinou, P. Tuning the Catalytic Activity of Graphene Nanosheets for Oxygen Reduction Reaction via Size and Thickness Reduction. ACS Appl. Mater. Interfaces 2014, 6, 19726–19736. [Google Scholar] [CrossRef]
  507. Kwak, D.; Khetan, A.; Noh, S.; Pitsch, H.; Han, B. First Principles Study of Morphology, Doping Level, and Water Solvation Effects on the Catalytic Mechanism of Nitrogen-Doped Graphene in the Oxygen Reduction Reaction. ChemCatChem 2014, 6, 2662–2670. [Google Scholar] [CrossRef]
  508. Li, M.T.; Zhang, L.P.; Xu, Q.; Niu, J.B.; Xia, Z.H. N-Doped Graphene as Catalysts for Oxygen Reduction and Oxygen Evolution Reactions: Theoretical Considerations. J. Catal. 2014, 314, 66–72. [Google Scholar] [CrossRef]
  509. Gao, Y.X.; Zhang, L.P.; Xia, Z.H.; Li, C.M.; Dai, L.M. Hole-Punching for Enhancing Electrocatalytic Activities of 2D Graphene Electrodes: Less Is More. J. Chem. Phys. 2020, 153, 074701. [Google Scholar] [CrossRef]
  510. Hassan, M.; Haque, E.; Reddy, K.R.; Minett, A.I.; Chen, J.; Gomes, V.G. Edge-Enriched Graphene Quantum Dots for Enhanced Photo-Luminescence and Supercapacitance. Nanoscale 2014, 6, 11988–11994. [Google Scholar] [CrossRef]
  511. Li, Q.Q.; Zhang, S.; Dai, L.M.; Li, L.S. Nitrogen-Doped Colloidal Graphene Quantum Dots and Their Size-Dependent Electrocatalytic Activity for the Oxygen Reduction Reaction. J. Am. Chem. Soc. 2012, 134, 18932–18935. [Google Scholar] [CrossRef]
  512. Favaro, M.; Ferrighi, L.; Fazio, G.; Colazzo, L.; Di Vaentin, C.; Durante, C.; Sedona, F.; Gennaro, A.; Agnoli, S.; Granozzi, G. Single and Multiple Doping in Graphene Quantum Dots: Unraveling the Origin of Selectivity in the Oxygen Reduction Reaction. ACS Catal. 2015, 5, 129–144. [Google Scholar] [CrossRef]
  513. Wang, M.R.; Fang, J.; Hu, L.T.; Lai, Y.Q.; Liu, Z.Y. Defects-Rich Graphene/Carbon Quantum Dot Composites as Highly Efficient Electrocatalysts for Aqueous Zinc/Air Batteries. Int. J. Hydrogen Energy 2017, 42, 21305–21310. [Google Scholar] [CrossRef]
  514. Wang, M.; Wu, Z.P.; Dai, L.M. Graphitic Carbon Nitrides Supported by Nitrogen-Doped Graphene as Efficient Metal-Free Electrocatalysts for Oxygen Reduction. J. Electroanal. Chem. 2015, 753, 16–20. [Google Scholar] [CrossRef]
  515. Chao, L.; Qin, Y.; Liu, Y.; Kong, Y.; Chu, F.Q. Electrochemically Exfoliating Graphite into N-Doped Graphene and Its Use as a High Efficient Electrocatalyst for Oxygen Reduction Reaction. J. Solid State Electrochem. 2017, 21, 1287–1295. [Google Scholar] [CrossRef]
  516. Kim, H.W.; Bukas, V.J.; Park, H.; Park, S.; Diederichsen, K.M.; Lim, J.; Cho, Y.H.; Kim, J.; Kim, W.; Han, T.H.; et al. Mechanisms of Two-Electron and Four-Electron Electrochemical Oxygen Reduction Reactions at Nitrogen-Doped Reduced Graphene Oxide. ACS Catal. 2020, 10, 852–863. [Google Scholar] [CrossRef]
  517. Jahan, M.; Li, K.; Zhao, G.L. Electric Field Poling Effect on the Electrocatalytic Properties of Nitrogen-Functionalized Graphene Nanosheets. Energy Technol. 2018, 6, 2408–2418. [Google Scholar] [CrossRef]
  518. Zhou, X.J.; Bai, Z.Y.; Wu, M.J.; Qiao, J.L.; Chen, Z.W. 3-Dimensional Porous N-Doped Graphene Foam as a Non-Precious Catalyst for the Oxygen Reduction Reaction. J. Mater. Chem. A Mater. 2015, 3, 3343–3350. [Google Scholar] [CrossRef]
  519. Mao, S.; Lu, G.H.; Chen, J.H. Three-Dimensional Graphene-Based Composites for Energy Applications. Nanoscale 2015, 7, 6924–6943. [Google Scholar] [CrossRef]
  520. Mo, Z.Y.; Zheng, R.P.; Peng, H.L.; Liang, H.G.; Liao, S.J. Nitrogen-Doped Graphene Prepared by a Transfer Doping Approach for the Oxygen Reduction Reaction Application. J. Power Sources 2014, 245, 801–807. [Google Scholar] [CrossRef]
  521. Zhao, Y.; Hu, C.G.; Hu, Y.; Cheng, H.H.; Shi, G.Q.; Qu, L.T. A Versatile, Ultralight, Nitrogen-Doped Graphene Framework. Angew. Chem. Int. Ed. 2012, 51, 11371–11375. [Google Scholar] [CrossRef]
  522. Liang, J.; Du, X.; Gibson, C.; Du, X.W.; Qiao, S.Z. N-Doped Graphene Natively Grown on Hierarchical Ordered Porous Carbon for Enhanced Oxygen Reduction. Adv. Mater. 2013, 25, 6226–6231. [Google Scholar] [CrossRef] [PubMed]
  523. Zhong, H.X.; Wang, J.; Zhang, Y.W.; Xu, W.L.; Xing, W.; Xu, D.; Zhang, Y.F.; Zhang, X.B. ZIF-8 Derived Graphene-Based Nitrogen-Doped Porous Carbon Sheets as Highly Efficient and Durable Oxygen Reduction Electrocatalysts. Angew. Chem. Int. Ed. 2014, 53, 14235–14239. [Google Scholar] [CrossRef] [PubMed]
  524. Lei, Y.P.; Shi, Q.; Han, C.; Wang, B.; Wu, N.; Wang, H.; Wang, Y.D. N-Doped Graphene Grown on Silk Cocoon-Derived Interconnected Carbon Fibers for Oxygen Reduction Reaction and Photocatalytic Hydrogen Production. Nano Res. 2016, 9, 2498–2509. [Google Scholar] [CrossRef]
  525. Jiang, Z.J.; Jiang, Z.Q. Reduction of the Oxygen Reduction Reaction Overpotential of Nitrogen-Doped Graphene by Designing It to a Microspherical Hollow Shape. J. Mater. Chem. A Mater. 2014, 2, 14071–14081. [Google Scholar] [CrossRef]
  526. Wang, J.; Jin, H.L.; He, Y.H.; Lin, D.J.; Liu, A.L.; Wang, S.; Wang, J.C. The Selective Formation of Graphene Ranging from Two-Dimensional Sheets to Three-Dimensional Mesoporous Nanospheres. Nanoscale 2014, 6, 7204–7208. [Google Scholar] [CrossRef]
  527. Chen, L.; Du, R.; Zhu, J.H.; Mao, Y.Y.; Xue, C.; Zhang, N.; Hou, Y.L.; Zhang, J.; Yi, T. Three-Dimensional Nitrogen-Doped Graphene Nanoribbons Aerogel as a Highly EffiCient Catalyst for the Oxygen Reduction Reaction. Small 2015, 11, 1423–1429. [Google Scholar] [CrossRef]
  528. Yang, Y.; Liu, T.Y.; Liao, Q.; Ye, D.D.; Zhu, X.; Li, J.; Zhang, P.Q.; Peng, Y.; Chen, S.W.; Li, Y. A Three-Dimensional Nitrogen-Doped Graphene Aerogel-Activated Carbon Composite Catalyst That Enables Low-Cost Microfluidic Microbial Fuel Cells with Superior Performance. J. Mater. Chem. A Mater. 2016, 4, 15913–15919. [Google Scholar] [CrossRef]
  529. Baskakov, S.A.; Manzhos, R.A.; Lobach, A.S.; Baskakova, Y.V.; Kulikov, A.V.; Martynenko, V.M.; Milovich, F.O.; Kumar, Y.; Michtchenko, A.; Kabachkov, E.N.; et al. Properties of a Granulated Nitrogen-Doped Graphene Oxide Aerogel. J. Non Cryst. Solids 2018, 498, 236–243. [Google Scholar] [CrossRef]
  530. Shin, Y.E.; Sa, Y.J.; Park, S.; Lee, J.; Shin, K.H.; Joo, S.H.; Ko, H. An Ice-Templated, PH-Tunable Self-Assembly Route to Hierarchically Porous Graphene Nanoscroll Networks. Nanoscale 2014, 6, 9734–9741. [Google Scholar] [CrossRef]
  531. Patra, S.; Choudhary, R.; Roy, E.; Madhuri, R.; Sharma, P.K. Heteroatom-Doped Graphene “Idli”: A Green and Foody Approach towards Development of Metal Free Bifunctional Catalyst for Rechargeable Zinc-Air Battery. Nano Energy 2016, 30, 118–129. [Google Scholar] [CrossRef]
  532. Feng, W.J.; Lin, Y.X.; Zhao, T.J.; Zhang, P.F.; Su, H.; Lv, L.B.; Li, X.H.; Chen, J.S. Direct Reduction of Oxygen Gas over Dendritic Carbons with Hierarchical Porosity: Beyond the Diffusion Limitation. Inorg. Chem. Front. 2018, 5, 2023–2030. [Google Scholar] [CrossRef]
  533. Begum, H.; Ahmed, M.S.; Cho, S.; Jeon, S. Simultaneous Reduction and Nitrogen Functionalization of Graphene Oxide Using Lemon for Metal-Free Oxygen Reduction Reaction. J. Power Sources 2017, 372, 116–124. [Google Scholar] [CrossRef]
  534. Shi, J.L.; Tian, G.L.; Zhang, Q.; Zhao, M.Q.; Wei, F. Customized Casting of Unstacked Graphene with High Surface Area (>1300 m(2) g(-1)) and Its Application in Oxygen Reduction Reaction. Carbon 2015, 93, 702–712. [Google Scholar] [CrossRef]
  535. Paton-Carrero, A.; de la Osa, A.; Sanchez, P.; Rodriguez-Gomez, A.; Romero, A. Towards New Routes to Increase the Electrocatalytic Activity for Oxygen Reduction Reaction of N-Doped Graphene Nanofibers. J. Electroanal. Chem. 2020, 878, 114631. [Google Scholar] [CrossRef]
  536. Geim, A.K.; Novoselov, K.S. The Rise of Graphene. Nat. Mater. 2007, 6, 183–191. [Google Scholar] [CrossRef]
  537. Kosynkin, D.V.; Higginbotham, A.L.; Sinitskii, A.; Lomeda, J.R.; Dimiev, A.; Price, B.K.; Tour, J.M. Longitudinal Unzipping of Carbon Nanotubes to Form Graphene Nanoribbons. Nature 2009, 458, 872–876. [Google Scholar] [CrossRef]
  538. Chen, P.; Xiao, T.Y.; Qian, Y.H.; Li, S.S.; Yu, S.H. A Nitrogen-Doped Graphene/Carbon Nanotube Nanocomposite with Synergistically Enhanced Electrochemical Activity. Adv. Mater. 2013, 25, 3192–3196. [Google Scholar] [CrossRef]
  539. Choi, C.H.; Chung, M.W.; Kwon, H.C.; Chung, J.H.; Woo, S.I. Nitrogen-Doped Graphene/Carbon Nanotube Self-Assembly for Efficient Oxygen Reduction Reaction in Acid Media. Appl. Catal. B Environ. 2014, 144, 760–766. [Google Scholar] [CrossRef]
  540. Ratso, S.; Kruusenberg, I.; Vikkisk, M.; Joost, U.; Shulga, E.; Kink, I.; Kallio, T.; Tammeveski, K. Highly Active Nitrogen-Doped Few-Layer Graphene/Carbon Nanotube Composite Electrocatalyst for Oxygen Reduction Reaction in Alkaline Media. Carbon 2014, 73, 361–370. [Google Scholar] [CrossRef]
  541. Tian, G.L.; Zhao, M.Q.; Yu, D.S.; Kong, X.Y.; Huang, J.Q.; Zhang, Q.; Wei, F. Nitrogen-Doped Graphene/Carbon Nanotube Hybrids: In Situ Formation on Bifunctional Catalysts and Their Superior Electrocatalytic Activity for Oxygen Evolution/Reduction Reaction. Small 2014, 10, 2251–2259. [Google Scholar] [CrossRef]
  542. Zhang, Y.; Jiang, W.J.; Zhang, X.; Guo, L.; Hu, J.S.; Wei, Z.D.; Wan, L.J. Engineering Self-Assembled N-Doped Graphene-Carbon Nanotube Composites towards Efficient Oxygen Reduction Electrocatalysts. Phys. Chem. Chem. Phys. 2014, 16, 13605–13609. [Google Scholar] [CrossRef]
  543. Shui, J.L.; Wang, M.; Du, F.; Dai, L.M. N-Doped Carbon Nanomaterials Are Durable Catalysts for Oxygen Reduction Reaction in Acidic Fuel Cells. Sci. Adv. 2015, 1, 1400129. [Google Scholar] [CrossRef] [PubMed]
  544. Qazzazie, D.; Halhouli, M.; Yurchenko, O.; Urban, G. Control over Fuel Cell Performance through Modulation of Pore Accessibility: Investigation and Modeling of Carbon Nanotubes Effects on Oxygen Reduction at N-Graphene-Based Nanocomposite. Nanotechnology 2016, 27, 475401. [Google Scholar] [CrossRef]
  545. Zhang, L.J.; Li, H.J.; Li, K.Z.; Wei, J.F.; Fu, Q.G. Synthesis of Hybrid Carbon Spheres@nitrogen-Doped Graphene/Carbon Nanotubes and Their Oxygen Reduction Activity Performance. RSC Adv. 2016, 6, 32661–32669. [Google Scholar] [CrossRef]
  546. Fu, Y.; Tian, C.G.; Liu, F.Y.; Wang, L.; Yan, H.J.; Yang, B. An Effective Poly(p-Phenylenevinylene) Polymer Adhesion Route toward Three-Dimensional Nitrogen-Doped Carbon Nanotube/Reduced Graphene Oxide Composite for Direct Electrocatalytic Oxygen Reduction. Nano Res. 2016, 9, 3364–3376. [Google Scholar] [CrossRef]
  547. Pham, D.T.; Li, B.; Lee, Y.H. Nitrogen-Doped Activated Graphene/SWCNT Hybrid for Oxygen Reduction Reaction. Curr. Appl. Phys. 2016, 16, 1242–1249. [Google Scholar] [CrossRef]
  548. Faisal, S.N.; Subramaniyam, C.M.; Haque, E.; Islam, M.M.; Noorbehesht, N.; Roy, A.K.; Islam, M.S.; Liu, H.K.; Dou, S.X.; Harris, A.T.; et al. Nanoarchitectured Nitrogen-Doped Graphene/Carbon Nanotube as High Performance Electrodes for Solid State Supercapacitors, Capacitive Deionization, Li-Ion Battery, and Metal-Free Bifunctional Electrocatalysis. ACS Appl. Energy Mater. 2018, 1, 5211–5223. [Google Scholar] [CrossRef]
  549. Xue, L.F.; Li, Y.C.; Liu, X.F.; Liu, Q.T.; Shang, J.X.; Duan, H.P.; Dai, L.M.; Shui, J.L. Zigzag Carbon as Efficient and Stable Oxygen Reduction Electrocatalyst for Proton Exchange Membrane Fuel Cells. Nat. Commun. 2018, 9, 3819. [Google Scholar] [CrossRef]
  550. Kong, F.T.; Qiao, Y.; Zhang, C.Q.; Fan, X.H.; Zhao, Q.B.; Kong, A.G.; Shan, Y.K. Oriented Synthesis of Pyridinic-N Dopant within the Highly Efficient Multifunction Carbon-Based Materials for Oxygen Transformation and Energy Storage. ACS Sustain. Chem. Eng. 2020, 8, 10431–10443. [Google Scholar] [CrossRef]
  551. McCreery, R.L. Advanced Carbon Electrode Materials for Molecular Electrochemistry. Chem. Rev. 2008, 108, 2646–2687. [Google Scholar] [CrossRef]
  552. Jee, Y.; Karimaghaloo, A.; MacedoAndrade, A.; Moon, H.; Li, Y.; Han, J.W.; Ji, S.; Ishihara, H.; Su, P.C.; Cha, S.W.; et al. Graphene-Based Oxygen Reduction Electrodes for Low Temperature Solid Oxide Fuel Cells. Fuel Cells 2017, 17, 344–352. [Google Scholar] [CrossRef]
  553. Farzaneh, A.; Goharshadi, E.K.; Gharibi, H.; Saghatoleslami, N.; Ahmadzadeh, H. Insights on the Superior Performance of Nanostructured Nitrogen-Doped Reduced Graphene Oxide in Comparison with Commercial Pt/C as Cathode Electrocatalyst Layer of Passive Direct Methanol Fuel Cell. Electrochim. Acta 2019, 306, 220–228. [Google Scholar] [CrossRef]
  554. Ma, G.X.; Zhao, J.H.; Zheng, J.F.; Zhu, Z.P. Synthesis of nitrogen-doped graphene and its catalytic activity for the oxygen reduction reaction in fuel cells. New Carbon Mater. 2012, 27, 258–265. [Google Scholar] [CrossRef]
  555. He, C.Y.; Li, Z.S.; Cai, M.L.; Cai, M.; Wang, J.Q.; Tian, Z.Q.; Zhang, X.; Shen, P.K. A Strategy for Mass Production of Self-Assembled Nitrogen-Doped Graphene as Catalytic Materials. J. Mater. Chem. A Mater. 2013, 1, 1401–1406. [Google Scholar] [CrossRef]
  556. Liu, X.F.; Antonietti, M. Moderating Black Powder Chemistry for the Synthesis of Doped and Highly Porous Graphene Nanoplatelets and Their Use in Electrocatalysis. Adv. Mater. 2013, 25, 6284–6290. [Google Scholar] [CrossRef]
  557. Du, D.H.; Li, P.C.; Ouyang, J.Y. Nitrogen-Doped Reduced Graphene Oxide Prepared by Simultaneous Thermal Reduction and Nitrogen Doping of Graphene Oxide in Air and Its Application as an Electrocatalyst. ACS Appl. Mater. Interfaces 2015, 7, 26952–26958. [Google Scholar] [CrossRef]
  558. Manju, V.; Vusa, C.S.R.; Arumugam, P.; Berchmans, S. A Facile and Versatile Electrochemical Tuning of Graphene for Oxygen Reduction Reaction in Acidic, Neutral and Alkali Media. ChemistrySelect 2017, 2, 8541–8552. [Google Scholar] [CrossRef]
  559. Chen, P.W.; Xu, C.X.; Yin, H.; Gao, X.; Qu, L.T. Shock Induced Conversion of Carbon Dioxide to Few Layer Graphene. Carbon 2017, 115, 471–476. [Google Scholar] [CrossRef]
  560. Tao, H.C.; Yan, C.; Robertson, A.W.; Gao, Y.N.; Ding, J.J.; Zhang, Y.Q.; Ma, T.; Sun, Z.Y. N-Doping of Graphene Oxide at Low Temperature for the Oxygen Reduction Reaction. Chem. Commun. 2017, 53, 873–876. [Google Scholar] [CrossRef]
  561. Kumar, S.; Gonen, S.; Friedman, A.; Elbaz, L.; Nessim, G.D. Doping and Reduction of Graphene Oxide Using Chitosan-Derived Volatile N-Heterocyclic Compounds for Metal-Free Oxygen Reduction Reaction. Carbon 2017, 120, 419–426. [Google Scholar] [CrossRef]
  562. Ustavytska, O.; Kurys, Y.; Koshechko, V.; Pokhodenko, V. One-Step Electrochemical Preparation of Multilayer Graphene Functionalized with Nitrogen. Nanoscale Res. Lett. 2017, 12, 175. [Google Scholar] [CrossRef]
  563. Wang, Y.Q.; Yu, F.; Zhu, M.Y.; Ma, C.H.; Zhao, D.; Wang, C.; Zhou, A.M.; Dai, B.; Ji, J.Y.; Guo, X.H. N-Doping of Plasma Exfoliated Graphene Oxide via Dielectric Barrier Discharge Plasma Treatment for the Oxygen Reduction Reaction. J. Mater. Chem. A Mater. 2018, 6, 2011–2017. [Google Scholar] [CrossRef]
  564. Lemes, G.; Sebastian, D.; Pastor, E.; Lazard, M.J. N-Doped Graphene Catalysts with High Nitrogen Concentration for the Oxygen Reduction Reaction. J. Power Sources 2019, 438, 227036. [Google Scholar] [CrossRef]
  565. Wang, K.; Wang, J.X.; Wu, Y.; Zhao, S.; Wang, Z.; Wang, S.C. Nitrogen-Doped Graphene Prepared by a Millisecond Photo-Thermal Process and Its Applications. Org Electron 2018, 56, 221–231. [Google Scholar] [CrossRef]
  566. Kakaei, K.; Ghadimi, G. A Green Method for Nitrogen-Doped Graphene and Its Application for Oxygen Reduction Reaction in Alkaline Media. Mater. Technol. 2021, 36, 46–53. [Google Scholar] [CrossRef]
  567. Dan, M.; Vulcu, A.; Porav, S.A.; Leostean, C.; Borodi, G.; Cadar, O.; Berghian-Grosan, C. Eco-Friendly Nitrogen-Doped Graphene Preparation and Design for the Oxygen Reduction Reaction. Molecules 2021, 26, 3858. [Google Scholar] [CrossRef]
  568. Shim, Y.; Han, J.; Sa, Y.J.; Lee, S.; Choi, K.; Oh, J.; Kim, S.; Joo, S.H.; Park, S. Electrocatalytic Performances of Heteroatom-Containing Functionalities in N-Doped Reduced Graphene Oxides. J. Ind. Eng. Chem. 2016, 42, 149–156. [Google Scholar] [CrossRef]
  569. Kim, S.; Choi, K.; Shim, Y.; Lee, S.; Park, S. The Effect of KOH Treatment on the Chemical Structure and Electrocatalytic Activity of Reduced Graphene Oxide Materials. Chem. A Eur. J. 2016, 22, 11435–11440. [Google Scholar] [CrossRef]
