Hao
Wan
abc,
Xiaohe
Liu
*b,
Haidong
Wang
c,
Renzhi
Ma
*a and
Takayoshi
Sasaki
*a
aInternational Center for Materials Nanoarchitectonics (WPI-MANA), National Institute for Materials Science (NIMS), Namiki 1-1, Tsukuba, Ibaraki 305-0044, Japan. E-mail: MA.Renzhi@nims.go.jp; SASAKI.Takayoshi@nims.go.jp
bState Key Laboratory of Powder Metallurgy and School of Materials Science and Engineering, Central South University, Changsha, Hunan 410083, P. R. China. E-mail: liuxh@csu.edu.cn
cSchool of Minerals Processing and Bioengineering, Central South University, Changsha, Hunan 410083, P. R. China
First published on 14th February 2019
Electrocatalysis for clean energy conversion, including water splitting (hydrogen and oxygen evolution) and oxygen reduction, has been considered as a pivotal strategy to alleviate the increasing energy crisis and environmental pollution derived from the overuse of nonrenewable fossil fuels. Since the current electrocatalysts are usually based on high-cost and scarce noble metal elements (Pt, Ru, and Ir), developing low-cost and earth-abundant catalysts is of great practical promise for realizing industry-scale applications. In this regard, electrocatalysts based on 3d transition metal elements (Mn, Fe, Co, and Ni, etc.) have been proposed as a class of prospective materials due to their abundance and high activity. In this work, recent advances in developing high-performance nanostructured electrocatalysts for sustainable clean energy conversion are briefly reviewed, with a particular focus on morphology design, composition tuning, surface engineering and metal coordination symmetry/geometry control. The latest studies indicate that carefully designed nanostructures based on 3d transition metal elements can attain comparable electrocatalytic performance to the commercial noble metal-based counterparts. This review may offer new insights into the rational design of nanostructures with further improved electrocatalytic activity and pathway selectivity to achieve the ultimate goal of realizing renewable electrochemical energy conversion.
Compounds based on 3d transition metals (TMs, such as Mn, Fe, Co and Ni) constitute a typical class of non-precious metal-based materials. Due to the unique electronic configurations ([Ar]3d5–84s2), they show facile redoxable properties and have been widely used in supercapacitors, batteries, sensors, etc.10–13 Furthermore, in the past decade, various kinds of TM-based materials, including pure metals, alloys, carbides, nitrides, bismuthides, chalcogenides, phosphides, oxides, hydroxides and/or their composites, have also been investigated for use as electrocatalysts.14–18 Although these TM-based catalysts have shown catalytic performance superior to that of non-TM-based materials (e.g. carbon) and even comparable to that of noble metal-based benchmarks, considerable challenges and issues still need to be addressed. Therefore, exploiting TM-based electrocatalyst candidates with high activity and selectivity remains a rigorous challenge.
There are some excellent reviews on electrocatalysts for highly efficient electrocatalysis. However, they mainly focus on metal-free nanostructures,19 materials based on specific species (e.g. oxides or dichalcogenides)20–22 or particular electrocatalytic processes (such as the OER).23,24 Few of them have paid sufficient attention to the versatile strategies for improving the activity of electrocatalysts. To achieve the high performance for industrial use, it is therefore indispensable to systematically survey newly emerging advances which will bring new methodologies.
Typically, two strategies are widely adopted for improving the performance of electrocatalysts.25 (i) The first approach is to expose more accessible active sites for catalytic reactions. (ii) The second way is to enhance the intrinsic reactivity of each active site. In practice, these two strategies are not mutually independent; sometimes, they can be applied simultaneously.
In this respect, we review several typical case investigations on recent advances in improving the electrocatalytic OER, HER and/or ORR performance of TM-based electrocatalysts, e.g., morphology design, composition tuning, surface engineering and metal coordination control. The versatile methodologies and rationales of 3d TM-based materials reviewed in this article may also be applicable for the fundamental investigations and practical uses of other precious metal-free electrocatalysts.
