4
Carbon‐Based Electrochemical Oxygen Reduction and Hydrogen Evolution Catalysts

Ji Liang1, Yao Zheng2, Anthony Vasileff2, and Shizhang Qiao2

1University of Wollongong, Institute for Superconducting & Electronic Materials, Australian Institute of Innovative Materials, Innovation Campus, Squires Way, North Wollongong, NSW, 2500, Australia

2The University of Adelaide, School of Chemical Engineering, Adelaide, SA, 5005, Australia

4.1 Carbon Materials for Electrochemical Oxygen Reduction Catalysis

Oxygen reduction reactions (ORRs) are the most widely existing (electro)chemical processes in the nature. These are spontaneously involved in aerobic combustion, galvanic corrosion, metabolism, and many other natural processes. Apart from these, the ORR is also the critical electrochemical process in various energy conversion and storage devices. For instance, in an oxygen sensor, the concentration of oxygen is determined by the ORR potential, whereas in the next‐generation metal–air batteries or fuel cells, ORR ability contributes to overall performance. On this basis, the ORR is the most important step during the operation of these devices, and it is of great scientific necessity to study this simple but critical reaction.

These devices are typically operated at ambient temperature and pressure. However, in these conditions, the ORR generally exhibits sluggish kinetics. Consequently, catalysts are necessary for achieving meaningful reaction rates. Currently, platinum‐based catalysts (commonly used as platinum nanoparticles loaded on carbon black, denoted as Pt/C) are the most commonly applied and commercialized materials in the market. Their catalytic activity toward the ORR is generally a result of the morphology and microstructures of the Pt nanoparticles, which is also closely linked to the structure and chemistry of the carbon substrates. As a result, extensive studies have been performed on various carbon substrates for Pt nanoparticles and other metals or even directly as metal‐free catalysts for the ORR. In this section, we focus on summarizing the latest developments of these carbon‐based materials, especially toward reaction engineering, and provide road map of their structural evolution.

4.1.1 Electrochemical Process in the Reduction of Oxygen

4.1.1.1 Electrochemical Process and Catalytic Mechanism of the ORR

The electrochemical ORR is the reduction of oxygen molecules that are dissolved in aqueous/organic solutions or solids, from 0 oxidation state to −1 or −2 oxidation state through either two 2‐electron or one 4‐electron processes. Generally, the ORR involves the transfer of multiple electrons and has different initial and final potentials and products in different electrolytes (acidic or basic). As a result of all of these factors, the ORR is an extremely complex reaction.

If the ORR takes place in an aqueous acidic electrolyte and is catalyzed by Pt/C catalysts (the most common configuration for a low‐temperature proton‐exchange fuel cell), every oxygen molecule (O2) accepts four electrons and combines with four protons (hydrogen cations, H+) to form water (two H2O molecules). This process can be described as follows:

4.1vol-2-c04-math-0001

where E0 represents the relative potential to the standard hydrogen electrode (SHE).

If a catalyst is absent or not active enough, the ORR can also go through two 2‐electron processes in acidic electrolytes. Here, hydrogen peroxide is the intermediate product of the first two‐electron process and water is the final product of the second two‐electron process. These two steps can be described as follows:

4.2vol-2-c04-math-0002
4.3vol-2-c04-math-0003

However, in basic electrolyte, similar four‐electron and two‐electron processes still occur, with OH being the final reduction product. The four‐electron ORR process in basic electrolytes can be described as follows:

4.4vol-2-c04-math-0004

For the two‐electron ORR processes in basic electrolyte, it is generally believed that an oxygen molecule first receives two electrons to produce HO22− and then another two electrons to produce OH, as shown in the following equations:

4.5vol-2-c04-math-0005
4.6vol-2-c04-math-0006

Apart from this, in some conditions, the ORR may also undergo a possible one‐electron process to produce free radicals [1]. However, this is not commonly observed in typical energy conversion and storage devices. Therefore, we will not discuss this aspect in detail in this chapter.

If the aqueous electrolyte is replaced by an organic electrolyte (e.g. in the cathode of a lithium–air battery), similar 4‐electron or two 2‐electron pathways still exist, but with lithium ions (Li+) incorporated:

4.7vol-2-c04-math-0007
4.8vol-2-c04-math-0008

And in certain aprotic electrolytes, oxygen molecules can also receive one electron to form superoxides:

4.9vol-2-c04-math-0009

For a fuel cell or lithium–air/oxygen battery, it is required that the electrode reactions proceed relatively easily. Consequently, a larger reaction current is desirable at a lower overpotential for these reactions, especially for the naturally sluggish ORR processes. The intrinsic catalytic activity of various metals (regardless of the effect of their microstructure, morphology, or mass diffusion effects) can be described and reflected by the adsorption energy of oxygen molecules on their surface. A volcano plot can be drawn versus their catalytic activity (Figure 4.1) [2].

Image described caption and surrounding text.

Figure 4.1 The intrinsic ORR catalytic activity of various metals.

Source: Nørskov et al. 2004 [2]. Copyright 2004. Reproduced with permission from American Chemical Society.

As illustrated, Pt, Pd, and other noble metals exhibit the highest intrinsic catalytic activity toward the ORR. The transition metal species (Ni, Co, Fe, Co, etc.) tend to bind too strongly with oxygen and form oxides, whereas Au and Ag hardly adsorb oxygen. As a result, these metals exhibit inferior ORR catalytic activity compared with Pt and Pd. Activity improvements still exist for these noble metals, and this can be achieved by alloying Pt with metals that bind oxygen more strongly (e.g. Fe, Co, Ni, Cr, etc.) to form an optimized structure (e.g. a single‐atom layered alloy outside a Pt core). Theoretical calculations have shown that the adoption energy of these alloys toward oxygen can be slightly improved compared with Pt (i.e. from 1.57 eV for pure Pt to 1.89, 2.00, and 2.06 eV for Pt/Ni, Pt/Co, and Pt/Fe alloys, respectively). Due to these improvements, their ORR catalytic activity can be improved above that of pure Pt [2].

For the metal‐free catalysts, it is generally believed that their catalytic activity originates from the nonuniform distribution of charges on their surface (e.g. caused by heteroatom dopants in carbon frameworks), which can adsorb oxygen molecules by static electric forces. Similar to their metallic counterparts, the activity of these metal‐free catalysts can also be determined by their adsorption energies toward oxygen, as illustrated in Figure 4.2 [3].

Volcano plot depicting Catalytic activity of different dopants on the surface of graphene with Pt parallel to the x-axis; N-G; P-G; O-G; B-G; X-G; S-G; G are plotted in squares.

Figure 4.2 Catalytic activity of different dopants on the surface of graphene.

Source: Jiao et al. 2014 [3]. Copyright 2014. Reproduced with permission from American Chemical Society.

According to the calculations, the adsorption energies of these metal‐free catalysts toward oxygen are far lower than those of metallic catalysts at present. However, similar to the metal‐based catalysts, these metal‐free catalysts can be improved by simultaneously doping with multiple elements. These metal‐free catalysts are also much lower in price, making them quite attractive for practical applications.

Compared with ORR behavior in ideal conditions mentioned above, the actual ORR process is largely related to the catalysts' surrounding environment. Taking the ORR process in aqueous electrolyte as an example, the overall reaction pathway is as follows: (i) dissolved oxygen molecules diffuse from the electrolyte to the catalyst surface, (ii) oxygen adsorbs on the catalysts surface, (iii) electron relocation from the electrode to the adsorbed oxygen molecules, (iv) desorption of the reduction products from the catalyst surface, and (v) diffusion of these products from the catalyst surface to the electrolyte (Figure 4.3). Consequently, the actual ORR process involves not only energy conversion but also mass transfer. The role of ORR catalysts is the promotion of one or several of these steps, in order to facilitate the kinetics of the overall reaction. Based on these considerations, an ideal ORR catalyst should possess the following characteristics: appropriate adsorption strength toward oxygen molecules, high electronic conductivity, suitable pore structures, and good stability in electrochemical environments. By far, the most commonly employed commercial catalysts are Pt, Pd, and other platinum group metals. These metals are often prepared as nanoparticles and loaded on highly conductive substrates (e.g. carbon black) in order to most effectively utilize their catalytic surface. For example, the catalytic activity of Pt‐based catalysts is largely related to their geometry and crystal facets (e.g. 110 facets are preferred in acidic electrolytes, but 100 and 111 facets are not) [4]. In addition, choosing suitable particle size is also important in exposing these facets to maximize the adsorption of oxygen molecules. This requires the particle size to be as small as possible, while maintaining good stability. For other metal or metal‐free catalysts, these principles are also applicable to their design and synthesis.

Schematic diagram depicting ORR process on the electrode surface in aqueous electrolytes with Desorption and diffusion, and Diffusion and Adsorption labeled on arrows in the reaction.

Figure 4.3 ORR process on the electrode surface in aqueous electrolytes.

In order to lower the surface energy, fine metal particles tend to spontaneously aggregate into larger ones. Simultaneously, it is difficult to form sufficient conductive networks with these particles due to vast interphase resistance. Therefore, these metal particles (i.e. active sites) have to be loaded onto various conductive substrates to achieve high dispersion and conductivity for practical applications. As a result, the final microstructure and the overall performance of these composite catalysts are heavily determined by such conductive substrates. For metal‐free catalysts, the catalytic activity originates from their surface chemistry (i.e. active sites in the carbon frameworks). Hence, the chemical composition and microstructure of these catalysts also directly influence their ORR activity.

Based on this, carbon materials not only act as highly conductive substrates in ORR catalysts but also significantly affect (i) the effectiveness of active site exposure, (ii) the ease of reactant diffusion to active sites, and (iii) the electrochemical stability of the active sites. All these factors strongly contribute to the overall performance of these ORR catalysts.

4.1.1.2 Applications of ORR and ORR Catalysis

The ORR commonly exists in applications related to energy conversion and storage, such as fuel cells or metal–air batteries. In these devices, the ORR is the main cathodic reaction, and its efficiency largely determines their performance. In this regard, developing highly active ORR catalysts is critical for achieving high‐performance devices that are suitable for commercialization and practical applications.

A fuel cell is a high‐efficiency and low emission energy generation device that directly converts the chemical energy in fuels (hydrogen, methanol, etc.) into electricity [5, 6]. Taking a low‐temperature fuel cell, for example, it is composed of a cathode, anode, electrolyte (commonly aqueous acidic/basic solutions), proton/anion exchange membrane, and circuit/gas/heat‐controlling systems (Figure 4.4) [7]. In an acidic electrolyte, hydrogen is catalytically oxidized to produce hydrogen cations (i.e. protons, H+), which then travel through the proton exchange membrane to the cathode side of the cell. The ORR takes place at the cathode side of the cell, and the product combines the protons with oxygen to yield water. As only protons, rather than electrons, travel through the proton exchange membrane, the electrons have to pass through the outer circuit to form electrical current. If the electrolytes were basic, then the ORR product at the cathode is OH anions, which again pass through the anion exchange membrane to the anode to combine with hydrogen to yield water.

Image described caption and surrounding text.

Figure 4.4 Structural diagram of a low‐temperature fuel cell.

Source: Kirubakaran et al. [7]. Copyright 2009. Reproduced with permission from Elsevier.

Despite the different respective onset potentials for the cathodic/anodic reactions in acidic/basic electrolytes, the actual output voltage of the whole cell, which is dependent on the potential difference between the anode and the cathode, is the same (1.23 V) regardless of electrolyte type. Here, the overall reaction can be described as the combination of hydrogen and oxygen to yield water as follows:

4.10vol-2-c04-math-0010

Depending on operating temperatures, fuel cells can be sorted into high‐temperature fuel cells (operated over 600 °C) and medium/low‐temperature fuel cells (operated below 200 °C) (Figure 4.5) [8]. High‐temperature fuel cells are primarily solid oxide fuel cells (SOFCs) and molten carbonate fuel cells (MCFCs), which generally use hydrogen, natural gas, hydrocarbons, or even coal as fuel. Due to their very high operating temperature (600–1000 °C), the electrode reactions can be carried out spontaneously and continuously even without the need of catalysts. However, they require large volumes and complex structures, which limit them to stationary applications, e.g. in a power grid.

Schematic diagram depicting Category of fuel cells assorted by their operating temperatures with Internal and External reforming marked with chemical formulas and anode anode, cathode, and electrolyte labeled.

Figure 4.5 Category of fuel cells assorted by their operating temperatures.

Source: Steele et al. 2001 [8]. Copyright 2001. Reproduced with permission from Nature Publishing Group.

Low‐temperature fuel cells often operate in the temperature range between room temperature to 200 °C. This includes polymeric electrolyte membrane/proton exchange membrane fuel cells (PEMFCs), direct methanol fuel cells (DMFCs), phosphoric acid fuel cells (PAFCs), and alkaline fuel cells (AFCs)/anion exchange membrane fuel cells (AEMFCs). Apart from their low operating temperatures, they have the advantages of smaller required volumes and higher energy density, which make them suitable for portable energy supply for applications like electric vehicles and mobile devices. However, in these types of devices, the ORR cannot occur spontaneously at a satisfactory rate for high power output, and as a result, catalysts are often inevitable in order to accelerate the electrode reactions.

Another important application of electrochemical ORR is metal–air batteries, using metals such as aluminum, magnesium, zinc, or lithium as the anode and air/oxygen as the cathode. The most widely studied example is the rechargeable lithium–air battery. Due to lithium's extremely high capacity, this type of battery could theoretically yield a very high energy density (up to 5200 Wh kg−1), giving it a great potential for high density and portable energy storage applications. As a result, they have attracted huge attention recently.

In a lithium–air battery, lithium metal is directly used as the anode‐active material, whereas air or oxygen is used as the cathode‐active material. During discharge, lithium is oxidized and loses electrons to form lithium ions (Li+), which pass through the electrolyte to the cathode to combine with reduced oxygen to form various lithium oxides. The electrons then travel though the outer circuit and generate a current [9]. The discharge process can be generally described by the following equation:

4.11vol-2-c04-math-0011

Some researchers also believe that lithium oxide, rather than peroxide, may be partially produced:

4.12vol-2-c04-math-0012

When aprotic solvent electrolytes are used, a pathway to lithium superoxide is also possible, as shown in Equation 4.9.

