Open access peer-reviewed chapter

Metal‐Organic Frameworks and their Applications in Hydrogen and Oxygen Evolution Reactions

Written By

Fengxiang Yin, Xiao Zhang, Xiaobo He and Hao Wang

Submitted: 20 November 2015 Reviewed: 20 June 2016 Published: 12 October 2016

DOI: 10.5772/64657

From the Edited Volume

Metal-Organic Frameworks

Edited by Fahmina Zafar and Eram Sharmin

Chapter metrics overview

2,793 Chapter Downloads

View Full Metrics

Abstract

The hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) play a vital role in many energy storage and conversion systems, including water splitting, rechargeable metal‐air batteries, and the unitized regenerative fuel cells. The noble‐metal catalysts based on Pt, Ir, and Au are the best electrocatalysts for the HER/OER, but they suffer from high price and scarcity problems. Therefore, it is urgently necessary to develop efficient, low‐cost, and environment‐friendly non‐noble metal electrocatalysts. Metal‐organic frameworks (MOFs) are crystalline materials with porous network structure. MOFs possess various compositions, large specific surface area, tunable pore structures, and they are easily functionalized. MOFs have been widely studied and applied in many fields, such as gas adsorption/separation, drug delivery, catalysis, magnetism, and optoelectronics. Recently, MOFs‐based electrocatalysts for HER/OER have been rapidly developed. These MOFs‐based catalysts exhibit excellent catalytic performance for HER/OER, demonstrating a promising application prospect in HER/OER. In this chapter, the concept, structure, category, and synthesis of MOFs will be first introduced briefly. Then, the applications of the MOFs‐based catalysts for HER/OER in recent years will be discussed in details. Specially, the synthesis, structure, and catalytic performance for HER/OER of the MOFs‐based catalysts will be emphatically discussed.

Keywords

  • metal‐organic frameworks
  • electrocatalytic water splitting
  • hydrogen evolution reaction
  • oxygen evolution reaction
  • electrocatalysts

1. Introduction

Electrocatalytic hydrogen evolution reaction (HER) and oxygen evolution reaction (OER) are the key reactions for many energy storage and conversion systems, including water splitting, rechargeable metal‐air batteries, and the unitized regenerative fuel cells [13]. However, the two reactions are kinetically sluggish, which leads to large overpotentials in the electrodes, decreases the performance of these systems, and makes them difficult to meet the requirements for commercialization at present. It has been reported that suitable electrocatalysts can effectively accelerate the HER/OER and reduce their overpotentials in the electrodes. At present, noble metal catalysts show the best catalytic performance for HER and OER. For example, carbon‐supported Pt catalysts exhibit the best catalytic performance for HER [4], while IrO2 and RuO2 are very effective for OER [5]. However, the reserves of these noble metals are extremely limited in nature resulting in a high price of these noble metal catalysts, which restricts their practical applications. Therefore, the development of highly efficient electrocatalysts with low cost for HER/OER is indispensable. In recent years, a large amount of non‐noble metal catalysts have been synthesized for HER/OER. The non‐noble metal catalysts for HER mainly include transition metal phosphides [68], such as CoP, NiP, transition metal sulfides [911], such as MoS2, CoS, Ni‐S, and their alloys [12, 13] such as Ni‐Mo, Ni‐Cu. As for OER, the non‐noble metal catalysts mainly focus on metal oxides [1416] such as Co3O4, MnOx, NixCo3-xO4, and metal sulfides [17, 18] such as Ni3S2, NiCo2S4. Although these non‐noble metal catalysts have shown excellent catalytic performance for HER and/or OER, to date, there is no one catalyst that can fully meet the demands of the practical applications. It is necessary to develop more economically viable electrocatalysts with higher catalytic performance for HER and/or OER, and further accelerate the commercial applications of these energy storage and conversion systems.

Metal‐organic frameworks (MOFs) are porous crystalline materials with topological structures. MOFs are formed by assembling metal nodes and organic ligands. The metal node precursors mainly come from metal nitrates or chlorides, while the organic ligands mainly include benzimidazolate, dicarboxylic acid, and others. MOFs can be synthesized by some simple methods under mild conditions such as solvothermal, diffusion, microwave, and ionothermal methods [1922]. MOFs possess unique structural characterizations such as high porosity, ultrahigh specific surface area (∼6240 m2 g−1) and tunable pore structure and easy functionalization [23], and they have been widely investigated and applied in gas adsorption/separation [24], magnetism [25], optoelectronics [26], and catalysis [27]. In recent years, MOFs have attracted great attention for use in energy conversion and storage systems such as water splitting, fuel cells, metal‐air batteries, and supercapacitors. More recently, the applications of MOFs have been expanded to the HER and OER catalysis. To date, some novel MOFs‐based catalysts for HER and OER have been synthesized. On the one hand, the active species can be introduced into the structures of MOFs, directly acting as effective catalysts for HER and/or OER; on the other hand, MOFs also can be used as supports to disperse the active species for HER and/or OER due to their porosity. In addition, the highly porous MOFs can be transformed to the derived catalysts by self‐sacrifice methods. These MOFs‐derived catalysts usually possess improved electronic conductivity and remarkable HER and/or OER catalytic performance. This chapter reviews recent progress of MOFs and MOFs‐derived electrocatalysts for HER and/or OER. The MOFs are first introduced briefly. The applications of MOFs‐based catalysts for HER and/or OER are the main contents and involve details. This work provides a helpful reference for researchers who are studying the MOFs‐based electrocatalysts for HER and/or OER.

Advertisement

2. Overview of metal‐organic frameworks

As a class of the coordination polymers, MOFs are crystalline porous materials. MOFs are formed through self‐assembling using the inorganic metals as nodes and organic ligands as linkers [28, 29]. They possess unique structure properties, such as ultrahigh specific surface area and high porosity, tunable pore structure, and easy functionalization. The specific surface area of MOFs are much higher than that of other porous materials such as activated carbon and zeolites. For example, the synthesized MOF‐201 possesses a high BET specific surface area of ∼6240 m2 g−1 [30], while zeolites possess specific surface area less than 600 m2 g−1. High porosity of MOFs also brings up a very low density (∼0.13 g cm−3) [23], which is important to the storage of fuels or the applications in energy conversion and storage. The pore structure of MOFs can be tuned by choosing different organic ligands or altering the length of organic ligands. It was reported that Deng et al. prepared a novel IRMOF (IRMOF‐74 XI) with large apertures of ∼9.8 nm, by expanding the length of its organic linker to ∼5 nm [31]. Apart from the tunable pore structure, MOFs can be easily functionalized by decorating the pores, surfaces, or introducing guests into MOFs, thus resulting in a large amount of MOFs with various physicochemical properties [32]. Due to these unique properties, MOFs have been widely developed and applied in gas adsorption/separation [24], magnetism [25], optoelectronics [26], and catalysis [27] in recent years.

According to topological structure features, MOFs can be generally divided into four categories: isoreticular metal‐organic frameworks (IRMOFs), zeoliticimidazolate frameworks (ZIFs), materials of institute Lavoisier frameworks (MILs) and pocket‐channel frameworks (PCNs) [33, 34].

