7
Functionalized Graphene‐Based, Metal‐Free Electrocatalysts for Oxygen Reduction Reaction in Fuel Cells

Nanjundan Ashok Kumar1, and Jong‐Beom Baek2

1The University of Queensland, School of Chemical Engineering, St Lucia, Brisbane, 4072, Australia

2Centre for Dimension‐Controllable Organic Frameworks, Ulsan National Institute of Science and Technology (UNIST), School of Energy and Chemical Engineering, 50 UNIST, Ulsan, 44919, South Korea

7.1 Introduction

The future global economy is expected to consume more energy than ever before leading to the energy crisis and detrimental environmental issues. Hence, there is rising interest in energy conversion and storage from sources that do not rely on fossil fuels due to their nonrenewable nature [1]. Potential technologies, such as batteries, supercapacitors, and fuel cells, have been acknowledged to be feasible and efficient, particularly, for stationary and nonstationary energy applications. Fuel cells that convert fuels (methanol or hydrogen) into electricity with zero emission are the most environment‐friendly devices for the future [2]. The working principle of polymer electrolyte membrane (PEM) fuel cell is depicted in Figure 7.1. It consists of two electrodes: a cathode, where oxygen reduction reaction (ORR) occurs, and an anode for fuel (methanol or hydrogen) oxidation, separated by a polymer membrane. ORR is more than six magnitudes slower than the fuel oxidation reaction, which limits the energy conversion efficiency of a fuel cell. In simple terms, ORR represents reaction involving reduction of oxygen molecules into various oxide components depending on the reaction kinetics or electrolyte composition (either aqueous alkaline or acidic electrolyte).

Schematic diagram depicting fuel cell components with working principle of polymer electrolyte membrane (PEM) fuel cell with labels for parts and a left arrow from the left - V and a right arrow from the right + V.

Figure 7.1 Fuel cell components with working principle of polymer electrolyte membrane (PEM) fuel cell.

Source: Reproduced with permission [3]. Copyright 2012, Macmillan Publishers Ltd.

ORR is a multistep reaction that proceeds either through an efficient four‐electron process or via a two‐step, two‐electron pathway. In a four‐electron pathway, either four electrons are inserted into an oxygen molecule (in acid media) or four hydroxyl groups (alkaline media) react at once to form a water molecule [4]. If an oxygen molecule receives only two electrons at first with a delayed supply of two electrons later, then a two‐electron pathway is predominant, which leads to the formation of reactive peroxide species: H2O2 (in an acidic medium) and HO2 (in an alkaline medium). The reduction mechanism becomes more complicated when multiple interactive processes (oxygen adsorption, charge transfer, and product dissociation) come into the picture. Also, there are dissociative and associative mechanisms for both four‐electron and two‐electron reduction pathways. In general, the high‐selectivity four‐electron reduction pathway is preferable, as it is devoid of surface‐damaging peroxidase groups and maximizes the generation of electricity [5]. The typical ORR in acid and alkaline media is given below [6].

Acidic media:

vol-2-c07-math-0001

Alkaline media:

vol-2-c07-math-0002

Platinum is considered the best ORR electrocatalyst, which has undergone extensive research since the first fuel cell invention in 1839 [7]. However, the rarity and cost of platinum (Pt) have driven researchers to consider alternatives as they pose serious concerns for commercializing fuel cells. Thus, developing robust, low‐cost, and highly active metal‐free ORR electrocatalysts are of utmost importance, if cheap fuel cell technology is to be realized.

7.2 Carbon Materials as ORR Electrocatalysts

At present, low‐temperature PEM fuel cells use Pt nanoparticles on a carbon (C) support (Pt/C) for ORR. Carbon black, which is produced from heavy petroleum product, was investigated as a fuel cell catalytic support, its ORR catalytic activity as compared to the commercially available metal‐based catalyst was meager [8, 9]. In addition, aggregation and detachment of platinum due to carbon corrosion could lead to decrease in the performance in the long run [10]. Even though Pt alloys have been considered an alternative catalyst to lower platinum content, large‐scale production would still require a significant amount of this Pt metal. Also, carbon‐supported metal oxides and other metal oxides have also been considered for ORR catalysts [5]. One major limitation of such electrocatalysts is their dissolution in acidic media, restricting their utility to alkaline media alone. However, novel carbon nanomaterials like carbon nanotubes (CNTs), fullerene, graphene, etc., with tunable physicochemical properties have shown promise as a metal‐free catalyst for ORR [9, 1115]. So far, such carbons are the best‐suited ORR catalysts developed to replace Pt, not only because of their wide availability but also for their high electrical conductivity, mechanical strength, thermal properties, large surface area, and low cost. Today, they are sought after alternatives to precious metal‐based electrocatalysts.

As a typical first example, Dai et al. prepared vertically aligned nitrogen‐doped CNTs as a metal‐free catalyst for ORR, which exhibited more activity and operational stability than commercial Pt/C electrodes in alkaline fuel cells (Figure 7.2a–d) [16]. In such systems, crossover effects from methanol were nil. Also, theoretical and experimental studies have demonstrated that nitrogen doping improves the catalytic effect. The presence of nitrogen could create a net positive charge on the nearby carbon atom in the carbon plane, to alter the mode of oxygen absorption and thereby attract electrons from anode for better ORR [16]. The vertically aligned nitrogen‐doped CNT electrode was established to be unresponsive to CO poisoning even after adding a significant amount (10%) of CO in oxygen (grey line in Figure 7.2b). The commercial Pt/C on glassy carbon (GC) electrode was, however, poisoned under the same conditions (black line in Figure 7.2b). This result shows the high tolerance of the vertically aligned nitrogen‐doped CNTs/GC electrode for ORR. The metal‐free, heteroatom‐doped, carbon‐based catalyst is one of the most effective materials for overcoming the problems that Pt catalysts are facing.

Image described caption and surrounding text.

Figure 7.2 (a) Scanning electron microscopy image of the vertically aligned nitrogen‐doped carbon nanotubes on a quartz substrate (inset: digital photograph of the nanotube array after being transferred onto a polystyrene‐nonaligned CNT conductive nanocomposite film). (b) CO poison effect on the it chronoamperometric response for the Pt‐C/GC and vertically aligned nitrogen‐doped carbon nanotubes/GC electrodes. The arrow indicates the addition of 55 ml min−1 CO gas into the 550 ml min−1 O2 flow; the mixture gas of ∼9% CO (v/v) was then introduced into the electrochemical cell. i0, initial current. (c) Calculated charge density distribution calculated for the nitrogen‐doped carbon nanotubes. (d) Schematic representations of possible adsorption modes of an oxygen molecule at the carbon nanotubes (top) and nitrogen‐doped carbon nanotubes (bottom).

Source: Gong et al. 2009 [16]. Copyright 2009. Reprinted with permission from American Academy of Advancement of Science.

Confirming the significant ORR activity of such carbon materials, researchers have ventured into developing new strategies to engineer efficient metal‐free ORR catalysts, which include graphene [17], heteroatom‐doped graphene [5], heteroatom‐doped graphite [18], and other carbon nanomaterials [1924] with superior performance, excellent durability, and methanol tolerance in fuel cells. Especially, graphene‐based, metal‐free catalysts have been demonstrated to show exceptional stability and high efficiency in fuel cells [11]. With well‐established and scalable synthesis protocols that are available today, utilizing graphene‐based electrocatalysts can help in cost‐cutting and contribute to maintaining economic viability for fuel cells and other energy devices, leading to substantial impact on the energy community. Forthcoming sections are aimed at providing an outlook of the significant developments that took place in utilizing metal‐free, graphene‐based electrocatalysts for ORR in fuel cells. In particular, the following sections will shed light on the surface engineering of graphene, mainly discussing the restoration of π conjugations and the importance of tailoring graphene and heteroatom doping, as a new class of promising metal‐free electrocatalysts for the ORR. The chapter takes the readers through the future perspectives and design considerations for a high‐performance, scalable ORR catalyst using metal‐free graphene.