Because lithium metal is directly used in lithium–air batteries, only nonaqueous electrolytes can be in direct contact with the lithium surface. However, there are more electrolyte choices available for the cathode side. According to these differences in electrolytes, lithium–air batteries can be divided into three main types [9]: (i) with lithium immersed in an electrolyte of dissolved lithium salts, in which condition, a solid electrolyte interphase is simultaneously or artificially formed on the surface of the lithium metal, preventing further corrosion of the lithium and providing a diffusion pathway for lithium ions (Figure 4.6a,b); (ii) with a solid‐phase electrolyte in between the anode and the cathode, which provides pathways for lithium ions and prevents internal short circuiting (Figure 4.6c); and (iii) with nonaqueous electrolytes in the anode and aqueous electrolyte in the cathode, employing a hydrophobic membrane between these two electrolytes for lithium transportation (Figure 4.6d). Despite the structural differences among these different lithium–air batteries, their cathodes are all highly air‐penetrable electrodes on which the ORR is carried out.

Image described caption and surrounding text.

Figure 4.6 Structural diagrams of various lithium–air batteries using different electrolytes.

Source: Girishkumar et al. 2010 [9]. Copyright 2010. Reproduced with permission from American Chemical Society.

Currently, the most frequently used commercial ORR catalysts are Pt or Pt alloy (Pt/Pd, Pt/Ru, etc.) nanoparticles loaded on highly conductive carbon black. However, due to the high cost and poor recyclability of Pt‐based catalysts, it is difficult to produce a commercially viable device for widespread use on these catalysts. In a PEMFC, the ORR process is much more sluggish compared with the anode reactions. Consequently, catalyst loading at the cathode side of a PEMFC largely overweighs the loading on the anode side. However, in a metal–air battery, all of the catalyst is used to accelerate the cathodic ORR, whereas its anode does not require any catalysts. In this regard, if the specific activity of these catalysts could be improved or they are replaced altogether by cheaper catalysts, then fuel cells and metal–air batteries could be significantly more viable from the economic point of view. According to the above discussions, because the actual performance of a catalyst is largely dependent on the microstructure and chemistry of the carbon (substrates), the recently emerged novel carbon structures and materials thus provide a number of possibilities for the development of inexpensive catalysts with high activity.

4.1.2 Carbons as Catalyst Supports for ORR

The initial research on Pt/C catalysts dates back to the 1970s during the development of the PAFC by the United Technologies Corporation [10]. In this cell, carbon fiber textiles are used as the current collector, on which Pt is also loaded as the catalyst. During the last 40 years, Pt has been the most widely used catalyst for the ORR, and many carbons with different microstructures and properties have been employed as supports in order to improve the performance of the metal nanoparticle catalysts. In this section, we will focus on recent developments of carbon materials used as catalyst supports for the ORR.

4.1.2.1 Composition of Metal Nanoparticles and Carbons

To prepare Pt/C composite catalysts, two main routes have been developed, namely the impregnation method and colloidal method. In the impregnation method, carbon substrates are dispersed in the solution containing the metal precursors (e.g. metal salts). Thereafter, reducing agents (e.g. sodium borohydride or hydrazine) are added to reduce the metal ions into the metal. The metal forms nanoparticles that attach to the surface of the carbon substrates to produce the composite catalysts. The metal precursor can also be soaked in the pores of the carbons by evaporating the solvents in the suspension containing carbon and metal precursor before the reduction. Following this, the metal nanoparticles can be obtained inside the pores of the carbons by reducing the metal precursors, e.g. by heating the mixture in reductive atmospheres. This method is especially suitable when loading metal nanoparticles on highly porous carbon substrates, which can provide a large surface for depositing nanoparticles. Based on this impregnation method, a modified method, called incipient wetness impregnation, has been developed specifically to suit the synthesis of catalyst particles inside the pores of carbon support. This method uses the same amount of precursor solution as the carbon pore volume, thus avoiding possible catalyst deposition outside of the pores. In the colloidal method, metal nanoparticles are first prepared to form a colloidal suspension in a solvent and carbons are then mixed with this colloidal suspension to adsorb the metal particles on its surface. The colloidal method can provide catalyst particles with uniform and small size, especially when the particles are stabilized by proper surfactants in the solution.

Apart from these two methods, other techniques, like electrochemical deposition and the scattering method, have been studied for the preparation of Pt/C catalysts. In electrochemical deposition, a potential is applied to the carbon substrates in order to reduce the metal species and form metal nanoparticles on the carbon surface, whereas in the scattering method, Pt metal is used as a target material and carbon as a substrate in order to load fine Pt nanoparticles on the carbon surface.

For newly emerged ORR catalysts based on non‐noble metals and their compounds, they can be composited with carbon through a variety of methods. This includes the hydrothermal method, which converts metal precursors into metal (oxides) of different structures and directly loads them onto the carbon surface. Other techniques explored involve the direct introduction of metal precursors with carbon precursors, followed by simultaneous carbonization and metal (oxide) formation. These will be discussed in detail in the following sections.

4.1.2.2 Conventional Carbons: Carbon Black and Graphite

Carbon black and natural graphite are typical conventional carbon materials used as catalyst substrates. They possess advantages such as low cost and high electrical conductivity, making them the most frequently used support for ORR catalysts. Carbon black powder is usually obtained by pyrolyzing hydrocarbons (e.g. natural gas, gasoline, and coal) in an oxygen‐deficient environment [11]. Microscopically, it is an onion‐like particle composed of stacked graphite microlayers (Figure 4.7a). The particle size of carbon black is in the range of several to tens of nanometers. Due to the lack of long‐range order in the structure of carbon black, they are generally regarded as amorphous carbon.

Image described caption and surrounding text.

Figure 4.7 (a) Diagram showing the microstructure of carbon black particles. (b) The morphology of Pt nanoparticles loaded on carbon black.

Source: Kim et al. 2006 [12]. Copyright 2006. Reproduced with permission from Elsevier.

Source: Wissler 2006 [11]. Copyright 2006. Reproduced with permission from Elsevier.

Based on synthesis method, carbon black can be categorized into furnace black, channel black, and acetylene black. Carbon black prepared by different methods may have greatly different morphologies and properties, as summarized in Table 4.1.

Table 4.1 Summary of the properties of different carbon blacks.

Carbon blacks Supplier Synthesis method Surface area (m2 g−1) Particle size (nm)
Vulcan XC‐72 Cabot Furnace black 250–260 20–50 [13]
Black Pearl 2000 Cabot Furnace black 1475–1500 15 [14, 15]
Denka Black Denka Acetylene black 58–65 40 [14, 15]
Shawinigan Black Chevron Acetylene black 70–90 40–50 [15, 16]
Conductex 975 Ultra Columbian Furnace black 250 24 [14]
3250/3750/3950 Mitsubishi 240/800/1500 28/28/16 [17]
Ketjen EC‐300 J Akzo Nobel Furnace black 800 [15] 30 [18]
Ketjen EC‐600 JD Akzo Nobel Furnace black 1270 30 [19]
Ketjen EC‐600 JD Cabot Furnace black 1274 30–35 [20]

Among these carbon black materials, Vulcan XC‐72 produced by Cabot is the most widely used catalyst support, commanding about 80% of the current market share [15]. Other carbon blacks, such as Black Pearl and Ketjen Black have also been widely used as catalysts supports. In order to maximize the utilization of expensive Pt, Pt is commonly synthesized into particles of 3–5 nm and then loaded onto the carbon black substrates (Figure 4.7b). However, these highly active metal particles can not only catalyze the reduction of oxygen but also accelerate the electrochemical oxidation of the amorphous carbon substrates, especially in corrosive electrolytes [21, 22].

Pristine carbon black can withstand air at over 400 °C without being notably oxidized. However, once it is loaded with Pt nanoparticles, oxidation can occur at temperatures as low as 125 °C. At higher temperatures, oxidation of carbon black becomes more rapid upon higher Pt loadings (Figure 4.8). Once oxidation takes place, the microstructure of the carbon black can be rapidly destroyed, which leads to severe detachment or aggregation of the catalyst nanoparticles and thus loss of catalytic activity. Compared with the oxidation in air, electrochemical corrosion of carbon takes place more easily in the corrosive electrolyte environments of fuel cells or metal–air batteries. This can be ascribed to two main factors: (i) the oxidation potential of carbon is as low as 0.2 V versus SHE, which means electrochemical oxidation of carbon may initiate at this low potential (being far lower than the ORR equilibrium potential of 1.23 V) [15, 24] and (ii) the water present in the electrolyte can also enhance the oxidation of carbon, combining with oxidized carbon to form either carbon monoxide or carbon dioxide in the medium‐ and low‐temperature fuel cells, respectively [25, 26].

Two line graphs depicting oxidation of carbon black with various Pt loadings at (a) 173C and (b) 195C in air with points plotted of wt%: 80; 60; 40; 30; 20; 10; 5.

Figure 4.8 Oxidation of carbon black with various Pt loadings at (a) 173 °C and (b) 195 °C in air.

Source: Stevens and Dahn 2005 [23]. Copyright 2005. Reproduced with permission from Elsevier.

In order to suppress the oxidation of the carbon substrates, graphitic carbon materials with higher crystallinity have been employed. However, graphite generally does not provide sufficient surface area to load high amount of Pt nanoparticles. To solve this problem, graphite has been finely ground to expose greater surface area. For example, after high‐energy ball milling for 50 h, the specific surface area of graphite increased significantly from 7 to 580 m2 g−1 [27]. However, this top–down treatment also simultaneously introduces a large number of defects into the graphite, which was evidenced by Raman spectroscopy. Figure 4.9 compares the Raman spectra of milled graphite for different mill times where an increase in the D peak indicates a higher degree of defects. This shows that simple ball milling severely compromised the desired high crystallinity of the graphite, and thus, it is not suitable for the synthesis of carbon substrates.

Line graph of Raman spectra of graphite upon different hours of ball milling with different points plotted to HSG-0; HSG-5; HSG-20; HSG-50.

Figure 4.9 Raman spectra of graphite upon different hours of ball milling.

Source: Li et al. 2008 [27]. Copyright 2008. Reproduced with permission from Elsevier.

Another route to synthesizing highly crystalline carbon is to graphitize amorphous carbons (e.g. carbon black) at extremely high temperatures to cause the rearrangement of carbon atoms with higher order. For example, by heating XC‐72 carbon black in high‐vacuum or inert atmosphere to over 2000 °C, its crystallinity can be greatly enhanced. This is supported by an increased Ig:Id ratio in the Raman spectra, which indicates the formation of long‐range graphitic structures. However, the rearrangement of carbon atoms also causes the disappearance of pores, especially micropores and mesopores, which are also a form of defect. Commonly, a loss of surface area up to 60% is expected to occur for carbons after such treatment [28, 29].

For these conventional carbon materials, it is extremely hard to simultaneously achieve both large surface area and high crystallinity. With the purpose of achieving higher nanoparticle loadings on carbons with higher graphitization for better overall catalytic performance, a number of novel carbon materials have been developed and employed for ORR catalysis applications.

4.1.2.3 Carbon Nanomaterials as Supports for ORR Catalysts

Carbon nanomaterials include carbons on which at least one dimension is in the nanoscale. For example, if all the three dimensions of a carbon material are in the nanoscale (e.g. carbon quantum dots or fullerenes), they can be regarded as 0‐dimensional nanocarbon. If one dimension of the materials is in the nanoscale (e.g. graphene), they can be called as two‐dimensional nanocarbons. If two dimensions of the materials are in nanoscale (e.g. carbon nanotubes (CNTs) or carbon nanofibers (CNFs)), then they belong to the group of one‐dimensional nanocarbons. Some research also categorizes carbons with nanopores into this type of nanocarbons; however, we will discuss them in a separate section.

Carbon nanomaterial formation can be regarded as the scissoring, bending, stacking, and/or reassembling of two‐dimensional graphene nanosheets, which is constructed of sp2‐hybridized carbon atoms (Figure 4.10) [30, 31]. For example, directly bending, rolling, and stitching of a ribbon‐like graphene sheet would result in a CNT, whereas stacking of single‐layered graphene would result in multilayered graphene or graphite, if thick enough.

“Schematic diagrams of processes of (a) formation of three carbon nanomaterials from graphene building blocks and (b) diagram and transmission electron microscopy (TEM) images of carbon nanohorns and nanocoils.”

Figure 4.10 (a) Formation of different carbon nanomaterials from graphene building blocks and (b) diagram and transmission electron microscopy (TEM) images of carbon nanohorns and nanocoils [3032].

Source: Reproduced with permission from Nature Publishing Group, copyright 2007; American Chemical Society, copyright 2003; and Elsevier, copyright 2002.

The 0‐dimensional fullerenes are usually very small in size (<1 nm). Consequently, they cannot be used to support metal nanoparticles or effectively form a highly conductive network. As a result, the most commonly used carbon nanomaterials to support Pt nanoparticles are the one‐dimensional CNTs/CNFs or two‐dimensional graphene.

CNTs can be divided into single‐walled carbon nanotubes (SWCNTs) with a typical diameter less than 1 nm and multiwalled carbon nanotubes (MWCNTs) with a diameter up to tens of nanometers. The length of CNTs may vary from tens of nanometers to several millimeters [32]. If structural defects exist on the walls of the CNTs (e.g. pentagonal rings instead of hexagonal rings), distortion of the tube walls may occur. This may lead to the formation of a tube end closure (i.e. a carbon nanohorn) or twisting of tubes (i.e. a carbon nanocoil) [31, 33].

