11
Heteroatom‐Doped, Carbon‐Supported Metal Catalysts for Electrochemical Energy Conversions

Tanyuan Wang1, Qing Li1,, and Gang Wu2,

1 Huazhong University of Science and Technology, School of Materials Science and Engineering, 1037 Luoyu Road, Wuhan, 430074, China

2 University at Buffalo, The State University of New York, Department of Chemical and Biological Engineering, Flint Road, Buffalo, NY, 14260, USA

11.1 Introduction

Electrochemical energy conversion techniques based on clean and sustainable energy sources are among the hottest research fields due to the increasing energy demands and environmental concerns [13]. To date, the basic principle for the design of energy conversion devices is to achieve the efficient conversion between chemical energy and electricity as electricity is a universal energy currency in our society. Based on it, a variety of devices like fuel cells [4, 5], water electrolyzers [6, 7], metal–air batteries [8, 9], and CO2 reduction systems [10] have been developed. Oxygen reduction reaction (ORR), oxygen evolution reaction (OER), hydrogen evolution reaction (HER), hydrogen oxidation reaction (HOR), methanol oxidation reaction (MOR), and CO2 reduction reaction (CO2RR) are the most important electrochemical reactions involved in these electrochemical energy conversion devices and determine the efficiency of them. Even though they possess different reaction mechanisms, they suffer from one common trouble with sluggish kinetics and concurrent large overpotentials, which greatly limit their performance. To solve this issue, well‐designed electrocatalysts with excellent activity, durability, and affordable cost are required to push forward the reactions and improve the efficiency of the devices.

Generally, there are three groups of electrocatalysts for the above‐mentioned reactions: (i) precious metal‐based materials such as Pt and Pt alloys [1113]; (ii) non‐precious metal materials such as non‐noble metal oxides [14, 15], sulfides [16, 17], and metal–nitrogen–carbon (M–N–C) catalysts [1820]; and (iii) metal‐free materials like heteroatom‐doped carbons [21, 22]. Usually, metal‐containing catalysts, especially in nanometer sizes, show better activity relative to their metal‐free counterparts due to the existence of inherently more active metal species. In most cases, nanocatalysts consist of metals that have to be loaded on supports to avoid dissolution, migration, and/or aggregation of metal nanostructures (spheres, rods, wires, cages, frames, etc.). Among various supporting materials, carbon‐based materials draw special attention due to their excellent electrical and physical properties, large specific surface area, and low cost [1, 2326]. Recently, heteroatom (N, B, S, and P)‐doped carbon nanomaterials are intensively investigated as viable supports for the metal nanostructures because of their ability to tune the electronic and geometric properties of carbon. The interaction between heteroatom and metal and the high conductivity and surface area of the doped carbon‐based materials are suggested to significantly enhance the catalytic performance of the hybrid structures. In this chapter, we will mainly focus on the structural tuning and catalytic performance of heteroatom‐doped, carbon‐supported metal catalysts. Metal nanostructures supported on N, B, S, P, and multi‐heteroatom‐doped carbon nanomaterials and their catalytic properties for ORR, OER, HER, HOR, CO2RR, and other reactions will be introduced systematically. The synthesis–structure–property correlations of the catalysts will be discussed. Finally, a brief conclusion and an outlook on the development of the heteroatom‐doped, carbon‐supported metal catalysts will be proposed.

11.2 N‐Doped, Carbon‐Supported Metal Catalysts

11.2.1 Design and Synthesis

Nitrogen is the mostly researched doping atom for carbon materials as it has atom size similar to carbon and can bond strongly with carbon. Because N is more electronegative than carbon, it would break the charge neutrality of the pure carbon structure after doping and result in electron deficiency for the near carbon, thereby tuning the catalytic performance of the materials [27, 28]. Moreover, N can be doped in various locations in the carbon lattice, thus leading to multiple possible configurations [29, 30]. As shown in Figure 11.1, N can exist as graphitic, pyridinic, pyrrolic, amino, and oxidized type in the doped carbon materials, and they can be distinguished by X‐ray photoelectron spectroscopy (XPS) [31]. In particular, pyridinic nitrogen is obtained by doping at the edge of the graphene layer, contributing one pπ electron to the graphitic π system. Meanwhile, graphitic nitrogen is the result of in‐plane doping and contributes two pπ electrons. Among the N‐bonding structures, graphitic‐N and pyridinic‐N have received more attention as they are suggested to dominate the electrocatalytic behaviors of the N‐doped carbons [24, 32].

Image described caption and surrounding text.

Figure 11.1 The N 1s XPS spectrum of N‐doped carbon (a). Schematic illustration for various “N” atoms represents the pyridinic‐N, pyrrolic‐N, graphitic‐N, and amino group in N‐doped carbon materials (b).

Source: Zhang et al. 2013 [31]. Copyright 2013. Elsevier.

The enhanced electrocatalytic performance of the N‐doped carbon, abundance of active sites and defects, makes it an ideal support for metal nanocatalysts compared with traditional undoped carbon materials. Specifically, the interaction between metal and the N‐center or N‐induced defect in the carbon structure would possibly improve the electrocatalytic performance of the materials [33]. The enhanced metal–supports interaction between metal and the N‐doped carbon would improve the stability of the catalysts [34]. This interaction may also promote the charge transfer at the metal–supports interface, thus further enhancing the catalytic activity of the N‐doped, carbon‐supported metal structure [35]. In addition, the CN sites are proved to be the preferred sites for the growth of the metal nanostructures [36], and single metal atom can even be stabilized by the N‐doped carbon structures [37, 38], which offers promising advantages for the design of novel N‐doped, carbon‐supported metal electrocatalysts compared with the traditional carbon‐supported metal.

The strategies for the synthesis of N‐doped, carbon‐supported metal catalysts can be generally categorized into three ways. First is a two‐step, post‐loading method. In this method, N‐doped carbon is first fabricated and then the metal nanostructures can be loaded on the N‐doped carbon by either in situ nucleation or self‐assembly of the prepared metal nanostructures [3942]. Second is the one‐step method in which N‐doping and nucleation of metal nanoparticles can be realized via one‐pot solvothermal synthesis [25, 43]. Usually, such a process is difficult to control the size and shape of metal nanostructures. The third way is the direct pyrolysis of N, metal, and carbon precursors to fabricate M–N–C catalysts [4446]. In this catalyst, metal atoms are embedded into the N‐doped carbon matrix to form M–Nx‐active sites, which will be discussed in detail in the following sections.

The key factor for the two‐step, post‐loading method is the synthesis of N‐doped carbons. N‐doped carbons can be fabricated by chemical vapor deposition (CVD) [47, 48] or pyrolysis [49, 50] of C and N precursors, arc discharge [51], or plasma treatment [52, 53] of carbon materials in N precursor such as NH3, thermal annealing [54, 55], or solvothermal treatment [56] of carbon nanostructure like graphene oxide (GO) with N precursor. Under these reacting conditions, N would bond with C and doped in the lattices of carbon planner. And the N precursor is suggested to play a key role in the bonding configuration of the N‐doped structures. For instance, Ruoff and coworkers prove that annealing of GO with polyaniline (PANI) would result in the generation of pyridinic‐N, while annealing of GO with polypyrrole (PPy) would tend to form pyrrolic‐N. If NH3 is used as the precursor, graphitic‐N and pyridinic‐N would be obtained for the doped structure [57].