IRMOFs are a large family of MOFs, which are built up by [Zn4O]6+ and organic hydroxyl acid as metal nodes and ligands, respectively. In the IRMOFs family, MOF‐5 is the typical one which was prepared by Li et al. as early as in 1999 [35]. MOF‐5 is highly porous, and the specific surface area and the density are about 2900 m2 g−1 and 0.59 g cm−3, respectively. Specially, the space access to guest substance reaches up to 55–61%.

The topological structures of pores of ZIFs are similar to aluminosilicate zeolite. Generally, imidazole or its derivatives and Zn or Co ions are used as ligands and metal nodes, respectively. The N atoms in imidazole and its derivatives can coordinate with Zn, Co, or other transitional metal ions. Compared with other kinds of MOFs, ZIFs show higher thermal and chemical stability. For example, as a typical ZIF material, ZIF‐8 can sustain its structures in inorganic or organic solvents or even in boiling water for 1–7 days [36].

Generally, there are two categories of MILs. One is formed by using lanthanides or transition metals as metal nodes and dicarboxylic acids as organic ligands. The other is obtained using trivalent metals such as aluminum or vanadium as metal nodes, and terephthalic acids or trimesic acids as organic ligands. MILs are highly porous and possess ultrahigh specific surface area. MIL‐100 and MIL‐101 are typical MILs. In MIL‐100, Cr3+ ions are metal nodes and BTC (1,3,5‐benzenetricarboxylate) is used as organic ligands. It possesses a high specific surface area of ∼3100 m2/g and mesopores with the pore size of ∼2.9 nm [37]. In addition, MIL‐100 exhibits excellent thermal stability below 275°C [37].

During the synthesis of PCNs, Cu ions or oxo‐clusters are generally used as metal nodes, and tricarboxylic acids, 4,4′,4″‐s‐triazine‐2,4,6‐triyltribenzoate (H3TATB) or s‐heptazine tribenzoate (HTB) is usually used as an organic ligands. PCNs have pocket and three‐dimensional orthogonal channels which are connected through small sized windows. As a typical PCN, Cu‐BTC is formed by using Cu2(COO)4 as metal nodes and BTC as organic ligands. Two typical channel structures can be observed in Cu‐BTC. One is a small octahedron pocket, while the other is the three-dimensional orthogonal channels (∼1 nm) [38].

To prepare MOFs, various synthetic methods have been proposed, such as solvothermal (hydrothermal) method, diffusion method, microwave‐assisted method and ionothermal method. During the solvothermal (hydrothermal) process, water or organic solvents are usually used as reaction medium in an airtight reactor. High temperature and pressure condition can be achieved through heating the airtight reactor. Thus, some compounds with poor solubility at room temperature and atmospheric pressure can be dissolved and re‐crystallized to obtain the desirable MOFs. In addition, high crystallinity and controllable particle size for the resulting MOFs can be obtained through the solvothermal (hydrothermal) process. Therefore, the solvothermal (hydrothermal) method is effective to produce MOFs with good orientation and perfect crystals [19, 39].

The diffusion methods include vapor diffusion, liquid diffusion, and gel diffusion methods. The products are obtained in the two‐phase interface through slow crystallization [20]. The diffusion methods have two advantages, including mild synthesis conditions and high quality crystal MOFs. However, the efficiency of this method is usually low, which requires a long synthesis time, up to 1 week. Furthermore, this method also requires precursors with good solubility.

In the microwave‐assisted process, the charging particles are placed in electromagnetic fields. These charging particles collide with high speed, forming the final products. Due to the high frequency of microwave, the microwave‐assisted method is considered as a high energy efficiency method [21]. It has many advantages, such as uniform heating, high reaction rate, selective heating, and no hysteresis effects.

For the ionothermal synthesis, ionic liquids and eutectic mixture are usually used as solvents [22]. Ionothermal synthesis can avoid the high‐pressure reaction condition, and produce final MOFs in an open condition, because ionic liquids have almost no vapor pressure.

In MOFs, the organic ligands cannot only be used as linkers, but also adsorb gas molecules through Van der Waals force. This is the reason that MOFs materials can be used for gas adsorption/separation [24]. Till date, most of the studies on MOFs still focus on gas adsorption/separation. Because some metals such as group VIII metals possess excellent magnetic properties, MOFs are also popular in magnetism field by using magnetic metal ions as nodes, and they possess paramagnetic or antiferromagnetic properties [25]. Similar regularity was observed of MOFs in the optoelectronics area. Luminescent lanthanide(III) ions which are utilized as nodes for MOFs are always designed as luminescent centers [26]. Due to the high specific surface area and tunable pore structure, MOFs have also been applied in electrocatalysis involving energy conversion and storage in recent years [32]. For instance, MOFs can be directly used as electrocatalysts in the applications for HER and/or OER. However, the electric conductivity of most MOFs is poor. The combination of MOFs and highly conductive substrates (such as graphene, carbon nanotubes, Ni foams and conductive glasses) is an efficient protocol to enhance the electrocatalytic performance of MOFs [27]. In addition, MOFs also can serve as precursors to form derived catalysts such as metal oxides, metal (oxides or carbides or sulfides or phosphides)/nanocarbon hybrids or porous nanocarbons, which is another effective way to take advantage of MOFs for HER and/or OER [40].

Advertisement

3. HER/OER

The growing concern of energy crisis and environment pollution promotes the development of highly effective clean energy storage and conversion systems, such as water splitting [1], metal‐air batteries [2], and fuel cells [3]. In such systems, oxygen‐ and hydrogen‐involving electrocatalytic processes, including oxygen reduction reaction (ORR), hydrogen oxidization reaction (HOR), hydrogen evolution reaction (HER), and oxygen evolution reaction (OER) are the most mentioned key components in determining the practical performance of these energy storage and conversion systems. The former two reactions consume oxygen and hydrogen to provide power, but the premise is that the available oxygen and hydrogen are generated from water splitting through OER and HER, respectively. The OER/HER involved in water splitting can be given in the following equations [9, 41]:

Hydrogen evolution reaction:

2H+ + 2e → H2 in acid solution

2H2O + 2e → H2 + 2OH- in alkaline solution

Oxygen evolution reaction:

2H2O - 4e → O2 + 4H+ in acid solution

4OH - 4e → O2 + 2H2O in alkaline solution

Water splitting can produce pure hydrogen and oxygen which can be directly utilized as raw materials for fuel cells or for other industrial processes individually. Apart from water splitting, HER or OER can be associated with other electrochemical processes. For instance, OER is also a key reaction for rechargeable metal‐air batteries [2] and regenerative fuel cells [3] when being charged. Rechargeable metal‐air batteries and regenerative fuel cells are seen as promising ways to produce electricity in the future and may effectively reduce our dependence on traditional fossil fuels and relieve environmental pollution.

Unlike HOR/ORR, HER/OER don’t maintain a constant value at high overpotential. In other words, HER/OER are not limited by the mass transfer rate and they obey Butler‐Volmer model even in high overpotential. Generally, the HER/OER catalytic mechanisms differ from each other. Even so, the descriptor (ΔGH*, hydrogen binding energy/adsorption free energy) is widely used to reflect HER activity for catalysts, while the descriptor (ΔGO*−ΔGOH*) is used to evaluate the OER activities [5]. HER involves the two‐electron transfer processes, while OER includes more complex processes with the multistep charge‐/proton‐transfer. These complex processes result in the low efficiencies of HER/ORR and make it difficult for them to satisfy practical applications. Therefore, the development of efficient electrocatalysts for HER/OER to accelerate the kinetics and reduce the overpotentials is of great significance for the commercialization of HER/OER‐related devices.