Due to their unique one‐dimensional and high‐crystalline characteristics, CNTs have very unique properties compared with conventional carbon materials, including high electrical conductivity (up to 104 versus 10 S m−1 for carbon black), very few surface defects to permit electrochemical corrosion, and good mechanical properties to maintain the structural stability during electrochemical processes. However, the very highly crystallized but passive surface of CNTs also has its disadvantages, namely limited deposition sites on which to anchor the metal nanoparticles or other active species for the ORR. To deal with this, artificial defects have to be intentionally introduced onto the surface of CNTs as anchoring sites for metal particle deposition and for better particle dispersion as well. A common method to achieve this is by surface oxidation of the CNTs in order to graft oxygen‐containing functional groups on their surface. For instance, soaking CNTs in nitric acid can graft carboxyl groups (COOH) on their surface [34], whereas harsher acidic treatments, such as refluxing CNTs in a mixture of sulfuric and nitric acids at 140 °C for several hours, will result in surface decoration with multiple functional groups (i.e. hydroxyl (OH), carboxyl (COOH), and carbonyl (CO) groups with an approximate molar ratio of c. 4 : 2 : 1) [35, 36]. Small particle size (2–5 nm) and very good particle dispersion have been achieved for Pt/Ru alloy nanoparticles when deposited on surface‐functionalized CNTs with a mass loading of 20 wt.% (Figure 4.11) [37].

Image described caption and surrounding text.

Figure 4.11 TEM images of Pt/Ru alloy nanoparticles deposited on the surface‐functionalized CNTs, (a) Pt:Ru = 7 : 3 and (b) Pr:Ru = 1 : 1.

Source: Prabhuram et al. 2006 [37]. Copyright 2007. Reproduced with permission from American Chemical Society.

Other than treating CNTs with strong and oxidative acids, there are also reports using alternative techniques to realize surface modification of CNTs. For example, after pre‐oxidation of CNTs using dilute nitride acid (5 M), further treatment with acetic acid to graft additional oxygen‐containing groups has been tested [38]. Compared with the treatments with harsh acids, this method tends to retain the smooth surface and structural integrity of the CNTs (Figure 4.12). Further, Pt nanoparticles with small size and high dispersion can still be deposited on CNTs treated in this way [38].

“TEM images of CNTs treated by nitric acid and acetic acid, and Pt/CNTs composites; and bar graphs of particle size distributions of Pt. (a, a1, a2) Pt loading of 0.11; (b, b1, b2) 0.24; (c, c1, c2) 0.42.”

Figure 4.12 TEM images of CNTs treated by nitric acid and acetic acid, and Pt/CNTs composites, as well as the particle size distributions of Pt. (a, a1, a2) Pt loading of 0.11 mg cm−2; (b, b1, b2) 0.24 mg cm−2; (c, c1, c2) 0.42 mg cm−2.

Source: Saha et al. 2008 [38]. Copyright 2008. Reproduced with permission from Elsevier.

Such two‐step oxidation method not only well maintains the structure of CNTs but also helps obtain highly dispersed Pt particles on the CNT surface. At a very high Pt loading of 0.42 mg cm−2, the Pt nanoparticle size can still be controlled to less than 5 nm with a uniform distribution on the CNT surface (Figure 4.12) [38]. Due to these advantages, this Pt/Ru/CNT composite catalyst achieves improved performance compared with conventional Pt/C.

Apart from the noble metals such as Pt, Ru, and Pd, other non‐noble metals or their compounds have also been recently used in order to develop more cost‐effective ORR catalysts. The ORR catalytic mechanism of these materials has also been studied in detail, and a number of new synthesis methods have also been developed.

For example, CNTs were slightly oxidized at low temperatures, and oxygen‐containing functional groups were formed on their surface without compromising their structures. Thereafter, the as‐treated CNTs were hydrothermally treated together with cobalt acetate and ammonia to finally form a CoO/CNT composite for ORR catalysis (Figure 4.13) [39]. Using this method, simultaneous N doping by ammonia and uniform distribution of CoO nanoparticles of very small size (∼5 nm) were achieved on the resultant CNT material. In terms of performance, this material showed comparative activity to commercial Pt/C catalysts in basic electrolyte. The origin of this excellent activity, as summarized by the author, can be attributed to the covalent bonding between the N‐doped CNTs and the CoO particles anchored on their surface (CoNC bonds). This bonding configuration significantly enhances the electron transition from the CNT surface to the CoO and thus facilitates the electrochemical ORR process.

(a) SEM and (b) TEM images of CoO nanoparticles loaded on CNTs with (220) marked in an inset in (b). In (b), arrows indicate NCNT and CoO.

Figure 4.13 (a) Scanning electron microscopy (SEM) and (b) TEM images of CoO nanoparticles loaded on CNTs.

Source: Liang et al. 2012 [39]. Copyright 2012. Reproduced with permission from American Chemical Society.

Non‐noble metal nanoparticles can also be formed during the formation of carbon nanomaterials. For instance, cobalt ferrocyanide (Co2Fe(CN)6) was ball milled with graphene nanosheets and then annealed at 600 °C in an inert atmosphere (Figure 4.14a). During this process, the graphene sheets curved to form CNTs, which wrapped the Co/Fe nanoparticles inside to form a capsule‐like hybrid material (Figure 4.14b–g) [40]. Compared with pure graphene, a significant improvement in ORR catalytic activity was achieved on this Co/Fe/CNT/graphene hybrid material. By simply changing the mass ratio of the two precursors (ferrocyanide and graphene), the CNT wall number could also be tuned. This enabled the researchers to study the effect of wall number on ORR mechanism and performance. The results show that with increasing wall numbers, the overall ORR activity of the hybrid decreases and that the optimum wall number is 1–3. A Fe/CNT composite with a similar capsule‐like structure can also be fabricated by directly annealing suitable precursors in an inert gas. For example, by pyrolyzing sodium azide (NaN3) and ferrocene (Fe(C5H5)2), a bamboo‐like CNT encapsulating iron (oxide) nanoparticles can be obtained, also showing stable ORR activity [41]. In spite of enhanced ORR performance to comparative analogs, the catalytic activities of these materials are still inferior to commercial Pt/C. This may be attributed to the capsule structure, which prevents the active species from effectively contacting the reactant oxygen species in the electrolyte.

(a) Schematic diagram depicting the synthesis of a Fe/Co alloy loaded on a CNT/graphene hybrid by (1) Ball milling, (2) Annealing, (3) Acid treat; (b) its SEM image; and (c-g) TEM images at different magnifications.

Figure 4.14 (a) Synthesis of a Fe/Co alloy loaded on a CNT/graphene hybrid, (b) its SEM image, and (c–g) TEM images at different magnifications.

Source: Deng et al. 2013 [40]. Copyright 2014. Reproduced with permission from Royal Society of Chemistry.

Some studies have shown that some organic polymers are also active for ORR catalysis and thus can be used as active materials and loaded on CNTs. One such polymer is poly(diallyldimethylammonium chloride) (PDDA), which can provide a positively charged surface when they are grafted onto CNTs and electrostatically adsorb the oxygen molecules in the electrolyte (Figure 4.15a) [42]. Despite the relatively poor ORR activity compared with that of conventional Pt/C, in terms of ORR onset potential and catalytic current, they possess superior ORR stability, far outperforming Pt/C, and show very good reaction selectivity.

Image described caption and surrounding text.

Figure 4.15 (a) Structural diagram of a CNT with surface decoration of poly(diallyldimethylammonium chloride) (PDDA) and (b) its linear scanning voltammetry curves for the ORR.

Source: Wang et al. 2011 [42]. Copyright 2011. Reproduced with permission from American Chemical Society.

Generally, CNTs have unique advantages of high electrical conductivity and structural stability. In order to best utilize these characteristics, future studies may be necessary to develop more facile and cost‐effective methods to achieve uniform nanoparticle loading without compromising their graphitic structures (i.e. maintaining its high conductivity and resistance against electrochemical corrosion). Attention should also be paid to the electron interface transition between the metal nanoparticles and the CNT substrates, which may contribute to high impedance if sufficient coupling is not achieved. This electron interface transition may be enhanced by creating strong interactions between these two species, like by the formation of covalent bonds between them.

The other typical and commonly applied one‐dimensional carbon nanomaterials are CNFs. Highly graphitized CNFs can be obtained by catalyst‐assisted (e.g. transition metals) chemical vapor deposition at 700–1200 K. The most frequently used precursors for CNFs are carbon‐containing gases, such as methane, ethane, and carbon monoxide [43, 44]. Another type of CNF is composed of amorphous carbon, which can be derived from the carbonization of precursor fibers (e.g. polyacrylonitrile (PAN) nanofibers) [45]. According to the stacking orientation of the graphite platelets in CNFs, they can be divided into three types: platelet, ribbon, and herringbone structure, as shown in Figure 4.16 [46]. The diameters of CNF can range from ten to several hundred nanometers based on the synthesis method.

Image described caption and surrounding text.

Figure 4.16 Diagram of CNFs with different microstructures.

Source: Bessel et al. 2001 [46]. Copyright 2001. Reproduced with permission from American Chemical Society.

Rather than loading the metal (compound) nanoparticles on the surface of the CNFs, as in the case of CNTs, these particles can also be homogeneously distributed throughout the CNF by directly incorporating the metal precursor to the CNF precursors (i.e. the polymer nanofibers) and then carbonizing the composite precursor fiber. This method has been used to obtain Fe‐containing CNFs. Specifically, polyvinylpyrrolidone (PVP) was chosen as the CNF precursor. It was then dissolved in ethanol and ferrous acetylacetonate (Fe(acac)2) was added as the iron precursor. The composite precursor fibers were fabricated by electrospinning, which laid over each other to form an interconnected network (Figure 4.17a). This composite precursor (PVP/Fe(acac)2) was then carbonized and converted into the Fe/CNFs composite to be used for ORR catalysis [47].

Image described caption and surrounding text.

Figure 4.17 (a) Schematic of the synthesis of Fe‐containing CNFs and (b, c) the SEM and TEM images of the composite CNFs (Fe–N/C NMs).

Source: Wu et al. 2015 [47]. Copyright 2015. Reproduced with permission from Nature Publishing Group.

By tuning the synthesis parameters (i.e. time, humidity, and the aging period in air), the structures of the final products can also be changed. For example, if the precursor was aged in air for 24 h with a humidity of 70% and then carbonized, a mat‐like network with individual fibers overlapping each other can be obtained (FeN/C NMs, Figure 4.17b,c), whereas direct carbonization of the precursor fibers without aging would lead to the formation of an interconnected structure with the fibers fused together at the connecting points (FeN/C NNs) [47]. In this material, the iron species exist in the forms of both metal and metal‐oxide nanoparticles (i.e. Fe and Fe3O4), which evenly distribute on the surface and inside of the CNFs due to the homogeneous distribution of the iron precursors. Better ORR catalytic performance was observed for FeN/C NNs compared with FeN/C NMs, which was explained to be due interconnectedness of the fused points among the CNFs, permitting improved electron transport throughout the network.

A similar electrospinning concept can be transferred to the fabrication of Co/CNF composites, using PAN as the CNF precursor and cobalt acetate as the Co precursor. By carbonizing the composite fibers containing these two components, a composite CNF loading a mixture of Co and its oxides (CoO and Co3O4) can be obtained with most of the metal species located near or on the surface of the CNFs (Figure 4.18) [48]. Similar to the previous case, the Co content in this material can be tuned by simply changing the ratio of the two precursors, and by increasing the Co content, ORR activity also improves. Even though its activity is not as good as Pt/C, enhanced ORR stability is achieved, as is the case for many other non‐noble metal ORR catalysts.

Image described caption and surrounding text.

Figure 4.18 The TEM images of Co‐containing CNFs from (a) side and (b) cross‐sectional views.

Source: Kim et al. 2015 [48]. Copyright 2015. Reproduced with permission from Royal Society of Chemistry.

Compared with CNTs, CNFs are more versatile from the aspects of their synthesis methods, which make it easier to load active metal species on them. However, these Pt‐free catalysts are still inferior to commercial Pt/C catalysts. The main reason for this is the intrinsically low ORR activity of these transition metal species. Additionally, these amorphous CNFs have poor conductivity, much lower than CNTs and carbon black. Therefore, finding a solution to improve the crystallinity of the CNFs, while maintaining their synthesis versatility, may be helpful to further improve their ORR catalytic performance. Apart from this, other research attention should be focused on decreasing the metal (oxides) particle size being loaded on the CNFs. Improving from the current tens of nanometers to a more desirable size, say 3–5 nm, could help to better utilize the metal species.

Graphene is a new member in the family of carbon materials, which has drawn tremendous attention since its discovery in 2004. To prepare graphene, a variety of technologies have been developed. It can be prepared by mechanical methods, like the mechanical exfoliation of natural graphite, or by electrochemical/chemical methods. Some of these include the electrochemical/sono‐assisted exfoliation of graphite in solvents [4954], sono‐assisted exfoliation of chemically oxidized graphite (i.e. graphite oxide) to prepare graphene oxide (GO) followed by reduction [55], and the thermal expansion technique whereby rapid heating of oxidized graphite exfoliates the graphene layers [56]. High‐quality and highly crystalline graphene can also be prepared by chemical vapor deposition of carbon‐containing gases on transition metal substrates (e.g. Cu or Ni) [57, 58]. Among the graphene materials produced by these methods, the synthesis route via reduction of GO is especially suitable for catalyst construction because it has abundant surface functional groups that can act as anchoring sites for active species. It is also a relatively cost‐effective method that is scalable for mass production.

Chloroplatinic acid (H2PtCl6) is the typical precursor for the preparation of Pt nanoparticles. To obtain a H2PtCl6/graphene composite precursor, H2PtCl6 was dissolved in acetone and mixed with graphene powders produced by the thermal expansion method. After heating this composite in hydrogen at 300 °C, the H2PtCl6 decomposed into Pt nanoparticles, which homogeneously distributed on the graphene surface. The Pt formed nanoparticles in the size range of 1.5–3.5 nm at a loading of 20 wt.% (Figure 4.19) [59]. Compared with commercial Pt/C catalysts, the Pt particles on the Pt/graphene composite are much smaller. As a result, this material has a larger electrochemical active surface available to adsorb oxygen (108 versus 72 m2 g−1), affording better ORR performance. Further, the rich surface functional groups are able to strongly adsorb the Pt nanoparticles on the graphene surface, which help improve the stability of the catalyst. After long‐term testing (5000 cycles), the particle size of Pt on the Pt/graphene composite is still smaller than that of the Pt/C catalyst (5.5 versus 6.9 nm). At a significantly higher Pt loading of 60 wt.%, monodispersed Pt nanoparticles can still be obtained on the graphene surface, with only a slightly larger particle size of c. 1.87 nm [60]. Besides Pt, this method is also applicable for loading other noble metal nanoparticles with very small sizes, such as Pd or Pt/Ru alloys [60].