7.3 Structurally Engineered Graphene as Metal‐Free Catalysts for ORR

Graphene is a vast carpet of the aromatic framework with tremendous opportunities for surface design. It is a one‐atom‐thick sheet with a densely packed honeycomb crystal lattice. Single‐layer graphene isolation was first reported by Geim and Novoselov in 2004 using a scotch tape method, which splits graphite galleries into thinner layers until a single layer was obtained [25]. Although mechanical peeling from graphite produces graphene, its yield is inherently low. In general, the preparation of graphene can be achieved either by the top‐down or by the bottom‐up approach. The bottom‐up approach uses chemical vapor deposition (CVD) of hydrocarbons or another chemical synthesis [26]. A top‐down method for preparation of graphene involves exfoliation of graphite by chemical means to form graphene oxide (GO) [27] followed by its reduction. GO itself possesses excellent activity toward many important transformations as an electrocatalyst. GO offers a wide range of carbonyl and epoxy groups, which can be selectively transformed to other functionalities to serve as an efficient catalyst. The alternation of carboxyl groups into other functionalities requires activation, which can then form covalent linkages with nucleophiles. In general, carboxyl groups are transformed into amide or ester groups by reaction with an amine or hydroxyl containing nucleophiles. Similarly, alteration of GO through epoxy is believed to happen via a ring‐opening reaction [26]. Such alternations restore the π conjugations of such reduced graphene oxide (rGO) with properties similar to those of graphene.

As discussed earlier, there is a predominant interest in the development of metal‐free catalysts for ORR due to their enhanced stability in alkaline media. Given the unique electrical and thermal properties, resource availability, corrosion resistance, and large specific surface area, graphene‐based materials are believed to be best suited as a metal‐free catalyst for ORR applications [2]. Graphene or its derivatives by virtue of itself can exhibit excellent catalytic properties [28]. Functionalizing graphene with other nitrogen‐containing groups like amine can significantly enhance the electrical conductivity. Hou and coworkers [29] used ammonia to functionalize graphene from GO, with the presence of 10.8 at.% nitrogen. The role of nitrogen was investigated for its superior catalytic performance. It was suggested that the graphitic and pyridinic nitrogen compositions were key factors in enhancing the current density. Following a similar strategy, Chen and coworkers [30] synthesized an amine‐functionalized holey graphene using GO and ammonia via hydrothermal method followed by etching. Interestingly, this functionalized graphene showed superior performance than nitrogen‐doped graphene, possibly due to the porous structure and the presence of electron‐donating group (e.g. amine) as well as large number of holes in its sheet plate, thus providing higher electrical conductivity and more active edge nitrogen atoms. Although a small overpotential was observed toward ORR, it was more favorable than the commercial Pt/C catalyst. Also, the functionalized graphene was immune toward methanol crossover and CO poisoning, making it desirable as a metal‐free catalyst. Later, Rezaei's group reported a pyridine‐functionalized rGO. Pyridine was attached to the epoxy group or hydroxyl group at the surface of GO through a thermal reduction strategy. The catalyst showed superior performance for oxygen reduction at lower overpotential with improved current density; with performance comparable with that of Pt catalyst [31]. In another instance, a layer‐by‐layer assembly technique was used to prepare a graphene film [32]. An unprecedented catalytic activity for ORR in alkaline media was observed. This was credited to the synergistic effect of poly(diallyl dimethyl ammonium chloride) and rGO used in the film preparation.

Baek's group reported a large‐scale production of edge‐selective functionalized graphene nanoplatelets (EFGnPs) in the presence of hydrogen, carbon dioxide, sulfur trioxide, or carbon dioxide/sulfur trioxide mixture through a simple ball‐milling technique to obtain hydrogen GnP (HGnP), carboxylic acid GnP (CGnP), sulfonic acid GnP (SGnP), and carboxylic acid/sulfonic acid functionalized GnP (CSGnP) (Figure 7.3) [33]. The functionalization in the mechanochemical ball‐milling process involves the mechanical cleavage of carbon bonds and gases in the ball mill equipment.

Image described caption and surrounding text.

Figure 7.3 Schematic of the mechanochemical reaction in a sealed ball mill crusher between the carbon species (generated in situ) and reactant gases. The cracking of graphite by ball milling in the presence of corresponding gases and subsequent exposure to air moisture resulted in the formation of edge‐functionalized graphene platelets. The red balls represent the reactant gases such as hydrogen, carbon dioxide, sulfur trioxide, and air moisture (oxygen and moisture). Rotating disk electrode voltammograms (below the schematic) in an oxygen‐saturated 0.1 M aq. KOH solution with a scan rate of 0.1 V s−1 at rotating rates of 400, 900, 1200, 1600, 2000, and 2500 rpm: (a) the pristine graphite; (b) HGnP; (c) CGnP; (d) SGnP; (e) CSGnP; and (f) Pt/C.

Source: Jeon et al. 2013 [33]. Copyright 2013. Reprinted with permission from American Chemical Society.

Some unreacted group in the graphite is terminated by exposure to air, leading to the formation of hydroxyl and carboxyl groups at the broken edges of the functionalized graphene nanoplatelets (GnPs) (Figure 7.3a). The reaction kinetics of edge‐functionalized graphene electrodes were studied using a rotating disk electrode (RDE). In comparison with that of the onset potentials of pristine graphite (∼−0.25 V), those with edge functionalization gradually decreased according to their polarity order of HGnP (−0.24 V) < CGnP (−0.22 V) < CSGnP (−0.16 V) < SGnP (−0.16 V) (Figure 7.3b–e), indicating that ORR mechanism was a mixture of a two‐electron ORR process at lower potential and a four‐electron ORR process at a higher potential with respect to functionality. The overall electrocatalytic activity is in the following order: SGnP > CSGnP > CGnP > HGnP > the pristine graphite. Among all the functionalized GnPs, the sulfonic‐acid‐functionalized GnP exhibited superior ORR performance when compared to the commercially available platinum‐based electrocatalyst.

Using similar same ball‐milling technique, the same group reported edge‐functionalized GnPs using halogens (Cl, Br, and I) [34]. A large specific surface area of 471–662 m2 g−1 was obtained, indicating successful exfoliation of XGnPs (X = Cl, Br, or I), with easy processability in various solvents. The samples showed high electrocatalytic activity toward ORR with a superior tolerance to methanol crossover/CO poisoning effects and high shelf life.

Edge/surface functionalization of GnPs is sought after as a promising method to tailor carbon materials. Still, there is plenty of room to improve producibility and performance depending on the tailoring technique. In addition, much care should be taken to differentiate functionalized and doped GO, as sometimes both can coexist. Therefore, region‐selective functionalization and/or doping (heteroatoms) seems to be a more efficient strategy that can provide clear outlines as to how a heteroatom can affect the ORR catalytic performance in fuel cells.

7.3.1 Heteroatom‐Doped Graphene

Heteroatom‐doped graphene materials are a promising class of metal‐free electrocatalysts for the ORR in fuel cell. In general, a carbon material is said to be doped, when a heteroatom replaces some of the carbon atoms in the graphitic lattice [11]. Heteroatom doping in carbon nanomaterials can be done either during material synthesis or post‐synthesis through heteroatom‐containing precursors. Post‐synthesis doping only alters the surface properties without affecting the bulk properties. In contrast, a homogeneous incorporation of dopants into the carbon sample could be achieved through in situ heteroatom doping. Surface engineering of graphene with heteroatoms, such as nitrogen, sulfur, phosphorous, and boron enables better tuning of the electron donor properties and consequently enhances the catalytic activity (Figure 7.4). Recently, investigations have revealed that the electrochemical performance of such metal‐free, carbon‐based catalysts arise from change in the charge distribution caused by the differences in the electron negativities between carbon and heteroatoms [23].

Image described caption and surrounding text.

Figure 7.4 (a) Predominantly doped heteroatoms; (b) graphitic lattice doped with heteroatoms after post‐treatment; (c) periodic table representing electronegativities of the elements.