Fig. 2 Electrocatalytic performance of unary 3d TM-based catalysts. Comparison of required overpotential for single-metal TM hydro(oxy)oxides in the OER. Reproduced with permission from ref. 26, Copyright 2012, Nature Publishing Group. (b) HER polarization curves of metal selenides. Reproduced with permission from ref. 30, Copyright 2013, the Royal Society of Chemistry. (c) ORR polarization curves and (d) the corresponding j@0.5 V vs. RHE, half-wave potential (E1/2), kinetic-limiting current density (Jk) and electron transfer number (n) of single-metal carbides and the Pt/C material (GNRs in (c) and (d) refers to graphene nanoribbons). Reproduced with permission from ref. 31, Copyright 2015, American Chemical Society. |
Moreover, to further improve the electrocatalytic performance of the TM-based materials, some advances have been achieved recently in terms of morphology, composition, surface and coordination, which will be reviewed below.
To date, diverse nanostructures have been developed for rapid electrochemical conversion processes. As summarized in Fig. 3, these nanostructures mainly have hollow interiors and low dimensions for more accessible active sites and better electrical conductivity. For 0-dimensional (0D) nanoparticles, when the sizes in three dimensions are reduced to the nanoscale, some structural advantages will emerge, resulting in high electrocatalytic activity. Some nanoarchitectures are therefore elaborately designed. A current-induced high-temperature (≈2470 K) rapid thermal shock (≈12 ms) was adopted on iron pyrite (FeS2) bulks, leading to the successful preparation of FeS2 nanoparticles stabilized on reduced graphene oxide (indicated as nano-FeS2-rGO). Compared with its microsized counterparts (denoted as micro-FeS2-rGO), nano-FeS2-rGO has shown a lower onset potential, and a lower overpotential at a current density of 10 mA cm−2 with a smaller Tafel slope (66 mV dec−1).33 The enhanced activity is ascribed to the morphological advantage of the ultrafine nanoparticles, which may provide more accessible active sites and faster electron transfer. Similarly, a progressive cation exchange reaction of the octahedral shaped oxide precursors was developed for the synthesis of Co9−xNixS8 octahedral nanocages with a size of ∼80 nm.34 Because of the structural merit, they needed a low overpotential of only 364 mV to reach the anodic current density of 10 mA cm−2, even surpassing the commercial RuO2 material. When the nanoparticles were further minimized to the atomic size, single-atom catalysts were achieved. Since the first report on single-atom Pt (Pt1) supported on FeOx by Zhang's group in 2011,35 single-atom catalysts have attracted great research interest. In contrast to the conventional electrocatalysts, each active metal site in single-atom nanostructures is isolated from others, resulting in higher atom utilization and catalytic activity. Different from the Pt1/FeOx material in the first report, the TM-based single-atom catalysts usually use electronically conductive carbon materials, e.g., graphene or carbon nanotubes, as supports derived from the corresponding metal–organic precursors.36,37 Due to the structural feature, i.e. all metal atoms are isolated with other congeneric sites and bridged by neighboring carbon and/or nitrogen atoms, a high utilization ratio of nearly 100% for the metal sites can be realized by the single-atom structures. However, on account of the structural characteristics, the loading mass of the TM sites in single-atom catalysts is generally low (<5 wt%). It will be a great challenge to increase the loading of active metal sites in single-atom nanostructures.
Fig. 3 Nanostructured catalysts for electrochemical conversions: 0D nanoparticles33 and single-atom catalysts;36 1D nanorods,40 nanowires41 and nanoneedles;42 2D nanoplates,47 and nanosheets (the inset shows the visible Tyndall effect of the nanosheet dispersion);48 hollow cuboids,52 spheres53 and cages.54 Adapted with permission: nanoparticles: ref. 33, Copyright 2017, hollow cuboids, ref. 52, Copyright 2016, hollow spheres, ref. 53, and hollow cages: ref. 54, Copyright 2017, John Wiley & Sons, Inc.; sing-atom catalysts: ref. 36, Copyright 2018, nanoplates: ref. 47, Copyright 2014, Nature Publishing Group; nanorods: ref. 40, Copyright 2012, nanowire: ref. 41, Copyright 2017, nanosheets: ref. 48, Copyright 2015, American Chemical Society. |
Although 0D nanoarchitectures have attained impressively high performance in electrocatalytic processes, substantial endeavors have also been made in exploiting 1-dimensional (1D) electrocatalysts. Compared with 0D electrocatalysts, 1D nanostructures (such as nanorods, nanofibers, and nanowires) are beneficial for the extensive exposure of specific facets due to the preferential crystal growth, which may be advantageous for highly efficient electrocatalysis. The 1D structural features may also bring about path-directing effects in catalyst electrodes, greatly enhancing the electron transport properties.38,39 Moreover, it is notable that for single-crystalline 1D motifs, almost no defects exist on the surfaces, which endows these materials with ideal long-term electrocatalytic durability. For example, a solvothermal process at an elevated temperature of 190 °C was adopted for