(a) Bright-field and dark-field (inset) TEM images of Pt nanoparticles loaded on graphene and (b) a histogram depicting the particle size distribution of Pt on graphene.

Figure 4.19 (a) Bright‐field and dark‐field (inset) TEM images of Pt nanoparticles loaded on graphene and (b) the particle size distribution of Pt on graphene.

Source: Kou et al. 2009 [59]. Copyright 2009. Reproduced with permission from Elsevier.

The colloid method can also be used to load metal nanoparticles on graphene. For example, Pt/Fe alloy nanoparticles (diameter of c. 7 nm) were first synthesized by the solvothermal treatment of platinum(II)acetylacetonate (Pt(acac)2) and iron carbonyl (Fe(CO)5) at 220 °C. A homogeneous suspension of graphene in tetrahydrofuran was then mixed with a suspension of the Pt/Fe alloy and dissolved in a mixture of tetrahydrofuran and ethanol. The resultant mixture was ultrasonicated and centrifuged in order to obtain the Pt/Fe/graphene composite ORR catalyst [61]. The same process was used to prepare Pt/Fe nanoparticles loaded on carbon black as a comparative analog material. In acidic electrolyte, both catalysts exhibit superior activity to commercial Pt/C for ORR. However, the graphene composite had improved catalytic activity over the carbon black composite. Additionally, this Pt/Fe/graphene composite catalyst also possesses extremely good electrochemical stability in acidic electrolyte and shows negligible performance attenuation after 10 000 cycles. Using a similar strategy, Co metal nanoparticles (diameter of c. 10 nm) were synthesized and their surface was spontaneously oxidized by air to form a Co/CoO core shell structure. A suspension of these particles was ultrasonically blended with graphene in order to obtain a Co/CoO/graphene composite catalyst for the ORR [62]. However, this catalyst does not have comparable ORR activity to commercial Pt/C, despite its good stability.

More methods have been developed to load non‐noble metals and their compounds onto the surface of graphene, and the most common one is the “one‐pot” hydrothermal treatment of GO and metal salts to convert them into graphene and metal oxides/hydroxides. For instance, by hydrothermally treating a suspension containing GO, cobalt acetate, and ammonia, a composite catalyst composed of reduced GO and cobalt oxide (Co3O4/graphene) can be obtained with a Co3O4 nanoparticle size of 4–8 nm (Figure 4.20a,b) [63]. In this process, the ammonia also dopes the graphene with nitrogen (N content of c. 4%), and the N dopants within the graphene network form covalent bonds with the Co3O4 particles, effectively anchoring them on graphene surface. This strong covalent bonding is able to facilitate electron transfer and thus significantly improves the material's ORR performance. Moreover, the ORR performance of this material further improves with higher concentrations of basic electrolyte.

“TEM images of graphene-based catalysts loaded with different metal compounds, Co3O4/graphene and the electron diffraction pattern (inset of b). Arrows point to Co3O4 on graphene (a), Co3O4 (b), and Graphene (b).”; TEM images of MnCo2O4/graphene and the electron diffraction pattern (inset of d) with arrows pointing to (400); (311); (220).

Figure 4.20 TEM images of graphene‐based catalysts loaded with different metal compounds, (a, b) Co3O4/graphene and the electron diffraction pattern (inset of b). (c, d) MnCo2O4/graphene and the electron diffraction pattern (inset of d).

Source: Liang et al. 2012 [64]. Copyright 2012. Reproduced with permission from American Chemical Society.

Source: Liang et al. 2011 [63]. Copyright 2011. Reproduced with permission from Nature Publishing Group.

Similarly, dual‐metal oxides (e.g. MnCo2O4) can also be synthesized and loaded on graphene by the hydrothermal method (Figure 4.20c,d) [64]. The obtained oxide particles (c. 5 nm) are homogeneously dispersed on the surface of graphene. In MnCo2O4, Co3+ ions in the Co3O4 crystal are replaced by Mn3+ ions, and as a result, the ORR activity of these oxide nanoparticles is improved. Interestingly, this dual‐metal oxide can directly reduce oxygen into OH in basic electrolyte through a four‐electron pathway that is similar to that of Pt.

During the hydrothermal treatment, the functional groups on the GO surface are partially removed, not only resulting in the reduction of GO but also leading to the reassembly of the graphene sheets due to the π–π interactions between them. In the case that the initial concentration of GO is high enough, they may go through a self‐assembly process to form bulky units (i.e. graphene hydrogel) after being reduced. During this process, the hetero particles in the suspension may be trapped inside the three‐dimensional network of the hydrogel to form a composite with the graphene. The solvents inside the hydrogel can be removed afterward by other techniques (e.g. by cyro/freeze‐drying) without destroying its three‐dimensional structure and finally yield a highly porous and lightweight graphene‐based composite aerogel.

This unique property of graphene has been utilized to prepare a Fe3O4/graphene composite aerogel as a free‐standing, highly porous, and bulky ORR catalyst [65]. A hydrogel containing Fe3O4/graphene was prepared by the hydrothermal treatment of an aqueous solution containing iron(II)acetate, GO, and pyrrole (Figure 4.21a). Water in the gel was then removed by freeze‐drying of the hydrogel to yield the Fe3O4/graphene aerogel. Interconnected macropores of several microns can be observed on the aerogel, whose walls are composed of graphene (Figure 4.21b,c). On the graphene walls, Fe3O4 particles of about 100 nm can be found to be evenly dispersed and act as the ORR‐active sites. Compared with Pt/C catalysts, this Fe3O4/graphene composite is still inferior, especially with respect to ORR onset potential and conversion efficiency (i.e. relatively low electron transfer numbers). Similar concepts for fabricating bulky graphene assemblies may also be transferred to the synthesis of materials with other metal species and may also find roles for other electrochemical/structural applications.

Image described caption and surrounding text.

Figure 4.21 (a) Diagram of the synthesis of a Fe3O4/graphene composite aerogel and (b, c) SEM images of the material at different magnifications.

Source: Wu et al. 2012 [65]. Copyright 2012. Reproduced with permission from American Chemical Society.

4.1.2.4 Porous Carbons as Catalyst Supports for ORR

Apart from conventional carbon and carbon nanomaterials, another very important group of carbon materials is porous carbons. According to the difference in pore size, porous carbons can be divided into three categories: microporous carbon (pore size < 2 nm), mesoporous carbon (2 nm < pore size < 50 nm), and macroporous carbon (pore size > 50 nm). The synthesis of porous carbon is also very versatile, and the most frequently adopted method for controllable synthesis is the templating method. This method utilizes “templates” with different structures to tailor the pore structures of the obtained carbons and the templates are then removed to expose the carbon pores.

Template Synthesis of Porous Carbons

Porous carbons can be synthesized using two types of templates: a soft template and a hard template. Hard templates are materials with certain “rigidity,” which does not significantly change its structure or morphology during the synthesis of the porous carbon. Hard templates that are usually employed include porous silica, porous alumina, silica microspheres, polystyrene (PS)/poly(methyl methacrylate) (PMMA) spheres, and many bio‐derived templates. Usually, hard templates are stable during the synthesis of carbons (especially the various oxide‐based templates), which means they may need to be removed by subsequent treatments in order to expose the pores on the carbons.

By choosing hard templates with different pore structures, a porous carbon with a reversed structure can be obtained. For example, zeolite templates will result in carbons with micropores, whereas the mesoporous template SBA‐15 with tubular pores (p6mm structure) is frequently used to synthesize mesoporous carbon composed of parallel carbon nanorods (CMK‐3). If silica or other polymer microspheres are chosen as the template, the obtained carbon would then show sphere‐like pores with interconnected necks. When these spheres are orderly assembled (i.e. by sedimentation in solvents), then the obtained carbon would show a reversed opal structure with honeycomb‐like ordered pores. By simply changing the sphere diameter, the pore size of the carbons can be tuned (Figure 4.22a) [66]. Silica templates can also be used to introduce pores into graphene. For example, a silica colloidal suspension (12 nm) was blended with GO solution and then dried to form a silica/GO composite with silica nanospheres among the GO sheets. Thereafter, the silica was removed by etching with hydrofluoric acid (HF), forming a mesoporous graphene with the same pore size as the silica particle diameter (Figure 4.22b) [67].

Image described caption and surrounding text.

Figure 4.22 (a) Scheme showing the synthesis of porous carbon with different hard templates. (b) A synthesis of porous graphene with silica nanospheres.

Liang et al. 2012 [67]. Copyright 2012. Reproduced with permission from John Wiley and Sons.

Source: Lee et al. 2006 [66]. Copyright 2006. Reproduced with permission from John Wiley and Sons.

However, soft templates are usually molecules with hydrophilic/hydrophobic terminals (i.e. surfactants; e.g. the tri‐block copolymer P123/F127). These molecules form micelles in the solvent through self‐assembly and bind with the carbon precursor molecules (e.g. low‐molecular‐weight phenolic resin) through hydrogen bonding or electrostatic interactions to form a micelle/resin precursor in the solvent (Figure 4.23). This composite precursor then goes through an evaporation‐induced self‐assembling (EISA) process in ambient atmosphere followed by heating–solidification in air or an assembly–solidification process during the hydrothermal treatment [6871].

“Flow diagram with schematics depicting synthesis of mesoporous carbon from tri-block copolymer F127 and low-molecular-weight resin. Ethanol solution goes through process including Evaporation, Thermal polymerization; Carbonization and Heating in nitrogen.”

Figure 4.23 Synthesis of mesoporous carbon from tri‐block copolymer F127 and low‐molecular‐weight resin.

Source: Meng et al. 2005 [68]. Copyright 2005. Reproduced with permission from John Wiley and Sons.

In some cases, hard templates and soft templates can be used together, in order to prepare carbon materials with special pore structures. For example, PMMA microspheres were assembled into a hexagonally ordered structure (Figure 4.24a), and silica was then cast into the voids among the spheres using tetraethyl orthosilicate (TEOS) as the precursor. The PMMA template was then removed during the calcination of the silica, which leads to the formation of porous silica with honeycomb‐like pores (Figure 4.24b). Subsequently, a solution containing a soft template (P123) and carbon precursor (phenolic resin) was impregnated into the spherical pores of the porous silica and then converted to carbon after solidification and carbonization at high temperatures. After removal of the silica template, carbon nanospheres with ordered mesopores were obtained (Figure 4.24c–e) [72].

Image described caption and surrounding text.

Figure 4.24 Synthesis of mesoporous carbons using dual templates, (a) SEM images of PMMA template as low and (inset) high magnifications; (b) SEM images of the silica template with a reversed opal structure at low and (inset) high magnifications; (c–e) SEM and TEM images of the obtained mesoporous carbon nanosphere at different magnifications and from different view directions; and (f) diagram of the synthesis of a hierarchical porous carbon with both macropores and mesopores.

Source: Schuster et al. 2012 [72]. Copyright 2012. Liang et al. 2013 [73]. Copyright 2013. Reproduced with permission from John Wiley and Sons.

Another example is the simultaneous use of PS nanospheres and F127 as hard and soft templates, respectively, to prepare carbon materials with both macropores and mesopores [73]. In this method, PS spheres were orderly packed by sedimentation in an aqueous suspension to form an opal structure as the hard template. An ethanol solution containing phenolic resin (i.e. the carbon precursor) and F127 (i.e. the soft template) was then impregnated into the voids among the PS spheres under vacuum. Following, a resin/F127 composite with mesostructure, which existed in the void of the PS spheres, was formed by the room temperature EISA process during the removal of ethanol. The templates were then thermally decomposed during the carbonization of the resin, resulting in a hierarchical porous carbon with ordered tubular mesopores on the walls of the interconnected spherical macropores (Figure 4.24f).

Apart from these commonly used template materials, people have also used other materials, such as bio‐derived or other natural material with fine structures as the template to synthesize mesoporous carbons [74, 75]. These newly emerged template materials not only result in a variety of new carbon materials with various pore structures but also help in the development of more environment‐friendly technologies to synthesize cost‐effective porous carbons.

Porous Carbons as Catalyst Supports for ORR

Mesoporous carbon with dual pore sizes (CMK‐5) was first utilized as a supporting material for Pt nanoparticles for ORR [76]. Similar to CMK‐3, CMK‐5 is also composed of hexagonally ordered carbon nanorods. However, these are hollow‐structured like a tube. As a result, this type of carbon usually gives a much larger specific surface area than other mesoporous carbons, making them very suitable for catalyst support applications. To utilize this unique property, the impregnation–reduction method was applied to prepare the Pt/CMK‐5 composite catalyst for the ORR. Specifically, chloroplatinic acid precursor was dissolved in solvent and impregnated into the pores of CMK‐5, followed by the removal of solvents and reduction of precursor into Pt nanoparticles. The resultant material comprises smaller Pt particle size (<5 nm) than commercial Pt/C catalysts. As a result, a higher specific activity toward ORR catalysis was achieved on this Pt/CMK‐5 catalyst. Another typical mesoporous carbon (CMK‐3) has also been used for this purpose, which exhibits superior ORR activity to Pt/C as well. This indicates the huge potential of mesoporous carbons for ORR applications due to their unique pore structures [77].

For mesoporous carbon prepared by hard templates, metal (compound) nanoparticles are often loaded inside their pores by posttreatment methods after the carbons are synthesized. In contrast, it is more versatile and convenient to prepare the mesoporous carbon/metal composite through a soft template method. For example, during the soft template synthesis of mesoporous carbons, as illustrated in Figure 4.23 [68], metal salt can be used as the metal precursor and added into the system to involve in the self‐assembling process and to form a metal‐containing resin with ordered meso‐structures. After the subsequent carbonization, ordered mesoporous carbon with metal (compound) nanoparticles embedded in their framework can be obtained.