Either metal nanoparticles or metal compounds can be supported on the N‐doped carbon. Chemical reduction of metal salts in the presence of N‐doped carbon materials is the mostly used strategy to support noble metal nanoparticles on N‐doped carbon as noble metal ions such as PtCl62− can be easily reduced to metal [1, 39]. N‐doped graphene and N‐doped carbon nanotubes (N‐CNTs) are good supports for noble metals [58, 59]. During the synthesis, the noble metal salts are first dispersed in these N‐doped carbon materials or in a solution than containing N‐doped carbon materials. Then H2, NaBH4, or some other reduction regents are used to reduce the metal ions to zero valence and produce N‐doped, carbon‐supported precious metal nanoparticles [6062]. The noble metal nanoparticles are more easily anchored on the N sites of the doped structure due to the metal–N coordination, which may stabilize the metal nanoparticles and benefit the catalytic activity and stability of the hybrid materials [33, 63]. In particular, the pyridinic edge sites, which carry an extra lone‐pair of electrons, may provide stronger interaction between the metal and the support, thus inhibiting agglomeration of the nanoparticles. Additionally, such interactions may also induce the alterations of the electronic structure of the catalyst. Arrigo et al. prove that the nature of the interaction between Pd and the N species in N‐CNT is of σ‐type donation from the filled π orbital of the N atom to the empty d orbital of the Pd atom and a π back‐donation from the filled Pd atomic d orbital to the π*‐anti‐bonding orbital of the N atom (Figure 11.2) [64], which offers deeper understanding for the rational design of N‐doped, carbon‐supported metal nanocatalysts.

Schematic diagram of chemical structures depicting σ-bond between an empty metal d orbital (negative lobe in blue and positive lobe in red) and a filled Π orbital (negative lobe in pale blue and positive lobe in purple) from he support (a), Π-back bond between filled metal d-orbital 4w5w532 and empty Π*-anti-bonding orbital of the support (b).

Figure 11.2 σ‐bond between an empty metal d orbital (negative lobe in pale and positive lobe in black) and a filled π orbital (negative lobe in bright gray and positive lobe in gray) from the support (a), π‐back bond between filled metal d‐orbital and empty π*‐anti‐bonding orbital of the support (b).

Source: Arrigo et al. 2015 [64]. Copyright 2015. American Chemical Society.

Different from precious nanoparticles, transition metal compounds are usually loaded on N‐doped carbon materials by solvothermal processes. Dai and coworkers developed a series of N‐doped, carbon (graphene or carbon nanotube)‐supported cobalt oxides with carbon nanomaterials, NH4OH, and cobalt acetate as the precursors [25, 65]. Cobalt acetate was mixed with carbon nanomaterials in ethanol, followed by the addition of NH4OH. Then, the mixture was heated to a certain temperature to obtain N‐doped, carbon‐supported cobalt oxide nanoparticles. They found that post‐annealing of N‐doped, carbon‐supported Co3O4 in NH3 would result in N‐CNT‐supported CoO that showed enhanced ORR performance (Figure 11.3). N‐doped, carbon‐supported nickel sulfide was prepared by the reaction of Ni2(CO3)(OH)2 and thioacetamide [66]. During the synthesis, Ni2(CO3)(OH)2 was first dispersed in water with N‐doped graphene, then thioacetamide was added, and the solution was heated at 200 °C for 12 h to obtain N‐doped, graphene‐supported Ni3S4. In addition, N‐doped, carbon‐supported metal can also be fabricated by simply loading as‐synthesized nanoparticles on N‐doped carbon [40].

TEM image (a), cartoon of CoO/N-CNT hybrid (b), and ORR polarization curves of CoO/N-CNT catalyst (c). The TEM image has an arrow pointing to CoO on CNTs.

Figure 11.3 TEM image (a), cartoon of CoO/N‐CNT hybrid (b), and ORR polarization curves of CoO/N‐CNT catalyst (c).

Source: Liang et al. 2012 [65]. Copyright 2012. American Chemical Society.

The direct pyrolysis of N, metal (Fe, Co, Ni, Mn, etc.), and carbon precursors is commonly used to prepare M–N–C type of N‐doped, carbon‐supported metal catalysts [44, 6769]. In this process, the metal atoms would coordinate strongly with N to form M–N–C structure. Therefore, macrocycle complexes become ideal precursors for this process as they have natural metal–N4 centers. Later study proved that pyrolyzing a mixture of metal salts and N/C sources could produce similar M–N–C structure [70, 71]. A variety of N‐containing molecules including cyanamide, 1,10‐phenanthroline, and PANI are employed as the N precursors, whereas inorganic metal salts such as chlorides and acetates are investigated as metal precursors. Zelenay and coworkers synthesized N‐doped, carbon‐supported iron catalysts with PANI as the N sources [72]. They found that the formation of FeN center would be influenced by the reaction temperature, and pyrolysis at 900 °C would lead to the largest FeN content. Moreover, they proved that the presence of sulfur‐based oxidant in the aniline polymerization could benefit the formation of FeN structure, which offered a new idea for the design of highly efficient, N‐doped, carbon‐supported metal catalysts. Recently, the metal–organic framework (MOF)‐derived carbon nanostructures formed by pyrolyzing MOFs become a new group of N‐doped, carbon‐supported metal catalysts and attract considerable attention [19, 20, 7375]. MOFs hold large surface areas, controllable compositions, and structures and can achieve high conductivity after pyrolysis. Importantly, the metal ions can coordinate with N in MOFs and a uniform distribution of metal, N, and C could be achieved, which would contribute to their catalytic activity after pyrolysis. Zeolitic imidazolate frameworks (ZIFs) have been investigated as ideal precursors because they are rich in M–N coordinations and can be easily prepared at low cost [76, 77]. ZIFs are usually prepared by the coordination between tetrahedral metal ions like Zn, Co, and imidazolate compounds. Bimetallic zeolitic imidazolate frameworks (BMZIFs) can also be prepared using two kinds of metals. Jiang and coworkers synthesized Co/Zn BMZIFs with varied Co:Zn ratios by adding 2‐methylimidazole in methanol that contains Zn(NO3)2·6H2O and Co(NO3)2·6H2O [78]. Then, they heated the BMZIFs at 900 °C in N2 atmosphere to obtain N‐doped, carbon‐supported Co (CNCo). The BMZIF‐derived CNCo demonstrated enhanced ORR performance compared with that obtained by the pyrolysis of ZIF‐8 and ZIF‐67, and its ORR activity can be further improved by P doping. Figure 11.4 displays the scanning electron microscopy (SEM) and transmission electron microscopy (TEM) images of BMZIF and CNCo. The CNCo particles keep the polyhedron structure of ZIF after pyrolysis. It is of note that Co nanoparticles are also formed during pyrolysis in addition to Co–N–C structures (Figure 11.4e).

Image described caption and surrounding text.