Currently, the HER/OER electrocatalysts mainly include precious metal catalysts and non‐precious metal catalysts [5, 42].The precious metal catalysts have been widely developed due to their extremely excellent catalytic performance. For instance, precious metal catalysts for HER are mainly dependent on Pt‐based catalysts since they exhibit the best HER electrocatalytic performance in basic solution with a low onset overpotential. IrO2 and RuO2 have been demonstrated as the benchmark OER electrocatalysts (Pt‐based catalysts barely show any OER catalytic activity), and even show remarkable stability in acid solution. However, although the precious metal catalysts possess excellent catalytic performance, their scarcity and high price hamper their large‐scale applications. Thus, non‐precious metal catalysts with low‐cost and earth‐abundance for HER/OER are urgently needed.

The non‐precious metal catalysts for HER mainly include transition‐metal sulfides, phosphides, alloys, etc. Through element selection, Co sulfides or phosphides are seen as the most promising HER catalysts, but they suffer from low stability. The alloys have been developed by integrating Ni with Mo or Cu [12, 13]. Diverse structures and composition are the advantages to those alloyed catalysts, but their restriction of activity limits their applications. The non‐precious metal catalysts for the OER mainly contain transition metal oxides [1416], sulfides [17, 18], hydroxides [43], and carbons [44], etc. Among them, metal oxides in single phase or in mixed phases are the most developed ones as efficient OER electrocatalysts due to their diverse compositions, superior stability, and remarkable OER performance. Thus, metal oxides are widely explored as promising OER catalysts, but the mechanism of OER is very sensitive to the compositions and structures of the catalysts, which requires further design, preparation, and optimization of OER catalysts. In spite of rapid development of the HER/OER electrocatalysts with relatively good performance, to date, no one catalyst can satisfy the practical demands. More researches for highly efficient HER and OER electrocatalysts are needed.

Advertisement

4. MOFs‐based catalysts for HER

In recent years, MOFs have been widely explored to store hydrogen fuels due to high specific surface area, tunable pore structures, and various functions [45]. Meanwhile, they also have been developed as electrocatalysts for the HER through electrochemical splitting of water. Generally, MOFs can be directly used as catalysts or served as precursors for the derived catalysts for HER. Recently, Gong et al. synthesized Cu3(Mo8O26)(H2O)2(OH)2(L1)4 (L1 = 4H‐4‐amino‐1,2,4‐triazole) and Ag4(Mo8O26)‐(L2)2.5(H2O) (L2 = 3,5‐dimethyl‐4‐amino‐4H‐1,2,4‐triazole) MOFs via a hydrothermal process [46]. The former had a chain‐like structure, while the latter possessed a 3D structure. Both of them showed electrocatalytic activity for HER in 0.5 M H2SO4 with low overpotentials. Cu3(Mo8O26)(H2O)2(OH)2(L1)4 had better HER activity than Ag4(Mo8O26)‐(L2)2.5(H2O). Their HER activities were related to the redox of [Mo8O26]4− polymolybdate anions. Qin et al. developed another polymolybdate (POM)‐based MOF for HER [47]. This novel 3D open structure was formed by connecting POM fragments as nodes and H3BTB (H3BTB = benzene tribenzoate) or H3BPT (H3BPT = [1,1′‐biphenyl]‐3,4′,5‐tricarboxylic acid) as ligands. POM‐based MOFs had excellent HER activity likely due to the combination of the redox of POM and the porosity of MOFs. Noticeably, the POM‐based MOFs were stable not only in air but also in acid or basic solutions. Among these catalysts, NENU‐500 ([TBA]3[ε‐PMoV8MoVI4O36(OH)4Zn4] [BTB]4/3·18H2O, TBA+= tetrabutylammonium ion) exhibited excellent HER activity with an onset overpotential of about 180 mV and Tafel slope of 96 mV dec−1 in 0.5 M H2SO4. What’s more, this catalyst could maintain its catalytic activity after 2000th cycles in 0.5 M H2SO4 solution.

Cobalt dithiolene species are highly efficient molecular catalysts for HER. Cloigh et al. integrated cobalt dithiolene into metal‐organic surface (MOS, [Co3(BHT)2]3+and [Co3(THT)2]3+ (BHT= benzenehexathiol, THT=triphenylene‐2,3,6,7,10,11‐hexathiol) materials to obtain effective catalysts for HER [48]. The prepared catalysts Co/MOS possessed a two‐dimensional layer structure with high surface/volume ratio, resulting in high charge transfer efficiency and high surface active site concentration. The catalysts, especially [Co3(BHT)2]3+, showed not only remarkable stability in acidic solutions, but also excellent HER catalytic activity under wide pH conditions, and the current densities at 0.5 V (vs. Standard Hydrogen Electrode, RHE) increased with decreasing electrolyte pH. The mechanism of HER of the MOS involved Co3+/Co2+ redox reactions, following protonation of the S sites on the ligands.

Recently, MOFs also have been combined with 2D materials, such as graphene and metal dichalcogenide nanosheets, to form nanocomposites with unique properties and wide applications [49]. A Cu‐MOF/graphene oxide (GO) nanocomposite catalyst for HER was successfully synthesized via a solvothermal process [27]. The Cu‐MOF was prepared using Cu(NO3)2, 1,4‐benzenedicarboxylic acid and triethylene‐diamine as precursors. The integration of Cu‐MOF with graphene oxide (GO) can effectively enhance the electron transfer, which further significantly improve HER activity. It was also found that the GO content affected the HER activity of the nanocomposite catalysts. The optimized GO content was about 8%. The HER current density of the (GO 8 wt%) Cu‐MOF was high up to -30 mA cm−2 at an overpotential of -0.2 V in N2‐saturated 0.5 M H2SO4, whereas the overpotential of 20 wt% Pt was -0.06 V at the current density of -30 mA cm−2.

Figure 1.

Schematic illustration of (A) synthesis of Co‐P/NC [50]; (B) formation of CoP NPCs and CoP NRCs derived from ZIF‐67‐Co [51]; (C) the space‐confined synthesis of MoS2/3D‐NPC composites and its application in HER [53].