For example, in the dual‐template synthesis of a hierarchical porous carbon, iron chlorides (metal precursor) and acetylacetone (acac, used as a chelating agent for better dispersion of the metal ions) were dissolved in ethanol and then added to the precursor solutions (i.e. phenolic resin as the carbon precursor and F127 as the soft template). This mixture precursor solution was then impregnated in the void of an orderly packed PS hard template, which then went through a room temperature EISA process, solidification at 100 °C, and carbonization at a high temperature. This process yielded a hierarchical porous carbon with both macropores and mesopores (Figure 4.25) [78].

Image described caption and surrounding text.

Figure 4.25 Diagram of the synthesis of the Fe‐containing carbon hybrid of CNTs and hierarchical porous carbon.

Source: Liang et al. 2014 [78]. Copyright 2014. Reproduced with permission from John Wiley and Sons.

Interestingly, in this material, the iron species not only act as the ORR‐active species but also perform as the catalyst for the growth of CNTs during the carbonization of the resin precursor (Figure 4.26). These CNTs thus enhance electron transport during the ORR process and greatly improve the ORR activity of the material. The ORR performance of this material is closely comparable with that of commercial Pt/C with respect to ORR onset potential, catalytic current, and reaction efficiency [78].

Image described caption and surrounding text.

Figure 4.26 SEM and TEM images of the CNTs grown in situ from hierarchical porous carbon blocks.

Source: Liang et al. 2014 [78]. Copyright 2014. Reproduced with permission from John Wiley and Sons.

Two major advantages exist for mesoporous carbons employed as ORR catalysts, especially for carbons with ordered mesopores. These are as follows: (i) their large specific surface area provides sufficient sites that contribute to the homogeneous anchoring of the metal (compound) nanoparticles and (ii) the uniform and ordered mesopores can provide continuous and effective mass diffusion for the reactants to travel from the electrolyte to the surface active sites. In order to improve the intrinsically low conductivity of mesoporous carbons, due to their relatively low graphitization degree, it would be necessary to incorporate various carbon nanomaterials with the porous carbons to form interconnected networks for facile electron transportation, which is essential for electrochemical processes [73, 78].

4.1.3 Carbon Materials as Metal‐Free Catalysts for the ORR

Apart from being the supporting materials for metal or metal compound nanoparticles, carbon materials can also be directly used as completely metal‐free catalysts upon certain surface modifications, for example, by doping with heteroatoms. Nitrogen has been the most frequently used element to dope carbons. In the carbon framework, N atoms change the charge density of the carbon surface and thus disrupt its electron neutrality. On the surface of N‐doped carbon materials, the negatively charged centers can attract oxygen molecules dissolved in the electrolyte by static electric forces, leading to its catalytic activity for the ORR. Other elements, when doped in carbon, have also been found to be able to change the spin density of the carbon, which further enhances the adsorption of oxygen on its surface [7981]. Apart from affecting the electronic structures of the carbon surface, recent density functional theory (DFT) calculations have shown that N doping can also narrow the bandgap of the carbon materials and thus enhance its electronic conductivity [82].

Doping of carbons with nitrogen is commonly achieved by annealing carbons in ammonia or other N‐containing gases (N2 is normally not effective for N‐doping), by introducing N‐containing precursors into the carbon precursors, or by directly using N‐containing carbon precursors. For example, poly(amidoamine) can be directly used as a N‐containing carbon precursor. This precursor was impregnated into the pores of mesoporous silica SBA‐15 and carbonized. Subsequent removal of template formed the desired N‐doped mesoporous carbon (Figure 4.27) [83]. It was found that increased carbonization temperature positively shifted the ORR onset potential, indicating a higher catalytic activity. The highest activity was achieved for materials prepared at a carbonization temperature of 800 °C. These materials achieved an ORR onset potential of +0.03 V versus Ag/AgCl, which is close to that of Pt/C catalysts.

Image described caption and surrounding text.

Figure 4.27 (a) Diagram of N dopants in the carbon frameworks of a mesoporous carbon, (b) TEM image of N‐doped mesoporous carbon, and (c) the ORR linear scanning voltammetry profiles of N‐doped mesoporous carbon prepared at different temperatures.

Source: Nagaiah et al. 2012 [83]. Copyright 2012. Reproduced with permission from John Wiley and Sons.

The doping configurations of N atoms in the carbon framework also remarkably affect its catalytic activity [84]. According to their binding energies (Eb) with carbon, N species in the carbon framework can generally be labeled as pyridinic‐N (Eb = 398.5 eV), imine/nitrile‐like N (Eb = 399.5 eV), pyrrolic‐N (Eb = 400.5 eV), graphitic‐N (Eb = 401.3 eV), and N oxides (Eb = 402–405 eV). Among these N species, pyridinic‐N and graphitic‐N are believed to possess the highest activity for ORR catalysis.

Similar to N‐doped carbons, graphitic carbon nitride (g‐C3N4) has gained much attention in recent years due to its very unique surface structure of alternate and interconnected N and C atoms, which provides very abundant surface active sites for the adsorption of oxygen molecules [85, 86]. To overcome its semiconductivity, g‐C3N4 can be incorporated with various carbon materials (e.g. CMK‐3) to form composite catalysts (Figure 4.28a,b) [86]. Compared with pure g‐C3N4, this composite catalyst exhibits much improved activity toward the ORR and also catalyzes the reaction directly to form OH in basic electrolytes through the highly efficient four‐electron pathway. Like many other metal‐free catalysts, this g‐C3N4/CMK‐3 is very stable during long‐term operation and is highly selective for ORR against other side reactions (Figure 4.28c–f).

Image described caption and surrounding text.

Figure 4.28 (a, b) TEM images of the g‐C3N4/CMK‐3 from different angles and at different magnifications; (b, d) cyclic voltammetry and linear scanning voltammetry curves of the materials; and (e, f) stability and selectivity tests of the materials.

Source: Zheng et al. 2011 [86]. Copyright 2011. Reproduced with permission from American Chemical Society.

For this type of metal‐free ORR catalyst based on g‐C3N4 or other semi/nonconductive materials, it is critical to introduce certain conductive networks. DFT calculations also show that the mesoporous carbon (CMK‐3) in the g‐C3N4/CMK‐3 composite can effectively alternate the ORR pathway from two 2‐electron to one 4‐electron pathway, by preventing the OOH intermediate from accumulating on the catalyst surface [86].

Apart from nitrogen, other elements have also been explored as dopant elements in the carbon frameworks to achieve certain ORR activities. For example, phosphorus acts similar to nitrogen in tuning the electronic structure of the doped carbon, and as a result, it can enhance ORR catalytic activity. Synthesis of P‐doped mesoporous carbon has been fabricated using triphenylphosphine (P(C6H5)3) as the phosphorus precursor, benzene as the carbon precursor, and SBA‐15 as the hard template (Figure 4.29) [87]. Although possessing good ORR selectivity, its ORR activity is still not as good as Pt/C.

Image described caption and surrounding text.

Figure 4.29 (a) Synthesis diagram of P‐doped mesoporous carbon, (b, c) its SEM and TEM images at different magnifications, and (d, e) the cyclic voltammetry profiles showing its ORR activity and selectivity.

Source: Yang et al. 2012 [87]. Copyright 2012. Reproduced with permission from American Chemical Society.

For N or P doping, it is believed that the high electronegativity of these elements draws electron from the neighboring carbon atoms and thus forms a negatively charged center that is able to attract oxygen molecules and facilitate ORR catalysis. However, in the case of S‐doped carbons, which has similar electronegativity to that of carbon, its ORR activity may result from other mechanisms [67, 88]. According to our DFT calculations, S dopants in the carbon framework can affect the material's surface electronic structure by changing the spin density of the neighboring carbon atoms, which thus leads to ORR activity. This ORR activity can be further enhanced if nitrogen is co‐doped with sulfur into the carbon framework, which leads to a synergistic improvement in material's ORR activity (Figure 4.30) [67].

Chemical structural schematics depicting the charge density and spin density of carbon atoms in S- and N-co-doped carbon frameworks with a table at the right with Atom number; Spin density; Charge density of C1 to C5 and N and S all labeled in the diagram.

Figure 4.30 The charge density and spin density of carbon atoms in S‐ and N‐co‐doped carbon frameworks.

Source: Liang et al. 2012 [67]. Copyright 2012. Reproduced with permission from John Wiley and Sons.

For further improvements to the ORR activity of metal‐free catalysts, structural optimization can also be utilized. This can effectively be achieved by integrating porous carbons with carbon nanomaterials, which have large surface area and high conductivity, respectively (Figure 4.31a–e) [73].

(a-e) SEM and TEM images of the hierarchical porous carbon/graphene hybrid doped with nitrogen with Bridging N-G encircled and N-OMMC marked and (f, g) the cyclic voltammetry and linear scanning voltammetry profiles of the material as well as its comparative analogs.

Figure 4.31 (a–e) SEM and TEM images of the hierarchical porous carbon/graphene hybrid doped with nitrogen and (f, g) the cyclic voltammetry and linear scanning voltammetry profiles of the material as well as its comparative analogs.

Source: Liang et al. 2013 [73]. Copyright 2013. Reproduced with permission from John Wiley and Sons.

It has been proven that by grafting in‐situ‐formed graphene nanosheets on hierarchical porous carbons, the advantages of both materials can be successfully integrated. In this hybrid, the mesopores provide large specific surface area for the effective exposure of nitrogen‐active sites, the interconnected macropores facilitate the diffusion of reactants, and the in‐situ‐formed graphene acts as the conductive network. By preserving the beneficial properties of both components, this hybrid material thus results in much improved ORR performance (Figure 4.31f,g).

4.1.4 Section Summary

It is interesting to find that carbon materials have gone through a very clear road map from conventional carbons to novel carbons and from pure catalyst supports to important functional components in the development of catalysts for the ORR. It is unlikely that one single‐carbon material can meet all criteria necessary for an ideal ORR catalyst. As a result, very recent studies have tried to combine many different types of carbons together to achieve better overall performance.

Some research has revealed certain highly active surface sites for the ORR, such as transition metal atom and nitrogen complex sites. These complex doping configurations on the carbon surface are much more active compared with single‐atom‐doping configurations. As a result, it is necessary to develop new and simple technologies to introduce such complex doping configurations onto various carbons. Furthermore, it is also necessary to look for other doping configurations, which may possess even higher ORR activity.

4.2 Carbon Materials for the Electrochemical Hydrogen Evolution Reaction

Electrocatalytic hydrogen evolution reaction (HER) is a fundamental process in electrocatalysis because of its simplicity and is considered to be a cornerstone reaction in exploring the mechanism of complex electron transfer processes [89]. At the same time, the HER is one of the most mature electrochemical processes and is still used to produce high‐purity hydrogen fuel for hydrogen‐based fuel cells and other hydrogen‐based energy sources. Therefore, HER studies are well documented in both electrochemistry and materials fields [90]. At present, there are tremendous opportunities in advancing electrochemical surface science at the atomic/molecular level by merging computational and experimental methodologies, which may lead to many breakthroughs in the research and development of HER electrocatalysts [90, 91]. Among these, the HER in acidic media is well understood, including the detailed configuration of hydrogen adsorption/absorption behaviors and the construction of descriptor‐based (usually hydrogen–metal bond energy, EM–H and hydrogen adsorption free energy, ΔGH*) volcano plots, etc. [89, 9294]. These plots allow the quick establishment of the activity trend and explain the HER reactivity origin on a variety of carbon materials. More importantly, they can also provide a qualitative argument for tuning catalytic activities to achieve the “best” material [89]. However, such significant success has not been realized for alkaline HER. In fact, even though the reaction mechanism is justified, the computation of energetics and kinetics of alkaline HER have largely been neglected by the computational chemistry community. As a result, much debate still exists among the experimentalists, e.g. whether hydrogen adsorption energy acts as the sole activity descriptor, as is the case for acidic environments, and whether or not the effects of extra water dissociation energy barriers influence the overall reaction rate [91, 9597]. The main aim of this section is to provide a comprehensive account, addressing both computational and experimental aspects of the acidic HER process, toward a better/deeper understanding of the electrochemistry fundamental to this process by revealing the activity origin of a variety of carbon‐based materials including heteroatom‐engineered graphene, defective graphene, and hybridized graphene materials.

4.2.1 Atomic‐Level Understanding of Single Heteroatom‐Doped Carbon Materials

In principle, engineering nanocarbon materials with heteroatoms can modify their electronic properties and, consequently, the adsorption behavior of intermediates during the electrocatalytic processes. As mentioned previously, nitrogen‐doped graphene (N graphene) materials have shown comparable activities with non‐precious metal or even precious metal counterparts in the field of ORR in alkaline media. However, a similar achievement for the HER remains unattained [98100], and the performance gap between carbon materials and metallic benchmarks is even larger than those for the ORR. To reveal why this is the case, Qiao and coworkers have evaluated HER activities on a series of heteroatom (B, N, O, S, or P)‐doped graphene materials by combining spectroscopic characterization, electrochemical measurements, and DFT calculations [89]. The computed ΔGH* on all models constructed (based on experimentally synthesized samples) all show largely positive values (Figure 4.32a), which indicates relatively weak hydrogen adsorption. This behavior is attributed to the nature of doped graphene materials related to their electronic structures (Figure 4.32b). The study also shows that, from both experimental and theoretical perspectives, the best performance on a single‐element‐doped graphene still cannot match that of the metallic counterparts, even of the non‐precious metal nanostructured MoS2 (Figure 4.32c). To address this, they proposed two strategies to enhance the apparent HER activity of carbon‐based materials without changing their intrinsic electronic structure. These strategies are to increase the heteroatom‐doping level and/or increase the material's surface area (Figure 4.32d,e). For example, it is predicted that a doped graphene material with an exchange current of i0 = 1.10 × 10−21 A per site (the value for the MoS2 benchmark is 5.25 × 10−21), 5% heteroatom‐doping level, and surface area of 1000 m2 g−1 will have an apparent HER activity (from the point of view of overpotential) that exceeds MoS2. As each proposed physicochemical character has been individually achieved on different graphene/carbon materials, this approach is believed to be feasible.