Source: Dai 2012 [35]. Copyright 2012. Reprinted with permission from American Chemical Society. (c) Adapted from http://butane.chem.uiuc.edu/cyerkes/chem102ae_fa08/homepage/Chem102AEFa07/Lecture_Notes_102/copy%20of%20Lecture%2012%20.htm.

7.3.1.1 Nitrogen‐Doped Graphene Structures as Metal‐Free Catalysts for ORR

Nitrogen‐doped graphene has received significant interest in the recent years for applications including photocatalysis and ORR. The introduction of nitrogen into the carbon backbone can alter the electrochemical and thermomechanical properties of graphene [36]. Doping heteroatoms like nitrogen into the carbon architecture could be achieved during graphene fabrication, either by chemical vapor deposition method or using hydrothermal process. Also, posttreatment of graphene using reducing agents like hydrazine and ammonia can enrich the graphene with nitrogen [37, 38]. The nitrogen configuration in graphene falls under the following categories: graphitic, quaternary, pyrrolic, and pyridinic, which is mainly evaluated by X‐ray photoelectron spectroscopy (XPS) (Figure 7.5) [5]. Graphitic nitrogen represents the nitrogen atom within the backbone of a hexagonal ring. Pyrrolic and pyridinic nitrogen form sp3‐ and sp2‐hybridized bonds, respectively, by donating electrons to the π conjugation.

Image described caption and surrounding text.

Figure 7.5 Schematic representation of nitrogen configuration in graphite.

Source: Zhang et al. 2015 [39]. Copyright 2015. Reprinted with permission from American Academy of Advancement of Science.

Dai and coworkers first reported nitrogen‐doped graphene as a useful catalyst for ORR [40]. A nitrogen‐containing gas mixture was used to dope graphene film that was grown on a nickel support. The nitrogen to carbon atomic ratio was maintained at 4%. The resulting flexible and transparent film (Figure 7.6A) showed exceptional performance as an electrocatalyst for ORR. In comparison to the natural graphite (exhibits a two‐electron process), the nitrogen‐doped graphene showed a single‐step four‐electron process. The steady‐state catalytic current density of the nitrogen‐doped graphene showed about three times higher current density than that of the commercial Pt/C electrode over a large potential range (Figure 7.6B). The number of electrons transferred per oxygen molecule at the graphene electrode was calculated to be 3.6–4 (potentials range −0.4 to −0.8 V). These results indicate that the nitrogen‐doped graphene electrode is as promising as a metal‐free catalyst for the ORR in an alkaline solution. In addition, the nitrogen‐doped graphene remained unchanged with the addition of methanol, was insensitive to CO, and exhibited stable performance over cycling even after 200 000 cycles (potential range: −1.0 and 0 V) in air‐saturated potassium hydroxide electrolyte. This material showed high stability during operation, superior selectivity, crossover effect tolerance, and posed the potential for other electrocatalytic applications.

Image described caption and surrounding text.

Figure 7.6 (A): (a) Digital image of a N‐doped graphene film floating on water; (b,c) atomic force microscope (AFM) images of the N‐doped graphene film and the corresponding height profile from AFM (c1−c3 in c). (B): (a) Rotating ring disk electrode (RRDE) voltammograms for ORR in air‐saturated 0.1 M KOH at the graphene electrode, commercial Pt/C electrode, and nitrogen‐doped graphene (rotating rate of electrode:1000 rpm; catalyst's mass: 7.5 μg; scan rate: 0.01 V s−1). (b) Current density versus time (jt) chronoamperometric responses at −0.4 V in air‐saturated 0.1 M KOH. (c) Current versus time chronoamperometric response to CO. The arrow indicates the addition of 10% (v/v) CO into air‐saturated 0.1 M KOH at −0.4 V; j0 defines the initial current. (d) Cyclic voltammograms of nitrogen‐doped graphene electrode in air‐saturated 0.1 M KOH before and after a continuous potentiodynamic sweep for 200 000 cycles at a scan rate of 0.1 V s−1.

Source: Qu et al. 2010 [40]. Copyright 2010. Reprinted with permission from American Chemical Society.

In another work by Chen et al. [41], an interesting strategy was followed to obtain nitrogen‐doped graphene nanoribbons. Here, a CNT was longitudinally unzipped to form graphene nanoribbons, which were doped with nitrogen using a hydrothermal treatment method followed by a freeze‐drying process. The obtained 3D macroscopic aerogel was fire‐resistant with ultralow density, large surface area, and superior conductivity. The nitrogen‐doped nanoribbons exhibited excellent ORR activity, which was much better than the commercial Pt/C electrodes. Also, the material showed superior performance in both alkaline and acidic media. Zhou et al. [42] used a hard templating (silica template) method to produce a foam of nitrogen‐doped graphene, which also performed well as superior ORR catalyst in both acidic and alkaline media. Sun and coworkers [43] pyrolyzed graphene in the presence of ammonia to dope 2.8% of nitrogen. The as‐prepared nitrogen‐doped graphene showed a 4e process with a very high electrocatalytic activity in alkaline media. The performance of the material was comparable or even better than the commercially available Pt/C electrodes. The XPS studies indicated the presence of graphitic nitrogen, which was attributed to the superior ORR performance. However, in another study, pyridinic and pyrrolic nitrogen groups were associated with this performance, which was contradictory to the previous results [44]. This study used a variety of nitrogen precursors to dope graphene with different configurations of nitrogen (Figure 7.7a).

Image described caption and surrounding text.

Figure 7.7 (a) Schematic for the preparation of nitrogen‐doped graphene with different configurations. (b) Linear sweep voltammograms for nitrogen‐doped reduced graphene oxide 1000 °C, bare GC electrode, polyaniline/rGO, polypyrrole/rGO, boron‐ and nitrogen‐doped‐rGO, bare Pt electrode, and 20% Pt/C electrodes in an oxygen‐saturated 0.1 M KOH (scan rate of 10 mV s−1, rotation rate of 2500 rpm). (c) Linear sweep voltammograms for nitrogen‐doped reduced graphene oxide 1000 °C with a rotation rate from 500 to 3000 rpm.

Source: Lai et al. 2012 [44]. Copyright 2012. Reprinted with permission from Royal Society of Chemistry.

The RDE voltammograms were used to investigate the electrocatalytic properties of nitrogen‐doped rGO, polyaniline rGO, and polypyrrole rGO composites and were compared to bare platinum, GC, and commercial Pt/C catalysts. The linear voltammetric scans of different catalysts were taken in an oxygen‐saturated 0.1 M aq. KOH electrolyte with a rotation rate of 2500 rpm. With regard to the limiting current density, nitrogen‐doped rGO had the highest value followed by others. The polarization curves of different potentials and rotation speeds are shown in the (Figure 7.7b,c). Furthermore, it was found that the total atomic percentage of nitrogen had no effect on the electrocatalytic activity for ORR in alkaline media. However, each nitrogen group improves a portion of activity in ORR. For instance, the increased content of pyridinic nitrogen directed the 2e pathway to a 4e one. It is still a matter of consideration as to how the configuration of nitrogen in graphene addresses the ORR activity.

Xia and coworkers used melamine as a nitrogen precursor to dope graphene, which showed superior performance in ORR in alkaline media, which was accredited to the pyridinic nitrogen [45]. Woo and coworkers [46] confirmed that pyridinic group also shows higher ORR catalytic activity rGO even in acidic media. To further decipher the arguments on the nitrogen configuration, simulations were carried out by calculating the energetics of nitrogen‐doped graphene by changing the nitrogen content [47, 48]. The presence of graphitic nitrogen was found to significantly lower the energy costs of oxygen reduction [49]. A graphene with pyridinic nitrogen doping in acidic media for ORR activity was investigated by Xia and coworkers using B3LYP hybrid density functional theory [50, 51]. It was found that doping with nitrogen delivered high catalytic activity due to a high atomic charge and spin density. Also, the catalytic process was spontaneous and corroborated with the experimental results, leading to a reversible potential of 1.5 V (versus standard hydrogen electrode).