Because MOFs possess structural diversity, high specific surface area, and large pore volume, MOFs are regarded as ideal precursors to prepare various inorganic nanomaterials as electrocatalysts, such as porous carbon, metal‐oxide nanoparticles/porous carbon or metallic nanoparticles/carbon. It has been reported that cobalt species are active for HER. Among them, cobalt phosphides as HER electrocatalysts have been widely investigated in recent years. You et al. prepared a Co‐P/NC catalyst by embedding CoPx nanoparticles into N‐doped carbon through carbonizing Co‐ZIF‐67 and a subsequent phosphidation (Figure 1A) [50]. After carbonization, the obtained Co‐ZIF‐67 inherited the morphology of polyhedron‐like ZIF‐67, in which the metallic Co nanoparticles were wrapped by the porous N‐doped carbon shells. Through the following phosphidation, the products also preserved the polyhedron‐like morphology of ZIF‐67, while the metallic Co nanoparticles were transformed into CoPx nanoparticles. The optimal Co‐P/NC catalyst afforded a high specific surface area of 183 m2 g−1and a high pore volume of 0.276 cm3 g−1. The overpotentials of the catalyst at 10, 20, and 100 mA cm−2 were ∼154, 173, and 234 mV respectively in 1 M KOH solution, indicating an outstanding HER activity. Such high HER activity can be attributed to its unique structures, such as abundant HER active species (CoPx and N‐doped carbon), 3D interconnected mesopores, and the porous N‐doped carbon shells. Jiang et al. synthesized several cobalt phosphides by directly phosphating Co‐ZIF‐67 under mild conditions [51]. The CoP nanorod assemblies (NRAs) were obtained by calcined precursor in N2 atmosphere, while the CoP nanoparticle assemblies (NPAs) were obtained by thermally treated precursor air atmosphere (Figure 1B). The CoP NRAs showed better HER electrocatalytic performance than CoP NPAs. The CoP NRAs only needed overpotential of about 181 mV at 10 mA cm−2 in 0.5 M H2SO4, whereas the overpotential for CoP NPAs was 393 mV at 10 mA cm−2. After 1000th CV cycles, CoP NRAs and CoP NPAs both have slight loss of initial current density. The high efficiency of HER for CoP might be related to the charge transfer from Co to P, similar to the charge transfer process between hydride‐acceptor and proton‐acceptor of hydrogenase.

In addition to Co species, Ni species are also considered to be active for HER. Nickel phosphides (Ni2P and Ni12P5) were prepared by phosphatizing Ni‐MOF (Ni‐BTC) in mild conditions [52]. During the preparation of Ni‐BTC, nickel nitrates and BTC (benzene‐1,3,5‐tricarboxylic acid) were used as Ni sources and organic ligands, respectively. Then Ni2P and Ni12P5 were obtained by phosphating from sodium hypophosphite at 275 and 325°C, respectively. The prepared Ni12P5 nanoparticles show similar morphology to the Ni2P nanoparticles. However, the average diameter of the Ni2P (25 nm) was smaller than that of the Ni12P5 (80 nm). The HER activity of the Ni2P was better than that of the Ni12P5 nanoparticles in 0.5 M H2SO4, due to the similar effects of CoP on the HER activity [51]. The durability of the Ni2P nanoparticles was further evaluated by chronoamperometric durability test, and current density was reduced to 75% of the original value after 6 h test.

The two‐step synthesis method was also utilized by Liu et al. to prepare MoS2‐based HER catalyst with 3D hierarchical structure, in which MoS2 nanosheets grew in the nanopores of MOFs‐derived 3D carbons (MoS2/3D‐NPC) (Figure 1C) [53]. An Al‐PCP was formed first by a solvothermal process, and a 3D‐NPC was obtained through calcination. The 3D‐NPC possessed randomly assembling nanopores, benefiting the intercalating of Mo precursors. After a solvothermal process, The MoS2/3D‐NPC catalyst was finally obtained, where MoS2 nanosheets were dispersed well in nanoporous 3D‐NPC. This unique 3D hierarchical structure can provide exposed active site and enhanced conductivity, thus resulting in the high performance for HER. The MoS2/3D‐NPC catalyst requires overpotentials of 180 and 210 mV to achieve current densities of 1 and 10 mA cm−2 in 0.5 M H2SO4 solution, respectively.

In addition to MoS2, MoCx has also been widely studied for HER. The porous MoCx nano‐octahedrons were prepared through an MOFs‐derived strategy [54]. Cu‐based MOF [HKUST‐1; Cu3 (BTC)2(H2O)3] was first used to hold the Mo‐based Keggin‐type POMs (H3PMo12O40), forming an MOF, i.e., NENU‐5 nano‐octahedrons. The MoCx‐Cu intermediate was prepared by carbonizing NENU‐5 in inert atmosphere, and the final MoCx porous nanoparticles were obtained by removing Cu by Fe3+ etching. The obtained porous MoCx catalysts were evaluated for HER in both acidic and basic aqueous solutions. The polarization curves showed that the overpotentials of this catalyst were only 142 and 151 mV to deliver a current density of 10 mA cm−2 in acidic and alkaline solutions, respectively.

In summary, the highly porous MOFs containing redox sites have been developed as HER catalysts. However, their low conductivity limits their electrocatalytic performance for HER. Their combination with highly conductive substrates is an efficient way to enhance their HER activities. Furthermore, many researches have demonstrated that the MOFs‐derived catalysts may exhibit greatly improved performance in HER. These MOFs‐derived catalysts were obtained mainly through carbonization and a following phosphidation or sulfuration of MOFs containing active metal species (usually Co, Ni and Mo) at appropriate temperatures in inert or air atmosphere. Due to the active species, the novel porous structures, the enhanced conductivity and the protection from the heteroatom doped carbons, these MOF‐derived catalysts usually exhibit higher activity and stability as compared to the pristine MOFs catalysts. Hence, because of these unique structures and properties, the MOF‐derived materials benefit the development of HER catalysts with low costs and high performance.

Advertisement

5. MOFs‐based catalysts for OER

MOFs possess high specific surface area and tunable pore structures, and they are easily functionalized by different metal centers and organic linkers, highlighting their great potentials in electrocatalysis. Similar to those utilized for HER, MOFs have already been developed for OER catalysis basically in two ways: direct and indirect ones. In the first case, MOFs will be received directly as electrocatalysts, and therefore element selection plays a key role to construct coordination modes which should facilitate oxygen‐species adsorption/desorption and water dissociation. While in the second case, MOFs will be transformed into other composite materials in which metal, in the form of metallic–, oxide– or nitrite–phase, is associated with carbon. Despite having potential, MOFs‐derived materials are at the expense of MOFs. For instance, the specific surface area decreased and the ordered porosities were destroyed. Therefore, structure modification is crucial to make the materials maintain higher specific surface area and hieratical pore structures through pyrolysis in order to maximize the catalytic site density and provide accessible channels for mass transfer.

Regarding element selection, metal‐O, metal‐N, or mixed N‐metal‐O coordination modes have already been developed to construct OER catalysts. In MOFs, the metal ion redox typically involves in the electrochemical process, and the electron‐accepting ability of the nitrogen and oxygen atoms in organic linkers may polarize the adjacent metal atoms to afford better catalytic performance.

Babu et al. [55] investigated the electrocatalytic activity of commercially available Fe(BTC) MOF (BasoliteTM F300), with BTC=bezene‐1,3,5‐tricarboxylate. When studying the hydroxide on the voltammetric responses for Fe(BTC), a well‐defined hydroxide oxidation response was clearly observed if NaOH was added to a 0.1 M KCl background electrolyte. Such an oxidation current was ascribed to OER current. However, such a current could be seen only by pre‐scanning the Fe(BTC)‐modified electrode lower than 0.5 V vs. saturated calomel electrode (SCE), a reduction potential zone where the hydrous iron oxide intermediates could be formed. Fe(III/II) played a key role in determining the electrochemical behavior of Fe(BTC).