Image described caption and surrounding text.

Figure 4.32 (a) The computed lowest ΔGH* for different graphene‐based models. The ΔGH* values on graphitic‐type doping models is labeled by solid bars, whereas those on edge doping models are labeled by shaded bars. (b) volcano plot of theoretically computed ΔGH* versus exchange current (itheory) on various doped graphene models. Open symbols represent experimentally measured iexp on various synthesized graphene materials. (c) Electronic structure origins of HER activity on doped graphene materials: the relationship between ΔGH* and the highest peak position of density of states (DOSs) of the active carbon atoms. (d, e) Calculated HER polarization curves of graphene‐based materials (set as i0 = 1.10 × 10−21 A per site) with different doping levels and different surface areas.

Source: Jiao et al. 2016 [89]. Copyright 2016. Reproduced with permission from Nature Publishing Group.

4.2.2 Atomic‐Level Understanding of Dual‐Heteroatom‐Doped Carbon Materials

Even though altering the physicochemical properties of carbon materials is effective for enhancing their performance, optimization of their electronic structures is more essential and direct. Based on the successes achieved for ORR catalysts, dual doping is also considered to be one of the best strategies for modifying (specifically, to strengthen) the hydrogen adsorption on the surface of the carbon‐based materials. Indeed, the HER free‐energy diagram shows that doping N graphene with a secondary element, e.g. S or P, can increase the hydrogen adsorption strength (Figure 4.33a,b) [89]. There are also many experimental studies that have prepared a range of N/S and N/P co‐doped carbons for the HER, using different approaches, which validate this prediction. Moreover, N/S‐co‐doped samples always show improved performances over N/P‐co‐doped samples [102107]. This is also supported by the computational results whereby the active carbon atom in N/S‐co‐doped materials has a higher location of anti‐bonding states, i.e. closer to the Fermi level [89]. It should be noted that as shown by both theoretical and experimental studies, a randomly selected two‐heteroatom‐doped graphene is not guaranteed to possess enhanced activity. For example, N/B‐co‐doped graphene, having higher ORR performance compared with N graphene [108], shows the contrary with decreased HER performance when compared with N graphene [89, 101].

Image described caption and surrounding text.

Figure 4.33 (a) Representative atomic configurations of three dual‐doped (N,B; N,S; and N,P) models. (b) The three‐state free energy diagram for the pure, single‐, and dual‐doped graphene models. (c) Scheme of N,B; N,S; and N,P dual‐doped carbon nanosheets fabricated by polydopamine chemistry. (d) The HER polarization curves of different single‐ and dual‐doped graphene samples in 0.5 M H2SO4 solutions.

Source: Qu et al. 2017 [101]. Copyright 2016. Reproduced with permission from Nature Publishing Group.

Compared with single doping, dual‐doping carbon/graphene materials are more challenging due to the difficulty of controlling the doping configurations and achieving a satisfactory doping level of the secondary dopants. This is not favorable for realizing enhanced apparent activity, but more importantly, it blocks the atomic model construction for fundamental studies like hydrogen adsorption energetic computation and identification of possible active sites. To address this, many advanced synthetic strategies have been developed, including the selection of one‐ or two‐step doping [109, 110], precursor choice [111, 112], and in situ or ex situ doping [105, 113]. A breakthrough has been achieved by Qiao and coworkers, who applied polydopamine (PDA) as a platform to achieve efficient and controllable co‐doping in nanocarbons [101]. The unique molecular structure of PDA offers a wide variety of interactions with different secondary heteroatom precursors. For example, as demonstrated in Figure 4.33c, the catechol groups of PDA can conjugate boric acid [114, 115]; PDA is highly reactive to thiol functional groups via Schiff base or Michael addition reaction [114, 116, 117]; the positively charged amino groups of PDA can bind to phosphate groups through electrostatic attraction. As a result, this newly developed dual‐doped material shows an increased HER performance compared with dual‐doped materials derived from traditional methods or even single‐element‐doped materials (Figure 4.33d). The favorable activity originates from the high doping level of both the primary N and secondary S or P dopants and the material's large surface area [101].

4.2.3 Atomic‐Level Understanding of Defective Graphene Materials

As a 2D material, there is always a certain amount of topological defects, either grown‐in or extrinsic, in graphene's lattice [118]. The latter always locate on the edge or curvature sites of the graphene, which may also significantly affect its adsorption behavior and electrocatalytic activity [119]. Therefore, both theoretical and experimental studies have reported that defective graphene without any heteroatom doping also shows favorable electrocatalytic activities [120122]. Defective graphene can also serve as a substrate to host metal nanoparticles. This is because it provides more efficient anchoring sites via strong π–π interaction to directly couple transition metal atom sites, leading to fast electron transfer kinetics and excellent stability [123, 124]. One of the featured examples was developed by Yao and coworkers [122], who introduced different types of topological defects into the graphene lattice by the removal of intentionally doped heteroatoms (Figure 4.34a,b). Unsurprisingly, the defective graphene shows an enhanced HER activity compared with pure and doped graphene (Figure 4.34c). The corresponding DFT computations also revealed the electrocatalytic activity origin of different types of topological defects (Figure 4.34d). This simple approach may import overwhelming advantages over the current complicated fabrication process of carbon‐based electrocatalysts like multistep co‐doping, functionalization and hybridization.

Image described caption and surrounding text.

Figure 4.34 (a) Schematic of the formation of defective graphene (DG) via N‐doped graphene (NG); (b) high‐angle annular dark‐field (HAADF) image of defective graphene, orange, green, blue, and red cycle indicates hexagons, pentagons, heptagons, and octagons carbon ring, respectively; (c) HER polarization curves of different carbon‐based materials in acidic media; (d) HER free energy diagram of different edge and defective graphene models.

Source: Jia et al. 2016 [122]. Copyright 2016. Reproduced with permission from John Wiley and Sons.

4.2.4 Atomic‐Level Understanding of Hybridized Carbon Materials

Besides engineering carbon with heteroatom doping and introducing defects, hybridizing carbon with different nonmetallic materials that are electrochemically non‐active or less active can also boost their HER activities. Generally, the promotion of carbons toward other nonmetal catalysts (host materials) can be reflected in three forms: (i) expanding surface area to enlarge the active sites on the host material, (ii) enhancing the electrical conductivity of the whole hybrid to facilitate the electron/charge transfer on host material's surface or interface, and (iii) fundamentally altering the electronic structure of host material to modify its surface adsorbing properties [125]. All the three ways, either individually or synergistically, could enhance the electrocatalytic activity of the resultant hybrids. The most featured example uses graphitic carbon nitride (g‐C3N4), which is known as a semiconductor and inert electrocatalyst, and can be engineered with carbon to boost its intrinsic activity [126130]. The carbon substrate can be varied with either single‐ or dual‐doped graphene layers and g‐C3N4 can also be in the form of 2D sheets [129], 3D porous films [130], or nanoribbons [128]. These hybrids all show significantly increased electrocatalytic activity in comparison with pristine g‐C3N4, and more importantly, their performances are close to or even higher than those of some classic metal electrocatalysts.

With the availability of DFT calculations, the origin of such synergistic activity enhancement can be uncovered at the atomic level. Taking g‐C3N4 hybridized with N graphene (C3N4@NG) nanosheets, for example [129], the ultrathin layers of g‐C3N4 and N graphene have strong chemical interaction with one another (Figure 4.35a,b). Consequently, the charge density in the hybrid's interlayers was redistributed in the form of electron transfer from the conductive N graphene to g‐C3N4, leading to a downshifting of the valance and conduction bands in g‐C3N4 (Figure 4.35c,d). As a result, the Fermi level crosses the conduction band of g‐C3N4, which facilitates enhanced electron mobility and consequently an increased electrocatalytic activity. Such theoretical predictions have been validated using experimental HER measurements as a C3N4@NG hybrid showed increased HER activity compared with its individual components (g‐C3N4 and N graphene) and their physical mixture counterparts (Figure 4.35e). DFT‐predicted values of ΔGH* (using different models) indicate that the adsorption on g‐C3N4 is too strong (large negative ΔGH* value) whereas on N graphene it is too weak (large positive ΔGH* value), both unfavorable for the HER. Chemical coupling of the two components into a uniform hybrid can result in an optimized ΔGH* value and, therefore, a mediated adsorption–desorption behavior to facilitate overall HER kinetics (Figure 4.35f). As shown in Figure 4.4g, metal‐free C3N4@NG catalyst perfectly follows the volcano plot trend along with a wide collection of metal catalysts. More importantly, its activity, judged on the basis of both electrocatalytic i0 and thermodynamic ΔGH* properties, is comparable with that of the state‐of‐the‐art nanostructured MoS2.

Image described caption and surrounding text.

Figure 4.35 (a) HAADF image of C3N4@NG hybrid. (b) Nitrogen K‐edge NEXAFS spectra of pure g‐C3N4 and C3N4@NG hybrid. (c) Electron transfer between g‐C3N4 and N graphene layers; (d) the projected DOS on pure g‐C3N4 and C3N4@NG hybrid models; (e) the HER polarization curves for four metal‐free electrocatalysts and 20% Pt/C reference in acidic media; (f) the calculated free energy diagram of HER at the equilibrium potential for three metal‐free catalysts and Pt reference; (g) volcano plots of i0 as a function of the ΔGH* for C3N4@NG (red triangle), common metal catalysts (open symbols), and a typical nanostructured MoS2 catalyst (closed symbol).

Source: Zheng et al. 2014 [129]. Copyright 2014. Reproduced with permission from Nature Publishing Group.

4.2.5 Section Summary

The promising nonmetal carbon‐based materials have demonstrated competitive properties as the next generation of highly efficient HER catalysts. DFT computations have shown their powerful predicting capability for these materials, paving the way toward catalyst molecular design for the HER. A combination of theoretical calculations and experimental data provides clear and solid evidence that, similarly to precious metals, their well‐designed metal‐free counterparts have also great potential for highly efficient electrocatalytic HER. Looking forward, there are also some weaknesses with carbon‐based materials. Namely, their alkaline HER performances are always worse compared with metal electrocatalysts [95, 96, 131133], probably due to their weak water dissociation abilities. Also, carbon‐based materials are always chemically/electrochemically unstable under oxidizing potentials for the electrocatalytic oxygen evolution reaction (OER), limiting their application from bifunctional materials for the overall water splitting.

4.3 Conclusion, Summary, and Perspective

The electrochemical oxygen reduction and hydrogen evolution reactions are essential processes in next‐generation energy storage and conversion devices. The commercial catalysts for these reactions are commonly based on noble metals, such as Pt, Ru, Pd, and others, in the form of nanoparticles loaded on the surfaces of various conductive substrates (e.g. carbon black). Despite their excellent performance, their high cost, and difficulty for recycling make them unsuitable for large‐scale applications and thus have hindered the commercialization of many advanced energy storage and conversion devices. With the rapid development of new carbon materials (e.g. CNTs, graphene, and various porous carbons), the research has also moved onto the utilization of these materials as novel catalyst supports or even completely metal‐free catalysts for the ORR and HER. It has shown that these carbon‐based materials, which are free of noble metals, can sometimes achieve very high catalytic activity for these reactions, which is even comparable with the commercial catalysts.

In these new materials, both the surface chemistry and the microstructure simultaneously control their overall performance: their activity originates from surface characteristics, whereas their microstructure (especially their pore and crystal structure) determines if these active sites can be effectively utilized. Based on these, it may be interesting and necessary for further studies to design materials that address both of these aspects. One possible solution could be the integration of various carbon materials, which complement each other, e.g. by growing highly graphitic CNTs from porous carbon blocks to achieve large surface area and high conductivity. Surface active sites can then be introduced onto these structurally optimized carbons by different surface modification techniques to render these carbon substrates with catalytic activities for the target electrochemical reactions.

Apart from the ORR and HER, these non‐noble metal and metal‐free catalysts have also been reported to be effective for other reactions, such as organic synthesis and photocatalysis. The concepts and principles in designing these ORR and HER catalysts may also be transferrable to these reactions, resulting new catalysts and new knowledge.

Acknowledgment

This work is supported by Discovery Early Career Researcher Award (DECRA) scheme (no. DE170100871).