7.3.1.2 Other Heteroatom‐Doped Graphene for ORR

Apart from nitrogen, sulfur, phosphorous, and boron are commonly sought heteroatom dopants to dope graphene/graphitic lattice. Doping graphene with sulfur has been widely reported to increase the ORR activity [6, 11, 36]. According to calculations, graphene could be either metallic or semiconducting in nature depending on the extent of doping [52]. Using DFT calculations, it was found that sulfur‐doped graphene could show ORR catalytic properties comparable to those of platinum [53]. Experimentally, sulfur can be doped to a graphene lattice using GO with sulfur‐containing gases and also through chemical means. Recently, many chemical modification strategies to dope sulfur have been reported as a potential metal‐free alternative electrocatalyst [5458]. Among the many possible strategies to dope sulfur, Baek et al. [59], produced edge‐selectively sulfurized GnPs using a simple ball‐milling technique (Figure 7.8). The ORR catalytic activity of the product was found to be efficient. In addition, the authors also found that sulfur‐doped graphene has a much slower rate of attenuation and no fuel crossover. Also, it was found that the ORR activity can in fact surpass the activity of commercially available platinum when the sample was oxidized. Sulfur‐oxide‐doped edges (OSO bonds) were found to promote ORR activity.

Image described caption and surrounding text.

Figure 7.8 (A): (a) Schematic of the ball milling process to produce edge‐sulfurized graphene. (b) SEM images of pristine the graphite. (c) Sulfur‐doped graphene. (d) Energy‐dispersive X‐ray spectra showing the presence of sulfur. (B): (a) Current–time (It) chronoamperometric responses of the pristine graphite, sulfur‐doped graphite, and commercial electrodes at −0.30 V in an oxygen‐saturated 0.1 M aq. KOH solution at a rotation rate of 1200 rpm and (b) current–time chronoamperometric responses for the sample electrodes at −0.40 V in the same solution upon addition of 3.0 M methanol at 300 s.

Source: Jeon et al. 2013 [59]. Copyright 2013. Reprinted with permission from John Wiley & Sons.

Very recently, tremendous progress was made in the field. For example, Huang and coworkers [60] reported a three‐dimensional (3D) porous graphene‐based nanosphere frameworks an efficient metal‐free electrocatalyst for ORR. The catalyst showed a four‐electron transfer reaction in alkaline media with better methanol tolerance and shelf life. Such a performance was attributed to the positive effect of sulfur doping and the presence of pores, thereby aiding diffusion of ions and electron transfer more feasible. In another instance, sulfur‐doped graphene was synthesized by thermally treating an exfoliated graphene with CS2 gas [54]. Depicting a higher current density than commercially available Pt/C catalyst (4.7 versus 6.99 mA cm−2 at −1.0 V), the sulfur‐doped graphene was very durable. To further understand the contribution of sulfur toward ORR, theoretical studies using density functional theory by various groups [53, 57]. Although it was found that sulfur could replace carbon atoms as sulfur or sulfur oxide or form a cluster of sulfur [53], it was also predicted that sulfur‐doped monovacancy graphene was stable and the sulfur dopant itself was probably the active center [57]. In most cases, calculations corroborated experimental results, suggesting that sulfur‐doped graphene can show superior performance as a metal‐free ORR catalyst.

Being adjacent to carbon in the periodic table, boron and nitrogen show positive catalytic effects when doped to graphene. Considerable efforts have been made in the recent past to dope boron to graphene as boron being deficient with one electron than that of carbon. Boron generally acts as a p‐type dopant forming sp2 hybridization within the lattice. In 1974, pyrolytic graphite enriched with boron showed an increased space charge capacitance [61]. The methodology of doping boron significantly affects the electrochemical performance. For instance, boron‐doped in‐plane substituting carbon in the graphene is more stable than bonding out of plane. Such a boron‐rich graphene could be synthesized by simply annealing graphene in the presence of boron oxide. Several research groups have recently attempted to synthesize boron‐doped graphene or even graphene quantum dots [6266]. In general, the electronegativity of boron is lesser compared with that of nitrogen. Nevertheless, they exhibited excellent catalytic activity for ORR in alkaline electrolytes, with better CO tolerance and long‐term stability [36]. In an interesting approach, inspired by a Chinese food named “Fried Ice,” a 3D graphene framework with hierarchical pores were synthesized using boric acid as the dopant (Figure 7.9) [66]. By controlling the reaction temperature, the boron content, specific surface area, and the morphology of the samples could be tailored. As a metal‐free catalyst with high ORR performance in alkaline media, this material showed a better onset potential (−0.12 V versus saturated calomel electrode (SCE)) than the commercial Pt/C catalyst. A four‐electron transfer process for oxygen reduction was deduced (Figure 7.9d). Its tolerance to methanol and carbon monoxide was tested against a commercial Pt/C catalyst. Although Pt/C depicted a distinct decrease in stability on the addition of 1 wt% methanol, the boron‐doped graphene, however, was highly stable (Figure 7.9e).

Image described caption and surrounding text.

Figure 7.9 (a) Schematic to the preparation of 3D boron‐doped porous graphene framework; (b) photo of the as‐prepared 3D graphene monoliths; (c) illustration of the interaction between GO and boric acid molecules during the thermal treatment; (d) electron transfer number versus applied potential (V); and (e) ORR stability in the presence of methanol (1 wt%) compared to that of the commercial Pt/C (40 wt%) in an oxygen‐saturated 0.1 M aq. KOH solution at −0.4 V at a rotation of 1000 rpm.

Source: Zuo et al. 2013 [66]. Copyright 2013. Reprinted with permission from Royal Society of Chemistry.

Phosphorous shares the same group in the periodic table and displays chemical properties similar to those of nitrogen despite of its larger atomic size. When doped to graphene, this causes distortion to the structure by transforming sp2 to sp3 carbons, thereby creating a defect‐based active site [67]. The defect caused by phosphorus substitution in graphene has a little energy of formation and can increase the energy bandgap. Although, phosphorous‐doped graphitic layers were found to be efficient ORR catalysts [68], only a handful reports are available as metal‐free ORR catalysts. Recently, a thermal annealing approach with triphenylphosphine as the reductant has been used to fabricate phosphorous‐doped graphene [69]. The phosphorus doping introduces a significant number of oxygen functional groups to the graphene skeleton. They show superior performance as a metal‐free catalyst for ORR as shown in Figure 7.10.

Image described caption and surrounding text.

Figure 7.10 (a) Cyclic voltammograms of phosphorous‐doped graphene in a nitrogen‐ and oxygen‐saturated 0.1 M aq. KOH solution (scan rate: 100 mV s−1); (b) linear sweep voltammetry curves of graphene, phosphorous‐doped graphene, and Pt/C in an oxygen saturated 0.1 M aq. KOH solution (scan rate: 10 mV s−1; rotation: 1600 rpm). Inset shows the onset potentials of the sample; (c) methanol crossover effect upon addition of 3 M methanol after about 10 min in an oxygen‐saturated 0.1 M aq. KOH solution at 0.7 V; and (d) stability evaluation for 20 000 s in an oxygen‐saturated 0.1 M aq. KOH solution at 0.7 V (rotation: 1600 rpm).

Source: Zhang et al. 2013 [69]. Copyright 2013. Reprinted with permission from John Wiley & Sons.

The cyclic voltammograms (Figure 7.10a) reveal a reduction process at ∼0.6 V in an oxygen‐saturated 0.1 M aq. KOH solution, whereas no peaks were found when nitrogen‐saturated electrolyte was used in the same potential range. The linear sweep voltammetry curves recorded in an oxygen‐saturated 0.1 M aq. KOH showed a much higher current for the phosphorous‐doped graphene (1.4 times at −0.2 V) in comparison to that of graphene over the entire potential range. A positive ORR onset potential of ∼0.92 V than that of graphene and commercial Pt/C was observed. When methanol was introduced in the electrolyte, no traceable change was observed in the ORR current as opposed to Pt/C, indicating its negligible crossover effect and methanol tolerance. In addition, a highly stable cycling was obtained for phosphorous‐doped graphene, indicating their superior performance as an ORR electrocatalyst.