  570. Kim, H.W.; Park, H.; Roh, J.S.; Shin, J.E.; Lee, T.H.; Zhang, L.; Cho, Y.H.; Yoon, H.W.; Bukas, V.J.; Guo, J.H.; et al. Carbon Defect Characterization of Nitrogen-Doped Reduced Graphene Oxide Electrocatalysts for the Two-Electron Oxygen Reduction Reaction. Chem. Mater. 2019, 31, 3967–3973. [Google Scholar] [CrossRef]
  571. Fan, X.F.; Zheng, W.T.; Kuo, J.L. Oxygen Reduction Reaction on Active Sites of Heteroatom-Doped Graphene. RSC Adv. 2013, 3, 5498–5505. [Google Scholar] [CrossRef]
  572. del Cueto, M.; Ocon, P.; Poyato, J.M.L. Comparative Study of Oxygen Reduction Reaction Mechanism on Nitrogen-, Phosphorus-, and Boron-Doped Graphene Surfaces for Fuel Cell Applications. J. Phys. Chem. C 2015, 119, 2004–2009. [Google Scholar] [CrossRef]
  573. Gracia-Espino, E. Behind the Synergistic Effect Observed on Phosphorus Nitrogen Codoped Graphene during the Oxygen Reduction Reaction. J. Phys. Chem. C 2016, 120, 27849–27857. [Google Scholar] [CrossRef]
  574. Flyagina, I.S.; Hughes, K.J.; Mielczarek, D.C.; Ingham, D.B.; Pourkashanian, M. Identifying the Catalytic Active Sites in Heteroatom-Doped Graphene for the Oxygen Reduction Reaction. Fuel Cells 2016, 16, 568–576. [Google Scholar] [CrossRef]
  575. Dobrota, A.S.; Pasti, I.A.; Mentus, S.V.; Skorodumova, N. V A DFT Study of the Interplay between Dopants and Oxygen Functional Groups over the Graphene Basal Plane—Implications in Energy-Related Applications. Phys. Chem. Chem. Phys. 2017, 19, 8530–8540. [Google Scholar] [CrossRef] [PubMed]
  576. Li, F.; Shu, H.B.; Liu, X.T.; Shi, Z.Y.; Liang, P.; Chen, X.S. Electrocatalytic Activity and Design Principles of Heteroatom-Doped Graphene Catalysts for Oxygen-Reduction Reaction. J. Phys. Chem. C 2017, 121, 14434–14442. [Google Scholar] [CrossRef]
  577. Sinthika, S.; Waghmare, U.V.; Thapa, R. Structural and Electronic Descriptors of Catalytic Activity of Graphene-Based Materials: First-Principles Theoretical Analysis. Small 2018, 14, 1703609. [Google Scholar] [CrossRef]
  578. Zou, X.L.; Wang, L.Q.; Yakobson, B.I. Mechanisms of the Oxygen Reduction Reaction on B- and/or N-Doped Carbon NanoMaterials with Curvature and Edge Effects. Nanoscale 2018, 10, 1129–1134. [Google Scholar] [CrossRef]
  579. Aguilar-Galindo, F.; Ocon, P.; Poyato, J.M.L. Exploring the Catalytic Efficiency of X-Doped (X = B, N, P) Graphene in Oxygen Reduction Reaction: Influence of Solvent and Border Effects. Int. J. Quantum Chem. 2018, 118, e25579. [Google Scholar] [CrossRef]
  580. Chen, X.; Ge, F.; Lai, N.J. N, O Co-Doped Graphene as a Potential Catalyst for the Oxygen Reduction Reaction. J. Electrochem. Soc. 2019, 166, F847–F851. [Google Scholar] [CrossRef]
  581. Yang, N.; Li, L.; Li, J.; Ding, W.; Wei, Z.D. Modulating the Oxygen Reduction Activity of Heteroatom-Doped Carbon Catalysts via the Triple Effect: Charge, Spin Density and Ligand Effect. Chem. Sci. 2018, 9, 5795–5804. [Google Scholar] [CrossRef]
  582. Zhang, X.M.; Xia, Z.X.; Li, H.Q.; Yu, S.S.; Wang, S.L.; Sun, G.Q. The Mechanism and Activity of Oxygen Reduction Reaction on Single Atom Doped Graphene: A DFT Method. RSC Adv. 2019, 9, 7086–7093. [Google Scholar] [CrossRef]
  583. Doronin, S.V.; Volykhov, A.A.; Inozemtseva, A.I.; Usachov, D.Y.; Yashina, L.V. Comparative Catalytic Activity of Graphene Imperfections in Oxygen Reduction Reaction. J. Phys. Chem. C 2020, 124, 6038–6053. [Google Scholar] [CrossRef]
  584. Cao, L.J.; Yang, M.Y.; Lu, Z.G.; Pan, H. Exploring an Effective Oxygen Reduction Reaction Catalyst via 4e(−) Process Based on Waved-Graphene. Sci. China Mater. 2017, 60, 739–746. [Google Scholar] [CrossRef]
  585. Xie, Y.; Wang, Z.W.; Zhu, T.Y.; Shu, D.J.; Hou, Z.F.; Terakura, K. Breaking the Scaling Relations for Oxygen Reduction Reaction on Nitrogen-Doped Graphene by Tensile Strain. Carbon 2018, 139, 129–136. [Google Scholar] [CrossRef]
  586. Yang, Z.; Yao, Z.; Li, G.F.; Fang, G.Y.; Nie, H.G.; Liu, Z.; Zhou, X.M.; Chen, X.; Huang, S.M. Sulfur-Doped Graphene as an Efficient Metal-Free Cathode Catalyst for Oxygen Reduction. ACS Nano 2012, 6, 205–211. [Google Scholar] [CrossRef]
  587. Jeon, I.Y.; Zhang, S.; Zhang, L.P.; Choi, H.J.; Seo, J.M.; Xia, Z.H.; Dai, L.M.; Baek, J.B. Edge-Selectively Sulfurized Graphene Nanoplatelets as Efficient Metal-Free Electrocatalysts for Oxygen Reduction Reaction: The Electron Spin Effect. Adv. Mater. 2013, 25, 6138–6145. [Google Scholar] [CrossRef]
  588. Stergiou, A.; Perivoliotis, D.K.; Tagmatarchis, N. (Photo)Electrocatalysis of Molecular Oxygen Reduction by S-Doped Graphene Decorated with a Star-Shaped Oligothiophene. Nanoscale 2019, 11, 7335–7346. [Google Scholar] [CrossRef]
  589. Zhang, L.P.; Niu, J.B.; Li, M.T.; Xia, Z.H. Catalytic Mechanisms of Sulfur-Doped Graphene as Efficient Oxygen Reduction Reaction Catalysts for Fuel Cells. J. Phys. Chem. C 2014, 118, 3545–3553. [Google Scholar] [CrossRef]
  590. Lu, Z.S.; Li, S.; Liu, C.; He, C.Z.; Yang, X.W.; Ma, D.W.; Xu, G.L.; Yang, Z.X. Sulfur Doped Graphene as a Promising Metal-Free Electrocatalyst for Oxygen Reduction Reaction: A DFT-D Study. RSC Adv. 2017, 7, 20398–20405. [Google Scholar] [CrossRef]
  591. Garcia, A.G.; Baltazar, S.E.; Castro, A.H.R.; Robles, J.F.P.; Rubio, A. Influence of S and P Doping in a Graphene Sheet. J. Comput. Theor. Nanosci. 2008, 5, 2221–2229. [Google Scholar] [CrossRef]
  592. Duan, J.J.; Chen, S.; Jaroniec, M.; Qiao, S.Z. Heteroatom-Doped Graphene-Based Materials for Energy-Relevant Electrocatalytic Processes. ACS Catal. 2015, 5, 5207–5234. [Google Scholar] [CrossRef]
  593. Xu, J.X.; Dong, G.F.; Jin, C.H.; Huang, M.H.; Guan, L.H. Sulfur and Nitrogen Co-Doped, Few-Layered Graphene Oxide as a Highly Efficient Electrocatalyst for the Oxygen-Reduction Reaction. ChemSusChem 2013, 6, 493–499. [Google Scholar] [CrossRef] [PubMed]
  594. Zhang, Y.J.; Chu, M.; Yang, L.; Deng, W.F.; Tan, Y.M.; Ma, M.; Xie, Q.J. Synthesis and Oxygen Reduction Properties of Three-Dimensional Sulfur-Doped Graphene Networks. Chem. Commun. 2014, 50, 6382–6385. [Google Scholar] [CrossRef] [PubMed]
  595. Wang, C.X.; Yang, F.; Xu, C.; Cao, Y.; Zhong, H.L.; Li, Y.F. Sulfur-Doped Porous Graphene Frameworks as an Efficient Metal-Free Electrocatalyst for Oxygen Reduction Reaction. Mater. Lett. 2018, 214, 209–212. [Google Scholar] [CrossRef]
  596. Liang, J.; Jiao, Y.; Jaroniec, M.; Qiao, S.Z. Sulfur and Nitrogen Dual-Doped Mesoporous Graphene Electrocatalyst for Oxygen Reduction with Synergistically Enhanced Performance. Angew. Chem. Int. Ed. 2012, 51, 11496–11500. [Google Scholar] [CrossRef]
  597. Poh, H.L.; Simek, P.; Sofer, Z.; Pumera, M. Sulfur-Doped Graphene via Thermal Exfoliation of Graphite Oxide in H2S, SO2, or CS2 Gas. ACS Nano 2013, 7, 5262–5272. [Google Scholar] [CrossRef]
  598. Higgins, D.C.; Hoque, M.A.; Hassan, F.; Choi, J.Y.; Kim, B.; Chen, Z.W. Oxygen Reduction on Graphene-Carbon Nanotube Composites Doped Sequentially with Nitrogen and Sulfur. ACS Catal. 2014, 4, 2734–2740. [Google Scholar] [CrossRef]
  599. Shervedani, R.K.; Amini, A. Carbon Black/Sulfur-Doped Graphene Composite Prepared by Pyrolysis of Graphene Oxide with Sodium Polysulfide for Oxygen Reduction Reaction. Electrochim. Acta 2014, 142, 51–60. [Google Scholar] [CrossRef]
  600. Chen, X.; Chen, X.H.; Xu, X.; Yang, Z.; Liu, Z.; Zhang, L.J.; Xu, X.J.; Chen, Y.; Huang, S.M. Sulfur-Doped Porous Reduced Graphene Oxide Hollow Nanosphere Frameworks as Metal-Free Electrocatalysts for Oxygen Reduction Reaction and as Supercapacitor Electrode Materials. Nanoscale 2014, 6, 13740–13747. [Google Scholar] [CrossRef]
  601. Park, J.E.; Jang, Y.J.; Kim, Y.J.; Song, M.S.; Yoon, S.; Kim, D.H.; Kim, S.J. Sulfur-Doped Graphene as a Potential Alternative Metal-Free Electrocatalyst and Pt-Catalyst Supporting Material for Oxygen Reduction Reaction. Phys. Chem. Chem. Phys. 2014, 16, 103–109. [Google Scholar] [CrossRef]
  602. Wang, J.C.; Ma, R.G.; Zhou, Z.Z.; Liu, G.H.; Liu, Q. Magnesiothermic Synthesis of Sulfur-Doped Graphene as an Efficient Metal-Free Electrocatalyst for Oxygen Reduction. Sci. Rep. 2015, 5, 9304. [Google Scholar] [CrossRef]
  603. Wu, X.B.; Xie, Z.Y.; Sun, M.; Lei, T.; Zuo, Z.M.; Xie, X.M.; Liang, Y.L.; Huang, Q.Z. Edge-Rich and (N, S)-Doped 3D Porous Graphene as an Efficient Metal-Free Electrocatalyst for the Oxygen Reduction Reaction. RSC Adv. 2016, 6, 90384–90387. [Google Scholar] [CrossRef]
  604. Yazdi, A.Z.; Roberts, E.P.L.; Sundararaj, U. Nitrogen/Sulfur Co-Doped Helical Graphene Nanoribbons for Efficient Oxygen Reduction in Alkaline and Acidic Electrolytes. Carbon 2016, 100, 99–108. [Google Scholar] [CrossRef]
  605. Legrand, U.; Meunier, J.L.; Berk, D. Addition of Sulphur to Graphene Nanoflakes Using Thermal Plasma for Oxygen Reduction Reaction in Alkaline Medium. Plasma Chem. Plasma Process. 2017, 37, 841–856. [Google Scholar] [CrossRef]
  606. Huang, Z.; Zhou, H.H.; Yang, W.J.; Fu, C.P.; Chen, L.; Kuang, Y.F. Three-Dimensional Hierarchical Porous Nitrogen and Sulfur-Codoped Graphene Nanosheets for Oxygen Reduction in Both Alkaline and Acidic Media. ChemCatChem 2017, 9, 987–996. [Google Scholar] [CrossRef]
  607. Yao, Y.; Guo, Y.S.; Du, W.; Tong, X.Y.; Zhang, X. In Situ Synthesis of Sulfur-Doped Graphene Quantum Dots Decorated Carbon Nanoparticles Hybrid as Metal-Free Electrocatalyst for Oxygen Reduction Reaction. J. Mater. Sci. Mater. Electron. 2018, 29, 17695–17705. [Google Scholar] [CrossRef]
  608. Zhang, J.; Song, X.M.; Li, P.; Wu, Z.X.; Wu, Y.M.; Wang, S.; Liu, X. Sulfur, Nitrogen Co-Doped Nanocomposite of Graphene and Carbon Nanotube as an Efficient Bifunctional Electrocatalyst for Oxygen Reduction and Evolution Reactions. J. Taiwan Inst. Chem. Eng. 2018, 93, 336–341. [Google Scholar] [CrossRef]
  609. Zhang, J.; Wang, J.; Wu, Z.X.; Wang, S.; Wu, Y.M.; Liu, X. Heteroatom (Nitrogen/Sulfur)-Doped Graphene as an Efficient Electrocatalyst for Oxygen Reduction and Evolution Reactions. Catalysts 2018, 8, 475. [Google Scholar] [CrossRef]
  610. Lee, J.; Noh, S.; Pham, N.D.; Shim, J.H. Top-down Synthesis of S-Doped Graphene Nanosheets by Electrochemical Exfoliation of Graphite: Metal-Free Bifunctional Catalysts for Oxygen Reduction and Evolution Reactions. Electrochim. Acta 2019, 313, 1–9. [Google Scholar] [CrossRef]
  611. Chen, W.Y.; Ge, C.; Li, J.T.; Beckham, J.L.; Yuan, Z.; Wyss, K.M.; Advincula, P.A.; Eddy, L.; Kittrell, C.; Chen, J.H.; et al. Heteroatom-Doped Flash Graphene. ACS Nano 2022, 16, 6646–6656. [Google Scholar] [CrossRef]
  612. Ma, Z.L.; Dou, S.; Shen, A.L.; Tao, L.; Dai, L.M.; Wang, S.Y. Sulfur-Doped Graphene Derived from Cycled Lithium-Sulfur Batteries as a Metal-Free Electrocatalyst for the Oxygen Reduction Reaction. Angew. Chem. Int. Ed. 2015, 54, 1888–1892. [Google Scholar] [CrossRef]
  613. Wong, C.H.A.; Sofer, Z.; Klimova, K.; Pumera, M. Microwave Exfoliation of Graphite Oxides in H2S Plasma for the Synthesis of Sulfur-Doped Graphenes as Oxygen Reduction Catalysts. ACS Appl. Mater. Interfaces 2016, 8, 31849–31855. [Google Scholar] [CrossRef] [PubMed]
  614. Bag, S.; Mondal, B.; Das, A.K.; Raj, C.R. Nitrogen and Sulfur Dual-Doped Reduced Graphene Oxide: Synergistic Effect of Dopants Towards Oxygen Reduction Reaction. Electrochim. Acta 2015, 163, 16–23. [Google Scholar] [CrossRef]
  615. Klingele, M.; Pham, C.; Vuyyuru, K.R.; Britton, B.; Holdcroft, S.; Fischer, A.; Thiele, S. Sulfur Doped Reduced Graphene Oxide as Metal-Free Catalyst for the Oxygen Reduction Reaction in Anion and Proton Exchange Fuel Cells. Electrochem. Commun. 2017, 77, 71–75. [Google Scholar] [CrossRef]
  616. Van Pham, C.; Klingele, M.; Britton, B.; Vuyyuru, K.R.; Unmuessig, T.; Holdcroft, S.; Fischer, A.; Thiele, S. Tridoped Reduced Graphene Oxide as a Metal-Free Catalyst for Oxygen Reduction Reaction Demonstrated in Acidic and Alkaline Polymer Electrolyte Fuel Cells. Adv. Sustain. Syst. 2017, 1, 1600038. [Google Scholar] [CrossRef]
  617. Chabu, J.M.; Wang, L.Q.; Tang, F.Y.; Zeng, K.; Sheng, J.P.; Walle, M.D.; Deng, L.; Liu, Y.N. Synthesis of Three-Dimensional Nitrogen and Sulfur Dual-Doped Graphene Aerogels as an Efficient Metal-Free Electrocatalyst for the Oxygen Reduction Reaction. ChemElectroChem 2017, 4, 1885–1890. [Google Scholar] [CrossRef]
  618. Liu, F.; Niu, F.S.; Chen, T.; Han, J.R.; Liu, Z.; Yang, W.R.; Xu, Y.H.; Liu, J.Q. One-Step Electrochemical Strategy for in-Situ Synthesis of S, N-Codoped Graphene as Metal-Free Catalyst for Oxygen Reduction Reaction. Carbon 2018, 134, 316–325. [Google Scholar] [CrossRef]
  619. Fan, T.J.; Zhang, G.X.; Jian, L.F.; Murtaza, I.; Meng, H.; Liu, Y.D.; Min, Y. Facile Synthesis of Defect-Rich Nitrogen and Sulfur Co-Doped Graphene Quantum Dots as Metal-Free Electrocatalyst for the Oxygen Reduction Reaction. J. Alloys Compd. 2019, 792, 844–850. [Google Scholar] [CrossRef]
  620. Chen, Y.; Li, J.; Mei, T.; Hu, X.G.; Liu, D.W.; Wang, J.C.; Hao, M.; Li, J.H.; Wang, J.Y.; Wang, X.B. Low-Temperature and One-Pot Synthesis of Sulfurized Graphene Nanosheets via in Situ Doping and Their Superior Electrocatalytic Activity for Oxygen Reduction Reaction. J. Mater. Chem. A Mater. 2014, 2, 20714–20722. [Google Scholar] [CrossRef]
  621. Chen, L.S.; Cui, X.Z.; Wang, Y.X.; Wang, M.; Qiu, R.H.; Shu, Z.; Zhang, L.X.; Hua, Z.L.; Cui, F.M.; Weia, C.Y.; et al. One-Step Synthesis of Sulfur Doped Graphene Foam for Oxygen Reduction Reactions. Dalton Trans. 2014, 43, 3420–3423. [Google Scholar] [CrossRef]
  622. Zhai, C.Y.; Sun, M.J.; Zhu, M.S.; Song, S.Q.; Jiang, S.J. A New Method to Synthesize Sulfur-Doped Graphene as Effective Metal-Free Electrocatalyst for Oxygen Reduction Reaction. Appl. Surf. Sci. 2017, 407, 503–508. [Google Scholar] [CrossRef]
  623. Zhang, P.; Wei, J.S.; Chen, X.B.; Xiong, H.M. Heteroatom-Doped Carbon Dots Based Catalysts for Oxygen Reduction Reactions. J. Colloid. Interface Sci. 2019, 537, 716–724. [Google Scholar] [CrossRef] [PubMed]
  624. Vasiliev, V.P.; Kotkin, A.S.; Kochergin, V.K.; Manzhos, R.A.; Krivenko, A.G. Oxygen Reduction Reaction at Few-Layer Graphene Structures Obtained via Plasma-Assisted Electrochemical Exfoliation of Graphite. J. Electroanal. Chem. 2019, 851, 113440. [Google Scholar] [CrossRef]
  625. Ahmed, M.S.; Han, H.S.; Jeon, S. One-Step Chemical Reduction of Graphene Oxide with Oligothiophene for Improved Electrocatalytic Oxygen Reduction Reactions. Carbon 2013, 61, 164–172. [Google Scholar] [CrossRef]
  626. Jeon, I.Y.; Choi, H.J.; Jung, S.M.; Seo, J.M.; Kim, M.J.; Dai, L.M.; Baek, J.B. Large-Scale Production of Edge-Selectively Functionalized Graphene Nanoplatelets via Ball Milling and Their Use as Metal-Free Electrocatalysts for Oxygen Reduction Reaction. J. Am. Chem. Soc. 2013, 135, 1386–1393. [Google Scholar] [CrossRef]
  627. Abdolmaleki, A.; Mallakpour, S.; Mahmoudian, M. Preparation and Evaluation of Edge Selective Sulfonated Graphene by Chlorosulfuric Acid as an Active Metal- Free Electrocatalyst for Oxygen Reduction Reaction in Alkaline Media. ChemistrySelect 2017, 2, 11211–11217. [Google Scholar] [CrossRef]
  628. Chua, C.K.; Pumera, M. Monothiolation and Reduction of Graphene Oxide via One-Pot Synthesis: Hybrid Catalyst for Oxygen Reduction. ACS Nano 2015, 9, 4193–4199. [Google Scholar] [CrossRef]
  629. Pal, S.; Bawari, S.; Vineesh, T.V.; Shyaga, N.; Narayanan, T.N. Selenium-Coupled Reduced Graphene Oxide as Single-Atom Site Catalyst for Direct Four-Electron Oxygen Reduction Reaction. ACS Appl. Energy Mater. 2019, 2, 3624–3632. [Google Scholar] [CrossRef]
  630. Agnoli, S.; Favaro, M. Doping Graphene with Boron: A Review of Synthesis Methods, Physicochemical Characterization, and Emerging Applications. J. Mater. Chem. A Mater. 2016, 4, 5002–5025. [Google Scholar] [CrossRef]
  631. Kaukonen, M.; Krasheninnikov, A.V.; Kauppinen, E.; Nieminen, R.M. Doped Graphene as a Material for Oxygen Reduction Reaction in Hydrogen Fuel Cells: A Computational Study. ACS Catal. 2013, 3, 159–165. [Google Scholar] [CrossRef]
  632. Fazio, G.; Ferrighi, L.; Di Valentin, C. Boron-Doped Graphene as Active Electrocatalyst for Oxygen Reduction Reaction at a Fuel-Cell Cathode. J. Catal. 2014, 318, 203–210. [Google Scholar] [CrossRef]
  633. Ferrighi, L.; Datteo, M.; Di Valentin, C. Boosting Graphene Reactivity with Oxygen by Boron Doping: Density Functional Theory Modeling of the Reaction Path. J. Phys. Chem. C 2014, 118, 223–230. [Google Scholar] [CrossRef]
  634. Wang, L.; Dong, H.L.; Guo, Z.Y.; Zhang, L.L.; Hou, T.J.; Li, Y.Y. Potential Application of Novel Boron-Doped Graphene Nanoribbon as Oxygen Reduction Reaction Catalyst. J. Phys. Chem. C 2016, 120, 17427–17434. [Google Scholar] [CrossRef]