Gong et al. [56] prepared two CoII metal‐organic frameworks, complex 1 (using 4,5‐di(4′‐carboxylphenyl) phthalic acid and 4,4′‐bipyridine as co‐ligands) and complex 2 (using 4,5‐di(4′‐carboxylphenyl)phthalic acid and azene as co‐ligands) as electrocatalysts. The OER catalytic activity of these two complexes in an aqueous buffer solution (pH = 6.8) was investigated. The results showed that complex 1 possessed an oxidation peak of water at a more negative potential of 1.02 V vs. SCE accompanied by a much higher OER current than bare glassy‐carbon (GC) electrode and ligand‐modified glassy‐carbon electrode. As confirmed by electrochemical impedance spectra, complex 1‐modified GC showed lower charge‐transfer resistance than the bare GC, indicating that the framework promoted the charge‐transfer, thus the OER catalytic activity. When applied complex 2, the 2‐modified GC showed lower OER overpotential and higher OER current than complex 1‐modified GC. Such an enhancement could be ascribed to the better charge build‐up of complex 2 than that of complex 1, further due to the different co‐linkers and frameworks in the two MOFs.

To verify the Co contributions for OER in MOFs, Yin’s group [57] first synthesized an MOF(Fe) catalyst by hydrothermal process using FeIII as metal precursor and 1,3,5‐BTC as organic ligand. The synthesized MOF(Fe) afforded a specific surface area up to 1600 m2 g−1. MOF(Fe) afforded excellent OER activity with a delivered current density of 2.30 mA·cm−2 at 0.90 V vs. Ag/AgCl in 0.1 M KOH, higher than the activated carbon (Super P). Further, they prepared an MOF(Fe/Co) as ORR/OER catalyst by a hydrothermal process using Fe and Co as mixed metal precursors, and the same organic ligand [58]. Surprisingly, MOF(Fe/Co) exhibited an improved OER activity with a delivered current density of 2.97 mA·cm−2 at 0.90 V vs. Ag/AgCl than the aforementioned MOF(Fe). Although MOF(Fe/Co) possessed lower specific surface area (∼1070.1 m2·g−1) than MOF(Fe), MOF(Fe/Co) exhibited better OER activity than MOF(Fe) in 0.1 M KOH electrolyte, primarily due to the aid of Co species.

Wang et al. [59] synthesized a cobalt‐based ZIF (Co‐ZIF‐9) by assembling Co ions with benzimidazolate ligands for water oxidation. Co‐ZIF‐9 had open‐framework structure, in which Co ions were coordinated by N atoms in benzimidazolate linkers. It had been found that this catalyst was thermodynamically feasible for catalytically oxidizing water molecule by density functional theory (DFT) calculation because Co ions had redox function and could bond ‐OH resulting in low activation barriers. In addition, the nearby benzimidazolate also promoted the OER reaction by accepting the eliminated ‐H atoms. Thus, Co‐ZIF‐9 showed obvious OER activity in wide pH basic solutions. Besides, the activity stability of Co‐ZIF‐9 was also investigated and there was no obvious deactivation in current density after 25 h test in potassium phosphate buffer.

Tan et al. [60] synthesized an MOF‐74 to study the water dissociation mechanism on open metal sites in MOFs through in situ IR spectroscopy and first‐principles calculations (Figure 2). The in situ IR spectroscopy provided a direct evidence of water reaction occurred on the metal centers. The water dissociation mechanism in MOF‐74 primarily depended on two aspects: (1) the covalent bond between water and metal center and (2) the hydrogen bonding between the O atoms of the linker and the H atoms of the water molecule. This work was of significance to demonstrate the coordinatively unsaturated metal centers as active sites for water dissociation.

Figure 2.

Schematic illustration of the OER catalytic mechanism on MOF‐74 [60].

Based on the above results, it has been theoretically or experimentally demonstrated that bare MOFs can be designed by element selection to show excellent OER catalytic activity. However, compared with the noble metal‐based and transition metal oxide‐based electrocatalysts, MOFs are still insufficient to catalyze OER with higher current probably due to their intrinsic poor electronic conductivity. Therefore, Loh et al. [27] prepared a Cu‐centered MOF from copper nitrate trihydrate, 1,4‐benzenedicarboxylic acid and triethylene‐diamine in the presence of GO. An optimal composition (GO 8 wt%) Cu‐MOF afforded an OER onset potential of 1.19 V vs. RHE in acid electrolytes, a ∼200 mV positive shift than that of pure Cu‐MOF. The 8 wt% GO‐incorporated Cu‐MOF also showed fast OER kinetics with a smaller Tafel slope of 65 mV·dec−1 than pure Cu‐MOF (89 mV·dec−1). The results demonstrated that the enhanced electrocatalytic properties and stability in acid of the GO‐MOF composite was due to the unique porous scaffold structure, improved charge transport, and synergistic interactions between the GO and MOF.

Because of the high specific surface area and the well‐defined porosities, MOFs are excellent catalyst support materials. Therefore, doping MOFs with other active species in the forms of oxides, ions, and complexes can open up a new route for MOFs in OER electrocatalysis.

Yin et al. [61] prepared α‐MnO2/MIL‐101(Cr) catalyst through a hydrothermal process. In this catalyst, α‐MnO2 particles were embedded in MIL‐101(Cr) matrix, resulting in strong interactions between α‐MnO2 and MIL‐101(Cr). The OER catalytic activity of the composite was tested using a carbon paper containing this catalyst as working electrode. The results revealed that α‐MnO2/MIL‐101(Cr) composite afforded an excellent OER catalytic activity with a delivered current density of 23.67 mA cm−2 at 0.9 V in 0.1 M KOH solution which was about two times higher than that of pure α‐MnO2 under the same conditions. The high specific surface area of MIL‐101(Cr) and abundant micropores of MIL‐101(Cr) were advantageous for the diffusion of electrolyte and the high dispersion of α‐MnO2 particles made it easy to contact the electrolyte, resulting in enhanced OER catalytic activity.

Subsequently, Yin’s group [62] decorated MIL‐101(Cr) with Co ions with various oxidization states through impregnation followed by post‐treatment under oxidant or reducing reagents. The Co species were highly dispersed on the MOF surface and showed various CoIII/CoII ratios. Primary results demonstrated that the OER activity is related to the CoIII contents since the catalyst showed better OER catalytic activity as the surface CoIII content increased. The reason might be the CoIII species could promote the OH- adsorbed onto the electrocatalyst surface. In addition, the porous and open structures of MIL‐101(Cr) support were in favor of the contact between oxygen species and the active Co sites.

Wang et al. [19] decorated a Zr‐MOF, namely UiO‐67, with 1‐3 Ir‐containing complexes to form MOF 1‐3, respectively. Many MOFs lack stability in water, but UiO‐67 is one of the exceptions. The OER performance of the samples was investigated in a pH = 1 solution with Ce4+ as an oxidant. The Ce4+ was reduced to Ce3+, while water was oxidized to form oxygen. The results showed that after the incorporation of Ir‐containing complexes in UiO‐67, MOF 1‐3 were the effective water oxidation catalysts with turnover frequencies of up to 4.8 h−1. The parent UiO‐67 showed negligible OER catalytic activity, demonstrating the significance of the Ir‐containing dopants as active centers.

The OER current density at 10 mA·cm−2 is a criterion used to judge an OER catalyst since such a current density is an important metric for practical solar fuel production. Unfortunately, most of the aforementioned MOF‐based OER catalysts fail to reach such a current density even at high overpotentials. Apart from directly being the OER catalysts, MOFs can be converted into other forms of materials through pyrolysis. Through pyrolysis, the organic linkers can be transformed into carbon materials, and the well‐dispersed metal centers in the MOFs’ frameworks can be transformed into phosphate‐ or oxide‐phase but still maintain excellent dispersion in carbons. Such a strategy is helpful to obtain materials with excellent charge‐transfer properties.