Very recently, Xu and coauthors [70] reported the synthesis of a metal‐free, phosphorus‐doped graphene nanosheets with a surface area of 496.67 m2 g−1 (phosphorous content = 1.16 at.%). GO and 1‐butyl‐3‐methlyimidazolium hexafluorophosphate were thermally annealed in an inert atmosphere. It was found that the phosphorous atoms were covalently bonded to the carbon atoms in graphene and the carbon framework was partially oxidized, which acted as active sites for the ORR. The performance was comparable to that of Pt/C in an alkaline medium.

7.3.1.3 Co‐doped Graphene Structures as Metal‐Free Catalysts for ORR

Doping multiple heteroatoms with different electronegativities (various combinations of nitrogen, boron, phosphorous, and sulfur) into graphitic carbons provides synergistic coupling effects, thereby enhancing the catalytic activity in ORR [71]. This effect makes such multiple atoms‐doped graphene more catalytically favorable than single‐variety heteroatom‐doped graphene. Also, first‐principles calculations predicted that the symbiotic effects of boron‐ and nitrogen‐doped graphene were excellent electrocatalysts for ORR [72, 73]. Hence, co‐doping strategies are being increasingly used for tuning the physical and electrochemical properties of graphene for various applications [12].

Recognizing the importance of co‐doping heteroatoms to graphene lattice, nitrogen‐ and sulfur‐dual‐doped graphene was prepared in a one‐pot synthesis procedure without the formation of any side products [74]. Later, the same group used a two‐step procedure to produce boron‐ and nitrogen‐doped graphene [75]. In both cases, it was found that the synergistic effect originating from the dopants was prime for the high‐performance ORR activity. In another report, Dai and coworkers [63] prepared a metal‐free BCN graphene with tunable boron and nitrogen levels as electrocatalyst for ORR in alkaline media. Modest boron and nitrogen levels in the graphene were found to be sufficient to obtain the best ORR activity. Later, Ajayan and coworkers demonstrated that nanoribbons of boron‐ and nitrogen‐substituted graphene can be used as efficient electrocatalysts for ORR [76]. Using GO as building blocks and boric acid and ammonia as boron‐ and nitrogen‐doping sources, porous 3D architectures with abundant edges and thin walls were prepared (Figure 7.11a). Rotating ring disk electrode (RRDE) voltammetry was performed in an O2‐saturated 0.1 M aq. KOH solution at a scanning rate of 10 mV s−1. It was found that the onset potential, half‐wave potential, saturated current density, and electron transfer number were strongly dependent on the doping concentrations in the nanoribbons and optimal doping was sufficient. The reported metal‐free catalysts showed superior performance compared to commercial Pt/C catalyst. In particular, the half‐wave potential (−0.03 V versus Ag/AgCl) was higher than other reported metal‐free catalysts in alkaline media and also higher than the commercial Pt/C catalyst (Figure 7.11b).

Image described caption and surrounding text.

Figure 7.11 (a) TEM image of the nanoribbons of boron‐ and nitrogen‐substituted graphene aerogels showing its 3D porous structure (inset shows the photograph of the aerogel) and (b) disk current densities of the RRDE versus potential derived from nanoribbons of boron‐ and nitrogen‐substituted graphene (BNC‐2), nitrogen‐doped graphene nanoribbon aerogels, and commercial Pt/C catalyst.

Source: Gong et al. 2015 [76]. Copyright 2015. Reprinted with permission from American Chemical Society.

Apart from doping boron and nitrogen, researchers have also investigated co‐doping sulfur or phosphorus with nitrogen [22, 7780]. Using a covalent functionalization strategy, a nitrogen–sulfur‐co‐doped graphene was synthesized using GO [81]. In addition to possessing superior electrocatalytic performance, the material also showed potential as anodes for lithium‐ion batteries. In another report, Cui and coworkers [82] produced a porous network of nitrogen–sulfur‐co‐doped graphene, maintaining a significant amount of catalytic active site. The pyrolysis temperature dictated the specific surface area and pore, with the optimum temperature at 900 °C. The doped graphene obtained at this temperature showed a four‐electron transfer pathway.

Another interesting variety of doping is triple doping where three different heteroatoms are doped into graphene [83, 84]. Thiele and coworkers [85] reported a fluorine‐, nitrogen‐, and sulfur‐doped graphene inexpensively synthesized by a simple route via pyrolysis of sulfur‐doped rGO at 600 °C in the presence of nafion and dimethyl formamide. The co‐doping of GO with fluorine and nitrogen created active sites due to the polarization of charge at CF and CN bonds. The presence of sulfur also provides additional catalytic active sites because of the unpaired electrons present at the defect sites of graphene. The synergistic combination of the three heteroatoms was credited to the enhanced performance in ORR. In an acidic and alkaline media, power densities of 14 and 46 mW cm−2 were obtained, respectively, in a PEM fuel cell, with the as‐prepared material as the cathode catalyst layer. In another report, Wang and coworkers developed a nitrogen‐, sulfur‐, and phosphorous‐tri‐doped graphene by annealing GO with acephate (O,S‐dimethyl acetylphosphoramidothioate) [86]. The tri‐doped graphene exhibited superior ORR catalytic activity along with excellent selectivity and durability as compared to commercial Pt/C fuel cells.

7.3.1.4 Graphene‐Based Composites as Metal‐Free Catalysts for ORR

Graphene‐based composites have been well established to provide superior performance in a variety of applications ranging from medicine to electronics. The combinations of two materials can provide synergistic effects to overcome the deficit property of the other. One significant problem associated with graphene synthesis is its aggregation, which lowers the effective surface area, thereby limiting the active sites and hampering catalytic ORR activity. A general approach followed is to sandwich other carbon materials such as carbon nanoparticles, CNTs, etc., between graphitic layers, so as to obtain superior surface area [87]. In this case, the carbon material acts as a spacer to limit the agglomeration and provide more active sites for ORR.

Yu and coworkers [88] reported the synthesis of nitrogen‐doped graphene and CNT composite via a hydrothermal approach using GO, CNTs, and ammonia as precursors. The composite delivered superior ORR catalytic activity via a four‐electron pathway. In another instance, Liu and coworkers [89] fabricated a nitrogen‐ and sulfur‐co‐doped mesoporous carbon/graphene composite using amino acid precursors. Excellent electrocatalytic activity, selectivity, and stability were obtained for ORR in alkaline media. Recently, Dai and coworkers [90] used a ball‐milling approach to prepare a composite of graphitic carbon nitride with nitrogen‐doped graphene. The composite showed excellent ORR activity with an onset potential of −0.002 V (versus SCE) and half‐wave potential at about −0.22 V (versus SCE). The performance was attributed to the combination of nitrogen‐rich carbon nitride and nitrogen‐doped graphene. To scale up the carbon material production for ORR, sustainable biomass precursors have been used. Pan and coworkers [87] recently reported a porous graphene and carbon composite using glucose as a precursor. A specific surface area of 1510.83 m2 g−1 was obtained. The presence of pyridinic and graphitic nitrogen helped in boosting the performance as an efficient metal‐free electrocatalyst, with an onset potential of 0.91 V (versus reversible hydrogen electrode (RHE)). A nearly four‐electron pathway for ORR in alkaline solution was observed. Although carbon/graphene composites are being reported, there are still some hurdles to overcome, such as weak interaction between the carbon material and graphene and feasibility of large‐scale productions owing to the cost of the other CNTs, especially, pure single‐walled CNTs. In addition to cost, CNTs can always have metal impurities, thus making the metal‐free catalyst topic less valid. Overall, almost all the reported studies have demonstrated excellent ORR activity with respect to onset potential, which might be directly and/or indirectly attributed to metallic residues, when compared to that of commercially available Pt/C such as Vulcan.