Figure 11.4 Scanning electron microscopy (SEM) images of (a) BMZIF and (b) CNCo. (c) Transmission electron microscopy (TEM) image of CNCo and the corresponding elemental mapping of C, N, and Co. (d, e) Enlarged TEM images of CNCo (inset in (e): high‐resolution transmission electron microscopy (HRTEM) image showing Co nanoparticle wrapped by the well‐developed graphitization layers).

Source: Chen et al. 2015 [78]. Copyright 2015. John Wiley & Sons.

11.2.2 N‐Doped, Carbon‐Supported Metal Electrocatalysts

N‐doped, carbon‐supported metal electrocatalysts have been under intensive investigations during recent years due to the combination of advanced carbon materials with metal‐active sites. The high surface area, good electronic conductivity, and abundant defects induced by N‐doping make various N‐doped carbons, e.g. graphene and carbon nanotube, promising supporting materials for metal nanocatalysts. A large variety of metal, metal oxides, metal sulfides, and many other metal‐based nanostructures have been supported on N‐doped carbons as catalysts [7988]. Their performance in catalyzing relevant electrochemical reactions including ORR, OER, HER, and CO2RR will be discussed in detail in this section.

11.2.2.1 Oxygen Electrocatalysis

Noble metal nanoparticles and transition metal compounds can be supported on N‐doped carbons to catalyze the electrochemical reduction or evolution of oxygen. N‐doped carbon is suggested to be a good support for Pt and Pt alloys when used in ORR electrocatalysis. Ramaprabhu and coworkers loaded Pt and Pt3Co nanoparticles on N‐doped, hydrogen‐exfoliated graphene (N‐HEG) by the reduction of H2PtCl6·6H2O and Co(NO3)2·6H2O in ethylene glycol/water mixture (Figure 11.5a–d) [79]. The N‐HEG‐supported Pt3Co exhibits enhanced ORR activity and four times higher power density than commercial Pt/C in proton exchange membrane fuel cells (Figure 11.5e,f), which should be attributed to the excellent dispersion of Pt3Co nanoparticles and the instinct ORR activity of the N‐doped graphene. The Pt nanowires can also be grown on N‐doped carbon in addition to Pt nanoparticles [89]. Using HCOOH as the reduction regent under room temperature, H2PtCl6·6H2O can be reduced to Pt nanowire with a diameter of 2.5 nm on N‐CNTs, which would exhibit enhanced ORR activity.

Image described caption and surrounding text.

Figure 11.5 SEM and TEM images of Pt/N‐HEG (a, c) and Pt3Co/N‐HEG (b, d). Insets in (c, d) show the electron diffraction patterns of Pt/N‐HEG and Pt3Co/N‐HEG. The scale bars correspond to 5 nm−1. (e) Theoretical fit of the polarization data for Pt/C, Pt3Co/C, Pt/N‐HEG, and Pt3Co/N‐HEG at 60 °C. (f) Stability study of the proton‐exchange‐membrane fuel cell (PEMFC) with cathode electrocatalyst containing Pt/N‐HEG and Pt3Co/N‐HEG at 60 °C without any back pressure.

Source: Vinayan et al. 2012 [79]. Copyright 2012, John Wiley & Sons.

In addition to precious metal‐based materials, N‐doped, carbon‐material–supported, non‐precious metal compound can also promote the reduction of oxygen. N‐doped, carbon‐supported cobalt oxides have been proved to be efficient ORR catalysts in an alkaline solution [25, 90]. Manganese oxide that supported on the mesoporous N‐doped carbon developed by the hydrothermal treatment of Mn salts and aniline followed by heat treatment can also catalyze the ORR efficiently [91]. Efficient OER electrocatalysts can also be designed with N‐doped, carbon‐supported transition metal (Fe, Co, Ni, and Mn) compounds [92]. Wang and coworkers fabricate a cobalt–cobalt oxide/N‐doped carbon hybrids (CoOx@CN) by a one‐pot thermal treatment method [93]. CoNO3·6H2O was first mixed with D(+)‐glucosamine hydrochloride and melamine in water. Then, the mixture was dried and calcined at 800 °C in N2 atmosphere for 1 h to obtain CoOx@CN. The fabricated hybrid demonstrated high OER activity with an overpotential of 26 mV to achieve the current density of 10 mA cm−2. N‐CNT‐supported NiCo2S4 (NiCo2S4/N‐CNT) is also synthesized to promote the reaction of oxygen reduction and oxygen evolution simultaneously [87]. The NiCo2S4/N‐CNT prepared by the solvothermal treatment of Ni(II), Co(II), and thiourea in the presence of CNT demonstrated comparable ORR activity to Pt/C and superior OER capability to RuO2, thus could reduce the charge‐discharge polarization (∼0.63 V) significantly. The interaction between sulfide spinels and N‐CNT and the homogenous distribution of NiCo2S4 nanocrystals are suggested to contribute to the high ORR/OER activity of NiCo2S4/N‐CNT.

11.2.2.2 HER Electrocatalysis

Recently, transition metal compounds including chalcogenides, phosphides, carbides, and nitrides have been studied as earth‐abundant and efficient HER catalysts [9496]. Therefore, the HER performance of the N‐doped, carbon‐supported metal nanomaterials have also been evaluated. Kim and coworkers prepared N‐CNT supported MoSx (MoSx/NCNT) with (NH4)2MoS4 and N‐CNT as the reactants (Figure 11.6) [97]. After being treated with HCl, amorphous MoSx would be decorated on the N‐CNT. The MoSx/NCNT shows excellent HER catalytic performance with an onset overpotential of 75 mV. Its Tafel slope is 40 mV dec−1 and can reach the current density of 10 mA cm−2 at the overpotential of 110 mV. The N‐CNT possesses more favorable interaction with precursor molecules compared with pristine CNT. This interaction and the amorphous structure of the MoSx are suggested to benefit the HER performance of the hybrid. After being annealed at 600 °C, MoSx would turn to crystal MoS2 (Figure 11.6g) and demonstrate inferior HER activity. You and coworkers synthesized MoS2‐nanosheets‐embedded, N‐doped carbon nanofibers by electrospinning of Na2MoO4 and polyacrylonitrile followed by carbonization at 900 °C [98]. This hybrid also shows enhanced HER activity with a Tafel slope of 48 mV dec−1 and an overpotential of 135 mV at the current density of 10 mA cm−2. In addition, molybdenum selenide and tungsten oxynitride can also exhibit improved HER activity when supported on N‐doped carbon [99, 100].

Image described caption and surrounding text.

Figure 11.6 MoSx/NCNT forest hybrid catalyst. (a) Schematic illustration of three‐dimensional hybrid catalyst synthesis. (b) Field emission scanning electron microscopy (FE‐SEM) image of bare NCNT forest. (c) FE‐SEM image of MoSx/NCNT forest hybrid. (d) Broad‐field FE‐SEM image of (c). (e) TEM image of bare NCNT. (f) TEM image of MoSx/NCNT. (g) TEM image of MoSx/NCNT after thermal annealing at 600 °C.

Source: Li et al. 2014 [97]. Copyright 2014. American Chemical Society.