You et al. [50] prepared a CoPx/NC catalyst derived from ZIF‐67 and its polyhedron‐like morphology survived from the pyrolysis. The specific surface area of CoPx/NC was up to 183 m2·g−1, and the pore volume is 0.276 m3·g−1. In OER activity test, the overpotential of CoPx/NC was about 354 mV at a current density of 10 mA cm−2, while IrO2 is about 368 mV at the same current density. Moreover, after 1000th continuous CV cycles, the overpotential of CoPx/NC at 10 mA cm−2 showed only slight difference, demonstrating the excellent OER activity durability of the catalyst.

Apart from CoPx, Co3O4 also has been reported to be active for OER, and can significantly improve the conductivity and stability properties of the catalysts when combined with carbon materials. The porous Co3O4‐based hybrids were typically obtained through a one‐step carbonization of Co‐MOFs. Li et al. [63] introduced MWCNT to MOFs and obtained the Co3O4@MWCNTs by carbonization and subsequent oxidation process. The thermal oxidation led to Co3O4‐N‐C active sites uniformly dispersed on MWCNTs (20–50 nm). And the introduction of MWCNT and the in‐situ N‐doped carbon carbonized from ligands could significantly improve the electronic conductivity of the catalysts. Thus, better OER activity was observed with an onset potential of only 1.5 V (vs. RHE). Ma et al. [64] synthesized MOF‐derived Co3O4‐carbon porous nanowire arrays. The Co3O4‐carbon were directly prepared on Cu foil as a working electrode. Since this electrode is binder‐free and carbon is formed in situ, the charge conductivity performance is greatly improved, resulting in excellent OER catalytic activity in 0.1 M KOH solution. It shows a sharp onset potential of 1.47 V (vs. RHE), very close to that of IrO2/C (1.45 V vs. RHE). The durability is also an important criterion for OER catalysts. The chronopotentiometric response at a current density of 10 mA·cm−2 was also recorded, and only 6.5% attenuation was observed within 30 h on Co3O4‐carbon, while that of IrO2/C is 4.7 times lager under the same condition.

The above results demonstrate that MOFs‐derived catalysts show much enhanced OER catalytic performance than pristine MOFs. There are likely several reasons for that. On one hand, carbon materials formed in situ or ex situ in catalysts can promote electronic conductivity and accelerate charge transfer. On the other hand, the OER active species can be well‐dispersed on carbon materials, resulting in improved OER activity. In addition, the strong interactions between OER active species and carbon materials stabilize the OER active site structures, thus leading to enhanced OER activity durability.

Advertisement

6. Conclusions

MOFs‐based electrocatalysts for HER and/or OER are rapidly developed in recent years due to unique structures of MOFs. These catalysts mainly include MOFs catalysts, MOFs supports for catalysts, and MOFs‐derived catalysts. Due to the fact that the pore structures and functions are tunable and devisable, it is convenient to directly design and construct the active sites for HER and/or OER in MOFs during the synthesis process. However, the vast majority of the synthesized MOFs suffer from poor electronic conductivity, leading to low electron transfer efficiency, which restricts catalytic performance. MOFs are highly porous materials and have ultrahigh specific surface area, thus they are regarded as the most promising support materials for catalysts. The active species for HER or OER can be well dispersed at the surfaces of MOFs or embedded in MOFs matrix, resulting in improved catalytic performance for HER or OER. However, MOFs are microporous materials with small aperture size (<2 nm).On one hand, the active species are difficult to be introduced in their pore channels. Hence, one cannot obtain the catalysts with high activity and stability for HER or OER. On the other hand, the accommodation of electrolyte in the pore channels is very limited. Thus, the active centers in MOFs cannot access the electrolytes with high efficiency. In addition, the poor electronic conductivity of MOFs is another drawback of these catalysts. Some post‐treatment methods can greatly improve the electrical conductivity of MOFs‐derived catalysts. However, the collapse of the pore structures of MOFs usually occurs during the preparation of the MOFs‐derived catalysts leading to a decreased specific surface area of the catalysts, which is adverse to the development of MOFs‐derived catalysts with high catalytic performance for HER or OER.

Although many MOFs‐based catalysts with high catalytic performance for HER and/or OER have been developed in recent years, there are still a lot of scientific and technical problems to solve before the developed MOFs‐based electrocatalysts can meet the requirements for commercialization. The problems mainly involve how to improve the electronic conductivity of MOFs, enlarge pore channels of MOFs to accommodate more electrolytes, limit the collapse of pore structure of MOFs, and maintain high specific surface area during pyrolysis. In addition, the reaction mechanisms of HER and OER, the transfer and diffusion properties of reactants and products, and the effects of electrolyte on the catalytic performance of MOFs‐based catalysts need to be clarified.

Advertisement

Acknowledgments

This work is supported by the National Natural Science Foundation of China (Project No. 21276018), the Natural Science Foundation of Jiangsu Province of China (Project No. BK20140268), the Fundamental Research Funds for the Central Universities (Project No. buctrc201526), and the Changzhou Sci&Tech Program (Project No. CJ20159006).