References

  1. 1. Damjanovic, A., Genshaw, M.A., and Bockris, J.O.M. (1967). The mechanism of oxygen reduction at platinum in alkaline solutions with special reference to  H2 O2. J. Electrochem. Soc. 114 (11): 1107–1112.
  2. 2. Nørskov, J.K., Rossmeisl, J., Logadottir, A. et al. (2004). Origin of the overpotential for oxygen reduction at a fuel‐cell cathode. J. Phys. Chem. B 108 (46): 17886–17892.
  3. 3. Jiao, Y., Zheng, Y., Jaroniec, M., and Qiao, S.Z. (2014). Origin of the electrocatalytic oxygen reduction activity of graphene‐based catalysts: a roadmap to achieve the best performance. J. Am. Chem. Soc. 136 (11): 4394–4403.
  4. 4. Marković, N.M., Adžić, R.R., Cahan, B.D., and Yeager, E.B. (1994). Structural effects in electrocatalysis: oxygen reduction on platinum low index single‐crystal surfaces in perchloric acid solutions. J. Electroanal. Chem. 377 (1–2): 249–259.
  5. 5. Dillon, R., Srinivasan, S., Aricò, A.S., and Antonucci, V. (2004). International activities in DMFC R&D: status of technologies and potential applications. J. Power Sources 127 (1–2): 112–126.
  6. 6. Costamagna, P. and Srinivasan, S. (2001). Quantum jumps in the PEMFC science and technology from the 1960s to the year 2000: Part I. Fundamental scientific aspects. J. Power Sources 102 (1–2): 242–252.
  7. 7. Kirubakaran, A., Jain, S., and Nema, R.K. (2009). A review on fuel cell technologies and power electronic interface. Renew. Sust. Energy Rev. 13 (9): 2430–2440.
  8. 8. Steele, B.C.H. and Heinzel, A. (2001). Materials for fuel‐cell technologies. Nature 414 (6861): 345–352.
  9. 9. Girishkumar, G., McCloskey, B., Luntz, A.C. et al. (2010). Lithium−air battery: promise and challenges. J. Phys. Chem. Lett. 1 (14): 2193–2203.
  10. 10. Landsman, D.A. and Luczak, F.J. (2010). Catalyst studies and coating technologies. In: Handbook of Fuel Cells. Wiley.
  11. 11. Wissler, M. (2006). Graphite and carbon powders for electrochemical applications. J. Power Sources 156 (2): 142–150.
  12. 12. Kim, M., Park, J.‐N., Kim, H. et al. (2006). The preparation of Pt/C catalysts using various carbon materials for the cathode of PEMFC. J. Power Sources 163 (1): 93–97.
  13. 13. Santiago, D., Rodríguez‐Calero, G.G., Rivera, H. et al. (2010). Platinum electrodeposition at high surface area carbon Vulcan‐XC‐72R material using a rotating disk‐slurry electrode technique. J. Electrochem. Soc. 157 (12): F189–F195.
  14. 14. Uchida, M., Aoyama, Y., Tanabe, M. et al. (1995). Influences of both carbon supports and heat‐treatment of supported catalyst on electrochemical oxidation of methanol. J. Electrochem. Soc. 142 (8): 2572–2576.
  15. 15. Tang, S., Sun, G., Qi, J. et al. (2010). Review of new carbon materials as catalyst supports in direct alcohol fuel cells. Chin. J. Catal. 31 (1): 12–17.
  16. 16. Antolini, E., Passos, R.R., and Ticianelli, E.A. (2002). Effects of the carbon powder characteristics in the cathode gas diffusion layer on the performance of polymer electrolyte fuel cells. J. Power Sources 109 (2): 477–482.
  17. 17. Kanno, K., Yoon, K.E., Fernandez, J.J. et al. (1994). Effects of carbon black addition on the carbonization of mesophase pitch. Carbon 32 (5): 801–807.
  18. 18. Wang, G., Sun, G., Wang, Q. et al. (2008). Improving the DMFC performance with Ketjen black EC 300J as the additive in the cathode catalyst layer. J. Power Sources 180 (1): 176–180.
  19. 19. Sakanishi, K., Hasuo, H.‐U., Mochida, I., and Okuma, O. (1995). Preparation of highly dispersed NiMo catalysts supported on hollow spherical carbon black particles. Energy Fuel 9 (6): 995–998.
  20. 20. Yu, J., Yoshikawa, Y., Matsuura, T. et al. (2005). Preparing gas‐diffusion layers of PEMFCs with a dry deposition technique. Electrochem. Solid‐State Lett. 8 (3): A152–A155.
  21. 21. Willsau, J. and Heitbaum, J. (1984). The influence of Pt‐activation on the corrosion of carbon in gas diffusion electrodes – a DEMS study. J. Electroanal. Chem. 161 (1): 93–101.
  22. 22. Roen, L.M., Paik, C.H., and Jarvi, T.D. (2004). Electrocatalytic corrosion of carbon support in PEMFC cathodes. Electrochem. Solid‐State Lett. 7 (1): A19–A22.
  23. 23. Stevens, D.A. and Dahn, J.R. (2005). Thermal degradation of the support in carbon‐supported platinum electrocatalysts for PEM fuel cells. Carbon 43 (1): 179–188.
  24. 24. Kinoshita, K. (1988). Carbon: Electrochemical and Physicochemical Properties. Wiley.
  25. 25. Thomas, J.M. (1970). Reactivity of carbon: some current problems and trends. Carbon 8 (4): 413–421.
  26. 26. Kinoshita, K. and Bett, J. (1973). Electrochemical oxidation of carbon black in concentrated phosphoric acid at 135 °C. Carbon 11 (3): 237–247.
  27. 27. Li, H.‐Q., Wang, Y.‐G., Wang, C.‐X., and Xia, Y.‐Y. (2008). A competitive candidate material for aqueous supercapacitors: high surface‐area graphite. J. Power Sources 185 (2): 1557–1562.
  28. 28. Hung, C.‐C., Lim, P.‐Y., Chen, J.‐R., and Shih, H.C. (2011). Corrosion of carbon support for PEM fuel cells by electrochemical quartz crystal microbalance. J. Power Sources 196 (1): 140–146.
  29. 29. Coloma, F., Sepulvedaescribano, A., and Rodriguezreinoso, F. (1995). Heat‐treated carbon‐blacks as supports for platinum catalysts. J. Catal. 154 (2): 299–305.
  30. 30. Geim, A.K. and Novoselov, K.S. (2007). The rise of graphene. Nat. Mater. 6 (3): 183–191.
  31. 31. Chen, X., Zhang, S., Dikin, D.A. et al. (2003). Mechanics of a carbon nanocoil. Nano Lett. 3 (9): 1299–1304.
  32. 32. Iijima, S. (2002). Carbon nanotubes: past, present, and future. Phys. B: Condens. Matter 323 (1–4): 1–5.
  33. 33. Iijima, S., Yudasaka, M., Yamada, R. et al. (1999). Nano‐aggregates of single‐walled graphitic carbon nano‐horns. Chem. Phys. Lett. 309 (3–4): 165–170.
  34. 34. Lordi, V., Yao, N., and Wei, J. (2001). Method for supporting platinum on single‐walled carbon nanotubes for a selective hydrogenation catalyst. Chem. Mater. 13 (3): 733–737.
  35. 35. Antolini, E. (2009). Carbon supports for low‐temperature fuel cell catalysts. Appl. Catal., B 88 (1–2): 1–24.
  36. 36. Hiura, H., Ebbesen, T.W., and Tanigaki, K. (1995). Opening and purification of carbon nanotubes in high yields. Adv. Mater. 7 (3): 275–276.
  37. 37. Prabhuram, J., Zhao, T.S., Tang, Z.K. et al. (2006). Multiwalled carbon nanotube supported PtRu for the anode of direct methanol fuel cells. J. Phys. Chem. B 110 (11): 5245–5252.
  38. 38. Saha, M.S., Li, R., and Sun, X. (2008). High loading and monodispersed Pt nanoparticles on multiwalled carbon nanotubes for high performance proton exchange membrane fuel cells. J. Power Sources 177 (2): 314–322.
  39. 39. Liang, Y., Wang, H., Diao, P. et al. (2012). Oxygen reduction electrocatalyst based on strongly coupled cobalt oxide nanocrystals and carbon nanotubes. J. Am. Chem. Soc. 134 (38): 15849–15857.
  40. 40. Deng, J., Yu, L., Deng, D. et al. (2013). Highly active reduction of oxygen on a FeCo alloy catalyst encapsulated in pod‐like carbon nanotubes with fewer walls. J. Mater. Chem. A 1 (47): 14868–14873.
  41. 41. Deng, D., Yu, L., Chen, X. et al. (2013). Iron encapsulated within pod‐like carbon nanotubes for oxygen reduction reaction. Angew. Chem. Int. Ed. 52 (1): 371–375.
  42. 42. Wang, S., Yu, D., and Dai, L. (2011). Polyelectrolyte functionalized carbon nanotubes as efficient metal‐free electrocatalysts for oxygen reduction. J. Am. Chem. Soc. 133 (14): 5182–5185.
  43. 43. Rodriguez, N.M., Kim, M.S., and Baker, R.T.K. (1993). Promotional effect of carbon monoxide on the decomposition of ethylene over an iron catalyst. J. Catal. 144 (1): 93–108.
  44. 44. De Jong, K.P. and Geus, J.W. (2000). Carbon nanofibers: catalytic synthesis and applications. Catal. Rev. 42 (4): 481–510.
  45. 45. Lee, S.‐H., Park, S.‐M., and Kim, Y. (2007). Effect of the concentration of sodium acetate (SA) on crosslinking of chitosan fiber by epichlorohydrin (ECH) in a wet spinning system. Carbohydr. Polym. 70 (1): 53–60.
  46. 46. Bessel, C.A., Laubernds, K., Rodriguez, N.M., and Baker, R.T.K. (2001). Graphite nanofibers as an electrode for fuel cell applications. J. Phys. Chem. B 105 (6): 1115–1118.
  47. 47. Wu, N., Wang, Y., Lei, Y. et al. (2015). Electrospun interconnected Fe‐N/C nanofiber networks as efficient electrocatalysts for oxygen reduction reaction in acidic media. Sci. Rep. 5: 17396.
  48. 48. Kim, M., Nam, D.‐H., Park, H.‐Y. et al. (2015). Cobalt‐carbon nanofibers as an efficient support‐free catalyst for oxygen reduction reaction with a systematic study of active site formation. J. Mater. Chem. A 3 (27): 14284–14290.
  49. 49. Edwards, R.S. and Coleman, K.S. (2013). Graphene synthesis: relationship to applications. Nanoscale 5 (1): 38–51.
  50. 50. Zhu, Y., Murali, S., Cai, W. et al. (2010). Graphene and graphene oxide: synthesis, properties, and applications. Adv. Mater. 22 (35): 3906–3924.
  51. 51. 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.
  52. 52. Wang, G., Wang, B., Park, J. et al. (2009). Highly efficient and large‐scale synthesis of graphene by electrolytic exfoliation. Carbon 47 (14): 3242–3246.
  53. 53. Blake, P., Brimicombe, P.D., Nair, R.R. et al. (2008). Graphene‐based liquid crystal device. Nano Lett. 8 (6): 1704–1708.
  54. 54. Hernandez, Y., Nicolosi, V., Lotya, M. et al. (2008). High‐yield production of graphene by liquid‐phase exfoliation of graphite. Nat. Nanotechnol. 3 (9): 563–568.
  55. 55. Park, S. and Ruoff, R.S. (2009). Chemical methods for the production of graphenes. Nat. Nanotechnol. 4 (4): 217–224.
  56. 56. Schniepp, H.C., Li, J.‐L., McAllister, M.J. et al. (2006). Functionalized single graphene sheets derived from splitting graphite oxide. J. Phys. Chem. B 110 (17): 8535–8539.
  57. 57. Li, X., Cai, W., An, J. et al. (2009). Large‐area synthesis of high‐quality and uniform graphene films on copper foils. Science 324 (5932): 1312–1314.
  58. 58. Chen, Z., Ren, W., Gao, L. et al. (2011). Three‐dimensional flexible and conductive interconnected graphene networks grown by chemical vapour deposition. Nat. Mater. 10 (6): 424–428.
  59. 59. Kou, R., Shao, Y., Wang, D. et al. (2009). Enhanced activity and stability of Pt catalysts on functionalized graphene sheets for electrocatalytic oxygen reduction. Electrochem. Commun. 11 (5): 954–957.
  60. 60. Seo, M.H., Choi, S.M., Kim, H.J., and Kim, W.B. (2011). The graphene‐supported Pd and Pt catalysts for highly active oxygen reduction reaction in an alkaline condition. Electrochem. Commun. 13 (2): 182–185.
  61. 61. Guo, S. and Sun, S. (2012). FePt nanoparticles assembled on graphene as enhanced catalyst for oxygen reduction reaction. J. Am. Chem. Soc. 134 (5): 2492–2495.
  62. 62. Guo, S., Zhang, S., Wu, L., and Sun, S. (2012). Co/CoO nanoparticles assembled on graphene for electrochemical reduction of oxygen. Angew. Chem. Int. Ed. 51 (47): 11770–11773.
  63. 63. Liang, Y., Li, Y., Wang, H. et al. (2011). Co3O4 nanocrystals on graphene as a synergistic catalyst for oxygen reduction reaction. Nat. Mater. 10 (10): 780–786.
  64. 64. Liang, Y., Wang, H., Zhou, J. et al. (2012). Covalent hybrid of spinel manganese–cobalt oxide and graphene as advanced oxygen reduction electrocatalysts. J. Am. Chem. Soc. 134 (7): 3517–3523.
  65. 65. Wu, Z.‐S., Yang, S., Sun, Y. et al. (2012). 3D nitrogen‐doped graphene aerogel‐supported Fe3O4 nanoparticles as efficient electrocatalysts for the oxygen reduction reaction. J. Am. Chem. Soc. 134 (22): 9082–9085.
  66. 66. Lee, J., Kim, J., and Hyeon, T. (2006). Recent progress in the synthesis of porous carbon materials. Adv. Mater. 18 (16): 2073–2094.
  67. 67. 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.
  68. 68. Meng, Y., Gu, D., Zhang, F. et al. (2005). Ordered mesoporous polymers and homologous carbon frameworks: amphiphilic surfactant templating and direct transformation. Angew. Chem. 117 (43): 7215–7221.
  69. 69. Fang, Y., Lv, Y., Che, R. et al. (2013). Two‐dimensional mesoporous carbon nanosheets and their derived graphene nanosheets: synthesis and efficient lithium ion storage. J. Am. Chem. Soc. 135 (4): 1524–1530.
  70. 70. Li, M., Ding, J., and Xue, J. (2013). Mesoporous carbon decorated graphene as an efficient electrode material for supercapacitors. J. Mater. Chem. A 1 (25): 7469–7476.