11.2.2.3 Other Electrocatalysis

PtRu/C is a commercially available catalyst for MOR, which is the anode reaction of direct methanol fuel cells (DMFCs). Similar to those observed in oxygen reduction electrocatalysis, N‐doped, carbon‐supported metal nanoparticles catalysts draw special attention in MOR due to the improved interaction between noble metal and the N‐doped substrates [39, 70]. The noble metal nanoparticles would usually show improved catalyst dispersion, lower onset potential, and increased current density when supported on N‐doped carbon [101103]. Moreover, some researchers suggest that the interaction between N and noble metal such as Pt will result in positive shift for its d‐band center, thereby increasing the chemisorptions of oxygen‐containing groups (e.g. OHads) and accelerating the CO oxidation during MOR [104]. O'Hayre and coworkers dispersed PtRu nanoparticles on N‐doped carbon via magnetron sputtering and proved that the N‐doped carbon could yield enhanced stability and performance of PtRu catalyst through metal–N interactions [105, 106]. Su et al. supported Pt nanoparticles by the ethylene glycol reduction method with H2PtCl6 as the precursor [61]. The obtained hybrid also showed enhanced mass activity compared with commercial Pt.

N‐doped, carbon‐supported metal nanoparticle can also be applied in the electrochemical reduction of CO2, which is critical for relieving the over‐produced CO2 and convert CO2 into a reusable form of carbons (e.g. CO, CH4, and C2H4). We synthesized uniformly distributed Cu nanoparticles with a diameter of 7 nm and assembled them on pyridinic‐N‐rich graphene (p‐NG) [40]. The p‐NG‐supported 7 nm Cu (p‐NG‐Cu‐7) exhibited high selectivity of 19% for the production of C2H4 at −0.9 V (versus reversible hydrogen electrode (RHE)) when used in the reduction of CO2 (Figure 11.7). The pyridinic‐N is suggested to function as CO2 and proton absorber, thus facilitating the hydrogenation and carbon–carbon coupling reactions on Cu nanoparticles.

Line graphs depicting  E (V versus RHE) versus Faradaic efficiency (a) with curves for CH4; C2H4; C2H6; formate. Bar graphs (b) with Selectivity on the y-axis depicting Formate and C2H4 for -0.8 V and CH4; C2H4; C2H6; formate for -0.9 V. Comparative bar graphs depicting Mass activity (c) and Selectivity (d) on the y-axes with bars for P-NG-Cu-7, GO-Cu-7, C-Cu-7 (c) and Ethylene, Other hydrocarbons (d).

Figure 11.7 (a) Reduction of potential‐dependent Faradic efficiency of the p‐NG‐Cu‐7‐catalyzed electrochemical reduction of CO2 to various hydrocarbons. (b, c) Product selectivity of the hydrocarbons generated from the p‐NG‐Cu‐7‐catalyzed reduction at (b) −0.8 V and (c) −0.9 V. (d) Comparison of mass activity of the p‐NG‐Cu‐7, C‐Cu‐7, and GO‐Cu‐7‐catalyzed reduction of CO2 to C2H4 at various reduction potentials. (e) Selectivity of C2H4 and all other hydrocarbons generated from the p‐NG‐Cu‐7‐catalyzed reduction of CO2 at various potentials.

Source: Li et al. 2016 [40]. Copyright 2016. Elsevier.

11.2.3 Metal–Nitrogen–Carbon Catalysts for Electrocatalysis

M–N–C structures are generally investigated as ORR electrocatalysts due to their exposed M–Nx center. Different from N‐doped carbon and transition metal compounds that usually work in alkali, M–N–C structures can demonstrate high ORR activity in not only alkali but also in acid with good stability. A variety of nitrogen precursors and transition metal salts along with carbon sources can be used to prepare M–N–C catalysts [44, 107, 108]. Generally, the mixture is heated at 700–1000 °C and acid leaching and second‐heat treatment are required to remove the undesired species and improve the ORR activity of the catalysts. PANI is a commonly studied N source for M–N–C structures. With Fe salt, Co salt, and PANI as the precursors, we fabricate high active ORR catalysts (PANI–FeCo–C) that can catalyze the ORR at potentials within 60 mV of that delivered by commercial Pt/C (Figure 11.8) [109]. The aniline was polymerized in 0.5 M HCl with the addition of ammonium peroxydisulfate as well as FeCl3 and Co(NO3)2·6H2O in the presence of carbon. After the complete polymerization of PANI, the mixture was dried and heated in N2 atmosphere. The heat‐treated sample was then pre‐leached in 0.5 M H2SO4 and re‐heated in N2 to obtain the final sample. Figure 11.8a,b proves that metal‐based nanoparticle still exists in the M–N–C structure. Compared with PANI–Fe–C, PANI–FeCo–C demonstrated enhanced ORR activity (Figure 11.8c,d). The heating temperature and steps are found to have significant influence on the ORR catalytic behavior of PANI‐derived catalysts [109111]. The PANI–Fe–C structure would display the most outstanding ORR activity after being treated at 900 °C.

Image described caption and surrounding text.

Figure 11.8 (a) TEM and (b) SEM images of a PANI‐FeCo‐C(1) catalyst. (c) Steady‐state ORR polarization plots (bottom) and H2O2 yield plots (top) measured with different PANI‐derived catalysts and reference materials: (1) as‐received carbon black (Ketjenblack EC‐300J); (2) heat‐treated carbon black; (3) heat‐treated PANI‐C; (4) PANI‐Co‐C; (5) PANI‐FeCo‐C(1); (6) PANI‐FeCo‐C(2); (7) PANI‐Fe‐C; and (8) E‐TEK Pt/C (20 μgPt cm−2). Electrolyte: O2‐saturated 0.5 M H2SO4 (0.1 M HClO4 in an experiment involving Pt catalysts (dashed lines)); temperature, 25 °C. Rotating ring disk electrode (RRDE) experiments were carried out at a constant ring potential of 1.2 V versus RHE. Rotating disk electrode (RDE)/RRDE rotating speed, 900 rpm; non‐precious metal catalyst loading, 0.6 mg cm−2. (d) Steady‐state ORR polarization plots (bottom) and H2O2 yield plots (top) measured with a PANI–Fe–C catalyst in 0.5 M H2SO4 electrolyte as a function of the heat treatment temperature: (1) 400 °C; (2) 600 °C; (3) 850 °C; (4) 900 °C; (5) 950 °C; and (6) 1000 °C.

Source: Wu et al. 2011 [109]. Copyright 2011. AAAS.