  635. Boukhvalov, D.W.; Zhidkov, I.S.; Kukharenko, A.I.; Slesarev, A.I.; Zatsepin, A.F.; Cholakh, S.O.; Kurmaev, E.Z. Stability of Boron-Doped Graphene/Copper Interface: DFT, XPS and OSEE Studies. Appl. Surf. Sci. 2018, 441, 978–983. [Google Scholar] [CrossRef]
  636. Sheng, Z.H.; Gao, H.L.; Bao, W.J.; Wang, F.B.; Xia, X.H. Synthesis of Boron Doped Graphene for Oxygen Reduction Reaction in Fuel Cells. J. Mater. Chem 2012, 22, 390–395. [Google Scholar] [CrossRef]
  637. Norskov, J.K.; Rossmeisl, J.; Logadottir, A.; Lindqvist, L.; Kitchin, J.R.; Bligaard, T.; Jónsson, H. Origin of the Overpotential for Oxygen Reduction at a Fuel-Cell Cathode. J. Phys. Chem. B 2004, 108, 17886–17892. [Google Scholar] [CrossRef]
  638. Tam, T.V.; Kang, S.G.; Kim, M.H.; Lee, S.G.; Hur, S.H.; Chung, J.S.; Choi, W.M. Novel Graphene Hydrogel/B-Doped Graphene Quantum Dots Composites as Trifunctional Electrocatalysts for Zn-Air Batteries and Overall Water Splitting. Adv. Energy Mater. 2019, 9, 1900945. [Google Scholar] [CrossRef]
  639. Zheng, Y.; Jiao, Y.; Ge, L.; Jaroniec, M.; Qiao, S.Z. Two-Step Boron and Nitrogen Doping in Graphene for Enhanced Synergistic Catalysis. Angew. Chem. Int. Ed. 2013, 52, 3110–3116. [Google Scholar] [CrossRef]
  640. Choi, C.H.; Chung, M.W.; Kwon, H.C.; Park, S.H.; Woo, S.I. B, N- and P, N-Doped Graphene as Highly Active Catalysts for Oxygen Reduction Reactions in Acidic Media. J. Mater. Chem. A Mater. 2013, 1, 3694–3699. [Google Scholar] [CrossRef]
  641. Zuo, Z.C.; Jiang, Z.Q.; Manthiram, A. Porous B-Doped Graphene Inspired by Fried-Ice for Supercapacitors and Metal-Free Catalysts. J. Mater. Chem. A Mater. 2013, 1, 13476–13483. [Google Scholar] [CrossRef]
  642. Jo, G.; Sanetuntikul, J.; Shanmugam, S. Boron and Phosphorous-Doped Graphene as a Metal-Free Electrocatalyst for the Oxygen Reduction Reaction in Alkaline Medium. RSC Adv. 2015, 5, 53637–53643. [Google Scholar] [CrossRef]
  643. Jiang, Z.Q.; Zhao, X.S.; Tian, X.N.; Luo, L.J.; Fang, J.H.; Gao, H.Q.; Jiang, Z.J. Hydrothermal Synthesis of Boron and Nitrogen Codoped Hollow Graphene Microspheres with Enhanced Electrocatalytic Activity for Oxygen Reduction Reaction. ACS Appl. Mater. Interfaces 2015, 7, 19398–19407. [Google Scholar] [CrossRef] [PubMed]
  644. Tien, H.N.; Kocabas, C.; Hur, S.H. One-Step Codoping of Reduced Graphene Oxide Using Boric and Nitric Acid Mixture and Its Use in Metal-Free Electrocatalyst. Mater. Lett. 2015, 143, 205–208. [Google Scholar] [CrossRef]
  645. Xu, C.C.; Su, Y.; Liu, D.J.; He, X.Q. Three-Dimensional N,B-Doped Graphene Aerogel as a Synergistically Enhanced Metal-Free Catalyst for the Oxygen Reduction Reaction. Phys. Chem. Chem. Phys. 2015, 17, 25440–25448. [Google Scholar] [CrossRef]
  646. Tabassum, H.; Zou, R.Q.; Mahmood, A.; Liang, Z.B.; Guo, S.J. A Catalyst-Free Synthesis of B, N Co-Doped Graphene Nanostructures with Tunable Dimensions as Highly Efficient Metal Free Dual Electrocatalysts. J. Mater. Chem. A Mater. 2016, 4, 16469–16475. [Google Scholar] [CrossRef]
  647. Jang, A.R.; Lee, Y.W.; Lee, S.S.; Hong, J.; Beak, S.H.; Pak, S.; Lee, J.; Shin, H.S.; Ahn, D.; Hong, W.K.; et al. Electrochemical and Electrocatalytic Reaction Characteristics of Boron-Incorporated Graphene via a Simple Spin-on Dopant Process. J. Mater. Chem. A Mater. 2018, 6, 7351–7356. [Google Scholar] [CrossRef]
  648. Gao, P.P.; Sun, M.; Wu, X.B.; Zhou, S.Z.; Deng, X.T.; Xie, Z.Y.; Xiao, L.; Jiang, L.H.; Huang, Q.Z. (B,N)-Doped 3D Porous Graphene-CNTs Synthesized by Chemical Vapor Deposition as a Bifunctional Catalyst for ORR and HER. RSC Adv. 2018, 8, 26934–26937. [Google Scholar] [CrossRef]
  649. Nazer, E.A.A.; Muthukrishnan, A. Synergistic Effect on BCN NanoMaterials for the Oxygen Reduction Reaction—A Kinetic and Mechanistic Analysis to Explore the Active Sites. Catal. Sci. Technol. 2020, 10, 6659–6668. [Google Scholar] [CrossRef]
  650. Han, T.H.; Huang, Y.K.; Tan, A.T.L.; Dravid, V.P.; Huang, J.X. Steam Etched Porous Graphene Oxide Network for Chemical Sensing. J. Am. Chem. Soc. 2011, 133, 15264–15267. [Google Scholar] [CrossRef]
  651. Zhou, Y.Z.; Yen, C.H.; Fu, S.F.; Yang, G.H.; Zhu, C.Z.; Du, D.; Wo, P.C.; Cheng, X.N.; Yang, J.; Wai, C.M.; et al. One-Pot Synthesis of B-Doped Three-Dimensional Reduced Graphene Oxide via Supercritical Fluid for Oxygen Reduction Reaction. Green Chem. 2015, 17, 3552–3560. [Google Scholar] [CrossRef]
  652. Van Tam, T.; Kang, S.G.; Babu, K.F.; Oh, E.S.; Lee, S.G.; Choi, W.M. Synthesis of B-Doped Graphene Quantum Dots as a Metal-Free Electrocatalyst for the Oxygen Reduction Reaction. J. Mater. Chem. A Mater. 2017, 5, 10537–10543. [Google Scholar] [CrossRef]
  653. Yao, Z.H.; Hu, M.C.; Iqbal, Z.; Wang, X.Q. N-8(-) Polynitrogen Stabilized on Boron-Doped Graphene as Metal-Free Electrocatalysts for Oxygen Reduction Reaction. ACS Catal. 2020, 10, 160–167. [Google Scholar] [CrossRef]
  654. Lan, R.S.; Liu, L.H.; Feng, H.; Chen, B.Y.; Shi, X.D.; Hong, J.M. Boron-Doped Reduced Graphene Oxide as an Efficient Cathode in Microbial Fuel Cells for Biological Toxicity Detection. Bioresour. Technol. 2024, 403, 9. [Google Scholar] [CrossRef] [PubMed]
  655. Wang, Z.J.; Cao, X.H.; Ping, J.F.; Wang, Y.X.; Lin, T.T.; Huang, X.; Ma, Q.L.; Wang, F.K.; He, C.B.; Zhang, H. Electrochemical Doping of Three-Dimensional Graphene Networks Used as Efficient Electrocatalysts for Oxygen Reduction Reaction. Nanoscale 2015, 7, 9394–9398. [Google Scholar] [CrossRef]
  656. Li, R.; Wei, Z.D.; Gou, X.L.; Xu, W. Phosphorus-Doped Graphene Nanosheets as Efficient Metal-Free Oxygen Reduction Electrocatalysts. RSC Adv. 2013, 3, 9978–9984. [Google Scholar] [CrossRef]
  657. Zhang, C.Z.; Mahmood, N.; Yin, H.; Liu, F.; Hou, Y.L. Synthesis of Phosphorus-Doped Graphene and Its Multifunctional Applications for Oxygen Reduction Reaction and Lithium Ion Batteries. Adv. Mater. 2013, 25, 4932–4937. [Google Scholar] [CrossRef]
  658. Razmjooei, F.; Singh, K.P.; Song, M.Y.; Yu, J.S. Enhanced Electrocatalytic Activity Due to Additional Phosphorous Doping in Nitrogen and Sulfur-Doped Graphene: A Comprehensive Study. Carbon 2014, 78, 257–267. [Google Scholar] [CrossRef]
  659. Balaji, S.S.; Ganesh, P.A.; Moorthy, M.; Sathish, M. Efficient Electrocatalytic Activity for Oxygen Reduction Reaction by Phosphorus-Doped Graphene Using Supercritical Fluid Processing. Bull. Mater. Sci. 2020, 43, 151. [Google Scholar] [CrossRef]
  660. Li, C.L.; Chen, Z.Y.; Kong, A.G.; Ni, Y.Y.; Kong, F.T.; Shan, Y.K. High-Rate Oxygen Electroreduction over Metal-Free Graphene Foams Embedding P-N Coupled Moieties in Acidic Media. J. Mater. Chem. A Mater. 2018, 6, 4145–4151. [Google Scholar] [CrossRef]
  661. Jang, D.; Lee, S.; Kim, S.; Choi, K.; Park, S.; Oh, J. Production of P, N Co-Doped Graphene-Based Materials by a Solution Process and Their Electrocatalytic Performance for Oxygen Reduction Reaction. ChemNanoMat 2018, 4, 118–123. [Google Scholar] [CrossRef]
  662. Ge, L.P.; Wang, D.; Yang, P.X.; Xu, H.; Xiao, L.H.; Zhang, G.X.; Lu, X.Y.; Duan, Z.Z.; Meng, F.; Zhang, J.Q.; et al. Graphite N-C-P Dominated Three-Dimensional Nitrogen and Phosphorus Co-Doped Holey Graphene Foams as High-Efficiency Electrocatalysts for Zn-Air Batteries. Nanoscale 2019, 11, 17010–17017. [Google Scholar] [CrossRef]
  663. Tong, X.; Cherif, M.; Zhang, G.X.; Zhan, X.X.; Ma, J.G.; Almesrati, A.; Vidal, F.; Song, Y.J.; Claverie, J.P.; Sun, S.H. N, P-Codoped Graphene Dots Supported on N-Doped 3D Graphene as Metal-Free Catalysts for Oxygen Reduction. ACS Appl. Mater. Interfaces 2021, 13, 30512–30523. [Google Scholar] [CrossRef]
  664. Poon, K.C.; Wan, W.Y.; Su, H.B.; Sato, H. One-Minute Synthesis via Electroless Reduction of Amorphous Phosphorus-Doped Graphene for Oxygen Reduction Reaction. ACS Appl. Energy Mater. 2021, 4, 5388–5391. [Google Scholar] [CrossRef]
  665. Ex 614 Zhang, X.L.; Lu, Z.S.; Fu, Z.M.; Tang, Y.A.; Ma, D.W.; Yang, Z.X. The Mechanisms of Oxygen Reduction Reaction on Phosphorus Doped Graphene: A First-Principles Study. J. Power Sources 2015, 276, 222–229. [Google Scholar] [CrossRef]
  666. Bai, X.W.; Zhao, E.J.; Li, K.; Wang, Y.; Jiao, M.G.; He, F.; Sun, X.X.; Sun, H.; Wu, Z.J. Theoretical Insights on the Reaction Pathways for Oxygen Reduction Reaction on Phosphorus Doped Graphene. Carbon 2016, 105, 214–223. [Google Scholar] [CrossRef]
  667. Yang, N.; Zheng, X.Q.; Li, L.; Li, J.; Wei, Z.D. Influence of Phosphorus Configuration on Electronic Structure and Oxygen Reduction Reactions of Phosphorus-Doped Graphene. J. Phys. Chem. C 2017, 121, 19321–19328. [Google Scholar] [CrossRef]
  668. Ho, K.I.; Huang, C.H.; Liao, J.H.; Zhang, W.J.; Li, L.J.; Lai, C.S.; Su, C.Y. Fluorinated Graphene as High Performance Dielectric Materials and the Applications for Graphene Nanoelectronics. Sci. Rep. 2014, 4, 7. [Google Scholar] [CrossRef]
  669. Vineesh, T.V.; Nazrullaa, M.A.; Krishnamoorthy, S.; Narayanan, T.N.; Alwarappan, S. Synergistic Effects of Dopants on the Spin Density of Catalytic Active Centres of N-Doped Fluorinated Graphene for Oxygen Reduction Reaction. Appl. Mater. Today 2015, 1, 74–79. [Google Scholar] [CrossRef]
  670. Jiang, S.; Sun, Y.J.; Dai, H.C.; Hu, J.T.; Ni, P.J.; Wang, Y.L.; Li, Z. Nitrogen and Fluorine Dual-Doped Mesoporous Graphene: A High-Performance Metal-Free ORR Electrocatalyst with a Super-Low HO2- Yield. Nanoscale 2015, 7, 10584–10589. [Google Scholar] [CrossRef]
  671. Kakaei, K.; Balavandi, A. Synthesis of Halogen-Doped Reduced Graphene Oxide Nanosheets as Highly Efficient Metal-Free Electrocatalyst for Oxygen Reduction Reaction. J. Colloid. Interface Sci. 2016, 463, 46–54. [Google Scholar] [CrossRef] [PubMed]
  672. Kakaei, K.; Balavandi, A. Hierarchically Porous Fluorine-Doped Graphene Nanosheets as Efficient Metal-Free Electrocatalyst for Oxygen Reduction in Gas Diffusion Electrode. J. Colloid. Interface Sci. 2017, 490, 819–824. [Google Scholar] [CrossRef]
  673. Guo, J.H.; Zhang, J.G.; Zhao, H.Q.; Fang, Y.S.; Ming, K.; Huang, H.; Chen, J.M.; Wang, X.C. Fluorine-Doped Graphene with an Outstanding Electrocatalytic Performance for Efficient Oxygen Reduction Reaction in Alkaline Solution. R Soc. Open. Sci. 2018, 5, 12. [Google Scholar] [CrossRef] [PubMed]
  674. Liu, A.L.; Li, W.; Jin, H.L.; Yu, X.C.; Bu, Y.F.; He, Y.H.; Huang, H.H.; Wang, S.; Wang, J.C. The Enhanced Electrocatalytic Activity of Graphene Co-Doped with Chlorine and Fluorine Atoms. Electrochim. Acta 2015, 177, 36–42. [Google Scholar] [CrossRef]
  675. Jeon, I.Y.; Choi, H.J.; Choi, M.; Seo, J.M.; Jung, S.M.; Kim, M.J.; Zhang, S.; Zhang, L.P.; Xia, Z.H.; Dai, L.M.; et al. Facile, Scalable Synthesis of Edge-Halogenated Graphene Nanoplatelets as Efficient Metal-Free Eletrocatalysts for Oxygen Reduction Reaction. Sci. Rep. 2013, 3, 1810. [Google Scholar] [CrossRef] [PubMed]
  676. Wu, K.H.; Wang, D.W.; Zeng, Q.C.; Li, Y.; Gentle, I.R. Solution Phase Synthesis of Halogenated Graphene and the Electrocatalytic Activity for Oxygen Reduction Reaction. Chin. J. Catal. 2014, 35, 884–890. [Google Scholar] [CrossRef]
  677. Wang, D.W.; Su, D.S. Heterogeneous NanoCarbon Materials for Oxygen Reduction Reaction. Energy Environ. Sci. 2014, 7, 576–591. [Google Scholar] [CrossRef]
  678. Zhan, Y.F.; Huang, J.L.; Lin, Z.P.; Yu, X.; Zeng, D.R.; Zhang, X.X.; Xie, F.Y.; Zhang, W.H.; Chen, J.; Meng, H. Iodine/Nitrogen Co-Doped Graphene as Metal Free Catalyst for Oxygen Reduction Reaction. Carbon 2015, 95, 930–939. [Google Scholar] [CrossRef]
  679. Marinoiu, A.; Raceanu, M.; Carcadea, E.; Varlam, M.; Balan, D.; Ion-Ebrasu, D.; Stefanescu, I.; Enachescu, M. Iodine-Doped Graphene for Enhanced Electrocatalytic Oxygen Reduction Reaction in Proton Exchange Membrane Fuel Cell Applications. J. Electrochem. Energy Convers. Storage 2017, 14, 031001. [Google Scholar] [CrossRef]
  680. Marinoiu, A.; Raceanu, M.; Carcadea, E.; Varlam, M.; Soare, A.; Stefanescu, I. Doped Graphene as Non-Metallic Catalyst for Fuel Cells. Mater. Sci. Medzg. 2017, 23, 108–113. [Google Scholar] [CrossRef]
  681. Marinoiu, A.; Raceanu, M.; Carcadea, E.; Varlam, M.; Stefanescu, L. Low Cost Iodine Intercalated Graphene for Fuel Cells Electrodes. Appl. Surf. Sci. 2017, 424, 93–100. [Google Scholar] [CrossRef]
  682. Jeon, I.Y.; Noh, H.J.; Baek, J.B. Direct and Efficient Conversion from Low-Quality Graphite to High-Quality Graphene Nanoplatelets. Flatchem 2018, 12, 10–16. [Google Scholar] [CrossRef]
  683. Marinoiu, A.; Raceanu, M.; Carcadea, E.; Varlam, M. Iodine-Doped Graphene—Catalyst Layer in PEM Fuel Cells. Appl. Surf. Sci. 2018, 456, 238–245. [Google Scholar] [CrossRef]
  684. Hoang, V.C.; Dinh, K.N.; Gomes, V.G. Iodine Doped Composite with Biomass Carbon Dots and Reduced Graphene Oxide: A Versatile Bifunctional Electrode for Energy Storage and Oxygen Reduction Reaction. J. Mater. Chem. A Mater. 2019, 7, 22650–22662. [Google Scholar] [CrossRef]
  685. Wang, D.; Hasegawa, S.; Shimizu, M.; Tanaka, J. Synthesis, Doping and Polarized Raman-Spectra of Highly Conducting Polyacetylene. Synth Met 1992, 46, 85–91. [Google Scholar] [CrossRef]
  686. Zeng, X.R.; Ko, T.M. Structure-Conductivity Relationships of Iodine-Doped Polyaniline. J. Polym. Sci. Part B Polym. Phys. 1997, 35, 1993–2001. [Google Scholar] [CrossRef]
  687. Wadekar, P.H.; Pethsangave, D.A.; Khose, R.V.; Some, S. Synthesis of Iodine-Functionalized Graphene Electrocatalyst Using Deep Eutectic Solvents for Oxygen Reduction Reaction and Supercapacitors. Energy Technol. 2021, 9, 2000750. [Google Scholar] [CrossRef]
  688. Marinoiu, A.; Ion-Ebrasu, D.; Soare, A.; Raceanu, M. Iodine-Doped Graphene Oxide: Fast Single-Stage Synthesis and Application as Electrocatalyst. Materials 2022, 15, 6174. [Google Scholar] [CrossRef]
  689. Hoyt, R.A.; Remillard, E.M.; Cubuk, E.D.; Vecitis, C.D.; Kaxiras, E. Polyiodide-Doped Graphene. J. Phys. Chem. C 2017, 121, 609–615. [Google Scholar] [CrossRef]
  690. Baek, J.Y.; Jeon, I.Y.; Baek, J.B. Edge-Iodine/Sulfonic Acid-Functionalized Graphene Nanoplatelets as Efficient Electrocatalysts for Oxygen Reduction Reaction. J. Mater. Chem. A Mater. 2014, 2, 8690–8695. [Google Scholar] [CrossRef]
  691. Lin, H.L.; Chu, L.; Wang, X.J.; Yao, Z.Q.; Liu, F.; Ai, Y.N.; Zhuang, X.D.; Han, S. Boron, Nitrogen, and Phosphorous Ternary Doped Graphene Aerogel with Hierarchically Porous Structures as Highly Efficient Electrocatalysts for Oxygen Reduction Reaction. New J. Chem. 2016, 40, 6022–6029. [Google Scholar] [CrossRef]
  692. Musico, Y.L.F.; Kakati, N.; Labata, M.F.M.; Ocon, J.D.; Chuang, P.Y.A. One-Pot Hydrothermal Synthesis of Heteroatom Co-Doped with Fluorine on Reduced Graphene Oxide for Enhanced ORR Activity and Stability in Alkaline Media. Mater. Chem. Phys. 2019, 236, 121804. [Google Scholar] [CrossRef]
  693. Raudsepp, R.; Türk, K.K.; Zarmehri, E.; Joost, U.; Rauwel, P.; Saar, R.; Mäeorg, U.; Dyck, A.; Bron, M.; Chen, Z.M.; et al. Boron and Fluorine Co-Doped Graphene/Few-Walled Carbon Nanotube Composite as Highly Active Electrocatalyst for Oxygen Reduction Reaction. ChemNanoMat 2024, 10, 11. [Google Scholar] [CrossRef]
  694. Ranganathan, A.; Pedireddi, V.R.; Rao, C.N.R. Hydrothermal Synthesis of Organic Channel Structures: 1:1 Hydrogen-Bonded Adducts of Melamine with Cyanuric and Trithiocyanuric Acids. J. Am. Chem. Soc. 1999, 121, 1752–1753. [Google Scholar] [CrossRef]
  695. Niu, W.H.; Li, L.G.; Liu, X.J.; Wang, N.; Liu, J.; Zhou, W.J.; Tang, Z.H.; Chen, S.W. Mesoporous N-Doped Carbons Prepared with Thermally Removable Nanoparticle Templates: An Efficient Electrocatalyst for Oxygen Reduction Reaction. J. Am. Chem. Soc. 2015, 137, 5555–5562. [Google Scholar] [CrossRef]
  696. Zhang, S.; Zhang, X.; Jiang, G.M.; Zhu, H.Y.; Guo, S.J.; Su, D.; Lu, G.; Sun, S.H. Tuning Nanoparticle Structure and Surface Strain for Catalysis Optimization. J. Am. Chem. Soc. 2014, 136, 7734–7739. [Google Scholar] [CrossRef]
  697. Feng, L.L.; Zou, Y.C.; Li, C.G.; Gao, S.; Zhou, L.J.; Sun, Q.S.; Fan, M.H.; Wang, H.J.; Wang, D.J.; Li, G.D.; et al. Nanoporous Sulfur-Doped Graphitic Carbon Nitride Microrods: A Durable Catalyst for Visible-Light-Driven H2 Evolution. Int. J. Hydrogen Energy 2014, 39, 15373–15379. [Google Scholar] [CrossRef]
  698. Gao, F.; Qu, J.Y.; Zhao, Z.B.; Zhou, Q.; Li, B.B.; Qiu, J.S. A Green Strategy for the Synthesis of Graphene Supported Mn3O4 Nanocomposites from Graphitized Coal and Their Supercapacitor Application. Carbon 2014, 80, 640–650. [Google Scholar] [CrossRef]