7.4 Conclusions and Perspectives

A paradigm shift from the use of commercial Pt/C catalyst to efficient and low‐cost, carbon‐based electrocatalyst for ORR seems indispensable. Therefore, the interest in developing metal‐free catalysts for ORR has multitudinously increased among the research community. Over the past decade, ample number of research groups have demonstrated the use of graphene‐based, metal‐free ORR catalysts in fuel cells. This chapter highlighted and summarized recently reported important works and breakthroughs carried out in this emerging energy arena. It is clear that functionalization of graphene and/or doping graphene sheets are considered to be effective strategies to tailor the properties of materials to obtain the best catalytic activity. Up to day, doping nitrogen atom to graphitic frameworks has been extensively studied. Despite the several advances made and the cost of the catalyst dramatically reduced compared to that of Pt/C, there are still unanswered questions and not all the reported works are on par with that of the commercially available Pt/C catalyst in terms of onset potential and reaction current. In addition to the difficulty in controlling the heteroatom configuration in graphitic structure, there is an ambiguity in deciphering the origin of ORR activity. This critically hampers the understanding of the clear catalytic mechanism. Therefore, steps to control the heteroatom‐doping level and configuration are paramount.

Although steps to dope multiple atoms to graphitic lattice are in its nascent stage, some progress has been achieved, the active centers and mechanism of catalytic activities still remain vague. Theoretical modeling and experimental validation would be beneficial to realize the full potential of these metal‐free catalysts. There has been considerable debate in research community about trace metal impurities that could contribute toward the ORR activity (more significantly to onset potential). Researchers working on making composites of CNTs and graphene should take utmost care to remove impurities in the nanotubes. Similarly, due to the harsh oxidation treatments, impurities can also originate in GO materials and thus they can directly affect the catalytic activity. An important aspect to note is that most of the metal‐free, graphene‐based catalysts are found to be less effective in acidic media. Therefore, more fundamental research toward developing such metal‐free catalyst would revolutionize the future energy technology.

Acknowledgements

This research was supported by the Creative Research Initiative (CRI, 2014R1A3A2069102), BK21 Plus (10Z20130011057), Science Research Center (SRC, 2016R1A5A1009405), and Climate Change (2016M1A2940910) programs through the National Research Foundation (NRF) of Korea.