high-crystalline 1D CaMn2O4 nanorods.40 Due to the structural features, CaMn2O4 nanorods have been characterized to exhibit comparable onset potential and exchange current density to the benchmark Pt/C material for the ORR. In addition, as a result of the high crystallinity, in a long-term electrocatalytic process, the current decay for CaMn2O4 nanorods is only 12.3%, lower than that for commercial Pt/C materials (17.7%), revealing a better durability. In particular, conductive matrices, such as carbon cloth (CC), have been widely used, which can promote directing growth along the orientated directions, resulting in 1D nanoarrays. Since these nanoarrays are stabilized on the substrates, aggregation is effectively avoided so that the electrocatalytic reactants and intermediates can move freely though the interspace, which has been well demonstrated by NiMoP2 nanowires/CC.41 Due to the excellent performance in both OER and HER, NiMoP2 nanowires/CC were adopted as bifunctional electrocatalysts for overall water splitting, achieving galvanostatic electrolysis at a current density of 10 mA cm−2 at a low potential of ∼1.5 V for over 24 h with no obvious potential fluctuation. Similarly, because of beneficial structural features, 1D NiCo2Px nanoneedles/carbon have also been proposed to show comparable HER activity towards commercial Pt/C catalysts in both acidic and alkaline electrolytes.42 The X-ray diffraction and microscopic analyses of the sample after a chronoamperometry measurement for up to 30 h further confirmed that both the composition and the 1D morphology were well preserved. These 1D nanoarrays with peculiar structural merits deliver considerable electrocatalytic activity, stability and durability, promising more possibilities as candidates for replacing commercial electrocatalysts.
For 2-dimensional (2D) nanomaterials, exfoliated from layered crystalline structures, a larger specific surface area is expected. When a dimensional confinement is imposed, these materials will exhibit facile electron and/or ion transfer properties in comparison to the bulk structures,43–45 which is beneficial for the electrocatalytic processes. Stimulated by this point, many 2D nanostructures have been synthesized, such as NiCo and/or CoCo layered double hydroxide (LDH) nanoplates via a topochemical approach46 or NiFe LDH through a hydrothermal process.47 When the LDH nanoplates were further exfoliated into single-layered nanosheets, the electrocatalytic activity increased sharply. The overpotential at an anodic exchange current density of 10 mA cm−2 was reduced to 0.30 V for the exfoliated NiFe LDH nanosheets, much lower than that for the bulk counterparts (0.35 V). As is well known, poor electronic conductivity limits the high activity of hydroxides. The combination of electrocatalytically active materials and conductive matrices is an effective approach for further improving the electrocatalytic performance. In common composites or hybrids, conductive matrices and the electrocatalytically active phases are in rather low contact areas, which is not favorable for instant electron or ion transfer. However, 2D superlattice nanostructures may provide a substantial advantage. In our previous work, as displayed in Fig. 4, positively charged single-layered NiFe LDH nanosheets were obtained by the exfoliation of the corresponding bulk counterparts.48 Meanwhile, negatively charged graphene oxide (GO) nanosheets were prepared by the delamination of natural graphite bulks using a modified Hummers' method.49 Then, the superlattice nanostructure was formed by the flocculation of electrocatalytically active NiFe LDH nanosheets and electronically conductive reduced GO (rGO) nanosheets according to an area-matching model based on electrostatic interactions. The nanostructured composites combined the characteristic advantages of redox-active NiFe LDH and conductive graphene-related materials. As a result, the NiFe LDH/rGO superlattice nanostructure reached an exchange current density of 10 mA cm−2 at a rather low overpotential of 0.217 V. Furthermore, a small Tafel slope of 40 mV dec−1 and a high turnover frequency of ∼0.1 s−1 at an overpotential of 300 mV were achieved, far surpassing those of the NiFe LDH/GO superlattice and NiFe LDH nanosheets and making this material the state-of-the-art NiFe-based electrocatalyst. Due to the excellent activity in the kinetically slow OER, overall water splitting, as shown in Fig. 4i, could readily be driven by a 1.5 V size AA battery using the NiFe LDH/rGO superlattice as a bifunctional HER and OER electrocatalyst. Similarly, exfoliated NiMn LDH nanosheets were also prepared, which yielded a current density of 10 mA cm−2 at an overpotential of 0.36 V, with a Tafel slope of 65 mV dec−1. Upon further flocculation with electronically conductive rGO nanosheets, the NiMn LDH/rGO superlattice required an overpotential of only 0.26 V to reach an exchange current density of 10 mA cm−2 in the OER, surpassing most Mn-based electrocatalysts.50 In addition, MoS2/rGO superlattices were prepared by modifying and reversing the charge state of rGO, which resulted in an onset potential of 88 mV vs. RHE,51 much lower than that for pure MoS2 nanosheets (258 mV vs. RHE) and superior to those of other reported MoS2-based HER electrocatalysts.