  699. Song, J.; Liu, T.F.; Ali, S.; Li, B.; Su, D.S. The Synergy Effect and Reaction Pathway in the Oxygen Reduction Reaction on the Sulfur and Nitrogen Dual Doped Graphene Catalyst. Chem. Phys. Lett. 2017, 677, 65–69. [Google Scholar] [CrossRef]
  700. Choi, C.H.; Chung, M.W.; Jun, Y.J.; Woo, S.I. Doping of Chalcogens (Sulfur and/or Selenium) in Nitrogen-Doped Graphene-CNT Self-Assembly for Enhanced Oxygen Reduction Activity in Acid Media. RSC Adv. 2013, 3, 12417–12422. [Google Scholar] [CrossRef]
  701. Huang, B.B.; Hu, X.; Liu, Y.C.; Qi, W.; Xie, Z.L. Biomolecule-Derived N/S Co-Doped CNT-Graphene Hybrids Exhibiting Excellent Electrochemical Activities. J. Power Sources 2019, 413, 408–417. [Google Scholar] [CrossRef]
  702. Villemson, K.M.; Kaare, K.; Raudsepp, R.; Kaambre, T.; Smits, K.; Wang, P.P.; Kuzmin, A.V.; Sutka, A.; Shainyan, B.A.; Kruusenberg, I. Identification of Active Sites for Oxygen Reduction Reaction on Nitrogen- and Sulfur-Codoped Carbon Catalysts. J. Phys. Chem. C 2019, 123, 16065–16074. [Google Scholar] [CrossRef]
  703. Su, Y.Z.; Zhang, Y.; Zhuang, X.D.; Li, S.; Wu, D.Q.; Zhang, F.; Feng, X.L. Low-Temperature Synthesis of Nitrogen/Sulfur Co-Doped Three-Dimensional Graphene Frameworks as Efficient Metal-Free Electrocatalyst for Oxygen Reduction Reaction. Carbon 2013, 62, 296–301. [Google Scholar] [CrossRef]
  704. Wang, X.; Wang, J.; Wang, D.L.; Dou, S.O.; Ma, Z.L.; Wu, J.H.; Tao, L.; Shen, A.L.; Ouyang, C.B.; Liu, Q.H.; et al. One-Pot Synthesis of Nitrogen and Sulfur Co-Doped Graphene as Efficient Metal-Free Electrocatalysts for the Oxygen Reduction Reaction. Chem. Commun. 2014, 50, 4839–4842. [Google Scholar] [CrossRef] [PubMed]
  705. Li, Y.F.; Li, M.; Jiang, L.Q.; Lin, L.; Cui, L.L.; He, X.Q. Advanced Oxygen Reduction Reaction Catalyst Based on Nitrogen and Sulfur Co-Doped Graphene in Alkaline Medium. Phys. Chem. Chem. Phys. 2014, 16, 23196–23205. [Google Scholar] [CrossRef] [PubMed]
  706. Ai, W.; Luo, Z.M.; Jiang, J.; Zhu, J.H.; Du, Z.Z.; Fan, Z.X.; Xie, L.H.; Zhang, H.; Huang, W.; Yu, T. Nitrogen and Sulfur Codoped Graphene: Multifunctional Electrode Materials for High-Performance Li-Ion Batteries and Oxygen Reduction Reaction. Adv. Mater. 2014, 26, 6186–6192. [Google Scholar] [CrossRef]
  707. Luo, Z.M.; Yang, D.L.; Qi, G.Q.; Shang, J.Z.; Yang, H.P.; Wang, Y.L.; Yuwen, L.H.; Yu, T.; Huang, W.; Wang, L.H. Microwave-Assisted Solvothermal Preparation of Nitrogen and Sulfur Co-Doped Reduced Graphene Oxide and Graphene Quantum Dots Hybrids for Highly Efficient Oxygen Reduction. J. Mater. Chem. A Mater. 2014, 2, 20605–20611. [Google Scholar] [CrossRef]
  708. Wu, M.; Wang, J.; Wu, Z.X.; Xin, H.L.L.; Wang, D.L. Synergistic Enhancement of Nitrogen and Sulfur Co-Doped Graphene with Carbon Nanosphere Insertion for the Electrocatalytic Oxygen Reduction Reaction. J. Mater. Chem. A Mater. 2015, 3, 7727–7731. [Google Scholar] [CrossRef]
  709. Zhang, H.H.; Liu, X.Q.; He, G.L.; Zhang, X.X.; Bao, S.J.; Hu, W.H. Bioinspired Synthesis of Nitrogen/Sulfur Co-Doped Graphene as an Efficient Electrocatalyst for Oxygen Reduction Reaction. J. Power Sources 2015, 279, 252–258. [Google Scholar] [CrossRef]
  710. Parvez, K.; Rincon, R.A.; Weber, N.E.; Cha, K.C.; Venkataraman, S.S. One-Step Electrochemical Synthesis of Nitrogen and Sulfur Co-Doped, High-Quality Graphene Oxide. Chem. Commun. 2016, 52, 5714–5717. [Google Scholar] [CrossRef]
  711. Akhter, T.; Islam, M.M.; Faisal, S.N.; Hague, E.; Minett, A.I.; Liu, H.K.; Konstantinov, K.; Dou, S.X. Self-Assembled N/S Codoped Flexible Graphene Paper for High Performance Energy Storage and Oxygen Reduction Reaction. ACS Appl. Mater. Interfaces 2016, 8, 2078–2087. [Google Scholar] [CrossRef]
  712. Amiinu, I.S.; Zhang, J.; Kou, Z.K.; Liu, X.B.; Asare, O.K.; Zhou, H.; Cheng, K.; Zhang, H.N.; Mai, L.Q.; Pan, M.; et al. Self-Organized 3D Porous Graphene Dual-Doped with Biomass-Sponsored Nitrogen and Sulfur for Oxygen Reduction and Evolution. ACS Appl. Mater. Interfaces 2016, 8, 29408–29418. [Google Scholar] [CrossRef]
  713. Pan, F.P.; Duan, Y.X.; Zhang, X.K.; Zhang, J.Y. A Facile Synthesis of Nitrogen/Sulfur Co-Doped Graphene for the Oxygen Reduction Reaction. ChemCatChem 2016, 8, 163–170. [Google Scholar] [CrossRef]
  714. Wu, M.; Dou, Z.Y.; Chang, J.J.; Cui, L.L. Nitrogen and Sulfur Co-Doped Graphene Aerogels as an Efficient Metal-Free Catalyst for Oxygen Reduction Reaction in an Alkaline Solution. RSC Adv. 2016, 6, 22781–22790. [Google Scholar] [CrossRef]
  715. Fan, M.M.; Huang, Y.; Yuan, F.S.; Hao, Q.L.; Yang, J.Z.; Sun, D.P. Effects of Multiple Heteroatom Species and Topographic Defects on Electrocatalytic and Capacitive Performances of Graphene. J. Power Sources 2017, 366, 143–150. [Google Scholar] [CrossRef]
  716. Gaidukevic, J.; Razumiene, J.; Sakinyte, I.; Rebelo, S.L.H.; Barkauskas, J. Study on the Structure and Electrocatalytic Activity of Graphene-Based Nanocomposite Materials Containing (SCN)(n). Carbon 2017, 118, 156–167. [Google Scholar] [CrossRef]
  717. Han, D.D.; Yan, Y.C.; Wei, J.S.; Wang, B.W.; Li, T.T.; Guo, G.N.; Yang, D.; Xie, S.H.; Dong, A.G. Fine-Tuning the Wall Thickness of Ordered Mesoporous Graphene by Exploiting Ligand Exchange of Colloidal Nanocrystals. Front. Chem. 2017, 5, 117. [Google Scholar] [CrossRef]
  718. Rivera, L.M.; Fajardo, S.; Arevalo, M.D.; García, G.; Pastor, E. S- and N-Doped Graphene NanoMaterials for the Oxygen Reduction Reaction. Catalysts 2017, 7, 278. [Google Scholar] [CrossRef]
  719. Zhang, H.H.; Niu, Y.L.; Hu, W.H. Nitrogen/Sulfur-Doping of Graphene with Cysteine as a Heteroatom Source for Oxygen Reduction Electrocatalysis. J. Colloid Interface Sci. 2017, 505, 32–37. [Google Scholar] [CrossRef]
  720. Arunchander, A.; Peera, S.G.; Panda, S.K.; Chellammal, S.; Sahu, A.K. Simultaneous Co-Doping of N and S by a Facile in-Situ Polymerization of 6-N, N-Dibutylamine-1,3,5-Triazine-2,4-Dithiol on Graphene Framework: An Efficient and Durable Oxygen Reduction Catalyst in Alkaline Medium. Carbon 2017, 118, 531–544. [Google Scholar] [CrossRef]
  721. Hassani, S.S.; Ganjali, M.R.; Samiee, L.; Rashidi, A.M.; Tasharrofi, S.; Yadegari, A.; Shoghi, F.; Martel, R. Comparative Study of Various Types of Metal-Free N and S Co-Doped Porous Graphene for High Performance Oxygen Reduction Reaction in Alkaline Solution. J. Nanosci. Nanotechnol. 2018, 18, 4565–4579. [Google Scholar] [CrossRef] [PubMed]
  722. Wu, D.L.; Wang, T.; Wang, L.X.; Jia, D.Z. Hydrothermal Synthesis of Nitrogen, Sulfur Co-Doped Graphene and Its High Performance in Supercapacitor and Oxygen Reduction Reaction. Microporous Mesoporous Mater. 2019, 290, 109556. [Google Scholar] [CrossRef]
  723. Verma, R.; Chakraborty, I.; Chowdhury, S.; Ghangrekar, M.M.; Balasubramanian, R. Nitrogen and Sulfur Codoped Graphene Macroassemblies as High-Performance Electrocatalysts for the Oxygen Reduction Reaction in Microbial Fuel Cells. ACS Sustain. Chem. Eng. 2020, 8, 16591–16599. [Google Scholar] [CrossRef]
  724. Zhang, W.L.; Zhang, Y.K.; Li, Y.D.; Yang, S.C.; Zhang, L.H.; Yu, F.S. Enhanced Oxygen Reduction Performance of Nitrogen and Sulfur Co-Doped Graphene Oxide by Immobilized Ionic Liquid. Chem. Eng. Sci. 2021, 236, 8. [Google Scholar] [CrossRef]
  725. Periyasamy, G.; Annamalai, K.; Patil, I.M.; Kakade, B. Sulfur and Nitrogen Co-Doped RGO Sheets as Efficient Electrocatalyst for Oxygen Reduction Reaction in Alkaline Medium. Diam. Relat. Mater. 2021, 114, 10. [Google Scholar] [CrossRef]
  726. Shah, M.; Lee, J.; Rauf, A.; Park, J.H.; Lim, B.; Yoo, P.J. Electrostatically Regulated Ternary-Doped Carbon Foams with Exposed Active Sites as Metal-Free Oxygen Reduction Electrocatalysts. Nanoscale 2018, 10, 19498–19508. [Google Scholar] [CrossRef]
  727. Sun, Y.B.; Deng, R.X.; Chi, C.; Chen, X.L.; Pan, Y.A.; Li, J.; Xia, X.H. One-Step Synthesis of S, N Dual-Element Doped RGO as an Efficient Electrocatalyst for ORR. J. Electroanal. Chem. 2023, 940, 9. [Google Scholar] [CrossRef]
  728. You, J.M.; Ahmed, M.S.; Han, H.S.; Choe, J.E.; Ustundag, Z.; Jeon, S. New Approach of Nitrogen and Sulfur-Doped Graphene Synthesis Using Dipyrrolemethane and Their Electrocatalytic Activity for Oxygen Reduction in Alkaline Media. J. Power Sources 2015, 275, 73–79. [Google Scholar] [CrossRef]
  729. Wu, M.B.; Liu, Y.; Zhu, Y.L.; Lin, J.; Liu, J.Y.; Hu, H.; Wang, Y.; Zhao, Q.S.; Lv, R.Q.; Qiu, J.S. Supramolecular Polymerization-Assisted Synthesis of Nitrogen and Sulfur Dual-Doped Porous Graphene Networks from Petroleum Coke as Efficient Metal-Free Electrocatalysts for the Oxygen Reduction Reaction. J. Mater. Chem. A Mater. 2017, 5, 11331–11339. [Google Scholar] [CrossRef]
  730. Zhao, Y.; Zhang, C.X.; Liu, T.; Fan, R.; Sun, Y.; Tao, H.J.; Xue, J.J. Low Temperature Green Synthesis of Sulfur-Nitrogen Co-Doped Graphene as Efficient Metal-Free Catalysts for Oxygen Reduction Reaction. Int. J. Electrochem. Sci. 2017, 12, 3537–3548. [Google Scholar] [CrossRef]
  731. Zhao, C.E.; Li, J.X.; Chen, Y.; Chen, J.Y. Nitrogen and Sulfur Dual-Doped Graphene as an Efficient Metal-Free Electrocatalyst for the Oxygen Reduction Reaction in Microbial Fuel Cells. N. J. Chem. 2019, 43, 9389–9395. [Google Scholar] [CrossRef]
  732. Siahkalroudi, Z.M.; Aghabarari, B.; Vaezi, M.; Rodríguez-Castellón, E.; Martínez-Huerta, M. V Effect of Secondary Heteroatom (S, P) in N-Doped Reduced Graphene Oxide Catalysts to Oxygen Reduction Reaction. Mol. Catal. 2021, 502, 9. [Google Scholar] [CrossRef]
  733. Zhang, X.R.; Wang, Y.Q.; Wang, K.; Huang, Y.L.; Lyu, D.D.; Yu, F.; Wang, S.B.; Tian, Z.Q.; Shen, P.K.; Jiang, S.P. Active Sites Engineering via Tuning Configuration between Graphitic-N and Thiophenic-S Dopants in One-Step Synthesized Graphene Nanosheets for Efficient Water-Cycled Electrocatalysis. Chem. Eng. J. 2021, 416, 11. [Google Scholar] [CrossRef]
  734. Zhang, X.R.; Wen, X.Y.; Pan, C.; Xiang, X.; Hao, C.; Meng, Q.H.; Tian, Z.Q.; Shen, P.K.; Jiang, S.P. N Species Tuning Strategy in N, S Co-Doped Graphene Nanosheets for Electrocatalytic Activity and Selectivity of Oxygen Redox Reactions. Chem. Eng. J. 2022, 431, 11. [Google Scholar] [CrossRef]
  735. Li, Y.; Yang, J.; Huang, J.P.; Zhou, Y.Z.; Xu, K.; Zhao, N.; Cheng, X.N. Soft Template-Assisted Method for Synthesis of Nitrogen and Sulfur Co-Doped Three-Dimensional Reduced Graphene Oxide as an Efficient Metal Free Catalyst for Oxygen Reduction Reaction. Carbon 2017, 122, 237–246. [Google Scholar] [CrossRef]
  736. Wang, S.Y.; Zhang, L.P.; Xia, Z.H.; Roy, A.; Chang, D.W.; Baek, J.B.; Dai, L.M. BCN Graphene as Efficient Metal-Free Electrocatalyst for the Oxygen Reduction Reaction. Angew. Chem. Int. Ed. 2012, 51, 4209–4212. [Google Scholar] [CrossRef]
  737. Kattel, S.; Atanassov, P.; Kiefer, B. Density Functional Theory Study of the Oxygen Reduction Reaction Mechanism in a BN Co-Doped Graphene Electrocatalyst. J. Mater. Chem. A Mater. 2014, 2, 10273–10279. [Google Scholar] [CrossRef]
  738. Sen, D.; Thapa, R.; Chattopadhyay, K.K. Rules of Boron-Nitrogen Doping in Defect Graphene Sheets: A First-Principles Investigation of Band-Gap Tuning and Oxygen Reduction Reaction Catalysis Capabilities. Chemphyschem 2014, 15, 2542–2549. [Google Scholar] [CrossRef]
  739. Gong, Y.J.; Fei, H.L.; Zou, X.L.; Zhou, W.; Yang, S.B.; Ye, G.L.; Liu, Z.; Peng, Z.W.; Lou, J.; Vajtai, R.; et al. Boron- and Nitrogen-Substituted Graphene Nanoribbons as Efficient Catalysts for Oxygen Reduction Reaction. Chem. Mater. 2015, 27, 1181–1186. [Google Scholar] [CrossRef]
  740. Tang, S.B.; Wu, W.H.; Liu, L.X.; Gu, J.J. Oxygen-Molecule Adsorption and Dissociation on BCN Graphene: A First-Principles Study. Chemphyschem 2017, 18, 101–110. [Google Scholar] [CrossRef]
  741. Ricca, C.; Labat, F.; Zavala, C.; Russo, N.; Adamo, C.; Merino, G.; Sicilia, E. B,N-Codoped Graphene as Catalyst for the Oxygen Reduction Reaction: Insights from Periodic and Cluster DFT Calculations. J. Comput. Chem. 2018, 39, 637–647. [Google Scholar] [CrossRef]
  742. Qin, L.; Wang, L.C.; Yang, X.; Ding, R.M.; Zheng, Z.F.; Chen, X.H.; Lv, B.L. Synergistic Enhancement of Oxygen Reduction Reaction with BC3 and Graphitic-N in Boron- and Nitrogen-Codoped Porous Graphene. J. Catal. 2018, 359, 242–250. [Google Scholar] [CrossRef]
  743. Larijani, H.T.; Khorshidian, M. Theoretical Insight into the Role of Pyridinic Nitrogen on the Catalytic Activity of Boron-Doped Graphene towards Oxygen Reduction Reaction. Appl. Surf. Sci. 2019, 492, 826–842. [Google Scholar] [CrossRef]
  744. Tai, J.P.; Hu, J.T.; Chen, Z.X.; Lu, H.B. Two-Step Synthesis of Boron and Nitrogen Codoped Graphene as a Synergistically Enhanced Catalyst for the Oxygen Reduction Reaction. RSC Adv. 2014, 4, 61437–61443. [Google Scholar] [CrossRef]
  745. Wu, W.M.; Leng, J.G.; Mei, H.L.; Yang, S.B. Defect-Rich, Boron-Nitrogen Bonds-Free and Dual-Doped Graphenes for Highly Efficient Oxygen Reduction Reaction. J. Colloid. Interface Sci. 2018, 521, 11–16. [Google Scholar] [CrossRef]
  746. Kang, G.S.; Lee, S.; Lee, D.C.; Yoon, C.W.; Joh, H.I. Edge-Enriched Graphene with Boron and Nitrogen Co-Doping for Enhanced Oxygen Reduction Reaction. Curr. Appl. Phys. 2020, 20, 456–461. [Google Scholar] [CrossRef]
  747. Fei, H.L.; Ye, R.Q.; Ye, G.L.; Gong, Y.J.; Peng, Z.W.; Fan, X.J.; Samuel, E.L.G.; Ajayan, P.M.; Tour, J.M. Boron- and Nitrogen-Doped Graphene Quantum Dots/Graphene Hybrid Nanoplatelets as Efficient Electrocatalysts for Oxygen Reduction. ACS Nano 2014, 8, 10837–10843. [Google Scholar] [CrossRef]
  748. Han, J.R.; Zhang, Y.L.; Niu, F.S.; Chen, T.; Liu, J.Q.; Xu, Y.H. Low-Cost and Highly Efficient Metal-Free Electrocatalysts for Oxygen Reduction Reaction: Environment-Friendly Three-Dimensional B, N Co-Doped Graphene Aerogels. Electrocatalysis 2019, 10, 56–62. [Google Scholar] [CrossRef]
  749. Byeon, A.; Lee, J.W. Electrocatalytic Activity of BN Codoped Graphene Oxide Derived from Carbon Dioxide. J. Phys. Chem. C 2013, 117, 24167–24173. [Google Scholar] [CrossRef]
  750. Zhu, J.L.; He, C.Y.; Li, Y.Y.; Kang, S.; Shen, P.K. One-Step Synthesis of Boron and Nitrogen-Dual-Self-Doped Graphene Sheets as Non-Metal Catalysts for Oxygen Reduction Reaction. J. Mater. Chem. A Mater. 2013, 1, 14700–14705. [Google Scholar] [CrossRef]
  751. Chen, W.; Xu, L.; Tian, Y.H.; Li, H.A.; Wang, K. Boron and Nitrogen Co-Doped Graphene Aerogels: Facile Preparation, Tunable Doping Contents and Bifunctional Oxygen Electrocatalysis. Carbon 2018, 137, 458–466. [Google Scholar] [CrossRef]
  752. Sapner, V.S.; Chavan, P.P.; Munde, A.V.; Sayyad, U.S.; Sathe, B.R. Heteroatom (N, O, and S)-Based Biomolecule-Functionalized Graphene Oxide: A Bifunctional Electrocatalyst for Enhancing Hydrazine Oxidation and Oxygen Reduction Reactions. Energy Fuels 2021, 35, 6823–6834. [Google Scholar] [CrossRef]
  753. Kumar, R.; Gopalakrishnan, K.; Ahmad, I.; Rao, C.N.R. BN-Graphene Composites Generated by Covalent Cross-Linking with Organic Linkers. Adv. Funct. Mater. 2015, 25, 5910–5917. [Google Scholar] [CrossRef]
  754. Yang, D.S.; Bhattacharjya, D.; Inamdar, S.; Park, J.; Yu, J.S. Phosphorus-Doped Ordered Mesoporous Carbons with Different Lengths as Efficient Metal-Free Electrocatalysts for Oxygen Reduction Reaction in Alkaline Media. J. Am. Chem. Soc. 2012, 134, 16127–16130. [Google Scholar] [CrossRef] [PubMed]
  755. Choi, C.H.; Park, S.H.; Woo, S.I. Phosphorus-Nitrogen Dual Doped Carbon as an Effective Catalyst for Oxygen Reduction Reaction in Acidic Media: Effects of the Amount of P-Doping on the Physical and Electrochemical Properties of Carbon. J. Mater. Chem. 2012, 22, 12107–12115. [Google Scholar] [CrossRef]
  756. Choi, C.H.; Park, S.H.; Woo, S.I. Heteroatom Doped Carbons Prepared by the Pyrolysis of Bio-Derived Amino Acids as Highly Active Catalysts for Oxygen Electro-Reduction Reactions. Green Chem. 2011, 13, 406–412. [Google Scholar] [CrossRef]
  757. Xue, X.X.; Tang, L.M.; Chen, K.Q.; Zhang, L.X.; Wang, E.G.; Feng, Y.X. Bifunctional Mechanism of N, P Co-Doped Graphene for Catalyzing Oxygen Reduction and Evolution Reactions. J. Chem. Phys. 2019, 150, 104701. [Google Scholar] [CrossRef]
  758. Yang, J.; Sun, H.Y.; Liang, H.Y.; Ji, H.X.; Song, L.; Gao, C.; Xu, H.X. A Highly Efficient Metal-Free Oxygen Reduction Electrocatalyst Assembled from Carbon Nanotubes and Graphene. Adv. Mater. 2016, 28, 4606–4613. [Google Scholar] [CrossRef]