MOF‐derived M–N–C materials are also fabricated as ORR electrocatalysts due to the unique properties of MOFs. These materials are generally prepared by the building of MOFs with metal ions and organic linkers, followed by pyrolysis of the MOF units. The morphology and structure of the final products are mainly dependent on the features of MOFs. Zou and coworkers prepared Co–N–C‐based polyhedron structures with ZIF‐67 as the precursor [112]. By controlling the size of ZIF‐67, the structure and the ORR catalytic performance of the Co–N–C materials can be tuned. The material derived from the smallest ZIF‐67 (300 nm) exhibited superior ORR activity with an onset potential of 0.86 V (versus RHE) and a half‐wave potential of 0.71 V in acid. Fe‐based MOFs can be synthesized through a reaction between FeCl2 and H3BTT·2HCl (BTT3− = 1,3,5‐benzenetristetrazolate) [113]. The Fe–BTT was then pyrolyzed, acid‐leached, and re‐heated in NH3 to obtain Fe–N–C structure with iron carbides. An onset potential of 0.91 V and a half‐wave potential of 0.81 V for ORR could be achieved for the catalysts. Recently, we fabricated M–N–C type of graphene tubes with a Co‐based, cage‐containing MOF as the template (Figure 11.9a,b) [114, 115]. Either benzene‐1,3,5‐tricarboxylic acid or 2,4,6‐tris(4‐pyridyl)‐1,3,5‐triazine was used as the ligand to prepare Co‐based MOF, and then dicyandiamide and iron(II) acetate were added. The mixture was heated at 1000 °C in N2 atmosphere to form graphene/graphene tube nanocomposites. The N‐doped graphene/graphene tube hybrid exhibits outstanding ORR activity with excellent cathode discharge capacity (Figure 11.9c) and acceptable cyclability in LiO2 batteries (Figure 11.9d). In addition, the M–N–C type of N‐doped graphene tubes (N‐GTs) developed using MIL‐100(Fe) MOF as a template has been used to support Pt nanoparticles (Pt/N‐GT). SEM and TEM images (Figure 11.9e,f) indicate that Pt nanoparticles are well dispersed on N‐GT supports. Compared with commercial Pt/C, the Pt/N‐GT exhibited superior activity to N‐GT and Pt/C catalysts (Figure 11.9g). In N 1s XPS spectra (Figure 11.9h), a shift in the N 1s peaks is observed when comparing Pt/N‐GT and N‐GT samples. These changes in binding energy are possibly due to an interaction and a coordination between pyridinic nitrogen and Pt particles, resulting from the richness in local electron density to facilitate the deposition of reduced Pt seeds on the support. Recently, N‐GTs as supports for FeCoNi alloy nanoparticles have been reported as efficient catalysts for OER in addition to ORR [116].

Image described caption and surrounding text.

Figure 11.9 (a) SEM and (b) TEM images of M–N–C‐based nanotubes. (c) Initial discharge performance for various catalysts at a current density of 50 mA gcat−1 in LiO2 battery tests. (d) Cycling test of the MOF‐derived N–Fe catalyst at a current density of 400 mA gcat−1 with voltage cutoff at 2.5 V (discharge) and 4.1 V (charge) [114]. (e) SEM and (f) TEM images of Fe–N–C‐nanotube‐supported Pt nanoparticles (Pt/N‐GT). (g) ORR polarization curves recorded with a rotating speed of 900 rpm for Pt/N‐GT, N‐GT, and Pt/C. (h) N 1s XPS spectra of original N‐GT and 20 wt% Pt/N‐GT catalysts.

Source: Li et al. 2015 [115]. Copyright 2015. John Wiley & Sons.

Source: Copyright 2014. Wiley‐VCH.

Very recently, metal nanoparticles are investigated as viable templates to fabricate structure‐defined M–N–C catalysts. Zhang and coworkers developed a facile in situ replication and polymerization strategy to fabricate FeN‐doped mesoporous carbon spheres (Fe‐NMCSs) with Fe3O4 microspheres as the templates (Figure 11.10) [117]. Pyrrole would penetrate into the mesoporous structure of the Fe3O4 particles and polymerize in the voids. The Fe3+ ions etched from the Fe3O4 could act as the oxidizer as well as the Fe sources. Thus, the materials would keep the initial structure after the in situ polymerization. After being heated at 950 °C for 1 h, the high ORR‐active Fe‐NMCSs could be obtained. From Fe 2p XPS spectrum, the Fe‐NMCSs show several kinds of Fe that originate from either Fe–N structure or Fe oxides (Figure 11.10c). In 0.1 M KOH, it could exhibit an onset potential of 1.027 V with a half‐wave potential of 0.86 V as well as four‐electron reaction route for ORR (Figure 11.10d,e).

Image described caption and surrounding text.

Figure 11.10 (a) Schematic representation of the preparation process of the Fe‐NMCS catalysts. (b) Schematic representation of nitrogen chemical states. (c) The Fe 2p XPS spectra of Fe‐NMCSs. (d) RDE polarization plots of Fe‐NMCSs, NMCSs, nitrogen carbon particles (NCPs), and commercial Pt/C catalysts at a scan rate of 10 mV s−1 and rotation speed of 1600 rpm in O2‐saturated 0.1 M KOH. (e) Voltammograms of Fe‐NMCSs at various speeds at a scan rate of 10 mV s−1; inset is the corresponding K–L plots at different potentials.

Source: Meng et al. 2016 [117]. Copyright 2016. John Wiley & Sons.

11.3 B‐Doped, Carbon‐Supported Metal Catalysts

11.3.1 Design and Synthesis

B doping is another way to tune the electronic and electrochemical properties of carbon materials [118121]. The introduction of B will induce a charge polarization in the carbon structure, which will benefit the electrocatalysis. Compared with N, B only has three valence electrons, which is less than carbon. Thus, B doping would result in p‐type behavior for the layered carbon materials. As B is lacks electron, it can hardly coordinate with transition metal to form metal–boron–carbon structure. Therefore, the reported B‐doped, carbon‐supported metal catalysts are mainly focusing on metal nanoparticles that supported on B‐doped carbon substrates.

To synthesize B‐doped, carbon‐supported metal catalysts, B‐doped carbon should be first prepared. CVD has been widely used to prepare B‐doped carbon with precursors containing C and B. Pasupathy and coworkers chemically deposited B‐doped graphene (BG) with CH4/B2H6 as the precursor and proved that B can be incorporated into the carbon lattice in graphitic form primarily and contributes ∼0.5 carriers into the graphene sheet per dopant, which was similar to the N‐doped graphene prepared by this method [122]. Cattelan et al. prepared large‐scale, B‐doped graphene with the same B and C sources [123] and demonstrated that the introduction of B would benefit the production of graphene with large area (Figure 11.11). The B‐doped carbon can also be prepared by microwave plasma and arc discharge with borane as the B sources [51, 124]. Thermal annealing of carbon with H3BO3 or B2O3 is preferred in preparing B‐doped carbon for electrocatalysis as it is more efficient to produce B‐doped carbon in large quantity [125, 126]. After obtaining B‐doped carbon, metal nanoparticles can be supported on it by either in situ nucleation or redispersion of synthesized nanoparticles.

Image described caption and surrounding text.

Figure 11.11 SEM micrographs of the CVD growth of B‐doped graphene layers on copper and schematic drawings of the growth: (a) first step of the synthesis showing the formation of large graphene domains almost covering about 80% of the copper substrates and (b) second step showing the further growth of graphene islands and the nucleation of some 3D clusters of nonstoichiometric boron carbide on bare copper areas (scale bar 3 μm).

Source: Cattelan et al. 2013 [123]. Copyright 2013. American Chemical Society.