  71. 71. Fang, Y., Gu, D., Zou, Y. et al. (2010). A low‐concentration hydrothermal synthesis of biocompatible ordered mesoporous carbon nanospheres with tunable and uniform size. Angew. Chem. Int. Ed. 49 (43): 7987–7991.
  72. 72. Schuster, J., He, G., Mandlmeier, B. et al. (2012). Spherical ordered mesoporous carbon nanoparticles with high porosity for lithium–sulfur batteries. Angew. Chem. Int. Ed. 51 (15): 3591–3595.
  73. 73. Liang, J., Du, X., Gibson, C. et al. (2013). N‐doped graphene natively grown on hierarchical ordered porous carbon for enhanced oxygen reduction. Adv. Mater. 25 (43): 6226–6231.
  74. 74. Liu, H.‐J., Wang, X.‐M., Cui, W.‐J. et al. (2010). Highly ordered mesoporous carbon nanofiber arrays from a crab shell biological template and its application in supercapacitors and fuel cells. J. Mater. Chem. 20 (20): 4223–4230.
  75. 75. Lyu, Z., Xu, D., Yang, L. et al. (2015). Hierarchical carbon nanocages confining high‐loading sulfur for high‐rate lithium–sulfur batteries. Nano Energy 12: 657–665.
  76. 76. Ryoo, R., Joo, S.H., Kruk, M., and Jaroniec, M. (2001). Ordered mesoporous carbons. Adv. Mater. 13 (9): 677–681.
  77. 77. Choi, W.C., Woo, S.I., Jeon, M.K. et al. (2005). Platinum nanoclusters studded in the microporous nanowalls of ordered mesoporous carbon. Adv. Mater. 17 (4): 446–451.
  78. 78. Liang, J., Zhou, R.F., Chen, X.M. et al. (2014). Fe–N decorated hybrids of CNTs grown on hierarchically porous carbon for high‐performance oxygen reduction. Adv. Mater. 26 (35): 6074–6079.
  79. 79. 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.
  80. 80. Kim, H., Lee, K., Woo, S.I., and Jung, Y. (2011). On the mechanism of enhanced oxygen reduction reaction in nitrogen‐doped graphene nanoribbons. Phys. Chem. Chem. Phys. 13 (39): 17505–17510.
  81. 81. Hu, X., Wu, Y., Li, H., and Zhang, Z. (2010). Adsorption and activation of O2 on nitrogen‐doped carbon nanotubes. J. Phys. Chem. 114: 9603–8607.
  82. 82. Strelko, V.V., Kuts, V.S., and Thrower, P.A. (2000). On the mechanism of possible influence of heteroatoms of nitrogen, boron and phosphorus in a carbon matrix on the catalytic activity of carbons in electron transfer reactions. Carbon 38 (10): 1499–1503.
  83. 83. Nagaiah, T.C., Bordoloi, A., Sánchez, M.D. et al. (2012). Mesoporous nitrogen‐rich carbon materials as catalysts for the oxygen reduction reaction in alkaline solution. ChemSusChem 5 (4): 637–641.
  84. 84. Wang, X., Lee, J.S., Zhu, Q. et al. (2010). Ammonia‐treated ordered mesoporous carbons as catalytic materials for oxygen reduction reaction. Chem. Mater. 22 (7): 2178–2180.
  85. 85. Liang, J., Zheng, Y., Chen, J. et al. (2012). Facile oxygen reduction on a three‐dimensionally ordered macroporous graphitic C3N4/carbon composite electrocatalyst. Angew. Chem. 124 (16): 3958–3962.
  86. 86. Zheng, Y., Jiao, Y., Chen, J. et al. (2011). Nanoporous graphitic‐C3N4@carbon metal‐free electrocatalysts for highly efficient oxygen reduction. J. Am. Chem. Soc. 133: 20116–20119.
  87. 87. Yang, D.‐S., Bhattacharjya, D., Inamdar, S. et al. (2012). Phosphorus‐doped ordered mesoporous carbons with different lengths as efficient metal‐free electrocatalysts for oxygen reduction reaction in alkaline media. J. Am. Chem. Soc. 134 (39): 16127–16130.
  88. 88. Wang, H., Bo, X., Zhang, Y., and Guo, L. (2013). Sulfur‐doped ordered mesoporous carbon with high electrocatalytic activity for oxygen reduction. Electrochim. Acta 108: 404–411.
  89. 89. Jiao, Y., Zheng, Y., Davey, K., and Qiao, S.‐Z. (2016). Activity origin and catalyst design principles for electrocatalytic hydrogen evolution on heteroatom‐doped graphene. Nat. Energy 1: 16130.
  90. 90. Zheng, Y., Jiao, Y., Jaroniec, M., and Qiao, S.Z. (2015). Advancing the electrochemistry of the hydrogen‐evolution reaction through combining experiment and theory. Angew. Chem. Int. Ed. 54 (1): 52–65.
  91. 91. Jiao, Y., Zheng, Y., Jaroniec, M., and Qiao, S.Z. (2015). Design of electrocatalysts for oxygen‐ and hydrogen‐involving energy conversion reactions. Chem. Soc. Rev. 44 (8): 2060–2086.
  92. 92. Nørskov, J.K., Bligaard, T., Logadottir, A. et al. (2005). Trends in the exchange current for hydrogen evolution. J. Electrochem. Soc. 152 (3): J23–J26.
  93. 93. Trasatti, S. (1972). Work function, electronegativity, and electrochemical behaviour of metals: III. Electrolytic hydrogen evolution in acid solutions. J. Electroanal. Chem. Interfacial Electrochem. 39 (1): 163–184.
  94. 94. Conway, B.E. and Tilak, B.V. (2002). Interfacial processes involving electrocatalytic evolution and oxidation of H2, and the role of chemisorbed H. Electrochim. Acta 47 (22–23): 3571–3594.
  95. 95. Strmcnik, D., Lopes, P.P., Genorio, B. et al. (2016). Design principles for hydrogen evolution reaction catalyst materials. Nano Energy 29: 29–36.
  96. 96. Sheng, W., Myint, M.N.Z., Chen, J.G., and Yan, Y. (2013). Correlating the hydrogen evolution reaction activity in alkaline electrolytes with the hydrogen binding energy on monometallic surfaces. Energy Environ. Sci. 6 (5): 1509–1512.
  97. 97. Zheng, J., Sheng, W., Zhuang, Z. et al. (2016). Universal dependence of hydrogen oxidation and evolution reaction activity of platinum‐group metals on pH and hydrogen binding energy. Sci. Adv. 2 (3): e1501602.
  98. 98. Huang, X., Zhao, Y., Ao, Z., and Wang, G. (2014). Micelle‐template synthesis of nitrogen‐doped mesoporous graphene as an efficient metal‐free electrocatalyst for hydrogen production. Sci. Rep. 4: 7557.
  99. 99. Sathe, B.R., Zou, X., and Asefa, T. (2014). Metal‐free B‐doped graphene with efficient electrocatalytic activity for hydrogen evolution reaction. Cat. Sci. Technol. 4 (7): 2023–2030.
  100. 100. Ge, J.‐M., Zhang, B., Lv, L.‐B. et al. (2015). Constructing holey graphene monoliths via supramolecular assembly: enriching nitrogen heteroatoms up to the theoretical limit for hydrogen evolution reaction. Nano Energy 15: 567–575.
  101. 101. Qu, K., Zheng, Y., Zhang, X. et al. (2017). Promotion of electrocatalytic hydrogen evolution reaction on nitrogen‐doped carbon nanosheets with secondary heteroatoms. J. Am. Chem. Soc. .
  102. 102. Ito, Y., Cong, W., Fujita, T. et al. (2015). High catalytic activity of nitrogen and sulfur co‐doped nanoporous graphene in the hydrogen evolution reaction. Angew. Chem. Int. Ed. 54 (7): 2131–2136.
  103. 103. Zhou, Y., Leng, Y., Zhou, W. et al. (2015). Sulfur and nitrogen self‐doped carbon nanosheets derived from peanut root nodules as high‐efficiency non‐metal electrocatalyst for hydrogen evolution reaction. Nano Energy 16: 357–366.
  104. 104. Liu, X., Zhou, W., Yang, L. et al. (2015). Nitrogen and sulfur co‐doped porous carbon derived from human hair as highly efficient metal‐free electrocatalysts for hydrogen evolution reactions. J. Mater. Chem. A 3 (16): 8840–8846.
  105. 105. Zheng, Y., Jiao, Y., Li, L., H. et al. (2014). Toward design of synergistically active carbon‐based catalysts for electrocatalytic hydrogen evolution. ACS Nano 8 (5): 5290–5296.
  106. 106. Jiang, H., Zhu, Y., Su, Y. et al. (2015). Highly dual‐doped multilayer nanoporous graphene: efficient metal‐free electrocatalysts for the hydrogen evolution reaction. J. Mater. Chem. A 3 (24): 12642–12645.
  107. 107. Wei, L., Karahan, H.E., Goh, K. et al. (2015). A high‐performance metal‐free hydrogen‐evolution reaction electrocatalyst from bacterium derived carbon. J. Mater. Chem. A 3 (14): 7210–7214.
  108. 108. Zheng, Y., Jiao, Y., Ge, L. et al. (2013). Two‐step boron and nitrogen doping in graphene for enhanced synergistic catalysis. Angew. Chem. 125 (11): 3192–3198.
  109. 109. 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.
  110. 110. 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.
  111. 111. Zhang, J., Zhao, Z., Xia, Z., and Dai, L. (2015). A metal‐free bifunctional electrocatalyst for oxygen reduction and oxygen evolution reactions. Nat. Nanotechnol. 10 (5): 444–452.
  112. 112. Qu, K., Zheng, Y., Dai, S., and Qiao, S.Z. (2016). Graphene oxide‐polydopamine derived N, S‐codoped carbon nanosheets as superior bifunctional electrocatalysts for oxygen reduction and evolution. Nano Energy 19: 373–381.
  113. 113. 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.
  114. 114. LaVoie, M.J., Ostaszewski, B.L., Weihofen, A. et al. (2005). Dopamine covalently modifies and functionally inactivates parkin. Nat. Med. 11 (11): 1214–1221.
  115. 115. Zhong, S., Zhou, L., Wu, L. et al. (2014). Nitrogen‐ and boron‐co‐doped core–shell carbon nanoparticles as efficient metal‐free catalysts for oxygen reduction reactions in microbial fuel cells. J. Power Sources 272: 344–350.
  116. 116. Lee, H., Dellatore, S.M., Miller, W.M., and Messersmith, P.B. (2007). Mussel‐inspired surface chemistry for multifunctional coatings. Science 318 (5849): 426–430.
  117. 117. Della Vecchia, N.F., Avolio, R., Alfè, M. et al. (2013). Building‐block diversity in polydopamine underpins a multifunctional eumelanin‐type platform tunable through a quinone control point. Adv. Funct. Mater. 23 (10): 1331–1340.
  118. 118. Banhart, F., Kotakoski, J., and Krasheninnikov, A.V. (2011). Structural defects in graphene. ACS Nano 5 (1): 26–41.
  119. 119. Tang, C. and Zhang, Q. (2017). Nanocarbon for oxygen reduction electrocatalysis: dopants, edges, and defects. Adv. Mater. 29 (13): 1604103.
  120. 120. Zhang, L., Xu, Q., Niu, J., and Xia, Z. (2015). Role of lattice defects in catalytic activities of graphene clusters for fuel cells. Phys. Chem. Chem. Phys. 17 (26): 16733–16743.
  121. 121. Yan, X., Jia, Y., Odedairo, T. et al. (2016). Activated carbon becomes active for oxygen reduction and hydrogen evolution reactions. Chem. Commun. 52 (52): 8156–8159.
  122. 122. Jia, Y., Zhang, L., Du, A. et al. (2016). Defect graphene as a trifunctional catalyst for electrochemical reactions. Adv. Mater. 28 (43): 9532–9538.
  123. 123. Jia, Y., Zhang, L., Gao, G. et al. (2017). A heterostructure coupling of exfoliated Ni–Fe hydroxide nanosheet and defective graphene as a bifunctional electrocatalyst for overall water splitting. Adv. Mater. 29 (17): 1700017.
  124. 124. Yan, X., Jia, Y., Chen, J. et al. (2016). Defective‐activated‐carbon‐supported Mn–co nanoparticles as a highly efficient electrocatalyst for oxygen reduction. Adv. Mater. 28 (39): 8771–8778.
  125. 125. Zheng, Y., Jiao, Y., and Qiao, S.Z. (2015). Engineering of carbon‐based electrocatalysts for emerging energy conversion: from fundamentality to functionality. Adv. Mater. 27 (36): 5372–5378.
  126. 126. Shinde, S.S., Sami, A., and Lee, J.‐H. (2015). Nitrogen‐ and phosphorus‐doped nanoporous graphene/graphitic carbon nitride hybrids as efficient electrocatalysts for hydrogen evolution. ChemCatChem 7 (23): 3873–3880.
  127. 127. Shinde, S.S., Sami, A., and Lee, J.‐H. (2015). Electrocatalytic hydrogen evolution using graphitic carbon nitride coupled with nanoporous graphene co‐doped by S and Se. J. Mater. Chem. A 3 (24): 12810–12819.
  128. 128. Zhao, Y., Zhao, F., Wang, X. et al. (2014). Graphitic carbon nitride nanoribbons: graphene‐assisted formation and synergic function for highly efficient hydrogen evolution. Angew. Chem. Int. Ed. 53 (50): 13934–13939.
  129. 129. Zheng, Y., Jiao, Y., Zhu, Y. et al. (2014). Hydrogen evolution by a metal‐free electrocatalyst. Nat. Commun. 5: 3783.
  130. 130. Duan, J., Chen, S., Jaroniec, M., and Qiao, S.Z. (2015). Porous C3N4 nanolayers@N‐graphene films as catalyst electrodes for highly efficient hydrogen evolution. ACS Nano 9 (1): 931–940.
  131. 131. Staszak‐Jirkovsky, J., Malliakas, C.D., Lopes, P.P. et al. (2016). Design of active and stable co–Mo–Sx chalcogels as pH‐universal catalysts for the hydrogen evolution reaction. Nat. Mater. 15 (2): 197–203.
  132. 132. Jin, H., Wang, J., Su, D. et al. (2015). In situ cobalt–cobalt oxide/N‐doped carbon hybrids as superior bifunctional electrocatalysts for hydrogen and oxygen evolution. J. Am. Chem. Soc. 137 (7): 2688–2694.
  133. 133. Gong, M., Wang, D.‐Y., Chen, C.‐C. et al. (2016). A mini review on nickel‐based electrocatalysts for alkaline hydrogen evolution reaction. Nano Res. 9 (1): 28–46.
..................Content has been hidden....................

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