References

  1. 1. Ursúa A, Gandia LM, Sanchis P. Hydrogen production from water electrolysis: Current status and future trends. Proceedings of the IEEE. 2012;100:410‐426. DOI: 10.1109/JPROC.2011.2156750
  2. 2. Suntivich J, Gasteiger HA, Yabuuchi N. Design principles for oxygen‐reduction activity on perovskite oxide catalysts for fuel cells and metal‐air batteries. Nature Chemistry. 2011;3:546‐550. DOI: 10.1038/nchem.1069
  3. 3. Bashyam R, Zelenay P. A class of non‐precious metal composite catalysts for fuel cells. Nature. 2006;443:63‐66. DOI: 10.1038/nature05118
  4. 4. Grigoriev SA, Millet P, Fateev VN. Evaluation of carbon‐supported Pt and Pd nanoparticles for the hydrogen evolution reaction in PEM water electrolysers. Journal of Power Sources. 2008;177:281‐285. DOI: 10.1016/j.jpowsour.2007.11.072
  5. 5. Jiao Y, Zheng Y, Jaroniec M. Design of electrocatalysts for oxygen‐ and hydrogen‐involving energy conversion reactions. Chemical Society Reviews. 2015;44:2060‐2086. DOI: 10.1039/C4CS00470A
  6. 6. Pu Z, Liu Q, Jiang P. CoP nanosheet arrays supported on a Ti plate: An efficient cathode for electrochemical hydrogen evolution. Chemistry of Materials. 2014;26:4326‐4329. DOI: 10.1021/cm501273s
  7. 7. Zhang Z, Lu B, Hao J. FeP nanoparticles grown on graphene sheets as highly active non‐precious‐metal electrocatalysts for hydrogen evolution reaction. Chemical Communications. 2014;50:11554‐11557. DOI: 10.1039/C4CC05285D
  8. 8. Gu S, Du H, Asiri AM. Three‐dimensional interconnected network of nanoporous CoP nanowires as an efficient hydrogen evolution cathode. Physical Chemistry Chemical Physics. 2014;16:16909‐16913. DOI: 10.1039/C4CP02613F
  9. 9. Xie J, Zhang J, LiS, Grote F. Controllable disorder engineering in oxygen‐incorporated MoS2 ultrathin nanosheets for efficient hydrogen evolution. Journal of the American Chemical Society. 2013;135:17881‐17888. DOI: 10.1021/ja408329q
  10. 10. Chen Z, Cummins D, Reinecke BN. Core‐shell MoO3‐MoS2 nanowires for hydrogen evolution: A functional design for electrocatalytic materials. Nano Letters. 2011;11:4168‐4175. DOI: 10.1021/nl2020476
  11. 11. Faber MS, Lukowski MA, Ding Q. Earth‐abundant metal pyrites (FeS2, CoS2, NiS2, and their alloys) for highly efficient hydrogen evolution and polysulfide reduction electrocatalysis. The Journal of Physical Chemistry C. 2014;118:21347‐21356. DOI: 10.1021/jp506288w
  12. 12. McKone JR, Sadtler BF, Werlang CA. Ni‐Mo nanopowders for efficient electrochemical hydrogen evolution. ACS Catalysis. 2013;3:166‐169. DOI: 10.1021/cs300691m
  13. 13. Yin Z, Chen F. A facile electrochemical fabrication of hierarchically structured nickel‐copper composite electrodes on nickel foam for hydrogen evolution reaction. Journal of Power Sources. 2014;265:273‐281. DOI: 10.1016/j.jpowsour.2014.04.123
  14. 14. Wang D, Chen X, Evans DG. Well‐dispersed Co3O4/Co2MnO4 nanocomposites as a synergistic bifunctional catalyst for oxygen reduction and oxygen evolution reactions. Nanoscale. 2013;5:5312‐5315. DOI: 10.1039/c3nr00444a
  15. 15. Gong M, Zhou W, Tsai MC. Nanoscale nickel oxide/nickel heterostructures for active hydrogen evolution electrocatalysis. Nature Communications. 2014;5:1‐6. DOI: 10.1038/ncomms5695
  16. 16. Li Y, Hasin P, Wu Y. NixCo3-xO4 nanowire arrays for electrocatalytic oxygen evolution. Advanced Materials. 2010;22:1926‐1929. DOI: 10.1002/adma.200903896
  17. 17. Feng LL, Yu G, Wu Y. High‐index faceted Ni3S2 nanosheet arrays as highly active and ultrastable electrocatalysts for water splitting. Journal of the American Chemical Society. 2015;137:14023‐14026. DOI: 10.1021/jacs.5b08186
  18. 18. Liu Q, Jin J, Zhang J. NiCo2S4@ graphene as a bifunctional electrocatalyst for oxygen reduction and evolution reactions. ACS Applied Materials and Interfaces. 2013;5:5002‐5008. DOI: 10.1021/am4007897
  19. 19. Wang C, Xie Z, deKrafft KE. Doping metal‐organic frameworks for water oxidation, carbon dioxide reduction, and organic photocatalysis. Journal of the American Chemical Society. 2011;133:13445‐13454. DOI: 10.1021/ja203564w
  20. 20. Chen X Y, Zhao B, Shi W. Microporous metal‐organic frameworks built on a Ln3 cluster as a six‐connecting node. Chemistry of Materials. 2005;17:2866‐2874. DOI: 10.1021/cm050526o
  21. 21. Polshettiwar V, Varma RS. Aqueous microwave chemistry: a clean and green synthetic tool for rapid drug discovery. Chemical Society Reviews. 2008;37:1546‐1557. DOI: 10.1039/B716534J
  22. 22. Jin K, Huang X, Pang L. [Cu(i)(bpp)] BF4: The first extended coordination network prepared solvothermally in an ionic liquid solvent. Chemical Communications. 2002;2872‐2873. DOI: 10.1039/B209937N
  23. 23. Furukawa H, Cordova KE, O’Keeffe M. The chemistry and applications of metal‐organic frameworks. Science. 2013;341:1230444. DOI: 10.1126/science.1230444
  24. 24. Rosi N L, Eckert J, Eddaoudi M. Hydrogen storage in microporous metal‐organic frameworks. Science. 2003;300:1127‐1129. DOI: 10.1126/science.1083440
  25. 25. Zhang WX, Yang YY, Zai SB. Syntheses, structures and magnetic properties of dinuclear copper (II)‐lanthanide (III) complexes bridged by 2‐hydroxymethyl‐1‐methylimidazole. European Journal of Inorganic Chemistry. 2008;2008:679‐685. DOI: 10.1002/ejic.200701041
  26. 26. Moore EG, Samuel APS, Raymond KN. From antenna to assay: Lessons learned in lanthanide luminescence. Accounts of Chemical Research. 2009;42:542‐552. DOI: 10.1021/ar800211j
  27. 27. Jahan M, Liu Z, Loh KP. A Graphene oxide and copper‐centered metal organic framework composite as a tri‐functional catalyst for HER, OER, and ORR. Advanced Functional Materials. 2013;23:5363‐5372. DOI: 10.1002/adfm.201300510
  28. 28. Kitagawa S. Metal‐organic frameworks (MOFs). Chemical Society Reviews. 2014;43:5415‐5418. DOI: 10.1039/C4CS90059F
  29. 29. Yaghi OM, Li H. Hydrothermal synthesis of a metal‐organic framework containing large rectangular channels. Journal of the American Chemical Society. 1995;117:10401‐10402. DOI: 10.1021/ja00146a033
  30. 30. Furukawa H, Ko N, Go YB. Ultrahigh porosity in metal‐organic frameworks. Science. 2010;329:424‐428. DOI: 10.1126/science.1192160
  31. 31. Deng H, Grunder S, Cordova KE. Large‐pore apertures in a series of metal‐organic frameworks. Science. 2012;336:1018‐1023. DOI: 10.1126/science.1220131
  32. 32. Xia W, Mahmood A, Zou R. Metal‐organic frameworks and their derived nanostructures for electrochemical energy storage and conversion. Energy and Environmental Science. 2015;8:1837‐1866. DOI: 10.1039/C5EE00762C
  33. 33. Fang QR, Makal TA, Young MD. Recent advances in the study of mesoporous metal‐organic frameworks. Comments on Inorganic Chemistry. 2010;31:165‐195. DOI: 10.1080/02603594.2010.520254
  34. 34. Sun D, Ma S, Ke Y. An interweaving MOF with high hydrogen uptake. Journal of the American Chemical Society. 2006;128:3896‐3897. DOI: 10.1021/ja058777l