11.3.2 B‐Doped, Carbon‐Supported Metal Nanoparticle Electrocatalysts

To date, the nanoparticles that supported on B‐doped carbon are mainly precious metal nanoparticles. Sun et al. loaded Pt nanoparticles on B‐doped graphene by a microwave‐assisted polyol process in ethylene glycol solution [127]. The Pt decorated on B‐doped graphene (Pt/BG) showed smaller size and narrower size distribution compared with that of Pt/G, thus demonstrating an enhanced catalytic behavior for MOR. The enhancement in activity is ascribed to the electronic interaction between B‐doped graphene and Pt nanoparticles [128]. Kim and coworkers also synthesized B‐doped graphene‐supported Pt with H2PtCl6 as the precursor (Figure 11.12) [129]. They found that carbon black (CB) is beneficial for the well dispersion of B‐doped, carbon‐supported Pt catalysts. The catalysts exhibit excellent stability in proton exchange membrane fuel cell due to the binding between Pt‐ and B‐doped graphene. Liu and coworkers embedded Au@AuPt core shell nanoparticle on B‐doped graphene by simple redisperse Au@AuPt nanoparticle on graphene substrate [126]. The hybrid show high electrochemical response to rutin and can achieve a detection limit (S/N = 3) of 2.84 × 10−10 M when constructed as an electrochemical sensor.

Schematic diagram depicting steps for the synthesis of B-doped, graphene-supported Pt nanoparticles: Graphite; Graphene oxide; Boron doped graphene; Pt-boron doped graphene with CB, Graphene sheet, Platinum; Boron; Oxygen in the legend.

Figure 11.12 Steps for the synthesis of B‐doped, graphene‐supported Pt nanoparticles.

Source: Yang et al. 2015 [129]. Copyright 2015. Elsevier.

11.4 Conclusions and Perspective

In conclusion, we summarized the recent progress of the heteroatom‐doped, carbon‐supported metal catalysts. The carbon supports that doped with N, B, S, P, and multi‐atoms are investigated separately for the deeper understanding of the heteroatom‐doping effect. With covalently bonding between heteroatom and C, the chemical and electrochemical properties of carbon materials would be tuned, which is beneficial for the loading of metal nanostructures and would result in an enhanced electrocatalytic performance. Among the heteroatom‐doping methods, N‐doping process shows the most outstanding behavior as it not only demonstrates affinity toward the binding of metal‐based nanoparticles but also coordinates with transition metal to form M–N–C structures. The strong interaction between metal and N significantly improves the electrocatalytic activity and stability of the N‐doped, carbon‐supported metal catalysts. Therefore, more efforts are needed in the investigation of coordination interaction between heteroatom and metal on the developing of novel heteroatom‐doped, carbon‐supported metal catalysts.

The general emphases for the design of highly efficient electrocatalysts are the amount of the catalytic sites, the activity of the catalytic sites, and the conductivity of the materials. The heteroatom‐doped, carbon‐supported metal materials usually demonstrate good conductivity. Therefore, further researches in design‐optimized, heteroatom‐doped, carbon‐supported metal electrocatalysts should be mainly focused on the improvement of the number and activity of the catalytic sites. Three‐dimensional structure and isolated single‐metal‐atom‐anchored structure may be beneficial for the heteroatom‐doped, carbon‐supported metal catalysts to expose more active sites, whereas multi‐metal and/or multi‐heteroatom (N, B, S, and P) center may result in more active catalytic sites. Moreover, theoretical studies are needed for the rational design of heteroatom‐doped, carbon‐supported metal structures. In addition, various reactions including ORR, OER, HER, and CO2RR can be catalyzed by different heteroatom‐doped, carbon‐supported metal catalysts; thus, bifunctional electrocatalysts that are able to catalyze two coupled reactions or reversible reactions such as OER–HER, OER–ORR, and OER–CO2RR are also attractive research fields for the development of novel heteroatom‐doped, carbon‐supported metal catalysts.