Fig. 4 NiFe LDH nanosheet/rGO superlattice with high OER activity. (a) Schematic procedure for preparing the NiFe LDH/rGO superlattice. (b) XRD patterns for NiFe LDH/GO and NiFe LDH/rGO superlattices. (c) TEM and (d) HRTEM images of NiFe LDH/rGO superlattices. (e–h) Polarization curves, Tafel slopes, and overpotentials at a current density of 10 mA cm−2 and calculated TOF values for NiFe LDH/GO, NiFe LDH/rGO superlattices and NiFe LDH nanosheets in the electrocatalytic water oxidation process. (i) Photograph of the overall water splitting device using the NiFe LDH/rGO superlattice as a bifunctional electrocatalyst. Reproduced with the permission of ref. 48, Copyright 2015, American Chemical Society. |
Nanomaterials with hollow interiors have also been developed for efficient electrocatalytic reactions. For electrocatalysts with hollow nanostructures, compared with their solid counterparts, more accessible active sites can be provided, and good electron and mass transfers are realized due to the nanoscale wall thickness. Therefore, hollow nanostructures can endow the electrocatalysts with favorable reaction kinetics, i.e., a small Tafel slope. Gao et al. adopted Ni–Co based hollow microcuboid precursors as sacrificial templates for the hierarchical NiCo2O4 hollow microcuboids.52 These microcuboids possess two kinds of porosities: one is the mesopores which could be described by the BET theory, and the other is macropores as the hollow interior. Due to the hollow characters, NiCo2O4 microcuboids showed high activity as a bifunctional electrocatalyst for both HER and OER, by which a low applied potential of only 1.74 V is required to achieve an overall water splitting current of 20 mA cm−2 with a Tafel slope of 49.7 mV dec−1. Similar to TM oxides, TM sulfide-based electrocatalysts with hollow nanostructures have also been extensively investigated for use in electrocatalysis. Triblock copolymer pluronic P123 was used as a structure-directing agent for the preparation of hollow MoOx/Ni3S2 composite nanospheres with an ultrathin wall of <1.5 nm, which have exhibited comparable HER and OER activity towards commercial Pt/C and IrO2/C materials, respectively, with excellent durability in a long-term chronoamperometry process for up to 200 h, markedly surpassing the abovementioned noble metal-based electrocatalysts.53 Hollow NiFe diselenide nanocages were derived from a site-selective ammonia etchant-treated prussian-blue analog nanocage precursor.54 As a result, NiFe diselenide nanocages showed an outstanding OER activity of 10 mA cm−2 at a low overpotential of 240 mV accompanied by a small Tafel slope of 24 mV dec−1.
In fact, these nanostructured features are not mutually exclusive and can be addressed synchronously, such as by combining hollow nanoprisms with 2D nanosheets.55,56 These advanced nanostructures are helpful in improving electrocatalytic performance. However, in some electrocatalysts, the metal centers may still suffer from intrinsically low activity, which may be improved by strategies such as composition tuning and surface engineering, as described below.