  759. Li, R.; Wei, Z.D.; Gou, X.L. Nitrogen and Phosphorus Dual-Doped Graphene/Carbon Nanosheets as Bifunctional Electrocatalysts for Oxygen Reduction and Evolution. ACS Catal. 2015, 5, 4133–4142. [Google Scholar] [CrossRef]
  760. Dong, L.Y.; Hu, C.G.; Huang, X.K.; Chen, N.; Qu, L.T. One-Pot Synthesis of Nitrogen and Phosphorus Co-Doped Graphene and Its Use as High-Performance Electrocatalyst for Oxygen Reduction Reaction. Chem. Asian J. 2015, 10, 2608–2613. [Google Scholar] [CrossRef]
  761. Qiao, X.C.; Liao, S.J.; You, C.H.; Chen, R. Phosphorus and Nitrogen Dual Doped and Simultaneously Reduced Graphene Oxide with High Surface Area as Efficient Metal-Free Electrocatalyst for Oxygen Reduction. Catalysts 2015, 5, 981–991. [Google Scholar] [CrossRef]
  762. Wang, J.; Wu, Z.X.; Han, L.L.; Liu, Y.Y.; Guo, J.P.; Xin, H.L.L.; Wang, D.L. Rational Design of Three-Dimensional Nitrogen and Phosphorus Co-Doped Graphene Nanoribbons/CNTs Composite for the Oxygen Reduction. Chin. Chem. Lett. 2016, 27, 597–601. [Google Scholar] [CrossRef]
  763. Zhao, Y.F.; Huang, S.F.; Xia, M.R.; Rehman, S.; Mu, S.C.; Kou, Z.K.; Zhang, Z.; Chen, Z.Y.; Gao, F.M.; Hou, Y.L. N-P-O Co-Doped High Performance 3D Graphene Prepared through Red Phosphorous-Assisted “Cutting-Thin” Technique: A Universal Synthesis and Multifunctional Applications. Nano Energy 2016, 28, 346–355. [Google Scholar] [CrossRef]
  764. Fang, B.; Yang, J.; Chen, C.; Zhang, C.X.; Chang, D.; Xu, H.X.; Gao, C. Carbon Nanotubes Loaded on Graphene Microfolds as Efficient Bifunctional Electrocatalysts for the Oxygen Reduction and Oxygen Evolution Reactions. ChemCatChem 2017, 9, 4520–4528. [Google Scholar] [CrossRef]
  765. Wang, X.W.; Liu, Y.; Wu, P.Y. Water-Soluble Triphenylphosphine-Derived Microgel as the Template towards in-Situ Nitrogen, Phosphorus Co-Doped Mesoporous Graphene Framework for Supercapacitor and Electrocatalytic Oxygen Reduction. Chem. Eng. J. 2017, 328, 417–427. [Google Scholar] [CrossRef]
  766. Zhou, L.J.; Zhang, C.Y.; Cai, X.Y.; Qian, Y.; Jiang, H.F.; Li, B.S.; Lai, L.F.; Shen, Z.X.; Huang, W. N, P Co-Doped Hierarchical Porous Graphene as a Metal-Free Bifunctional Air Cathode for Zn-Air Batteries. ChemElectroChem 2018, 5, 1811–1816. [Google Scholar] [CrossRef]
  767. Cheng, C.; Li, Y.; Maouche, C.; Li, B.; Zhou, Y.Z.; Wang, S.; Cheng, X.N.; Yang, J. Green Synthesis of N, P-Co Doped Porous Reduced Graphene Oxide as an Active Metal-Free Electrocatalyst toward Oxygen Reduction Reaction. J. Electroanal. Chem. 2021, 883, 11. [Google Scholar] [CrossRef]
  768. Chai, G.L.; Qiu, K.P.; Qiao, M.; Titirici, M.M.; Shang, C.X.; Guo, Z.X. Active Sites Engineering Leads to Exceptional ORR and OER Bifunctionality in P,N Co-Doped Graphene Frameworks. Energy Environ. Sci. 2017, 10, 1186–1195. [Google Scholar] [CrossRef]
  769. Zhang, X.R.; Zhang, X.; Xiang, X.; Pan, C.; Meng, Q.H.; Hao, C.; Tian, Z.Q.; Shen, P.K.; Jiang, S.P. Nitrogen and Phosphate Co-Doped Graphene as Efficient Bifunctional Electrocatalysts by Precursor Modulation Strategy for Oxygen Reduction and Evolution Reactions. ChemElectroChem 2021, 8, 3262–3272. [Google Scholar] [CrossRef]
  770. Sahraie, N.R.; Paraknowitsch, J.P.; Göbel, C.; Thomas, A.; Strasser, P. Noble-Metal-Free Electrocatalysts with Enhanced ORR Performance by Task-Specific Functionalization of Carbon Using Ionic Liquid Precursor Systems. J. Am. Chem. Soc. 2014, 136, 14486–14497. [Google Scholar] [CrossRef]
  771. Aijaz, A.; Akita, T.; Yang, H.; Xu, Q. From Ionic-Liquid@metal-Organic Framework Composites to Heteroatom-Decorated Large-Surface Area Carbons: Superior CO2 and H2 Uptake. Chem. Commun. 2014, 50, 6498–6501. [Google Scholar] [CrossRef]
  772. Hassan, M.; Haque, E.; Minett, A.I.; Gomes, V.G. Co-Doping of Activated Graphene for Synergistically Enhanced Electrocatalytic Oxygen Reduction Reaction. ChemSusChem 2015, 8, 4040–4048. [Google Scholar] [CrossRef]
  773. Lee, C.H.; Kang, G.S.; Lee, Y.K.; Lee, S.; Jo, S.M.; Yoo, S.J.; Jang, J.H.; Lee, D.C.; Joh, H.I. Synthesis and Properties of Nitrogen and Iodine Co-Functionalized Graphene Oxide and Its Electrochemical Applications. Sci. Adv. Mater. 2016, 8, 28–33. [Google Scholar] [CrossRef]
  774. Zhao, F.G.; Pan, B.Y.G.; Kong, Y.T.; Dong, L.; Hu, C.M.; Sang, Y.J.; Zhou, X.J.; Zuo, B.; Dong, X.P.; Li, B.X.; et al. Nonmainstream Out-Plane Fluoro- and Amino-Cofunctionalized Graphene for a Striking Electrocatalyst: Programming Substitutive/Reductive Defluorination toward Graphite Fluoride. Adv. Mater. Interfaces 2019, 6, 1801699. [Google Scholar] [CrossRef]
  775. Yan, R.; Wu, H.; Zheng, Q.; Wang, J.Y.; Huang, J.L.; Ding, K.J.; Guo, Q.G.; Wang, J.Z. Graphene Quantum Dots Cut from Graphene Flakes: High Electrocatalytic Activity for Oxygen Reduction and Low Cytotoxicity. RSC Adv. 2014, 4, 23097–23106. [Google Scholar] [CrossRef]
  776. Qiao, X.C.; Liao, S.J.; Wang, G.H.; Zheng, R.P.; Song, H.Y.; Li, X.H. Simultaneous Doping of Nitrogen and Fluorine into Reduced Graphene Oxide: A Highly Active Metal-Free Electrocatalyst for Oxygen Reduction. Carbon 2016, 99, 272–279. [Google Scholar] [CrossRef]
  777. Posudievsky, O.Y.; Kondratyuk, A.S.; Kozarenko, O.A.; Cherepanov, V.V.; Dovbeshko, G.I.; Koshechko, V.G.; Pokhodenko, V.D. Facile Mechanochemical Preparation of Nitrogen and Fluorine Co-Doped Graphene and Its Electrocatalytic Performance. Carbon 2019, 152, 274–283. [Google Scholar] [CrossRef]
  778. Li, Y.; Wen, H.J.; Yang, J.; Zhou, Y.Z.; Cheng, X.N. Boosting Oxygen Reduction Catalysis with N, F, and S Tri-Doped Porous Graphene: Tertiary N-Precursors Regulates the Constitution of Catalytic Active Sites. Carbon 2019, 142, 1–12. [Google Scholar] [CrossRef]
  779. Akula, S.; Sahu, A.K. Structurally Modulated Graphitic Carbon Nanofiber and Heteroatom (N,F) Engineering toward Metal-Free ORR Electrocatalysts for Polymer Electrolyte Membrane Fuel Cells. ACS Appl. Mater. Interfaces 2020, 12, 11438–11449. [Google Scholar] [CrossRef]
  780. Wu, J.; Zheng, X.J.; Jin, C.; Tian, J.H.; Yang, R.Z. Ternary Doping of Phosphorus, Nitrogen, and Sulfur into Porous Carbon for Enhancing Electrocatalytic Oxygen Reduction. Carbon 2015, 92, 327–338. [Google Scholar] [CrossRef]
  781. Choi, C.H.; Park, S.H.; Woo, S.I. Binary and Ternary Doping of Nitrogen, Boron, and Phosphorus into Carbon for Enhancing Electrochemical Oxygen Reduction Activity. ACS Nano 2012, 6, 7084–7091. [Google Scholar] [CrossRef]
  782. Wang, L.; Sofer, Z.; Pumera, M. Will Any Crap We Put into Graphene Increase Its Electrocatalytic Effect? ACS Nano 2020, 14. [Google Scholar] [CrossRef]
  783. Dou, S.; Shen, A.L.; Ma, Z.L.; Wu, J.H.; Tao, L.; Wang, S.Y. N-, P- and S-Tridoped Graphene as Metal-Free Electrocatalyst for Oxygen Reduction Reaction. J. Electroanal. Chem. 2015, 753, 21–27. [Google Scholar] [CrossRef]
  784. Ying, Y.D.; Ren, J.T.; Liu, Y.P.; Li, W.; Yuan, Z.Y. Facile Synthesis of Nitrogen, Phosphorus and Sulfur Tri-Doped Carbon Nanosheets as Efficient Oxygen Electrocatalyst for Rechargeable Zn-Air Batteries. Mater. Sci. Eng. B Adv. Funct. Solid State Mater. 2021, 273, 7. [Google Scholar] [CrossRef]
  785. Cai, Q.Y.; Ye, S.S.; Sun, H.Y.; Liao, Y.L.; Chen, H. N, S, P Tri-Doped Porous Graphene Electrocatalyst with Excellent Oxygen Reduction Activity for Zinc-Air Battery. Mater. Sci. Technol. 2024, 13, 294–306. [Google Scholar] [CrossRef]
  786. Routh, P.; Shin, S.H.; Jung, S.M.; Choi, H.J.; Jeon, I.Y.; Baek, J.B. Boron-Nitrogen-Phosphorous Doped Graphene Nanoplatelets for Enhanced Electrocatalytic Activity. Eur. Polym. J. 2018, 99, 511–517. [Google Scholar] [CrossRef]
  787. Hassani, S.S.; Ganjali, M.R.; Samiee, L.; Rashidi, A.M. Use of Grape Leaves for Producing Graphene for Use as an Oxygen Reduction Electrocatalyst. Int. J. Electrochem. Sci. 2020, 15, 4754–4773. [Google Scholar] [CrossRef]
  788. Tang, L.H.; Wang, Y.; Li, Y.M.; Feng, H.B.; Lu, J.; Li, J.H. Preparation, Structure, and Electrochemical Properties of Reduced Graphene Sheet Films. Adv. Funct. Mater. 2009, 19, 2782–2789. [Google Scholar] [CrossRef]
  789. Shao, Y.Y.; Wang, J.; Engelhard, M.; Wang, C.M.; Lin, Y.H. Facile and Controllable Electrochemical Reduction of Graphene Oxide and Its Applications. J. Mater. Chem 2010, 20, 743–748. [Google Scholar] [CrossRef]
  790. Wu, J.J.; Wang, Y.; Zhang, D.; Hou, B.R. Studies on the Electrochemical Reduction of Oxygen Catalyzed by Reduced Graphene Sheets in Neutral Media. J. Power Sources 2011, 196, 1141–1144. [Google Scholar] [CrossRef]
  791. Zhou, Y.F.; Zhang, G.Q.; Chen, J.; Yuan, G.E.; Xu, L.; Liu, L.F.; Yang, F.L. A Facile Two-Step Electroreductive Synthesis of Anthraquinone/Graphene Nanocomposites as Efficient Electrocatalyst for O-2 Reduction in Neutral Medium. Electrochem. Commun. 2012, 22, 69–72. [Google Scholar] [CrossRef]
  792. Liu, H.Y.; Zhang, G.Q.; Zhou, Y.F.; Gao, M.M.; Yang, F.L. One-Step Potentiodynamic Synthesis of Poly(1,5-Diaminoanthraquinone)/Reduced Graphene Oxide Nanohybrid with Improved Electrocatalytic Activity. J. Mater. Chem. A Mater. 2013, 1, 13902–13913. [Google Scholar] [CrossRef]
  793. Mooste, M.; Kibena-Poldsepp, E.; Ossonon, B.D.; Belanger, D.; Tammeveski, K. Oxygen Reduction on Graphene Sheets Functionalised by Anthraquinone Diazonium Compound during Electrochemical Exfoliation of Graphite. Electrochim. Acta 2018, 267, 246–254. [Google Scholar] [CrossRef]
  794. Guan, J.; Chen, X.; Wei, T.; Liu, F.P.; Wang, S.; Yang, Q.; Lu, Y.L.; Yang, S.F. Directly Bonded Hybrid of Graphene Nanoplatelets and Fullerene: Facile Solid-State Mechanochemical Synthesis and Application as Carbon-Based Electrocatalyst for Oxygen Reduction Reaction. J. Mater. Chem. A Mater. 2015, 3, 4139–4146. [Google Scholar] [CrossRef]
  795. Deng, D.H.; Yu, L.; Pan, X.L.; Wang, S.; Chen, X.Q.; Hu, P.; Sun, L.X.; Bao, X.H. Size Effect of Graphene on Electrocatalytic Activation of Oxygen. Chem. Commun. 2011, 47, 10016–10018. [Google Scholar] [CrossRef]
  796. 7Yuan, W.J.; Zhou, Y.; Li, Y.R.; Li, C.; Peng, H.L.; Zhang, J.; Liu, Z.F.; Dai, L.M.; Shi, G.Q. The Edge- and Basal-Plane-Specific Electrochemistry of a Single-Layer Graphene Sheet. Sci. Rep. 2013, 3, 2248. [Google Scholar] [CrossRef]
  797. Tao, L.; Wang, Q.; Dou, S.; Ma, Z.L.; Huo, J.; Wang, S.Y.; Dai, L.M. Edge-Rich and Dopant-Free Graphene as a Highly Efficient Metal-Free Electrocatalyst for the Oxygen Reduction Reaction. Chem. Commun. 2016, 52, 2764–2767. [Google Scholar] [CrossRef]
  798. Shi, J.L.; Wang, H.F.; Zhu, X.L.; Chen, C.M.; Huang, X.; Zhang, X.D.; Li, B.Q.; Tang, C.; Zhang, Q. The Nanostructure Preservation of 3D Porous Graphene: New Insights into the Graphitization and Surface Chemistry of Non-Stacked Double-Layer Templated Graphene after High-Temperature Treatment. Carbon 2016, 103, 36–44. [Google Scholar] [CrossRef]
  799. Lv, R.J.; Wang, H.J.; Yu, H.; Peng, F. Controllable Preparation of Holey Graphene and Electrocatalytic Performance for Oxygen Reduction Reaction. Electrochim. Acta 2017, 228, 203–213. [Google Scholar] [CrossRef]
  800. Xu, Z.W.; Fan, X.L.; Li, H.J.; Fu, H.; Lau, W.M.; Zhao, X.N. Edges of Graphene and Carbon Nanotubes with High Catalytic Performance for the Oxygen Reduction Reaction. Phys. Chem. Chem. Phys. 2017, 19, 21003–21011. [Google Scholar] [CrossRef]
  801. Zhang, H.J.; Zhang, X.; Yao, S.W.; Hu, H.; Ma, Z.F.; Yang, J.H. Thermal Evolution of the Structure and Activity of Non-Doped Graphene as Metal-Free Oxygen Reduction Electrocatalysts. J. Electrochem. Soc. 2018, 165, F526–F532. [Google Scholar] [CrossRef]
  802. Zhang, J.B.; Ren, M.Q.; Wang, L.Q.; Li, Y.L.; Yakobson, B.I.; Tour, J.M. Oxidized Laser-Induced Graphene for Efficient Oxygen Electrocatalysis. Adv. Mater. 2018, 30, 1707319. [Google Scholar] [CrossRef]
  803. San Roman, D.; Krishnamurthy, D.; Garg, R.; Hafiz, H.; Lamparski, M.; Nuhfer, N.T.; Meunier, V.; Viswanathan, V.; Cohen-Karni, T. Engineering Three-Dimensional (3D) Out-of-Plane Graphene Edge Sites for Highly Selective Two-Electron Oxygen Reduction Electrocatalysis. ACS Catal. 2020, 10, 1993–2008. [Google Scholar] [CrossRef]
  804. Wang, M.R.; Fang, Z.; Zhang, K.; Fang, J.; Qin, F.R.; Zhang, Z.A.; Li, J.; Liu, Y.X.; Lai, Y.Q. Synergistically Enhanced Activity of Graphene Quantum Dots/Graphene Hydrogel Composites: A Novel All-Carbon Hybrid Electrocatalyst for Metal/Air Batteries. Nanoscale 2016, 8, 11398–11402. [Google Scholar] [CrossRef] [PubMed]
  805. Lu, B.; Lv, L.X.; Zhang, X.Q.; Zhao, Y.; Chen, Q.; Cheng, H.H.; Qu, L.T. Highly Defective, Doping-Free Graphene Framework: A Rapid One-Step Formation Avenue. J. Power Sources 2021, 497, 7. [Google Scholar] [CrossRef]
  806. Yuasa, M.; Tanaka, M.; Shimizu, M.; Yoshida, M. Oxygen Reduction/Evolution Activity of a Mechanochemically Synthesized Multilayer Graphene. J. Electrochem. Soc. 2021, 168, 10. [Google Scholar] [CrossRef]
  807. Zhao, Y.L.; Wang, X.Y.; Guo, X.M.; Cheng, D.D.; Zhou, H.; Saito, N.; Fan, T.X. Synergetic Design of Dopant-Free Defect-Enriched 3D Interconnected Hierarchical Porous Graphene Mesh for Boosting Oxygen Reduction Reaction. Carbon 2021, 184, 609–617. [Google Scholar] [CrossRef]
  808. Kibena, E.; Mooste, M.; Kozlova, J.; Marandi, M.; Sammelselg, V.; Tammeveski, K. Surface and Electrochemical Characterisation of CVD Grown Graphene Sheets. Electrochem. Commun. 2013, 35, 26–29. [Google Scholar] [CrossRef]
  809. Zhang, L.P.; Xu, Q.; Niu, J.B.; Xia, Z.H. Role of Lattice Defects in Catalytic Activities of Graphene Clusters for Fuel Cells. Phys. Chem. Chem. Phys. 2015, 17, 16733–16743. [Google Scholar] [CrossRef]
  810. Jia, Y.; Zhang, L.Z.; Du, A.J.; Gao, G.P.; Chen, J.; Yan, X.C.; Brown, C.L.; Yao, X.D. Defect Graphene as a Trifunctional Catalyst for Electrochemical Reactions. Adv. Mater. 2016, 28, 9532–9538. [Google Scholar] [CrossRef]
  811. Ly, Q.; Merinov, B.V.; Xiao, H.; Goddard, W.A.; Yu, T.H. The Oxygen Reduction Reaction on Graphene from Quantum Mechanics: Comparing Armchair and Zigzag Carbon Edges. J. Phys. Chem. C 2017, 121, 24408–24417. [Google Scholar] [CrossRef]
  812. Bikkarolla, S.K.; Cumpson, P.; Joseph, P.; Papakonstantinou, P. Oxygen Reduction Reaction by Electrochemically Reduced Graphene Oxide. Faraday Discuss 2014, 173, 415–428. [Google Scholar] [CrossRef]
  813. Lilloja, J.; Kibena-Poldsepp, E.; Merisalu, M.; Rauwel, P.; Matisen, L.; Niilisk, A.; Cardoso, E.S.F.; Maia, G.; Sammelselg, V.; Tammeveski, K. An Oxygen Reduction Study of Graphene-Based NanoMaterials of Different Origin. Catalysts 2016, 6, 108. [Google Scholar] [CrossRef]
  814. Zhang, G.X.; Wei, Q.L.; Yang, X.H.; Tavares, A.C.; Sun, S.H. RRDE Experiments on Noble-Metal and Noble-Metal-Free Catalysts: Impact of Loading on the Activity and Selectivity of Oxygen Reduction Reaction in Alkaline Solution. Appl. Catal. B Environ. 2017, 206, 115–126. [Google Scholar] [CrossRef]
  815. Zhang, L.Y.; Liu, Z.; Xu, B.H.; Liu, H.D. Thermal Treated 3D Graphene as a Highly Efficient Metal-Free Electrocatalyst toward Oxygen Reduction Reaction. Int. J. Hydrogen Energy 2017, 42, 28278–28286. [Google Scholar] [CrossRef]
  816. Zhang, H.J.; Geng, J.; Yao, S.W.; Wang, J.C.; Ma, Z.F.; Yang, J.H. Nature of Carbon Materials Used as Nondoped Electrodes for Oxygen Reduction Reaction and Supercapacitor Applications. J. Electrochem. Soc. 2019, 166, F1–F8. [Google Scholar] [CrossRef]
  817. Nagappan, S.; Duraivel, M.; Han, S.H.; Yusuf, M.; Mahadadalkar, M.; Park, K.; Dhakshinamoorthy, A.; Prabakar, K.; Park, S.; Ha, C.S.; et al. Electrocatalytic Oxygen Reduction Reaction of Graphene Oxide and Metal-Free Graphene in an Alkaline Medium. Nanomaterials 2023, 13, 1315. [Google Scholar] [CrossRef]
  818. Wu, Y.; Muthukrishnan, A.; Nagata, S.; Nabae, Y. Kinetic Understanding of the Reduction of Oxygen to Hydrogen Peroxide over an N-Doped Carbon Electrocatalyst. J. Phys. Chem. C 2019, 123, 4590–4596. [Google Scholar] [CrossRef]
  819. Yan, Y.; Xia, B.Y.; Zhao, B.; Wang, X. A Review on Noble-Metal-Free Bifunctional Heterogeneous Catalysts for Overall Electrochemical Water Splitting. J. Mater. Chem. A Mater. 2016, 4, 17587–17603. [Google Scholar] [CrossRef]
  820. Lin, J.Y.; Xi, C.; Li, Z.; Feng, Y.; Wu, D.Y.; Dong, C.K.; Yao, P.; Liu, H.; Du, X.W. Lattice-Strained Palladium Nanoparticles as Active Catalysts for the Oxygen Reduction Reaction. Chem. Commun. 2019, 55, 3121–3123. [Google Scholar] [CrossRef]
  821. Choi, S.; Do, H.W.; Jin, D.; Kim, S.; Lee, J.; Soon, A.; Moon, J.; Shim, W. Revisiting the Role of the Triple-Phase Boundary in Promoting the Oxygen Reduction Reaction in Aluminum-Air Batteries. Adv. Funct. Mater. 2021, 31, 9. [Google Scholar] [CrossRef]