References

  1. 1 Liu, M.M., Zhang, R.Z., and Chen, W. (2014). Chem. Rev. 114: 5117–5160.
  2. 2 Nie, Y., Li, L., and Wei, Z.D. (2015). Chem. Soc. Rev. 44: 2168–2201.
  3. 3 Dai, L.M., Chang, D.W., Baek, J.B., and Lu, W. (2012). Small 8: 1130–1166.
  4. 4 Dai, L.M., Xue, Y.H., Qu, L.T. et al. (2015). Chem. Rev. 115: 4823–4892.
  5. 5 Zhu, C.Z., Li, H., Fu, S.F. et al. (2016). Chem. Soc. Rev. 45: 517–531.
  6. 6 Jiao, Y., Zheng, Y., Jaroniec, M.T., and Qiao, S.Z. (2015). Chem. Soc. Rev. 44: 2060–2086.
  7. 7 Hisatomi, T., Kubota, J., and Domen, K. (2014). Chem. Soc. Rev. 43: 7520–7535.
  8. 8 Wang, Z.L., Xu, D., Xu, J.J., and Zhang, X.B. (2014). Chem. Soc. Rev. 43: 7746–7786.
  9. 9 Li, Y.G. and Dai, H.J. (2014). Chem. Soc. Rev. 43: 5257–5275.
  10. 10 Appel, A.M., Bercaw, J.E., Bocarsly, A.B. et al. (2013). Chem. Rev. 113: 6621–6658.
  11. 11 Huang, X.Q., Zhao, Z.P., Cao, L. et al. (2015). Science 348: 1230–1234.
  12. 12 Shao, M.H., Chang, Q.W., Dodelet, J.P., and Chenitz, R. (2016). Chem. Rev. 116: 3594–3657.
  13. 13 Zhang, L., Roling, L.T., Wang, X. et al. (2015). Science 349: 412–416.
  14. 14 Hong, W.T., Risch, M., Stoerzinger, K.A. et al. (2015). Energy Environ. Sci. 8: 1404–1427.
  15. 15 Gong, M. and Dai, H.J. (2015). Nano Res. 8: 23–39.
  16. 16 Wang, D.Y., Gong, M., Chou, H.L. et al. (2015). J. Am. Chem. Soc. 137: 1587–1592.
  17. 17 Zou, X.X. and Zhang, Y. (2015). Chem. Soc. Rev. 44: 5148–5180.
  18. 18 Wu, G., Santandreu, A., Kellogg, W. et al. (2016). Nano Energy 29: 83–110.
  19. 19 Xia, B.Y., Yan, Y., Li, N. et al. (2016). Nat. Energy 1: 15006.
  20. 20 Xia, W., Mahmood, A., Zou, R.Q., and Xu, Q. (2015). Energy Environ. Sci. 8: 1837–1866.
  21. 21 Zhang, J.T., Zhao, Z.H., Xia, Z.H., and Dai, L.M. (2015). Nat. Nanotechnol. 10: 444–452.
  22. 22 Wang, D.W. and Su, D.S. (2014). Energy Environ. Sci. 7: 576–591.
  23. 23 Wang, Y.J., Wilkinson, D.P., and Zhang, J.J. (2011). Chem. Rev. 111: 7625–7651.
  24. 24 Li, Q., Mahmood, N., Zhu, J.H. et al. (2014). Nano Today 9: 668–683.
  25. 25 Liang, Y.Y., Li, Y.G., Wang, H.L. et al. (2011). Nat. Mater. 10: 780–786.
  26. 26 Trogadas, P., Fuller, T.F., and Strasser, P. (2014). Carbon 75: 5–42.
  27. 27 Zhao, Y., Nakamura, R., Kamiya, K. et al. (2013). Nat. Commun. 4: 2390.
  28. 28 Wu, G., Mack, N.H., Gao, W. et al. (2012). ACS Nano 6: 9764–9776.
  29. 29 Majeed, S., Zhao, J.M., Zhang, L. et al. (2013). Nanotechnol. Rev. 2: 615–635.
  30. 30 Zhang, L.P. and Xia, Z.H. (2011). J. Phys. Chem. C 115: 11170–11176.
  31. 31 Zhang, C.Z., Hao, R., Liao, H.B., and Hou, Y.L. (2013). Nano Energy 2: 88–97.
  32. 32 Guo, D.H., Shibuya, R., Akiba, C. et al. (2016). Science 351: 361–365.
  33. 33 Lepro, X., Terres, E., Vega‐Cantu, Y. et al. (2008). Chem. Phys. Lett. 463: 124–129.
  34. 34 Xia, W. (2016). Catal. Sci. Technol. 6: 630–644.
  35. 35 Su, D.S., Perathoner, S., and Centi, G. (2013). Chem. Rev. 113: 5782–5816.
  36. 36 Li, Y.H., Hung, T.H., and Chen, C.W. (2009). Carbon 47: 850–855.
  37. 37 Zhang, L.L., Wang, A.Q., Wang, W.T. et al. (2015). ACS Catal. 5: 6563–6572.
  38. 38 Chen, Y., Ji, S., Wang, Y. et al. (2017). Angew. Chem. Int. Ed. 56: 6937–6941.
  39. 39 Zhou, Y.K., Neyerlin, K., Olson, T.S. et al. (2010). Energy Environ. Sci. 3: 1437–1446.
  40. 40 Li, Q., Zhu, W.L., Fu, J.J. et al. (2016). Nano Energy 24: 1–9.
  41. 41 Mahyari, M. and Shaabani, A. (2014). J. Mater. Chem. A 2: 16652–16659.
  42. 42 Xiong, B., Zhou, Y.K., Zhao, Y.Y. et al. (2013). Carbon 52: 181–192.
  43. 43 Liang, Y.Y., Li, Y.G., Wang, H.L., and Dai, H.J. (2013). J. Am. Chem. Soc. 135: 2013–2036.
  44. 44 Zhang, H.G., Osgood, H., Xie, X.H. et al. (2017). Nano Energy 31: 331–350.
  45. 45 Wu, G. and Zelenay, P. (2013). Acc. Chem. Res. 46: 1878–1889.
  46. 46 Lee, J., Farha, O.K., Roberts, J. et al. (2009). Chem. Soc. Rev. 38: 1450–1459.
  47. 47 Qu, L.T., Liu, Y., Baek, J.B., and Dai, L.M. (2010). ACS Nano 4: 1321–1326.
  48. 48 Reddy, A.L.M., Srivastava, A., Gowda, S.R. et al. (2010). ACS Nano 4: 6337–6342.
  49. 49 Lin, Z.Y., Waller, G.H., Liu, Y. et al. (2013). Nano Energy 2: 241–248.
  50. 50 Pan, F.P., Jin, J.T., Fu, X.G. et al. (2013). ACS Appl. Mater. Interfaces 5: 11108–11114.
  51. 51 Panchokarla, L.S., Subrahmanyam, K.S., Saha, S.K. et al. (2009). Adv. Mater. 21: 4726–4730.
  52. 52 Jeong, H.M., Lee, J.W., Shin, W.H. et al. (2011). Nano Lett. 11: 2472–2477.
  53. 53 Kumar, N.A., Nolan, H., McEvoy, N. et al. (2013). J. Mater. Chem. A 1: 4431–4435.
  54. 54 Guo, B.D., Liu, Q.A., Chen, E.D. et al. (2010). Nano Lett. 10: 4975–4980.
  55. 55 Geng, D.S., Chen, Y., Chen, Y.G. et al. (2011). Energy Environ. Sci. 4: 760–764.
  56. 56 Long, D.H., Li, W., Ling, L.C. et al. (2010). Langmuir 26: 16096–16102.
  57. 57 Lai, L.F., Potts, J.R., Zhan, D. et al. (2012). Energy Environ. Sci. 5: 7936–7942.
  58. 58 Bai, J.C., Zhu, Q.Q., Lv, Z.X. et al. (2013). Int. J. Hydrogen Energy 38: 1413–1418.
  59. 59 Chetty, R., Kundu, S., Xia, W. et al. (2009). Electrochim. Acta 54: 4208–4215.
  60. 60 Wong, W.Y., Daud, W.R.W., Mohamad, A.B. et al. (2013). Int. J. Hydrogen Energy 38: 9370–9386.
  61. 61 Su, F.B., Tian, Z.Q., Poh, C.K. et al. (2010). Chem. Mater. 22: 832–839.
  62. 62 Yang, G.H., Li, Y.J., Rana, R.K., and Zhu, J.J. (2013). J. Mater. Chem. A 1: 1754–1762.
  63. 63 Li, X.G., Park, S., and Popov, B.N. (2010). J. Power Sources 195: 445–452.
  64. 64 Arrigo, R., Schuster, M.E., Xie, Z.L. et al. (2015). ACS Catal. 5: 2740–2753.
  65. 65 Liang, Y.Y., Wang, H.L., Diao, P. et al. (2012). J. Am. Chem. Soc. 134: 15849–15857.
  66. 66 Mahmood, N., Zhang, C.Z., and Hou, Y.L. (2013). Small 9: 1321–1328.
  67. 67 Chen, Z.W., Higgins, D., Yu, A.P. et al. (2011). Energy Environ. Sci. 4: 3167–3192.