References

  1. 1. Daems, N., Sheng, X., Vankelecom, I.F.J., and Pescarmona, P.P. (2014). Metal‐free doped carbon materials as electrocatalysts for the oxygen reduction reaction. J. Mater. Chem. A 2 (12): 4085–4110.
  2. 2. Zhou, X.J., Qiao, J.L., Yang, L., and Zhang, J.J. (2014). 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. 4 (8): 1301523.
  3. 3. Debe, M.K. (2012). Electrocatalyst approaches and challenges for automotive fuel cells. Nature 486 (7401): 43–51.
  4. 4. Zhang, J. (2008). PEM Fuel Cell Electrocatalysts and Catalyst Layers: Fundamentals and Applications. Springer Science + Business Media.
  5. 5. Wu, K.H., Wang, D.W., Su, D.S., and Gentle, I.R. (2015). A discussion on the activity origin in metal‐free nitrogen‐doped carbons for oxygen reduction reaction and their mechanisms. ChemSusChem 8 (17): 2772–2788.
  6. 6. Zhang, J.T. and Dai, L.M. (2015). Heteroatom‐doped graphitic carbon catalysts for efficient electrocatalysis of oxygen reduction reaction. ACS Catal. 5 (12): 7244–7253.
  7. 7. Hoogers, G. (2002). Fuel Cell Technology Handbook. CRC Press.
  8. 8. Martin, E., Tartakovsky, B., and Savadogo, O. (2011). Cathode materials evaluation in microbial fuel cells: a comparison of carbon, Mn2O3, Fe2O3 and platinum materials. Electrochim. Acta 58: 58–66.
  9. 9. Sawant, S.Y., Han, T.H., and Cho, M.H. (2016). Metal‐free carbon‐based materials: promising electrocatalysts for oxygen reduction reaction in microbial fuel cells. Int. J. Mol. Sci. 18 (1).
  10. 10. Zhu, C., Li, H., Fu, S. et al. (2016). Highly efficient nonprecious metal catalysts towards oxygen reduction reaction based on three‐dimensional porous carbon nanostructures. Chem. Soc. Rev. 45 (3): 517–531.
  11. 11. Dai, L., Xue, Y., Qu, L. et al. (2015). Metal‐free catalysts for oxygen reduction reaction. Chem. Rev. 115 (11): 4823–4892.
  12. 12. Hu, C. and Dai, L. (2016). Carbon‐based metal‐free catalysts for electrocatalysis beyond the ORR. Angew. Chem. Int. Ed. 55 (39): 11736–11758.
  13. 13. Liu, M., Zhang, R., and Chen, W. (2014). Graphene‐supported nanoelectrocatalysts for fuel cells: synthesis, properties, and applications. Chem. Rev. 114 (10): 5117–5160.
  14. 14. Liu, X. and Dai, L.M. (2016). Carbon‐based metal‐free catalysts. Nat. Rev. Mater. 1 (11): 16064.
  15. 15. Wu, Z.Y., Iqbal, Z., and Wang, X.Q. (2015). Metal‐free, carbon‐based catalysts for oxygen reduction reactions. Front. Chem. Sci. Eng. 9 (3): 280–294.
  16. 16. Gong, K., Du, F., Xia, Z. et al. (2009). Nitrogen‐doped carbon nanotube arrays with high electrocatalytic activity for oxygen reduction. Science 323 (5915): 760–764.
  17. 17. Zhu, C. and Dong, S. (2013). Recent progress in graphene‐based nanomaterials as advanced electrocatalysts towards oxygen reduction reaction. Nanoscale 5 (5): 1753–1767.
  18. 18. Liu, Z.W., Peng, F., Wang, H.J. et al. (2011). Phosphorus‐doped graphite layers with high electrocatalytic activity for the O2 reduction in an alkaline medium. Angew. Chem. Int. Ed. 123 (14): 3315–3319.
  19. 19. Silva, R., Voiry, D., Chhowalla, M., and Asefa, T. (2013). Efficient metal‐free electrocatalysts for oxygen reduction: polyaniline‐derived N‐ and O‐doped mesoporous carbons. J. Am. Chem. Soc. 135 (21): 7823–7826.
  20. 20. Wei, W., Liang, H.W., Parvez, K. et al. (2014). Nitrogen‐doped carbon nanosheets with size‐defined mesopores as highly efficient metal‐free catalyst for the oxygen reduction reaction. Angew. Chem. Int. Ed. 53 (6): 1570–1574.
  21. 21. Zheng, R.P., Mo, Z.Y., Liao, S.J. et al. (2014). Heteroatom‐doped carbon nanorods with improved electrocatalytic activity toward oxygen reduction in an acidic medium. Carbon 69 (0): 132–141.
  22. 22. Gao, S.Y., Liu, H.Y., Geng, K.R., and Wei, X.J. (2015). Honeysuckles‐derived porous nitrogen, sulfur, dual‐doped carbon as high‐performance metal‐free oxygen electroreduction catalyst. Nano Energy 12: 785–793.
  23. 23. Hao, L., Zhang, S., Liu, R. et al. (2015). Bottom‐up construction of triazine‐based frameworks as metal‐free electrocatalysts for oxygen reduction reaction. Adv. Mater. 27 (20): 3190–3195.
  24. 24. Hu, C. and Dai, L. (2017). Multifunctional carbon‐based metal‐free electrocatalysts for simultaneous oxygen reduction, oxygen evolution, and hydrogen evolution. Adv. Mater. 29 (9).
  25. 25. Novoselov, K.S., Geim, A.K., Morozov, S.V. et al. (2004). Electric field effect in atomically thin carbon films. Science 306 (5696): 666–669.
  26. 26. Haag, D. and Kung, H.H. (2013). Metal free graphene based catalysts: a review. Top. Catal. 57 (6–9): 762–773.
  27. 27. Hummers, W.S. Jr. and Offeman, R.E. (1958). Preparation of graphitic oxide. J. Am. Chem. Soc. 80 (6): 1339–1339.
  28. 28. Wang, H., Feng, H., and Li, J. (2014). Graphene and graphene‐like layered transition metal dichalcogenides in energy conversion and storage. Small 10 (11): 2165–2181.
  29. 29. Zhang, C.Z., Hao, R., Liao, H.B., and Hou, Y.L. (2013). Synthesis of amino‐functionalized graphene as metal‐free catalyst and exploration of the roles of various nitrogen states in oxygen reduction reaction. Nano Energy 2 (1): 88–97.
  30. 30. Jiang, Z.Q., Jiang, Z.J., Tian, X.N., and Chen, W.H. (2014). Amine‐functionalized holey graphene as a highly active metal‐free catalyst for the oxygen reduction reaction. J. Mater. Chem. A 2 (2): 441–450.
  31. 31. Ensafi, A.A., Jafari‐Asl, M., and Rezaei, B. (2016). Pyridine‐functionalized graphene oxide, an efficient metal free electrocatalyst for oxygen reduction reaction. Electrochim. Acta 194: 95–103.
  32. 32. Huang, D.K., Zhang, B.Y., Zhang, Y.B. et al. (2013). Electrochemically reduced graphene oxide multilayer films as metal‐free electrocatalysts for oxygen reduction. J. Mater. Chem. A 1 (4): 1415–1420.
  33. 33. Jeon, I.Y., Choi, H.J., Jung, S.M. et al. (2013). 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. 135 (4): 1386–1393.
  34. 34. Jeon, I.Y., Choi, H.J., Choi, M. et al. (2013). Facile, scalable synthesis of edge‐halogenated graphene nanoplatelets as efficient metal‐free eletrocatalysts for oxygen reduction reaction. Sci. Rep. 3: 1810.
  35. 35. Dai, L. (2012). Functionalization of graphene for efficient energy conversion and storage. Accounts of Chemical Research 46 (1): 31–42.
  36. 36. Kong, X.K., Chen, C.L., and Chen, Q.W. (2014). Doped graphene for metal‐free catalysis. Chem. Soc. Rev. 43 (8): 2841–2857.
  37. 37. Wang, R., Wang, Y., Xu, C. et al. (2013). Facile one‐step hydrazine‐assisted solvothermal synthesis of nitrogen‐doped reduced graphene oxide: reduction effect and mechanisms. RSC Adv. 3 (4): 1194–1200.
  38. 38. Li, X., Wang, H., Robinson, J.T. et al. (2009). Simultaneous nitrogen doping and reduction of graphene oxide. J. Am. Chem. Soc. 131 (43): 15939–15944.
  39. 39. Zhang, J., Xia, Z., and Dai, L. (2015). Carbon‐based electrocatalysts for advanced energy conversion and storage. Sci. Adv. 1 (7): e1500564.
  40. 40. Qu, L., Liu, Y., Baek, J.B., and Dai, L. (2010). Nitrogen‐doped graphene as efficient metal‐free electrocatalyst for oxygen reduction in fuel cells. ACS Nano 4 (3): 1321–1326.
  41. 41. Chen, L., Du, R., Zhu, J. et al. (2015). Three‐dimensional nitrogen‐doped graphene nanoribbons aerogel as a highly efficient catalyst for the oxygen reduction reaction. Small 11 (12): 1423–1429.
  42. 42. Zhou, X.J., Bai, Z.Y., Wu, M.J. et al. (2015). 3‐dimensional porous N‐doped graphene foam as a non‐precious catalyst for the oxygen reduction reaction. J. Mater. Chem. A 3 (7): 3343–3350.
  43. 43. Geng, D., Chen, Y., Chen, Y. et al. (2011). High oxygen‐reduction activity and durability of nitrogen‐doped graphene. Energy Environ. Sci. 4 (3): 760–764.
  44. 44. Lai, L., Potts, J.R., Zhan, D. et al. (2012). Exploration of the active center structure of nitrogen‐doped graphene‐based catalysts for oxygen reduction reaction. Energy Environ. Sci. 5 (7): 7936–7942.
  45. 45. Sheng, Z.‐H., Shao, L., Chen, J.‐J. et al. (2011). Catalyst‐free synthesis of nitrogen‐doped graphene via thermal annealing graphite oxide with melamine and its excellent electrocatalysis. ACS Nano 5 (6): 4350–4358.
  46. 46. Lee, K.R., Lee, K.U., Lee, J.W. et al. (2010). Electrochemical oxygen reduction on nitrogen doped graphene sheets in acid media. Electrochem. Commun. 12 (8): 1052–1055.
  47. 47. Kurak, K.A. and Anderson, A.B. (2009). Nitrogen‐treated graphite and oxygen electroreduction on pyridinic edge sites. J. Phys. Chem. C 113 (16): 6730–6734.
  48. 48. Okamoto, Y. (2009). First‐principles molecular dynamics simulation of O2 reduction on nitrogen‐doped carbon. Appl. Surf. Sci. 256 (1): 335–341.
  49. 49. Boukhvalov, D.W. and Son, Y.‐W. (2012). Oxygen reduction reactions on pure and nitrogen‐doped graphene: a first‐principles modeling. Nanoscale 4 (2): 417–420.