  822. Zavala, M.A.L.; Gutierrez, I.C.C. Effects of External Resistance, New Electrode Material, and Catholyte Type on the Energy Generation and Performance of Dual-Chamber Microbial Fuel Cells. Fermentation 2023, 9, 344. [Google Scholar] [CrossRef]
  823. Yang, Y.; Chang, H.L. Multi-Scale Porous Graphene/Activated Carbon Aerogel Enables Lightweight Carbonaceous Catalysts for Oxygen Reduction Reaction. J. Mater. Res. 2018, 33, 1247–1257. [Google Scholar] [CrossRef]
  824. Yang, Y.; Liu, T.Y.; Wang, H.Y.; Zhu, X.; Ye, D.D.; Liao, Q.; Liu, K.; Chen, S.W.; Li, Y. Reduced Graphene Oxide Modified Activated Carbon for Improving Power Generation of Air-Cathode Microbial Fuel Cells. J. Mater. Res. 2018, 33, 1279–1287. [Google Scholar] [CrossRef]
  825. Roy, P.; Ravindranath, R.; Periasamy, A.P.; Lien, C.W.; Liang, C.T.; Chang, H.T. Green Synthesis of Si-GQD Nanocomposites as Cost-Effective Catalysts for Oxygen Reduction Reaction. RSC Adv. 2016, 6, 108941–108947. [Google Scholar] [CrossRef]
  826. Wang, H.F.; Tang, C.; Wang, B.; Li, B.Q.; Cui, X.Y.; Zhang, Q. Defect-Rich Carbon Fiber Electrocatalysts with Porous Graphene Skin for Flexible Solid-State Zinc-Air Batteries. Energy Storage Mater. 2018, 15, 124–130. [Google Scholar] [CrossRef]
  827. Jiang, M.Y.; Zhang, Z.Q.; Chen, C.K.; Ma, W.C.; Han, S.J.; Li, X.; Lu, S.H.; Hu, X.J. High Efficient Oxygen Reduced Reaction Electrodes by Constructing Vertical Graphene Sheets on Separated Papillary Granules Formed Nanocrystalline Diamond Films. Carbon 2020, 168, 536–545. [Google Scholar] [CrossRef]
  828. He, Z.M.; Guo, Z.Q.; Guo, K.; Akasaka, T.; Lu, X. Compositing Fullerene-Derived Porous Carbon Fibers with Reduced Graphene Oxide for Enhanced ORR Catalytic Performance. C J. Carbon Res. 2022, 8, 13. [Google Scholar] [CrossRef]
  829. Wang, M.R.; Li, Y.; Fang, J.; Villa, C.J.; Xu, Y.B.; Hao, S.Q.; Li, J.; Liu, Y.X.; Wolverton, C.; Chen, X.Q.; et al. Superior Oxygen Reduction Reaction on Phosphorus-Doped Carbon Dot/Graphene Aerogel for All-Solid-State Flexible Al-Air Batteries. Adv. Energy Mater. 2020, 10, 1902736. [Google Scholar] [CrossRef]
  830. Suo, N.; Huang, H.; Wang, X.; Hou, X.D.; Shao, Z.G.; Zhang, G.F. Facile Synthesis and Electrocatalytic Performance for Oxygen Reduction of Boron-Doped Carbon Catalysts on Graphene Sheets. Fuel Cells 2021, 21, 328–336. [Google Scholar] [CrossRef]
  831. Qiao, M.; Tang, C.; He, G.; Qiu, K.; Binions, R.; Parkin, I.P.; Zhang, Q.; Guo, Z.; Titirici, M.M. Graphene/Nitrogen-Doped Porous Carbon Sandwiches for the Metal-Free Oxygen Reduction Reaction: Conductivity versus Active Sites. J. Mater. Chem. A Mater. 2016, 4, 12658–12666. [Google Scholar] [CrossRef]
  832. Ma, Y.W.; Zhang, L.R.; Li, J.J.; Ni, H.T.; Li, M.; Zhang, J.L.; Feng, X.M.; Fan, Q.L.; Hu, Z.; Huang, W. Carbon-Nitrogen/Graphene Composite as Metal-Free Electrocatalyst for the Oxygen Reduction Reaction. Chin. Sci. Bull. 2011, 56, 3583–3589. [Google Scholar] [CrossRef]
  833. Sun, Y.Q.; Li, C.; Shi, G.Q. Nanoporous Nitrogen Doped Carbon Modified Graphene as Electrocatalyst for Oxygen Reduction Reaction. J. Mater. Chem. 2012, 22, 12810–12816. [Google Scholar] [CrossRef]
  834. Zhuang, X.D.; Zhang, F.; Wu, D.Q.; Forler, N.; Liang, H.W.; Wagner, M.; Gehrig, D.; Hansen, M.R.; Laquai, F.; Feng, X.L. Two-Dimensional Sandwich-Type, Graphene-Based Conjugated Microporous Polymers. Angew. Chem. Int. Ed. 2013, 52, 9668–9672. [Google Scholar] [CrossRef] [PubMed]
  835. Zhang, Z.H.; Wu, P.Y. A Facile One-Pot Route towards Three-Dimensional Graphene-Based Microporous N-Doped Carbon Composites. RSC Adv. 2014, 4, 45619–45624. [Google Scholar] [CrossRef]
  836. Babu, K.F.; Rajagopalan, B.; Chung, J.S.; Choi, W.M. Facile Synthesis of Graphene/N-Doped Carbon Nanowire Composites as an Effective Electrocatalyst for the Oxygen Reduction Reaction. Int. J. Hydrogen Energy 2015, 40, 6827–6834. [Google Scholar] [CrossRef]
  837. Qu, K.G.; Zheng, Y.; Dai, S.; Qiao, S.Z. Polydopamine-Graphene Oxide Derived Mesoporous Carbon Nanosheets for Enhanced Oxygen Reduction. Nanoscale 2015, 7, 12598–12605. [Google Scholar] [CrossRef]
  838. Liu, S.W.; Zhang, H.M.; Zhao, Q.; Zhang, X.; Liu, R.R.; Ge, X.; Wang, G.Z.; Zhao, H.J.; Cai, W.P. Metal-Organic Framework Derived Nitrogen-Doped Porous Carbon@graphene Sandwich-like Structured Composites as Bifunctional Electrocatalysts for Oxygen Reduction and Evolution Reactions. Carbon 2016, 106, 74–83. [Google Scholar] [CrossRef]
  839. Wang, B.; Li, S.M.; Wu, X.Y.; Liu, J.H.; Chen, J. Biomass Chitin-Derived Honeycomb-like Nitrogen-Doped Carbon/Graphene Nanosheet Networks for Applications in Efficient Oxygen Reduction and Robust Lithium Storage. J. Mater. Chem. A Mater. 2016, 4, 11789–11799. [Google Scholar] [CrossRef]
  840. Niu, W.H.; Li, L.G.; Liu, J.; Wang, N.; Li, W.; Tang, Z.H.; Zhou, W.J.; Chen, S.W. Graphene-Supported Mesoporous Carbons Prepared with Thermally Removable Templates as Efficient Catalysts for Oxygen Electroreduction. Small 2016, 12, 1900–1908. [Google Scholar] [CrossRef]
  841. Qin, L.; Yuan, Y.F.; Wei, W.; Lv, W.; Niu, S.Z.; He, Y.B.; Zhai, D.Y.; Kang, F.Y.; Kim, J.K.; Yang, Q.H.; et al. Graphene-Directed Formation of a Nitrogen-Doped Porous Carbon Sheet with High Catalytic Performance for the Oxygen Reduction Reaction. J. Phys. Chem. C 2018, 122, 13508–13514. [Google Scholar] [CrossRef]
  842. Sun, J.; Lowe, S.E.; Zhang, L.J.; Wang, Y.Z.; Pang, K.L.; Wang, Y.; Zhong, Y.L.; Liu, P.R.; Zhao, K.; Tang, Z.Y.; et al. Ultrathin Nitrogen-Doped Holey Carbon@Graphene Bifunctional Electrocatalyst for Oxygen Reduction and Evolution Reactions in Alkaline and Acidic Media. Angew. Chem. Int. Ed. 2018, 57, 16511–16515. [Google Scholar] [CrossRef]
  843. Wang, N.; Tian, H.; Zhu, S.Y.; Yan, D.Y.; Mai, Y.Y. Two-Dimensional Nitrogen-Doped Mesoporous Carbon/Graphene Nanocomposites from the Self-Assembly of Block Copolymer Micelles in Solution. Chin. J. Polym. Sci. 2018, 36, 266–272. [Google Scholar] [CrossRef]
  844. Hou, D.; Zhang, J.C.; Tian, H.; Li, Q.; Li, C.; Mai, Y.Y. Pore Engineering of 2D Mesoporous Nitrogen-Doped Carbon on Graphene through Block Copolymer Self-Assembly. Adv. Mater. Interfaces 2019, 6, 1901476. [Google Scholar] [CrossRef]
  845. Sheng, K.; Yi, Q.F.; Hou, L.F.; Chen, A.L. Metal-Free Graphene Modified Nitrogen-Doped Ultra-Thin Hollow Carbon Spheres as Superior Cathodic Catalysts of Zn-Air Battery. J. Electrochem. Soc. 2020, 167, 070560. [Google Scholar] [CrossRef]
  846. Begum, H.; Ahmed, M.S.; Jung, S. Template-Free Synthesis of Polyacrylonitrile-Derived Porous Carbon Nanoballs on Graphene for Efficient Oxygen Reduction in Zinc-Air Batteries. J. Mater. Chem. A Mater. 2021, 9, 9644–9654. [Google Scholar] [CrossRef]
  847. Gai, H.Y.; Xue, S.; Wang, X.K.; Zhou, J.; Jiang, H.Q.; Huang, M.H. Sandwich-like Hierarchical Porous Dual-Carbon Catalyst with More Accessible Sites for Boosting Oxygen Reduction Reaction. Mater. Today Energy 2021, 21, 8. [Google Scholar] [CrossRef]
  848. Ilnicka, A.; Skorupska, M.; Tyc, M.; Kowalska, K.; Kamedulski, P.; Zielinski, W.; Lukaszewicz, J.P. Green Algae and Gelatine Derived Nitrogen Rich Carbon as an Outstanding Competitor to Pt Loaded Carbon Catalysts. Sci. Rep. 2021, 11, 13. [Google Scholar] [CrossRef]
  849. Kim, I.Y.; Kim, S.; Jin, X.Y.; Premkumar, S.; Chandra, G.; Lee, N.S.; Mane, G.P.; Hwang, S.J.; Umapathy, S.; Vinu, A. Ordered Mesoporous C3N5 with a Combined Triazole and Triazine Framework and Its Graphene Hybrids for the Oxygen Reduction Reaction (ORR). Angew. Chem. Int. Ed. 2018, 57, 17135–17140. [Google Scholar] [CrossRef]
  850. Xu, J.Y.; Liu, B. Intrinsic Properties of Nitrogen-Rich Carbon Nitride for Oxygen Reduction Reaction. Appl. Surf. Sci. 2020, 500, 144020. [Google Scholar] [CrossRef]
  851. Sun, Y.Q.; Li, C.; Xu, Y.X.; Bai, H.; Yao, Z.Y.; Shi, G.Q. Chemically Converted Graphene as Substrate for Immobilizing and Enhancing the Activity of a Polymeric Catalyst. Chem. Commun. 2010, 46, 4740–4742. [Google Scholar] [CrossRef]
  852. Yang, S.B.; Feng, X.L.; Wang, X.C.; Mullen, K. Graphene-Based Carbon Nitride Nanosheets as Efficient Metal-Free Electrocatalysts for Oxygen Reduction Reactions. Angew. Chem. Int. Ed. 2011, 50, 5339–5343. [Google Scholar] [CrossRef]
  853. Qin, Y.; Li, J.; Yuan, J.; Kong, Y.; Tao, Y.X.; Lin, F.R.; Li, S. Hollow Mesoporous Carbon Nitride Nanosphere/Three-Dimensional Graphene Composite as High Efficient Electrocatalyst for Oxygen Reduction Reaction. J. Power Sources 2014, 272, 696–702. [Google Scholar] [CrossRef]
  854. Qiu, K.P.; Guo, Z.X. Hierarchically Porous Graphene Sheets and Graphitic Carbon Nitride Intercalated Composites for Enhanced Oxygen Reduction Reaction. J. Mater. Chem. A Mater. 2014, 2, 3209–3215. [Google Scholar] [CrossRef]
  855. Tian, J.Q.; Ning, R.; Liu, Q.; Asiri, A.M.; Al-Youbi, A.O.; Sun, X.P. Three-Dimensional Porous Supramolecular Architecture from Ultrathin g-C3N4 Nanosheets and Reduced Graphene Oxide: Solution Self-Assembly Construction and Application as a Highly Efficient Metal-Free Electrocatalyst for Oxygen Reduction Reaction. ACS Appl. Mater. Interfaces 2014, 6, 1011–1017. [Google Scholar] [CrossRef]
  856. Wang, X.P.; Wang, L.X.; Zhao, F.; Hu, C.G.; Zhao, Y.; Zhang, Z.P.; Chen, S.L.; Shi, G.Q.; Qu, L.T. Monoatomic-Thick Graphitic Carbon Nitride Dots on Graphene Sheets as an Efficient Catalyst in the Oxygen Reduction Reaction. Nanoscale 2015, 7, 3035–3042. [Google Scholar] [CrossRef]
  857. Li, C.X.; Li, X.S.; Zhang, X.Y.; Yang, X.J.; Wang, L.Y.; Lu, W. The G-C3N4 Quantum Dot Decorated g-C3N4 Sheet/Reduced Graphene Oxide Composite as Efficient Metal-Free Electrocatalyst for Oxygen Reduction Reaction. J. Electrochem. Soc. 2020, 167, 100534. [Google Scholar] [CrossRef]
  858. Mane, R.S.; Periyasamy, G.; Jha, N. Sunlight Assisted Lewis Base Enriched 2D/2D g-C3N4-Graphene Oxide Composite as an Efficient ORR Electrocatalyst. Electrochim. Acta 2024, 481, 12. [Google Scholar] [CrossRef]
  859. Garcia, J.L.; Miyao, T.; Inukai, J.; Tongol, B.J. V Graphitic Carbon Nitride on Reduced Graphene Oxide Prepared via Semi-Closed Pyrolysis as Electrocatalyst for Oxygen Reduction Reaction. Mater. Chem. Phys. 2022, 288, 9. [Google Scholar] [CrossRef]
  860. Hu, C.; Yu, C.; Li, M.Y.; Wang, X.N.; Dong, Q.; Wang, G.; Qiu, J.S. Nitrogen-Doped Carbon Dots Decorated on Graphene: A Novel All-Carbon Hybrid Electrocatalyst for Enhanced Oxygen Reduction Reaction. Chem. Commun. 2015, 51, 3419–3422. [Google Scholar] [CrossRef]
  861. Samantara, A.K.; Sahu, S.C.; Ghosh, A.; Jena, B.K. Sandwiched Graphene with Nitrogen, Sulphur Co-Doped CQDs: An Efficient Metal-Free Material for Energy Storage and Conversion Applications. J. Mater. Chem. A Mater. 2015, 3, 16961–16970. [Google Scholar] [CrossRef]
  862. Zhou, L.H.; Fu, P.; Wang, Y.Q.; Sun, L.H.; Yuan, Y. Microbe-Engaged Synthesis of Carbon Dot-Decorated Reduced Graphene Oxide as High-Performance Oxygen Reduction Catalysts. J. Mater. Chem. A Mater. 2016, 4, 7222–7229. [Google Scholar] [CrossRef]
  863. Shin, J.H.; Guo, J.; Zhao, T.T.; Guo, Z.X. Functionalized Carbon Dots on Graphene as Outstanding Non-Metal Bifunctional Oxygen Electrocatalyst. Small 2019, 15, 9. [Google Scholar] [CrossRef] [PubMed]
  864. Ding, H.; Wei, J.S.; Xiong, H.M. Nitrogen and Sulfur Co-Doped Carbon Dots with Strong Blue Luminescence. Nanoscale 2014, 6, 13817–13823. [Google Scholar] [CrossRef] [PubMed]
  865. Xu, P.M.; Wu, D.Q.; Wan, L.; Hu, P.F.; Liu, R.L. Heteroatom Doped Mesoporous Carbon/Graphene Nanosheets as Highly Efficient Electrocatalysts for Oxygen Reduction. J. Colloid. Interface Sci. 2014, 421, 160–164. [Google Scholar] [CrossRef]
  866. Ma, Y.W.; Sun, L.Y.; Huang, W.; Zhang, L.R.; Zhao, J.; Fan, Q.L. Three-Dimensional Nitrogen-Doped Carbon Nanotubes/Graphene Structure Used as a Metal-Free Electrocatalyst for the Oxygen Reduction Reaction. J. Phys. Chem. C 2011, 115, 24592–24597. [Google Scholar] [CrossRef]
  867. Wang, X.W.; Ai, W.; Li, N.; Yu, T.; Chen, P. Graphene-Bacteria Composite for Oxygen Reduction and Lithium Ion Batteries. J. Mater. Chem. A Mater. 2015, 3, 12873–12879. [Google Scholar] [CrossRef]
  868. Qu, K.G.; Zheng, Y.; Dai, S.; Qiao, S.Z. Graphene Oxide-Polydopamine Derived N, S-Codoped Carbon Nanosheets as Superior Bifunctional Electrocatalysts for Oxygen Reduction and Evolution. Nano Energy 2016, 19, 373–381. [Google Scholar] [CrossRef]
  869. Liu, G.L.; Liu, Z.M.; Li, J.L.; Zeng, M.; Li, Z.Y.; He, L.; Li, F.W. Chitosan/Phytic Acid Hydrogel as a Platform for Facile Synthesis of Heteroatom-Doped Porous Carbon Frameworks for Electrocatalytic Oxygen Reduction. Carbon 2018, 137, 68–77. [Google Scholar] [CrossRef]
  870. Tan, H.B.; Zhao, Y.J.; Xia, W.; Zhao, J.C.; Xu, X.T.; Wood, K.; Sugahara, Y.; Yamauchi, Y.; Tang, J. Phosphorus- and Nitrogen-Doped Carbon Nanosheets Constructed with Monolayered Mesoporous Architectures. Chem. Mater. 2020, 32, 4248–4256. [Google Scholar] [CrossRef]
  871. Wang, S.Y.; Yu, D.S.; Dai, L.M.; Chang, D.W.; Baek, J.B. Polyelectrolyte-Functionalized Graphene as Metal-Free Electrocatalysts for Oxygen Reduction. ACS Nano 2011, 5, 6202–6209. [Google Scholar] [CrossRef]
  872. Huang, D.K.; Zhang, B.Y.; Zhang, Y.B.; Zhan, F.; Xu, X.B.; Shen, Y.; Wang, M.K. Electrochemically Reduced Graphene Oxide Multilayer Films as Metal-Free Electrocatalysts for Oxygen Reduction. J. Mater. Chem. A Mater. 2013, 1, 1415–1420. [Google Scholar] [CrossRef]
  873. Lee, J.S.; Jo, K.; Lee, T.; Yun, T.; Cho, J.; Kim, B.S. Facile Synthesis of Hybrid Graphene and Carbon Nanotubes as a Metal-Free Electrocatalyst with Active Dual Interfaces for Efficient Oxygen Reduction Reaction. J. Mater. Chem. A Mater. 2013, 1, 9603–9607. [Google Scholar] [CrossRef]
  874. Xiao, X.; Wang, T.J.; Bai, J.; Li, F.M.; Ma, T.Y.; Chen, Y. Enhancing the Selectivity of H2O2 Electrogeneration by Steric Hindrance Effect. ACS Appl. Mater. Interfaces 2018, 10, 42534–42541. [Google Scholar] [CrossRef]
  875. Getachew, T.; Addis, F.; Beyene, T.; Mehretie, S.; Admassie, S. Amino-Substituted Naphthalene Sulfonic Acid/Graphene Composite as Metal-Free Catalysts for Oxygen Reduction Reactions. Bull. Chem. Soc. Ethiop. 2019, 33, 359–372. [Google Scholar] [CrossRef]
  876. Wang, D.D.; Gao, X.L.; Zhao, L.M.; Zhou, J.; Zhuo, S.P.; Yan, Z.F.; Xing, W. Polydopamine-Coated Graphene Nanosheets as Efficient Electrocatalysts for Oxygen Reduction Reaction. RSC Adv. 2018, 8, 16044–16051. [Google Scholar] [CrossRef]
  877. Xu, Y.Z.; Chen, C.L.; Zhou, M.; Fu, G.Y.; Zhao, Y.Y.; Chen, Y.H. Improved Oxygen Reduction Activity of Carbon Nanotubes and Graphene through Adenine Functionalization. RSC Adv. 2017, 7, 26722–26728. [Google Scholar] [CrossRef]
  878. Zhang, M.; Yuan, W.J.; Yao, B.W.; Li, C.; Shi, G.Q. Solution-Processed PEDOT:PSS/Graphene Composites as the Electrocatalyst for Oxygen Reduction Reaction. ACS Appl. Mater. Interfaces 2014, 6, 3587–3593. [Google Scholar] [CrossRef]
  879. Long, X.J.; Li, D.H.; Wang, B.B.; Jiang, Z.J.; Xu, W.J.; Yang, D.J.; Xia, Y.Z. Heterocyclization Strategy for Construction of Linear Conjugated Polymers: Efficient Metal-Free Electrocatalysts for Oxygen Reduction. Angew. Chem. Int. Ed. 2019, 58, 11369–11373. [Google Scholar] [CrossRef]
  880. Shervedani, R.K.; Amini, A. Preparation of Graphene/Nile Blue Nanocomposite: Application for Oxygen Reduction Reaction and Biosensing. Electrochim. Acta 2015, 173, 354–363. [Google Scholar] [CrossRef]
  881. Cossio, J.J.B.; Pena, P.A.; Hernandez-Gordillo, A.; Garcia, L.F.D.; Santiago, A.R.P.; Echegoyen, L.; Reguera, E. In Situ Aniline-Polymerized Interfaces on GO-PVA Nanoplatforms as Bifunctional Supercapacitors and PH-Universal ORR Electrodes. ACS Appl. Energy Mater. 2020, 3, 4727–4737. [Google Scholar] [CrossRef]
  882. Pattanayak, P.; Pramanik, N.; Papiya, F.; Kumar, V.; Kundu, P.P. Metal-Free Keratin Modified Poly (Pyrrole-Co-Aniline)-Reduced Graphene Oxide Based Nanocomposite Materials: A Promising Cathode Catalyst in Microbial Fuel Cell Application. J. Environ. Chem. Eng. 2020, 8, 103813. [Google Scholar] [CrossRef]
  883. Wei, W.; Tao, Y.; Lv, W.; Su, F.Y.; Ke, L.; Li, J.; Wang, D.W.; Li, B.H.; Kang, F.Y.; Yang, Q.H. Unusual High Oxygen Reduction Performance in All-Carbon Electrocatalysts. Sci. Rep. 2014, 4, 6289. [Google Scholar] [CrossRef] [PubMed]