  35. 35. Li H, Eddaoudi M, O’Keeffe M. Design and synthesis of an exceptionally stable and highly porous metal‐organic framework. Nature. 1999;402:276‐279. DOI: 10.1038/46248
  36. 36. Park KS, Ni Z, Côté AP. Exceptional chemical and thermal stability of zeoliticimidazolate frameworks. Proceedings of the National Academy of Science of the United States of America. 2006;103:10186‐10191. DOI: 10.1073/pnas.0602439103
  37. 37. Mellot‐Draznieks C, Girard S, Férey G. Novel inorganic frameworks constructed from double‐four‐ring (D4R) units: Computational design, structures, and lattice energies of silicate, aluminophosphate, and gallophosphate candidates. Journal of the American Chemical Society. 2002;124:15326‐15335. DOI: 10.1021/ja020999l
  38. 38. Chui SSY, Lo SMF, Charmant JPH. A chemically functionalizable nanoporous material [Cu3(TMA)2(H2O)3]n. Science. 1999;283:1148‐1150. DOI: 10.1126/science.283.5405.1148
  39. 39. Qi Y, Luo F, Che Y. Hydrothermal synthesis of metal‐organic frameworks based on aromatic polycarboxylate and flexible bis (imidazole) ligands. Crystal Growth and Design. 2007;8:606‐611. DOI: 10.1021/cg700758c
  40. 40. Sun J K, Xu Q. Functional materials derived from open framework templates/precursors: synthesis and applications. Energy and Environmental Science. 2014 ;7:2071‐210. DOI: 10.1039/C4EE00517A
  41. 41. Huang SY, Ganesan P, Popov BN. Electrocatalytic activity and stability of Titania‐supported platinum‐palladium electrocatalysts for polymer electrolyte membrane fuel cell. ACS Catalysis. 2012;2:825‐831. DOI: 10.1021/cs300088n
  42. 42. Zou X, Zhang Y. Noble metal‐free hydrogen evolution catalysts for water splitting. Chemical Society Reviews. 2015;44:5148‐5180. DOI: 10.1039/C4CS00448E
  43. 43. Gong M, Dai H. A mini review of NiFe‐based materials as highly active oxygen evolution reaction electrocatalysts. Nano Research. 2015;8:23‐39. DOI: 10.1007/s12274‐014‐0591‐z
  44. 44. Zhang J, Zhao Z, Xia Z. A metal‐free bifunctional electrocatalyst for oxygen reduction and oxygen evolution reactions. Nature Nanotechnology. 2015;10:444‐452. DOI: 10.1038/nnano.2015.48
  45. 45. Silva P, Vilela SMF, Tome JPC. Multifunctional metal‐organic frameworks: From academia to industrial applications. Chemical Society Reviews. 2015;44:6774‐6803. DOI: 10.1039/C5CS00307E
  46. 46. Gong Y, Wu T, Jiang PG. Octamolybdate‐based metal‐organic framework with unsaturated coordinated metal center as electrocatalyst for generating hydrogen from water. Inorganic Chemistry. 2012;52:777‐784. DOI: 10.1021/ic3018858
  47. 47. Qin J S, Du DY, Guan W. Ultrastable polymolybdate‐based metal‐organic frameworks as highly active electrocatalysts for hydrogen generation from water. Journal of the American Chemical Society. 2015;137:7169‐7177. DOI: 10.1021/jacs.5b02688
  48. 48. Clough AJ, Yoo JW, Mecklenburg MH. Two‐dimensional metal‐organic surfaces for efficient hydrogen evolution from water. Journal of the American Chemical Society. 2014;137:118‐121. DOI: 10.1021/ja5116937
  49. 49. Li S, Yang K, Tan C. Preparation and applications of novel composites composed of metal‐organic frameworks and two‐dimensional materials. Chemical Communications. 2016;52:1555‐1562. DOI: 10.1039/c5cc09127f
  50. 50. You B, Jiang N, Sheng M. High‐performance overall water splitting electrocatalysts derived from cobalt‐based metal‐organic frameworks. Chemistry of Materials. 2015;27:7636‐7642. DOI: 10.1021/acs.chemmater.5b02877
  51. 51. Li L, Li X, Ai L. MOF‐derived nanostructured cobalt phosphide assemblies for efficient hydrogen evolution reaction. RSC Advances. 2015;5:90265‐90271. DOI: 10.1039/C5RA17427A
  52. 52. Tian T, Ai L, Jiang J. Metal‐organic framework‐derived nickel phosphides as efficient electrocatalysts toward sustainable hydrogen generation from water splitting. RSC Advances. 2015;5:10290‐10295. DOI: 10.1039/C4RA15680C
  53. 53. Liu Y, Zhou X, Ding T. 3D architecture constructed via the confined growth of MoS2nanosheets in nanoporous carbon derived from metal‐organic frameworks for efficient hydrogen production. Nanoscale. 2015;7:18004‐18009. DOI: 10.1039/C5NR03810C
  54. 54. Wu HB, Xia BY, Yu L. Porous molybdenum carbide nano‐octahedrons synthesized via confined carburization in metal‐organic frameworks for efficient hydrogen production. Nature Communications. 2015;6:6512. DOI: 10.1038/ncomms7512
  55. 55. Babu KF, Kulandainathan MA, Katsounaros I. Electrocatalytic activity of BasoliteTM F300 metal‐organic‐framework structures. Electrochemistry Communications. 2010;12:632‐635. DOI: 10.1016/j.elecom.2010.02.017
  56. 56. Gong Y, Hao Z, Meng J. Two CoII metal‐organic frameworks based on a multicarboxylate ligand as electrocatalysts for water splitting. ChemPlusChem. 2014;79:266‐277. DOI: 10.1002/cplu.201300334
  57. 57. Song G, Wang Z, Wang L. Preparation of MOF (Fe) and its catalytic activity for oxygen reduction reaction in an alkaline electrolyte. Chinese Journal of Catalysis. 2014;35:185‐195. DOI: 10.1016/S1872‐2067(12)60729‐3
  58. 58. Wang H, Yin F, Li G. Preparation, characterization and bifunctional catalytic properties of MOF (Fe/Co) catalyst for oxygen reduction/evolution reactions in alkaline electrolyte. International Journal of Hydrogen Energy. 2014;39:16179‐16186. DOI: 10.1016/j.ijhydene.2013.12.120
  59. 59. Wang S, Hou Y, Lin S. Water oxidation electrocatalysis by a zeolitic imidazolate framework. Nanoscale. 2014;6:9930‐9934. DOI: 10.1039/C4NR02399D
  60. 60. Tan K, Zuluaga S, Gong Q. Water reaction mechanism in metal organic frameworks with coordinatively unsaturated metal ions: MOF‐74. Chemistry of Materials. 2014;26:6886‐6895. DOI: 10.1021/cm5038183
  61. 61. Yin F, Li G, Wang H. Hydrothermal synthesis of α‐MnO2/MIL‐101 (Cr) composite and its bifunctional electrocatalytic activity for oxygen reduction/evolution reactions. Catalysis Communications. 2014;54:17‐21. DOI: 10.1016/j.catcom.2014.05.006
  62. 62. He X, Yin F, Li G. A Co/metal‐organic‐framework bifunctional electrocatalyst: The effect of the surface cobalt oxidation state on oxygen evolution/reduction reactions in an alkaline electrolyte. International Journal of Hydrogen Energy. 2015;40:9713‐9722. DOI: 10.1016/j.ijhydene.2015.06.027
  63. 63. Li X, Fang Y, Lin X. MOF derived Co3O4 nanoparticles embedded in N‐doped mesoporous carbon layer/MWCNT hybrids: extraordinary bi‐functional electrocatalysts for OER and ORR. Journal of Materials Chemistry A. 2015;3:17392‐17402. DOI: 10.1039/C5TA03900B
  64. 64. Ma TY, Dai S, Jaroniec M. Metal‐organic framework derived hybrid Co3O4‐carbon porous nanowire arrays as reversible oxygen evolution electrodes. Journal of the American Chemical Society. 2014;136:13925‐13931. DOI: 10.1021/ja5082553

Written By

Fengxiang Yin, Xiao Zhang, Xiaobo He and Hao Wang

Submitted: 20 November 2015 Reviewed: 20 June 2016 Published: 12 October 2016