  50. 50. Zhang, L. and Xia, Z. (2011). Mechanisms of oxygen reduction reaction on nitrogen‐doped graphene for fuel cells. J. Phys. Chem. C 115 (22): 11170–11176.
  51. 51. Zhang, L., Niu, J., Dai, L., and Xia, Z. (2012). Effect of microstructure of nitrogen‐doped graphene on oxygen reduction activity in fuel cells. Langmuir 28 (19): 7542–7550.
  52. 52. Denis, P.A., Faccio, R., and Mombru, A.W. (2009). Is it possible to dope single‐walled carbon nanotubes and graphene with sulfur? ChemPhysChem 10 (4): 715–722.
  53. 53. Zhang, L.P., Niu, J.B., Li, M.T., and Xia, Z.H. (2014). Catalytic mechanisms of sulfur‐doped graphene as efficient oxygen reduction reaction catalysts for fuel cells. J. Phys. Chem. C 118 (7): 3545–3553.
  54. 54. Park, J.E., Jang, Y.J., Kim, Y.J. et al. (2014). Sulfur‐doped graphene as a potential alternative metal‐free electrocatalyst and Pt‐catalyst supporting material for oxygen reduction reaction. Phys. Chem. Chem. Phys. 16 (1): 103–109.
  55. 55. Wang, J., Ma, R., Zhou, Z. et al. (2015). Magnesiothermic synthesis of sulfur‐doped graphene as an efficient metal‐free electrocatalyst for oxygen reduction. Sci. Rep. 5: 9304.
  56. 56. Klingele, M., Pham, C., Vuyyuru, K.R. et al. (2017). Sulfur doped reduced graphene oxide as metal‐free catalyst for the oxygen reduction reaction in anion and proton exchange fuel cells. Electrochem. Commun. 77: 71–75.
  57. 57. Lu, Z.S., Li, S., Liu, C. et al. (2017). Sulfur doped graphene as a promising metal‐free electrocatalyst for oxygen reduction reaction: a DFT‐D study. RSC Adv. 7 (33): 20398–20405.
  58. 58. Zhai, C.Y., Sun, M.J., Zhu, M.S. et al. (2017). A new method to synthesize sulfur‐doped graphene as effective metal‐free electrocatalyst for oxygen reduction reaction. Appl. Surf. Sci. 407: 503–508.
  59. 59. Jeon, I.Y., Zhang, S., Zhang, L. et al. (2013). Edge‐selectively sulfurized graphene nanoplatelets as efficient metal‐free electrocatalysts for oxygen reduction reaction: the electron spin effect. Adv. Mater. 25 (42): 6138–6145.
  60. 60. Chen, X., Chen, X., Xu, X. et al. (2014). Sulfur‐doped porous reduced graphene oxide hollow nanosphere frameworks as metal‐free electrocatalysts for oxygen reduction reaction and as supercapacitor electrode materials. Nanoscale 6 (22): 13740–13747.
  61. 61. Randin, J.‐P. and Yeager, E. (1974). Effect of boron addition on the differential capacitance of stress‐annealed pyrolytic graphite. J. Electroanal. Chem. Interfacial Electrochem. 54 (1): 93–100.
  62. 62. Rao, C.N.R., Gopalakrishnan, K., and Govindaraj, A. (2014). Synthesis, properties and applications of graphene doped with boron, nitrogen and other elements. Nano Today 9 (3): 324–343.
  63. 63. Wang, S., Zhang, L., Xia, Z. et al. (2012). BCN graphene as efficient metal‐free electrocatalyst for the oxygen reduction reaction. Angew. Chem. Int. Ed. 51 (17): 4209–4212.
  64. 64. Zhou, Y.Z., Yen, C.H., Fu, S.F. et al. (2015). One‐pot synthesis of B‐doped three‐dimensional reduced graphene oxide via supercritical fluid for oxygen reduction reaction. Green Chem. 17 (6): 3552–3560.
  65. 65. Tam, T.V., Kang, S.G., Kadumudi, F.B. et al. (2017). Synthesis of B‐doped graphene quantum dots as metal‐free electrocatalyst for oxygen reduction reaction. J. Mater. Chem. A.
  66. 66. Zuo, Z.C., Jiang, Z.Q., and Manthiram, A. (2013). Porous B‐doped graphene inspired by fried‐ice for supercapacitors and metal‐free catalysts. J. Mater. Chem. A 1 (43): 13476–13483.
  67. 67. Hulicova‐Jurcakova, D., Puziy, A.M., Poddubnaya, O.I. et al. (2009). Highly stable performance of supercapacitors from phosphorus‐enriched carbons. J. Am. Chem. Soci. 131 (14): 5026–5027.
  68. 68. Liu, Z.W., Peng, F., Wang, H.J. et al. (2011). Phosphorus‐doped graphite layers with high electrocatalytic activity for the O2 reduction in an alkaline medium. Angew. Chem. Int. Ed. 50 (14): 3257–3261.
  69. 69. Zhang, C., Mahmood, N., Yin, H. et al. (2013). Synthesis of phosphorus‐doped graphene and its multifunctional applications for oxygen reduction reaction and lithium ion batteries. Adv. Mater. 25 (35): 4932–4937.
  70. 70. Li, R., Wei, Z., Gou, X., and Xu, W. (2013). Phosphorus‐doped graphene nanosheets as efficient metal‐free oxygen reduction electrocatalysts. RSC Adv. 3 (2013): 9978.
  71. 71. Mulyadi, A., Zhang, Z., Dutzer, M. et al. (2017). Facile approach for synthesis of doped carbon electrocatalyst from cellulose nanofibrils toward high‐performance metal‐free oxygen reduction and hydrogen evolution. Nano Energy 32: 336–346.
  72. 72. Feng, Y., Li, F., Hu, Z. et al. (2012). Tuning the catalytic property of nitrogen‐doped graphene for cathode oxygen reduction reaction. Phys. Rev. B 85 (15): 155454.
  73. 73. Huang, S.‐F., Terakura, K., Ozaki, T. et al. (2009). 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 80 (23): 235410.
  74. 74. Liang, J., Jiao, Y., Jaroniec, M., and Qiao, S.Z. (2012). Sulfur and nitrogen dual‐doped mesoporous graphene electrocatalyst for oxygen reduction with synergistically enhanced performance. Angew. Chem. Int. Ed. 51 (46): 11496–11500.
  75. 75. Zheng, Y., Jiao, Y., Ge, L. et al. (2013). Two‐step boron and nitrogen doping in graphene for enhanced synergistic catalysis. Angew. Chem. Int. Ed. 52 (11): 3110–3116.
  76. 76. Gong, Y.J., Fei, H.L., Zou, X.L. et al. (2015). Boron‐ and nitrogen‐substituted graphene nanoribbons as efficient catalysts for oxygen reduction reaction. Chem. Mater. 27 (4): 1181–1186.
  77. 77. Duan, X., O'Donnell, K., Sun, H. et al. (2015). Sulfur and nitrogen Co‐doped graphene for metal‐free catalytic oxidation reactions. Small 11 (25): 3036–3044.
  78. 78. Xu, C., Su, Y., Liu, D., and He, X. (2015). Three‐dimensional N,B‐doped graphene aerogel as a synergistically enhanced metal‐free catalyst for the oxygen reduction reaction. Phys. Chem. Chem. Phys. 17 (38): 25440–25448.
  79. 79. Kim, I.T., Song, M.J., Kim, Y.B., and Shin, M.W. (2016). Microwave‐hydrothermal synthesis of boron/nitrogen co‐doped graphene as an efficient metal‐free electrocatalyst for oxygen reduction reaction. Int. J. Hydrogen Energy 41 (47): 22026–22033.
  80. 80. Zhang, J., Qu, L., Shi, G. et al. (2016). N,P‐Codoped carbon networks as efficient metal‐free bifunctional catalysts for oxygen reduction and hydrogen evolution reactions. Angew. Chem. Int. Ed. 55 (6): 2230–2234.
  81. 81. Ai, W., Luo, Z., Jiang, J. et al. (2014). Nitrogen and sulfur codoped graphene: multifunctional electrode materials for high‐performance li‐ion batteries and oxygen reduction reaction. Adv. Mater. 26 (35): 6186–6192.
  82. 82. Wu, M., Dou, Z.Y., Chang, J.J., and Cui, L.L. (2016). Nitrogen and sulfur co‐doped graphene aerogels as an efficient metal‐free catalyst for oxygen reduction reaction in an alkaline solution. RSC Adv. 6 (27): 22781–22790.
  83. 83. Razmjooei, F., Singh, K.P., Song, M.Y., and Yu, J.S. (2014). Enhanced electrocatalytic activity due to additional phosphorous doping in nitrogen and sulfur‐doped graphene: a comprehensive study. Carbon 78 (0): 257–267.
  84. 84. Duan, X., Indrawirawan, S., Sun, H., and Wang, S. (2015). Effects of nitrogen‐, boron‐, and phosphorus‐doping or codoping on metal‐free graphene catalysis. Catal. Today 249: 184–191.
  85. 85. Van Pham, C., Klingele, M., Britton, B. et al. (2017). Tridoped reduced graphene oxide as a metal‐free catalyst for oxygen reduction reaction demonstrated in acidic and alkaline polymer electrolyte fuel cells. Adv. Sustainable Syst. 1600038.
  86. 86. Dou, S., Shen, A.L., Ma, Z.L. et al. (2015). N‐, P‐ and S‐tridoped graphene as metal‐free electrocatalyst for oxygen reduction reaction. J. Electroanal. Chem. 753: 21–27.
  87. 87. Men, B., Sun, Y., Li, M. et al. (2016). Hierarchical metal‐free nitrogen‐doped porous graphene/carbon composites as an efficient oxygen reduction reaction catalyst. ACS Appl. Mater. Interfaces 8 (2): 1415–1423.
  88. 88. Chen, P., Xiao, T.Y., Qian, Y.H. et al. (2013). A nitrogen‐doped graphene/carbon nanotube nanocomposite with synergistically enhanced electrochemical activity. Adv. Mater. 25 (23): 3192–3196.
  89. 89. Xu, P., Wu, D., Wan, L. et al. (2014). Heteroatom doped mesoporous carbon/graphene nanosheets as highly efficient electrocatalysts for oxygen reduction. J. Colloid. Interface Sci. 421 (0): 160–164.
  90. 90. Wang, M., Wu, Z.P., and Dai, L.M. (2015). Graphitic carbon nitrides supported by nitrogen‐doped graphene as efficient metal‐free electrocatalysts for oxygen reduction. J. Electroanal. Chem. 753: 16–20.
..................Content has been hidden....................

You can't read the all page of ebook, please click here login for view all page.
Reset