  884. Cai, X.Y.; Xia, B.Y.; Franklin, J.; Li, B.S.; Wang, X.; Wang, Z.; Chen, L.W.; Lin, J.Y.; Lai, L.F.; Shen, Z.X. Free-Standing Vertically-Aligned Nitrogen-Doped Carbon Nanotube Arrays/Graphene as Air-Breathing Electrodes for Rechargeable Zinc-Air Batteries. J. Mater. Chem. A Mater. 2017, 5, 2488–2495. [Google Scholar] [CrossRef]
  885. Luo, J.M.; Yang, L.M.; Li, T.; Yang, L.X.; Luo, X.B.; Crittenden, J.C. Three-Dimensional Electrode Interface Assembled from RGO Nanosheets and Carbon Nanotubes for Highly Electrocatalytic Oxygen Reduction. Chem. Eng. J. 2019, 378, 122127. [Google Scholar] [CrossRef]
  886. Kong, F.T.; Qiao, Y.; Zhang, C.Q.; Fan, X.H.; Kong, A.G.; Shan, Y.K. Unadulterated Carbon as Robust Multifunctional Electrocatalyst for Overall Water Splitting and Oxygen Transformation. Nano Res. 2020, 13, 401–411. [Google Scholar] [CrossRef]
  887. Abreu, B.; Rocha, M.; Nunes, M.; Freire, C.; Marques, E.F. Carbon Nanotube/Graphene Nanocomposites Built via Surfactant-Mediated Colloid Assembly as Metal-Free Catalysts for the Oxygen Reduction Reaction. J. Mater. Sci. 2021, 56, 19512–19527. [Google Scholar] [CrossRef]
  888. Patil, I.M.; Lokanathan, M.; Kakade, B. Three Dimensional Nanocomposite of Reduced Graphene Oxide and Hexagonal Boron Nitride as an Efficient Metal-Free Catalyst for Oxygen Electroreduction. J. Mater. Chem. A Mater. 2016, 4, 4506–4515. [Google Scholar] [CrossRef]
  889. Yuan, G.E.; Zhang, G.Q.; Chen, J.; Fu, L.; Xu, L.; Yang, F.L. The Electrochemical Activities of Anthraquinone Monosulfonate Adsorbed on the Basal Plane of Reduced Graphene Oxide by Pi-Pi Stacking Interaction. J. Solid State Electrochem. 2013, 17, 2711–2719. [Google Scholar] [CrossRef]
  890. Shen, A.L.; Xia, W.J.; Zhang, L.P.; Dou, S.; Xia, Z.H.; Wang, S.Y. Charge Transfer Induced Activity of Graphene for Oxygen Reduction. Nanotechnology 2016, 27, 185402. [Google Scholar] [CrossRef]
  891. Wu, Y.; Liang, L.D.; Zhou, C.; Zheng, X.; Chen, Y.G.; Rittmann, B.E. Biogenic Melanin-Modified Graphene as a Cathode Catalyst Yields Greater Bioelectrochemical Performances by Stimulating Oxygen-Reduction and Microbial Electron Transfer. Acs EsT Water 2023, 8, 3369–3376. [Google Scholar] [CrossRef]
  892. Di Noto, V.; Negro, E.; Vezzù, K.; Bertasi, F.; Nawn, G. Origins, Developments, and Perspectives of Carbon Nitride-Based Electrocatalysts for Application in Low-Temperature FCs. Electrochem. Soc. Interface 2015, 24, 59–63. [Google Scholar] [CrossRef]
Figure 1. Schematic representation of (A) PEMFC and (B) AEMFC: the reactions occurring at both anode and cathode during their operation with H2 are reported [13]. © 2023 The Chemical Society of Japan and Wiley-VCH GmbH.
Figure 1. Schematic representation of (A) PEMFC and (B) AEMFC: the reactions occurring at both anode and cathode during their operation with H2 are reported [13]. © 2023 The Chemical Society of Japan and Wiley-VCH GmbH.
Molecules 30 02248 g001
Figure 2. Microbial fuel cell scheme from [14]. © 2016 Elsevier Ltd. (2016).
Figure 2. Microbial fuel cell scheme from [14]. © 2016 Elsevier Ltd. (2016).
Molecules 30 02248 g002
Figure 3. Schematic operating process of aqueous metal–air batteries from [16]. © 2019 Elsevier B.V. (2019).
Figure 3. Schematic operating process of aqueous metal–air batteries from [16]. © 2019 Elsevier B.V. (2019).
Molecules 30 02248 g003
Figure 4. Schematic illustration of solvothermal-assisted exfoliation and dispersion of graphene sheets in acetonitrile: (a) pristine expandable graphite; (b) ExG obtained by heating at 1000 °C under an atmosphere of forming gas (5% H2 and 95% Ar); (c) insertion of acetonitrile molecules into the interlayers of ExG; (d) exfoliated graphene sheets dispersed in acetonitrile; (e) optical images of four samples obtained under different processes and centrifugation conditions [22].
Figure 4. Schematic illustration of solvothermal-assisted exfoliation and dispersion of graphene sheets in acetonitrile: (a) pristine expandable graphite; (b) ExG obtained by heating at 1000 °C under an atmosphere of forming gas (5% H2 and 95% Ar); (c) insertion of acetonitrile molecules into the interlayers of ExG; (d) exfoliated graphene sheets dispersed in acetonitrile; (e) optical images of four samples obtained under different processes and centrifugation conditions [22].
Molecules 30 02248 g004
Figure 5. Exfoliation and N doping using supercritical ammonia. Reprinted with permission from [64]. Copyright {2016} American Chemical Society.
Figure 5. Exfoliation and N doping using supercritical ammonia. Reprinted with permission from [64]. Copyright {2016} American Chemical Society.
Molecules 30 02248 g005
Figure 6. Structure of GO [66] reprinted with permission from Springer Nature Chemistry © 2009 Macmillan Publishers Limited.
Figure 6. Structure of GO [66] reprinted with permission from Springer Nature Chemistry © 2009 Macmillan Publishers Limited.
Molecules 30 02248 g006
Figure 7. Representation of the procedures followed, starting with graphite flakes. Images of the oxidized hydrophobic carbon material collected during purification using IGO, HGO, and HGO+ (in the glass bottles) are presented. Reprinted with permission from [69]. Copyright {2010} American Chemical Society.
Figure 7. Representation of the procedures followed, starting with graphite flakes. Images of the oxidized hydrophobic carbon material collected during purification using IGO, HGO, and HGO+ (in the glass bottles) are presented. Reprinted with permission from [69]. Copyright {2010} American Chemical Society.
Molecules 30 02248 g007
Figure 8. Illustration of the preparation of GO based on a newly improved Hummers’ method from [76].
Figure 8. Illustration of the preparation of GO based on a newly improved Hummers’ method from [76].
Molecules 30 02248 g008
Scheme 1. Schematic representation of the major methods of oxidation of graphite to graphene oxide and the chemical reduction of graphene oxide by some reductants from [82]. © 2017 Elsevier Ltd. (2017).
Scheme 1. Schematic representation of the major methods of oxidation of graphite to graphene oxide and the chemical reduction of graphene oxide by some reductants from [82]. © 2017 Elsevier Ltd. (2017).
Molecules 30 02248 sch001
Figure 9. Ring opening of epoxides induced by hydrazine reactions and incorporation of N in the aromatic skeleton. Adapted with permission from [121]. Copyright {2010} American Chemical Society.
Figure 9. Ring opening of epoxides induced by hydrazine reactions and incorporation of N in the aromatic skeleton. Adapted with permission from [121]. Copyright {2010} American Chemical Society.
Molecules 30 02248 g009
Scheme 2. Illustration of the nitrogen doping process of melamine into graphite oxide layers, adapted with permission from [127]. Copyright {2011} American Chemical Society.
Scheme 2. Illustration of the nitrogen doping process of melamine into graphite oxide layers, adapted with permission from [127]. Copyright {2011} American Chemical Society.
Molecules 30 02248 sch002
Figure 10. Schematic diagram of the DC-arc discharge set-up [136].
Figure 10. Schematic diagram of the DC-arc discharge set-up [136].
Molecules 30 02248 g010
Figure 11. Illustration of the Cu-based graphene growth mechanism and graphene samples grown under different conditions. (a) Scheme for the Cu-based graphene growth mechanism: red hexagons are used to symbolize the active sites of the Cu surface, and blue spots signify active carbon species (CHx<4). (bg) Typical scanning electron microscopy (SEM) images of graphene synthesized under different growth conditions: (b) (E1), (c) (E2), (d) (E3), (e) (E4), (f) (E5), and (g) (E6). Reprinted with permission from [166]. Copyright {2012} American Chemical Society.
Figure 11. Illustration of the Cu-based graphene growth mechanism and graphene samples grown under different conditions. (a) Scheme for the Cu-based graphene growth mechanism: red hexagons are used to symbolize the active sites of the Cu surface, and blue spots signify active carbon species (CHx<4). (bg) Typical scanning electron microscopy (SEM) images of graphene synthesized under different growth conditions: (b) (E1), (c) (E2), (d) (E3), (e) (E4), (f) (E5), and (g) (E6). Reprinted with permission from [166]. Copyright {2012} American Chemical Society.
Molecules 30 02248 g011
Scheme 3. Scheme of a proposed mechanism for solvothermal synthesis of N-doped graphene via reaction of CCl4 and Li3N, where gray balls represent C atoms, blue represent N, green represent Cl, and purple represent Li. Reprinted with permission from [205]. Copyright {2011} American Chemical Society.
Scheme 3. Scheme of a proposed mechanism for solvothermal synthesis of N-doped graphene via reaction of CCl4 and Li3N, where gray balls represent C atoms, blue represent N, green represent Cl, and purple represent Li. Reprinted with permission from [205]. Copyright {2011} American Chemical Society.
Molecules 30 02248 sch003
Figure 12. Formation of C-graphene by solvothermal reaction between CCl4 and potassium. The photograph is of the autoclave after the reaction, showing the formation of C-graphene (black) and potassium chloride (KCl; white) [207] © 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 12. Formation of C-graphene by solvothermal reaction between CCl4 and potassium. The photograph is of the autoclave after the reaction, showing the formation of C-graphene (black) and potassium chloride (KCl; white) [207] © 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Molecules 30 02248 g012
Figure 13. Schematic representation of the bottom-up synthesis of C-NPG from molecular precursors 1 and 2 (anti-configuration) and structure of the C-NPG anti-configuration with the interface (rubicene) groups highlighted in red and the 2D unit cell indicated by a rectangle. Reprinted with permission from [216]. Copyright {2020} American Chemical Society.
Figure 13. Schematic representation of the bottom-up synthesis of C-NPG from molecular precursors 1 and 2 (anti-configuration) and structure of the C-NPG anti-configuration with the interface (rubicene) groups highlighted in red and the 2D unit cell indicated by a rectangle. Reprinted with permission from [216]. Copyright {2020} American Chemical Society.
Molecules 30 02248 g013
Figure 14. Reaction pathway for the formation of 1D porous graphene nanoribbons on Ag(111) and 2D porous graphene nanosheets on Au(111), starting from TBDBTP precursors. Color code: Ag, blue; Br, orange; C, gray; H, white [218] © 2023 WILEY-VCH Verlag GmbH.
Figure 14. Reaction pathway for the formation of 1D porous graphene nanoribbons on Ag(111) and 2D porous graphene nanosheets on Au(111), starting from TBDBTP precursors. Color code: Ag, blue; Br, orange; C, gray; H, white [218] © 2023 WILEY-VCH Verlag GmbH.
Molecules 30 02248 g014
Figure 15. LIG formation under controlled atmosphere. (a) Scheme and (b) photograph of the apparatus for the fabrication of LIG under a controlled atmosphere. Contact angles of LIG prepared in (c) ambient air, (d) air (chamber), (e) O2 (chamber), (f) Ar (chamber), (g) H2 (chamber), and (h) SF6 (chamber). Reprinted with permission from [224]. Copyright {2018} American Chemical Society.
Figure 15. LIG formation under controlled atmosphere. (a) Scheme and (b) photograph of the apparatus for the fabrication of LIG under a controlled atmosphere. Contact angles of LIG prepared in (c) ambient air, (d) air (chamber), (e) O2 (chamber), (f) Ar (chamber), (g) H2 (chamber), and (h) SF6 (chamber). Reprinted with permission from [224]. Copyright {2018} American Chemical Society.
Molecules 30 02248 g015
Figure 16. Schematic of the formation of LIPG from the PI/APP film. (a) Preparation of the PI/APP film from PAA containing the APP solution; (b) LIPG fabricated on PI/APP film by laser induction [233].
Figure 16. Schematic of the formation of LIPG from the PI/APP film. (a) Preparation of the PI/APP film from PAA containing the APP solution; (b) LIPG fabricated on PI/APP film by laser induction [233].
Molecules 30 02248 g016
Figure 17. Schematic outline of the general solid-state pyrolytic conversion strategy for the preparation of monolayer graphene. The blue spheres represent the Na2CO3 salt, the large green sphere represents Na+ in the solid sodium carboxylate precursor, and the small green spheres represent the Na2CO3 crystals generated in situ; reprinted with permission from Springer Nano Research©Tsinghua University Press and Springer-Verlag Berlin Heidelberg 2017.
Figure 17. Schematic outline of the general solid-state pyrolytic conversion strategy for the preparation of monolayer graphene. The blue spheres represent the Na2CO3 salt, the large green sphere represents Na+ in the solid sodium carboxylate precursor, and the small green spheres represent the Na2CO3 crystals generated in situ; reprinted with permission from Springer Nano Research©Tsinghua University Press and Springer-Verlag Berlin Heidelberg 2017.
Molecules 30 02248 g017
Figure 18. Pyrolytic apparatus and green materials used for the preparation of graphene [243].
Figure 18. Pyrolytic apparatus and green materials used for the preparation of graphene [243].
Molecules 30 02248 g018
Figure 19. Preparation scheme and SEM images of chlorinated VG. (a) Schematic illustration of VG grown on CC using the PECVD method, and subsequent chlorination. SEM images of (b) CC; (c) pristine CC@VG; (d) chlorinated CC@VG for 2 min; (e) chlorinated CC@VG for 4 min; (f) chlorinated CC@VG for 6 min; (g) chlorinated CC@VG for 8 min; (h) chlorinated CC@VG for 12 min; and (i) chlorinated CC@VG for 20 min, from [245]. © 2019 Elsevier Inc. (2019).
Figure 19. Preparation scheme and SEM images of chlorinated VG. (a) Schematic illustration of VG grown on CC using the PECVD method, and subsequent chlorination. SEM images of (b) CC; (c) pristine CC@VG; (d) chlorinated CC@VG for 2 min; (e) chlorinated CC@VG for 4 min; (f) chlorinated CC@VG for 6 min; (g) chlorinated CC@VG for 8 min; (h) chlorinated CC@VG for 12 min; and (i) chlorinated CC@VG for 20 min, from [245]. © 2019 Elsevier Inc. (2019).
Molecules 30 02248 g019
Figure 20. TEM image of the (A) Graphene-g-PPy; (B) mWCnT-g-PPy; (C) carbon nanohorn-g-PPy; (D) and the selected area electron diffraction (SAED) patterns of pure graphene (right, top) [246].
Figure 20. TEM image of the (A) Graphene-g-PPy; (B) mWCnT-g-PPy; (C) carbon nanohorn-g-PPy; (D) and the selected area electron diffraction (SAED) patterns of pure graphene (right, top) [246].
Molecules 30 02248 g020
Figure 21. Atomic resolution, aberration-corrected TEM image of a single-layer reduced-graphene oxide membrane. (a) Original image and (b) with color added to highlight the different features. Adapted with permission from [247]. Copyright {2010} American Chemical Society.
Figure 21. Atomic resolution, aberration-corrected TEM image of a single-layer reduced-graphene oxide membrane. (a) Original image and (b) with color added to highlight the different features. Adapted with permission from [247]. Copyright {2010} American Chemical Society.
Molecules 30 02248 g021
Figure 22. ADF-STEM images of graphene crystals. (a) SEM image of graphene transferred onto a TEM grid with over 90% coverage using novel, high-yield methods. Scale bar, 5 μm. (b) ADF-STEM image showing the defect-free hexagonal lattice inside a graphene grain. (c) Two grains (bottom left, top right) intersect with a 27° relative rotation. An aperiodic line of defects stitches the two grains together. (d) The image from c with the pentagons (blue), heptagons (red), and distorted hexagons (green) of the grain boundary outlined. (bd) Images were low-pass-filtered to remove noise; scale bars, 5 Å. Reprinted with permission from Springer Nature© 2011 Macmillan Publishers Limited.
Figure 22. ADF-STEM images of graphene crystals. (a) SEM image of graphene transferred onto a TEM grid with over 90% coverage using novel, high-yield methods. Scale bar, 5 μm. (b) ADF-STEM image showing the defect-free hexagonal lattice inside a graphene grain. (c) Two grains (bottom left, top right) intersect with a 27° relative rotation. An aperiodic line of defects stitches the two grains together. (d) The image from c with the pentagons (blue), heptagons (red), and distorted hexagons (green) of the grain boundary outlined. (bd) Images were low-pass-filtered to remove noise; scale bars, 5 Å. Reprinted with permission from Springer Nature© 2011 Macmillan Publishers Limited.