Fig. 5 Binary and/or ternary 3d TM-based materials for improved electrocatalytic performance. (a) Experimental and theoretical OER activity of TM-doped NiOOH models. Reproduced with permission from ref. 57, Copyright 2015, American Chemical Society. (b) OER polarization curves and Tafel slopes for binary Ni0.75V0.25-LDH and Ni0.75Fe0.25-LDH. (c and d) Comparison of electrocatalytic OER activity for NiFe LDH nanosheets with different Ni/Fe ratios. Reproduced with permission from ref. 48, Copyright 2015, American Chemical Society. (e) ΔJ of NiV-LDH with various Ni contents plotted against the CV scan rate; (f) electrochemical active surface area of NiV-LDH with various Ni contents; (g) anodic current density at an overpotential of 350 mV plotted against the electrochemical surface area of NiV-LDH with incremental Ni contents for the electrocatalytic OER. (b and e–g) Reproduced with permission from ref. 58, Copyright 2016, Nature Publishing Group. (h) Calculated adsorption energies on –O and –OH by pure and doped oxyhydroxides and oxides. Reproduced with permission from ref. 61, Copyright 2016, the American Association for the Advancement of Science. (i) Gibbs free energy change for Ni2+ → Ni3+ → Ni4+. Reproduced with permission from ref. 64, Copyright 2018, Nature Publishing Group. |
Furthermore, many studies on binary TM-based electrocatalysts have shown that the electrocatalytic performance can be further significantly optimized by tuning the component proportions. For instance, as shown in Fig. 5c and d, it was found that Ni2/3Fe1/3 LDH nanosheets exhibited a higher anodic current density and a smaller Tafel slope than both Ni3/4Fe1/4 and Ni4/5Fe1/5 LDHs in the electrocatalytic OER.48 The largest electrochemical active area was obtained by optimizing the molar ratio of Ni/V to 3, as shown in Fig. 5e–g, resulting in the highest electrocatalytic OER activity for binary NiV-LDH.58 NiCo2Se4 nanosheets have also been demonstrated to show higher reactivity than Ni2CoSe4 and Ni1.5Co1.5Se4 in electrocatalytic water oxidation.59
In addition to metal substitution, nonmetal elemental incorporation can also effectively tune the electronic structure of the materials, enhancing the electrocatalytic activity. Pristine Co3O4 displays semiconductive characteristics with a band gap of ∼1.5 eV. Xiao et al. employed Ar plasma etching for the formation of oxygen vacancies in Co3O4 (herein, the product is labeled as VO–Co3O4).60 The octahedral Co3+–O bonds were therefore destroyed. They further treated the VO–Co3O4 sample with Ar plasma in the presence of a phosphorus (P) precursor to obtain a P-doped Co3O4 (denoted as P-Co3O4) sample. The coordination analyses showed that after the subsequent plasma treatment, P atoms filled in the generated oxygen vacancies. The computational electronic properties showed that P filling can further lower the bandgap to ∼0.3 eV in comparison with ∼0.8 eV for VO–Co3O4, implying a more positive electrical conductivity. When these three materials were applied in the electrocatalytic HER, the onset potential for P-Co3O4, ∼0.05 V vs. RHE, was far lower than the values of 0.36 and 0.35 V vs. RHE for pristine Co3O4 and VO–Co3O4, respectively, indicating the activity improvement by nonmetal P doping.
The simultaneous incorporation of metal and nonmetal elements has also arisen recently as another type of co-substitution strategy. High-valence TM ions, such as Ni4+, in metal oxides have been revealed to contribute a higher activity for electrocatalytic water oxidation.63,64 However, in neutral electrolytes, the TM ions prefer to be in a relatively low-valence oxidation state, such as Ni2+. To reach the desired high oxidation state for the TM sites, as displayed in Fig. 5i, Zheng et al. developed a codoping strategy of incorporating Co, Fe and nonmetal P.65 The NiCoFeP oxyhydroxide showed a current density of 1 mA cm−2 at a rather low overpotential of 276 mV, far lower than the values of 390 mV for NiP oxyhydroxide and 330 mV for NiCoP oxyhydroxide, indicating the highest reactivity for the electrocatalytic OER in a neutral electrolyte. The DFT+U calculations show that substitutional doping of Ni by P substantially lowers the Gibbs free energy change during the Ni2+/Ni3+ oxidation process, while the introduction of Co and Fe will further decrease the transformation energy from Ni3+ to Ni4+. As a result of the codoping of Co, Fe and nonmetal P, the formation of Ni4+ was consistently promoted by reducing the required transition energy, resulting in the high activity of NiCoFeP oxyhydroxide.