  68. 68 Xia, Z.H., An, L., Chen, P.K., and Xia, D.G. (2016). Adv. Energy Mater. 6: 1600458.
  69. 69 Xia, W., Mahmood, A., Liang, Z.B. et al. (2016). Angew. Chem. Int. Ed. 55: 2650–2676.
  70. 70 Wood, K.N., O'Hayre, R., and Pylypenko, S. (2014). Energy Environ. Sci. 7: 1212–1249.
  71. 71 He, L., Weniger, F., Neumann, H., and Beller, M. (2016). Angew. Chem. Int. Ed. 55: 12582–12594.
  72. 72 Ferrandon, M., Kropf, A.J., Myers, D.J. et al. (2012). J. Phys. Chem. C 116: 16001–16013.
  73. 73 Shen, M.X., Wei, C.T., Ai, K.L., and Lu, L.H. (2017). Nano Res. 10: 1449–1470.
  74. 74 Zitolo, A., Goellner, V., Armel, V. et al. (2015). Nat. Mater. 14: 937–942.
  75. 75 Shui, J.L., Chen, C., Grabstanowicz, L. et al. (2015). Proc. Natl. Acad. Sci. U.S.A. 112: 10629–10634.
  76. 76 Zhang, L.J., Su, Z.X., Jiang, F.L. et al. (2014). Nanoscale 6: 6590–6602.
  77. 77 Wang, X.J., Fan, X.X., Lin, H.H. et al. (2016). RSC Adv. 6: 37965–37973.
  78. 78 Chen, Y.Z., Wang, C.M., Wu, Z.Y. et al. (2015). Adv. Mater. 27: 5010–5016.
  79. 79 Vinayan, B.P., Nagar, R., Rajalakshmi, N., and Ramaprabhu, S. (2012). Adv. Funct. Mater. 22: 3519–3526.
  80. 80 Castillejos, E., Bachiller‐Baeza, B., Asedegbega‐Nieto, E. et al. (2015). RSC Adv. 5: 81583–81598.
  81. 81 Chen, P.R., Chew, L.M., and Xia, W. (2013). J. Catal. 307: 84–93.
  82. 82 Castillejos, E., Chico, R., Bacsa, R. et al. (2010). Eur. J. Inorg. Chem. 5096–5102.
  83. 83 Haniff, M.A.S.M., Hafiz, S.M., Wahid, K.A. et al. (2017). J. Mater. Sci. 52: 6280–6290.
  84. 84 Zhang, S., Gao, H., Huang, M., and Zhou, J. (2017). J. Alloys Compd. 705: 801–805.
  85. 85 Zhang, L.L., Ge, D.H., Qu, G.L. et al. (2017). Nanoscale 9: 5451–5457.
  86. 86 Li, X.Y., Fu, N.Q., Zou, J.Z. et al. (2017). Electrochim. Acta 225: 137–142.
  87. 87 Han, X.P., Wu, X.Y., Zhong, C. et al. (2017). Nano Energy 31: 541–550.
  88. 88 Zhang, Y.F., Zuo, L.Z., Zhang, L.S. et al. (2016). Nano Res. 9: 2747–2759.
  89. 89 Sun, S.H., Zhang, G.X., Zhong, Y. et al. (2009). Chem. Commun. 7048–7050.
  90. 90 Kim, H., Kim, Y., Noh, Y. et al. (2017). ChemCatChem 9: 1503–1510.
  91. 91 Tan, Y.M., Xu, C.F., Chen, G.X. et al. (2012). Adv. Funct. Mater. 22: 4584–4591.
  92. 92 Morales‐Guio, C.G., Liardet, L., and Hu, X.L. (2016). J. Am. Chem. Soc. 138: 8946–8957.
  93. 93 Jin, H.Y., Wang, J., Su, D.F. et al. (2015). J. Am. Chem. Soc. 137: 2688–2694.
  94. 94 Morales‐Guio, C.G., Stern, L.A., and Hu, X.L. (2014). Chem. Soc. Rev. 43: 6555–6569.
  95. 95 Faber, M.S. and Jin, S. (2014). Energy Environ. Sci. 7: 3519–3542.
  96. 96 Yan, Y., Xia, B.Y., Xu, Z.C., and Wang, X. (2014). ACS Catal. 4: 1693–1705.
  97. 97 Li, D.J., Maiti, U.N., Lim, J. et al. (2014). Nano Lett. 14: 1228–1233.
  98. 98 Guo, Y.X., Zhang, X.Y., Zhang, X.P., and You, T.Y. (2015). J. Mater. Chem. A 3: 15927–15934.
  99. 99 Qu, B., Li, C.Y., Zhu, C.L. et al. (2016). Nanoscale 8: 16886–16893.
  100. 100 Li, Q., Cui, W., Tian, J.Q. et al. (2015). ChemSusChem 8: 2487–2491.
  101. 101 Guo, Y.X., He, J.P., Wang, T. et al. (2011). J. Power Sources 196: 9299–9307.
  102. 102 Lei, Z.B., Zhao, M.Y., Dang, L.Q. et al. (2009). J. Mater. Chem. 19: 5985–5995.
  103. 103 Lei, Z.B., An, L.Z., Dang, L.Q. et al. (2009). Microporous Mesoporous Mater. 119: 30–38.
  104. 104 Hammer, B. and Norskov, J.K. (2000). Adv. Catal. 45: 71–129.
  105. 105 Corpuz, A.R., Olson, T.S., Joghee, P. et al. (2012). J. Power Sources 217: 142–151.
  106. 106 Olson, T.S., Dameron, A.A., Wood, K. et al. (2013). J. Electrochem. Soc. 160: F389–F394.
  107. 107 Bashyam, R. and Zelenay, P. (2006). Nature 443: 63–66.
  108. 108 Wu, G., Nelson, M.A., Mack, N.H. et al. (2010). Chem. Commun. 46: 7489–7491.
  109. 109 Wu, G., More, K.L., Johnston, C.M., and Zelenay, P. (2011). Science 332: 443–447.
  110. 110 Li, Q., Wu, G., Cullen, D.A. et al. (2014). ACS Catal. 4: 3193–3200.
  111. 111 Wu, G., Johnston, C.M., Mack, N.H. et al. (2011). J. Mater. Chem. 21: 11392–11405.
  112. 112 Xia, W., Zhu, J.H., Guo, W.H. et al. (2014). J. Mater. Chem. A 2: 11606–11613.
  113. 113 Afsahi, F. and Kaliaguine, S. (2014). J. Mater. Chem. A 2: 12270–12279.
  114. 114 Li, Q., Xu, P., Gao, W. et al. (2014). Adv. Mater. 26: 1378–1386.
  115. 115 Li, Q., Pan, H.Y., Higgins, D. et al. (2015). Small 11: 1443–1452.
  116. 116 Gupta, S., Qiao, L., Zhao, S. et al. (2016). Adv. Energy Mater. 6: 201601198.
  117. 117 Meng, F.L., Wang, Z.L., Zhong, H.X. et al. (2016). Adv. Mater. 28: 7948–7955.
  118. 118 Li, X., Fan, L.L., Li, Z. et al. (2012). Adv. Energy Mater. 2: 425–429.
  119. 119 Duan, J.J., Chen, S., Jaroniec, M., and Qiao, S.Z. (2015). ACS Catal. 5: 5207–5234.
  120. 120 Wang, X.W., Sun, G.Z., Routh, P. et al. (2014). Chem. Soc. Rev. 43: 7067–7098.
  121. 121 Yen, W.C., Medina, H., Huang, J.S. et al. (2014). J. Phys. Chem. C 118: 25089–25096.
  122. 122 Zhao, L.Y., Levendorf, M., Goncher, S. et al. (2013). Nano Lett. 13: 4659–4665.
  123. 123 Cattelan, M., Agnoli, S., Favaro, M. et al. (2013). Chem. Mater. 25: 1490–1495.
  124. 124 Tang, Y.B., Yin, L.C., Yang, Y. et al. (2012). ACS Nano 6: 1970–1978.
  125. 125 Kim, Y.A., Fujisawa, K., Muramatsu, H. et al. (2012). ACS Nano 6: 6293–6300.
  126. 126 Chen, X.L., Yang, G.M., Feng, S.P. et al. (2017). Appl. Surf. Sci. 402: 232–244.
  127. 127 Sun, Y.R., Du, C.Y., An, M.C. et al. (2015). J. Power Sources 300: 245–253.
  128. 128 Muhich, C.L., Westcott, J.Y., Morris, T.C. et al. (2013). J. Phys. Chem. C 117: 10523–10535.
  129. 129 Yang, H.N., Lee, D.C., Park, K.W., and Kim, W.J. (2015). Energy 89: 500–510.
..................Content has been hidden....................

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