Molecules 30 02248 g022
Figure 23. (a) AFM image of the GO films. (b) Histogram of the AFM-depth intensities obtained from the dotted area of the image (a). The histogram is fitted with a linear combination of four Gaussian functions representing each peak. Adapted with permission from [263]. Copyright {2009} American Chemical Society.
Figure 23. (a) AFM image of the GO films. (b) Histogram of the AFM-depth intensities obtained from the dotted area of the image (a). The histogram is fitted with a linear combination of four Gaussian functions representing each peak. Adapted with permission from [263]. Copyright {2009} American Chemical Society.
Molecules 30 02248 g023
Figure 24. Series of STM images of the carbon-saturated layer taken during the phase transition of the disordered carbon layer into graphene at 920 K. The motion of the hole on the central terrace transforms the flat, disordered carbon layer into the graphene Moiré structure. 1100 Å × 1100 Å, tunneling current 0.07 nA, tunneling voltage 1.24 V. Reprinted with permission from [265]. Copyright {2013} American Chemical Society.
Figure 24. Series of STM images of the carbon-saturated layer taken during the phase transition of the disordered carbon layer into graphene at 920 K. The motion of the hole on the central terrace transforms the flat, disordered carbon layer into the graphene Moiré structure. 1100 Å × 1100 Å, tunneling current 0.07 nA, tunneling voltage 1.24 V. Reprinted with permission from [265]. Copyright {2013} American Chemical Society.
Molecules 30 02248 g024
Figure 25. Nanometer- and atomic-scale STM images for the highly reduced graphene oxide sheets prior to thermal treatment (a,b) after annealing at 1773 K (c,d), 1923 K (e,f), 2073 K (g,h) and 2223 K (i,j). The inset to (h) is a detailed 2 × 2 nm2 image showing a different image of the same type of defect. Imaging conditions: 0.3 nA (tunneling current) and 500 mV (bias voltage) (a,c,e,g,i); 1–4 nA and 5–10 mV (b,d,f,j); 0.5 nA and 80 mV (h); 0.8 nA and 50 mV (inset). Adapted from [266] from the Royal Society of Chemistry. https://pubs.rsc.org/en/content/articlelanding/2015/nr/c4nr05816j (accessed on 1 March 2025).
Figure 25. Nanometer- and atomic-scale STM images for the highly reduced graphene oxide sheets prior to thermal treatment (a,b) after annealing at 1773 K (c,d), 1923 K (e,f), 2073 K (g,h) and 2223 K (i,j). The inset to (h) is a detailed 2 × 2 nm2 image showing a different image of the same type of defect. Imaging conditions: 0.3 nA (tunneling current) and 500 mV (bias voltage) (a,c,e,g,i); 1–4 nA and 5–10 mV (b,d,f,j); 0.5 nA and 80 mV (h); 0.8 nA and 50 mV (inset). Adapted from [266] from the Royal Society of Chemistry. https://pubs.rsc.org/en/content/articlelanding/2015/nr/c4nr05816j (accessed on 1 March 2025).
Molecules 30 02248 g025aMolecules 30 02248 g025b
Figure 26. Raman spectra of five ion bombarded single layer graphene (SLG) measured at laser energy (EL) = 2.41 eV (λL = 514.5 nm), where the main Raman peaks are labeled. The distance between defects (LD) is independently measured following the procedure of Lucchese [271]. The respective ID/IG values are indicated for each spectrum. The notations within parentheses [e.g., 2D(G’)] indicate two commonly used notations for the same peak (2D and G’) [267,269]. Adapted with permission from [270]. Copyright {2011} American Chemical Society.
Figure 26. Raman spectra of five ion bombarded single layer graphene (SLG) measured at laser energy (EL) = 2.41 eV (λL = 514.5 nm), where the main Raman peaks are labeled. The distance between defects (LD) is independently measured following the procedure of Lucchese [271]. The respective ID/IG values are indicated for each spectrum. The notations within parentheses [e.g., 2D(G’)] indicate two commonly used notations for the same peak (2D and G’) [267,269]. Adapted with permission from [270]. Copyright {2011} American Chemical Society.
Molecules 30 02248 g026
Figure 27. (a) Raman spectra of five ion-bombarded SLGs measured with laser excitation energy, EL, of 1.58 eV. (b) Raman spectra of an ion-bombarded SLG with LD = 7 nm obtained using the three excitation energies reported in the table on the right. The ∼1450 cm−1 band in the Raman spectra at 785 nm is a third-order peak of the silicon substrate. Adapted with permission from [270] Copyright {2011}. American Chemical Society.
Figure 27. (a) Raman spectra of five ion-bombarded SLGs measured with laser excitation energy, EL, of 1.58 eV. (b) Raman spectra of an ion-bombarded SLG with LD = 7 nm obtained using the three excitation energies reported in the table on the right. The ∼1450 cm−1 band in the Raman spectra at 785 nm is a third-order peak of the silicon substrate. Adapted with permission from [270] Copyright {2011}. American Chemical Society.
Molecules 30 02248 g027
Figure 28. (a) Evolution of the spectra at 514 nm with the number of layers. (b) Evolution of the Raman spectra at 633 nm with the number of layers. (c) The four components of the 2D band in 2-layer graphene at 514 and 633 nm [268].
Figure 28. (a) Evolution of the spectra at 514 nm with the number of layers. (b) Evolution of the Raman spectra at 633 nm with the number of layers. (c) The four components of the 2D band in 2-layer graphene at 514 and 633 nm [268].
Molecules 30 02248 g028
Figure 29. Raman spectra of the prepared catalysts were obtained using an excitation laser wavelength of 532 nm. The calculated ID/IG values are also presented [273].
Figure 29. Raman spectra of the prepared catalysts were obtained using an excitation laser wavelength of 532 nm. The calculated ID/IG values are also presented [273].
Molecules 30 02248 g029
Figure 30. Wide-scan X-ray photoelectron spectra of (a) G-ST, (b) G-HO, and (c) G-HU, and high-resolution X-ray photoelectron spectra of the C1s peak of (d) G-ST, (e) G-HO, and (f) G-HU. Adapted from [274] with permission from the Royal Society of Chemistry. https://pubs.rsc.org/en/content/articlelanding/2012/nr/c2nr30490b (accessed on 1 March 2025).
Figure 30. Wide-scan X-ray photoelectron spectra of (a) G-ST, (b) G-HO, and (c) G-HU, and high-resolution X-ray photoelectron spectra of the C1s peak of (d) G-ST, (e) G-HO, and (f) G-HU. Adapted from [274] with permission from the Royal Society of Chemistry. https://pubs.rsc.org/en/content/articlelanding/2012/nr/c2nr30490b (accessed on 1 March 2025).
Molecules 30 02248 g030
Figure 31. XPS profiles for samples (a) NG−1, (b) NG-2, (c) NG-3, and (d) NG-4. Reproduced with permission from IOP Publishing. From [275]. © 2017 The Electrochemical Society.
Figure 31. XPS profiles for samples (a) NG−1, (b) NG-2, (c) NG-3, and (d) NG-4. Reproduced with permission from IOP Publishing. From [275]. © 2017 The Electrochemical Society.
Molecules 30 02248 g031
Figure 32. XRD patterns of GO, graphite (G), and reduced GOs obtained using various molar concentrations of NaBH4. A table with the interlayer distance is reported beside the graph [87] © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 32. XRD patterns of GO, graphite (G), and reduced GOs obtained using various molar concentrations of NaBH4. A table with the interlayer distance is reported beside the graph [87] © 2009 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Molecules 30 02248 g032
Figure 33. (A) XRD spectra of graphite, GO, and RGOHI–AcOH [277]. Adapted with permission from Springer Nature Communications © 2010 Macmillan Publishers Limited. (B) XRD spectra of graphite, GO, and RGONa-NH3 powder [279]. Adapted with permission from Springer Nature Communications © 2013 Macmillan Publishers Limited.
Figure 33. (A) XRD spectra of graphite, GO, and RGOHI–AcOH [277]. Adapted with permission from Springer Nature Communications © 2010 Macmillan Publishers Limited. (B) XRD spectra of graphite, GO, and RGONa-NH3 powder [279]. Adapted with permission from Springer Nature Communications © 2013 Macmillan Publishers Limited.
Molecules 30 02248 g033
Figure 34. (A) Optical absorption spectrum of GO and rGO with different reduction times. Zoomed image of the optical absorption spectrum (inset) [280]. (B) The UV-Vis absorption spectra of monolayer graphene and bilayer graphene [280]; peaks are labelled with a wavelength of maximum absorption and the value of maximum absorption. The UV transmittance (T in %) is measured at 550 nm. Adapted with permission from Springer Nature Communications © 2010 Macmillan Publishers Limited.
Figure 34. (A) Optical absorption spectrum of GO and rGO with different reduction times. Zoomed image of the optical absorption spectrum (inset) [280]. (B) The UV-Vis absorption spectra of monolayer graphene and bilayer graphene [280]; peaks are labelled with a wavelength of maximum absorption and the value of maximum absorption. The UV transmittance (T in %) is measured at 550 nm. Adapted with permission from Springer Nature Communications © 2010 Macmillan Publishers Limited.
Molecules 30 02248 g034
Figure 35. UV−Vis absorption of NGqDot; in the inset is the photo of the corresponding solution in water under 365 nm UV irradiation. Adapted with permission from [283]. Copyright {2012} American Chemical Society.
Figure 35. UV−Vis absorption of NGqDot; in the inset is the photo of the corresponding solution in water under 365 nm UV irradiation. Adapted with permission from [283]. Copyright {2012} American Chemical Society.
Molecules 30 02248 g035
Figure 36. NEXAFS (total electron yield) of pristine graphene (PG) and N-doped graphene (NG) on copper foil at the N K-edge. (Inset) XPS data for pristine and N-doped graphene, showing a higher-binding energy component (black arrow) for the doped sample [178]. Copyright © 2011, The American Association for the Advancement of Science.
Figure 36. NEXAFS (total electron yield) of pristine graphene (PG) and N-doped graphene (NG) on copper foil at the N K-edge. (Inset) XPS data for pristine and N-doped graphene, showing a higher-binding energy component (black arrow) for the doped sample [178]. Copyright © 2011, The American Association for the Advancement of Science.
Molecules 30 02248 g036
Figure 37. (A) Nitrogen species doped in the graphene layer. (B) N K-edge XANES spectra of NG-300, NG-500, NG-700, and NG-800. (C) C K-edge XANES spectra of NG-300, NG-500, NG-700, and NG-800. Adapted from [292] with permission from the Royal Society of Chemistry. https://pubs.rsc.org/en/content/articlelanding/2015/ta/c5ta01561h (accessed on 1 March 2025).
Figure 37. (A) Nitrogen species doped in the graphene layer. (B) N K-edge XANES spectra of NG-300, NG-500, NG-700, and NG-800. (C) C K-edge XANES spectra of NG-300, NG-500, NG-700, and NG-800. Adapted from [292] with permission from the Royal Society of Chemistry. https://pubs.rsc.org/en/content/articlelanding/2015/ta/c5ta01561h (accessed on 1 March 2025).
Molecules 30 02248 g037
Figure 38. Various configurations of N atoms in a nitrogen-doped graphene layer. (A) graphitic-N, pyridinic-N, and pyrrolic-N are identified with different colors. (B) A detailed description of various N atom configurations from [303]. © 2018 Elsevier Ltd. (2018).
Figure 38. Various configurations of N atoms in a nitrogen-doped graphene layer. (A) graphitic-N, pyridinic-N, and pyrrolic-N are identified with different colors. (B) A detailed description of various N atom configurations from [303]. © 2018 Elsevier Ltd. (2018).
Molecules 30 02248 g038
Scheme 4. Selective modification of the acetyl group on the N atom and ortho-C atom of the pyridinic ring for a N-doped graphene catalyst, inspired by selective substitution for the quinoline derivative in organic synthesis. AcCl = acetylchloride, DCM = dichloromethane. Reprinted with permission from [334]. Copyright {2018} American Chemical Society.
Scheme 4. Selective modification of the acetyl group on the N atom and ortho-C atom of the pyridinic ring for a N-doped graphene catalyst, inspired by selective substitution for the quinoline derivative in organic synthesis. AcCl = acetylchloride, DCM = dichloromethane. Reprinted with permission from [334]. Copyright {2018} American Chemical Society.
Molecules 30 02248 sch004
Figure 39. Schematic representation of NG synthesis showing the selectivity inside and outside of MMT during NG synthesis [452] © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 39. Schematic representation of NG synthesis showing the selectivity inside and outside of MMT during NG synthesis [452] © 2013 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Molecules 30 02248 g039
Figure 40. Scheme of the preparation procedure of GO precursors with different densities of in-plane lattice defects. Reprinted with permission from [499]. Copyright {2019} American Chemical Society.
Figure 40. Scheme of the preparation procedure of GO precursors with different densities of in-plane lattice defects. Reprinted with permission from [499]. Copyright {2019} American Chemical Society.
Molecules 30 02248 g040
Figure 41. Schematic illustration of the fabrication process of electrochemically exfoliated N-doped graphene. Reprinted from [515] with permission from Springer Journal of Solid State Electrochemistry © Springer-Verlag Berlin Heidelberg 2016.
Figure 41. Schematic illustration of the fabrication process of electrochemically exfoliated N-doped graphene. Reprinted from [515] with permission from Springer Journal of Solid State Electrochemistry © Springer-Verlag Berlin Heidelberg 2016.
Molecules 30 02248 g041
Figure 42. Several possible S-doped graphene clusters: (a) sulfur atoms adsorbed on the surface of the graphene cluster; substituting sulfur atoms at (b) zigzag and (c) armchair edges; SO2 substituted at (d) zigzag and (e) armchair edges; and (f) sulfur ring cluster connecting two pieces of graphene. The structures of the graphene are shown only partially to highlight the doping structures. Small white, gray, yellow, and red balls represent hydrogen, carbon, sulfur, and oxygen atoms, respectively. Reprinted with permission from [589]. Copyright {2014} American Chemical Society.
Figure 42. Several possible S-doped graphene clusters: (a) sulfur atoms adsorbed on the surface of the graphene cluster; substituting sulfur atoms at (b) zigzag and (c) armchair edges; SO2 substituted at (d) zigzag and (e) armchair edges; and (f) sulfur ring cluster connecting two pieces of graphene. The structures of the graphene are shown only partially to highlight the doping structures. Small white, gray, yellow, and red balls represent hydrogen, carbon, sulfur, and oxygen atoms, respectively. Reprinted with permission from [589]. Copyright {2014} American Chemical Society.
Molecules 30 02248 g042
Figure 43. The schematic illustration indicating the production of S-doped graphene by magnesiothermic reduction of CO32- ions in the presence of SO42- ions in an Ar flow [602].
Figure 43. The schematic illustration indicating the production of S-doped graphene by magnesiothermic reduction of CO32- ions in the presence of SO42- ions in an Ar flow [602].
Molecules 30 02248 g043
Figure 44. Schematic of the synthesis of rGOSe through a two-step procedure where GO layers are linked by selenium. Reprinted with permission from [629]. Copyright {2019} American Chemical Society.
Figure 44. Schematic of the synthesis of rGOSe through a two-step procedure where GO layers are linked by selenium. Reprinted with permission from [629]. Copyright {2019} American Chemical Society.
Molecules 30 02248 g044
Figure 45. Schematic of the synthesis of B-doped graphene quantum dots [638], © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 45. Schematic of the synthesis of B-doped graphene quantum dots [638], © 2019 WILEY-VCH Verlag GmbH & Co. KGaA, Weinheim.
Molecules 30 02248 g045
Figure 46. Top and side views of the different models of pristine and P-doped graphene: (a) pristine graphene (G), (b) PC3G, (c) OPC3G, (d) PC4G, and (e) OPC4G. The gray, white, red, and pink spheres represent the C, H, O, and P atoms, respectively. Reprinted with permission from [667]. Copyright {2017} American Chemical Society.
Figure 46. Top and side views of the different models of pristine and P-doped graphene: (a) pristine graphene (G), (b) PC3G, (c) OPC3G, (d) PC4G, and (e) OPC4G. The gray, white, red, and pink spheres represent the C, H, O, and P atoms, respectively. Reprinted with permission from [667]. Copyright {2017} American Chemical Society.
Molecules 30 02248 g046
Scheme 5. Illustration of the synthesis route for CnT/N,S/graphene3D. Adapted from [701]. © 2019 Elsevier B.V (2019).
Scheme 5. Illustration of the synthesis route for CnT/N,S/graphene3D. Adapted from [701]. © 2019 Elsevier B.V (2019).
Molecules 30 02248 sch005
Figure 47. Schematic illustration of the fabricated NS-3DrGO products: (a) crosslinking of GO by the addition of melamine and formaldehyde using a hydrothermal method; (b) homogenous mixing of benzyl disulfide with the as-prepared hybrid hydrogel; (c) pyrolysis of the hybrid xerogel to synthesize the N,S co-doped 3D rGO catalyst from [735]. © 2017 Elsevier Ltd. (2017).
Figure 47. Schematic illustration of the fabricated NS-3DrGO products: (a) crosslinking of GO by the addition of melamine and formaldehyde using a hydrothermal method; (b) homogenous mixing of benzyl disulfide with the as-prepared hybrid hydrogel; (c) pyrolysis of the hybrid xerogel to synthesize the N,S co-doped 3D rGO catalyst from [735]. © 2017 Elsevier Ltd. (2017).
Molecules 30 02248 g047
Figure 48. Schematic illustration of the synthesis procedures for 3D porous carbon nanotube-graphene hybrid foam with embedded N,P-coupled active species. Adapted from [660] with permission from the Royal Society of Chemistry. https://pubs.rsc.org/en/content/articlelanding/2018/ta/c7ta08186c (accessed on 1 March 2025).
Figure 48. Schematic illustration of the synthesis procedures for 3D porous carbon nanotube-graphene hybrid foam with embedded N,P-coupled active species. Adapted from [660] with permission from the Royal Society of Chemistry. https://pubs.rsc.org/en/content/articlelanding/2018/ta/c7ta08186c (accessed on 1 March 2025).
Molecules 30 02248 g048
Scheme 6. Proposed mechanism for the synthesis of N, X (B, P, or S) dual-doped graphene. Adapted from [208]. © 2016 Elsevier Ltd. (2016).
Scheme 6. Proposed mechanism for the synthesis of N, X (B, P, or S) dual-doped graphene. Adapted from [208]. © 2016 Elsevier Ltd. (2016).
Molecules 30 02248 sch006
Figure 49. Schematic representation of the synthesis of iodine and nitrogen co-doped graphene [772] © 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Figure 49. Schematic representation of the synthesis of iodine and nitrogen co-doped graphene [772] © 2014 Wiley-VCH Verlag GmbH & Co. KGaA, Weinheim.
Molecules 30 02248 g049
Figure 50. Schematic illustration of the preparation of N,S,P-doped graphene. Adapted from [715]. © 2017 Elsevier B.V. (2017).
Figure 50. Schematic illustration of the preparation of N,S,P-doped graphene. Adapted from [715]. © 2017 Elsevier B.V. (2017).
Molecules 30 02248 g050
Figure 51. Schematic plot of the mechanism of formation of holey graphene. Adapted from [799]. © 2017 Elsevier Ltd. (2017).
Figure 51. Schematic plot of the mechanism of formation of holey graphene. Adapted from [799]. © 2017 Elsevier Ltd. (2017).
Molecules 30 02248 g051
Figure 52. (A) Schematic of the structure of an aluminum–air battery. Structure of a layered spark-reduced graphene oxide (sGO) cathode in the inset. (B) SEM images of electrolyte-filled sGO [821] © 2021 Wiley GmbH.
Figure 52. (A) Schematic of the structure of an aluminum–air battery. Structure of a layered spark-reduced graphene oxide (sGO) cathode in the inset. (B) SEM images of electrolyte-filled sGO [821] © 2021 Wiley GmbH.
Molecules 30 02248 g052
Scheme 7. Proposed synthetic protocol for the monoatomic-thick g-C3N4 dots@graphene (MTCG). Adapted from [856] with permission from the Royal Society of Chemistry. https://pubs.rsc.org/en/content/articlelanding/2015/nr/c4nr05343e (accessed on 1 March 2025).
Scheme 7. Proposed synthetic protocol for the monoatomic-thick g-C3N4 dots@graphene (MTCG). Adapted from [856] with permission from the Royal Society of Chemistry. https://pubs.rsc.org/en/content/articlelanding/2015/nr/c4nr05343e (accessed on 1 March 2025).
Molecules 30 02248 sch007
Scheme 8. Illustration of the synthetic processes for nitrogen-doped carbon rGO@N/C and phosphorus- and nitrogen-doped carbon (rGO@PN/C). Reprinted with permission from [870]. Copyright {2020} American Chemical Society.
Scheme 8. Illustration of the synthetic processes for nitrogen-doped carbon rGO@N/C and phosphorus- and nitrogen-doped carbon (rGO@PN/C). Reprinted with permission from [870]. Copyright {2020} American Chemical Society.
Molecules 30 02248 sch008
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

Crociani, L. Metal-Free Graphene-Based Derivatives as Oxygen Reduction Reaction Electrocatalysts in Energy Conversion and Storage Systems: An Overview. Molecules 2025, 30, 2248. https://doi.org/10.3390/molecules30102248

AMA Style

Crociani L. Metal-Free Graphene-Based Derivatives as Oxygen Reduction Reaction Electrocatalysts in Energy Conversion and Storage Systems: An Overview. Molecules. 2025; 30(10):2248. https://doi.org/10.3390/molecules30102248

Chicago/Turabian Style

Crociani, Laura. 2025. "Metal-Free Graphene-Based Derivatives as Oxygen Reduction Reaction Electrocatalysts in Energy Conversion and Storage Systems: An Overview" Molecules 30, no. 10: 2248. https://doi.org/10.3390/molecules30102248

APA Style

Crociani, L. (2025). Metal-Free Graphene-Based Derivatives as Oxygen Reduction Reaction Electrocatalysts in Energy Conversion and Storage Systems: An Overview. Molecules, 30(10), 2248. https://doi.org/10.3390/molecules30102248

Article Metrics

Back to TopTop