Fig. 6 Heterostructures for electrocatalytic processes. (a) TEM image and (b) distribution of the electron densities of CoNi clusters@graphene core@shell nanostructures. Green and purple balls represent Ni and Co atoms, respectively. (c) Gibbs free energy profile of HER and (d) HER polarization curves of carbon nanotubes (CNTs), CoNi@graphene (or CoNi@carbon) and commercial Pt/C electrocatalysts. (a, b and d) Reproduced with permission from ref. 66, Copyright 2016, Nature Publishing Group and (c) reproduced with permission from ref. 67, Copyright 2015, John Wiley & Sons, Inc. (e) TEM image of hexagonal Ni nanoplates/rGO composites, (f) LSV curves for OER and (g) Tafel slopes of NiAl LDH precursors, pure Ni, NiAl LDH/GO, Ni/rGO and Ni@Pt composites. Reproduced with permission from ref. 68, Copyright 2015, American Chemical Society. (h) TEM and (i) HRTEM images of Fe2P/NGO composites, (j) LSV curves of NGO, Fe2P, Fe2P/NGO composites, the Fe2P + NGO mixture, and the commercial Pt/C electrocatalyst. Reproduced with permission from ref. 69, Copyright 2015, Elsevier Inc. (k) Polarization curves of Ni(OH)2/MoS2 composites, Ni(OH)2, MoS2 and commercial Pt/C electrocatalysts. The inset is the TEM image of Ni(OH)2/MoS2 composite. (l) Isosurfaces of the local charge density difference of the Ni(OH)2/MoS2 heterointerface; red, white, blue, purple and orange balls stand for H, O, Ni, Mo and S atoms, respectively. (m) Free energy diagram for HER on the MoS2 and/or Ni(OH)2 side, and the Ni(OH)2/MoS2 heterointerface. Reproduced with permission from ref. 70, Copyright 2017, Elsevier Inc. |
Fig. 7 Single-crystalline CoO nanorods (SC CoO NRs) with oxygen vacancies for OER and ORR. (a) SEM, (b) TEM and (c) high-resolution HAADF-STEM images of SC CoO NRs. (d and e) Atomic model of a nanopyramid. (f and g) Experimental and simulated HAADF-STEM images of the pyramidal structure, respectively. (h and i) Intensity profiles taken from the orange and gray lines in (f and g), respectively. Comparison of ORR and OER activity for SC CoO NCs and polycrystalline CoO nanocrystals (PC CoO NCs) (j and l) before and (k and m) after normalization. Reproduced with permission from ref. 79, Copyright 2016, Nature Publishing Group. |
Fig. 8 Phosphorylation treatment on NiFe LDH for better wettability in electrocatalytic water oxidation. (a) Schematic of NiFe/NiFe:Pi nanostructures fabricated by electrodeposition followed by a phosphorylation process. (b) OER polarization curves, (c) Tafel slopes, (d) CV scans and (e) electrochemical impedance spectra of NiFe LDH and NiFe NiFe/NiFe:Pi electrocatalysts. Reproduced with permission from ref. 87, Copyright 2016, American Chemical Society. |
NaCo4(PO4)3 is highly evaluated as a unique compound since all Co sites are in low-symmetry 5-coordination sites. As shown in Fig. 9a and b, different from conventional trigonal bipyramidal [CoO5] configurations with Co atoms in the body center, the [CoO5] structure in this compound is close to tetragonal pyramidal in shape, with each Co active site located on the bottom surface of the coordination geometry, which may lead to a more favorable adsorption on electronegatively charged oxygen atoms from water molecules. As indicated in Fig. 9c–f, benefitting from the peculiar [CoO5] coordination environments, NaCo4(PO4)3 nanoribbons achieved markedly higher OER activity and more favorable water oxidation kinetics than traditional Na2CoP2O7 with all Co atoms in tetrahedral [CoO4] forms. The half wave at a relatively low potential of 1.41 V vs. RHE refers to the adsorption of water molecules by the [CoO5] geometries of NaCo4(PO4)3 nanoribbons, accompanied by the oxidation of Co(II) to Co(III). When being normalized by the electrochemical active surface area, the electrocatalytic activity of each [CoO5] configuration at a given potential, such as 1.7 V vs. RHE, was calculated to be twice that of the [CoO4] tetrahedron. As a merit of the unique coordination, intrinsic electrical conductivity and nanoribbon morphology, NaCo4(PO4)3 yielded current densities of 1.0 and 11.0 mA cm−2 at low overpotentials of 0.373 and 0.570 V, respectively. The high-activity NaCo4(PO4)3 nanoribbons were even comparable to the benchmark RuO2 nanoparticles, outperforming most of the recently reported TM-based catalysts, such as atomically thin Co3S4 nanosheets and congeneric cobalt phosphate/GO composites.
Fig. 9 NaCo4(PO4)3 with [CoO5] pyramids for the electrocatalytic OER. Crystal structure models of (a) NaCo4(PO4)3 and (b) Na2CoP2O7 phases. The brown, red, green and blue balls represent Na, O, P and Co atoms, respectively, while the white balls represent partially occupied Na sites. A possible OER mechanism for (c) NaCo4(PO4)3 and (d) Na2CoP2O7 products in the electrocatalytic OER. The yellow and light pink balls represent adsorbed oxygen and hydrogen atoms, respectively. (e) Polarization curves of NaCo4(PO4)3, Na2CoP2O7 and benchmark RuO2 nanoparticles. (f) Comparison of OER activity between NaCo4(PO4)3 nanoribbons and recently reported electrocatalysts. Reproduced and modified with permission from ref. 97, Copyright 2018, American Chemical Society. |
Fig. 10 Single-atom Fe–N–C electrocatalysts with planar tetragonal [FeN4] configurations for ORR. (a and b) HRTEM images of FeN4/GN-2.7; (c) atomic model and the corresponding HRTEM simulation image for the Fe–N–C structure in (b). (e) STM image for FeN4/GN-2.7; (f) simulated STM image for the Fe–N–C structure in (e). (g) dI/dV spectra obtained along the white line in the inset image. The gray, blue and light blue balls in (c), (e), and (f) represent C, N, and Fe atoms, respectively. (a–g) Reproduced with permission from ref. 99, Copyright 2015, the American Association for the Advancement of Science. (h) ORR polarization curves for FeN4/GN, Pt/C and graphite; (i) ORR polarization curves for the FeN4/GN-2.7 sample before and after continuous CV scan; (j) electron transfer number and the H2O2 yield ratio for FeN4/GN-2.7, Pt/C, GN and FePc. (h–j) Reproduced with permission from ref. 100, Copyright 2017, Elsevier Inc. |
Another kind of TM–N coordination has also been demonstrated as [TMN2] moieties, exhibiting a higher ORR activity than [TMN4] counterparts. Yin and his coworkers have prepared two different Co–N coordination sites by tuning the calcination temperature of the bimetal ZnCo-MOF precursor (herein, the Zn content evaporates and is removed under high temperature conditions).101 The dominant Co single-atom sites are [CoN4] and [CoN2] for the Co–N–C species obtained at 800 and 900 °C (denoted as Co SAs/N–C (800) and Co SAs/N–C (900)), respectively. As indicated in Fig. 11, as a result of the high reactivity and high content of [CoN2] moieties, Co SAs/N–C (900) showed higher ORR activity than both Co SAs/N–C (800) and the Pt/C catalyst. The Tafel slope of 75 mV dec−1, lower than 79 mV dec−1 for Co SAs/N–C (800) and 96 mV dec−1 for Pt/C, revealed more favorable reaction kinetics for Co SAs/N–C (900) in the electrocatalytic ORR. Furthermore, Co SAs/N–C (900) possessed higher values of kinetic-limiting current density Jk and half-wave potential E1/2 than Co SAs/N–C (800) and the Pt/C catalyst, verifying the superior activity. Similarly, Guo's group proposed Fe atoms in the form of [FeN2] in N-doped ordered mesoporous carbon (NOMC).102 Due to the unique coordination configurations, the FeN2/NOMC material yielded a comparable polarization curve to the commercial Pt/C material in the electrocatalytic ORR. The E1/2 value, 0.863 V vs. RHE for FeN2/NOMC, was ∼30 mV more positive than that of the commercial Pt/C material. Furthermore, the Jk value was 45.2 mA cm−2, far superior to 16.2 mA cm−2 for the Pt/C catalyst, revealing higher ORR activity. Moreover, DFT calculations indicated that due to the lower interaction with *O2 and *OH intermediates in comparison with [FeN4] sites, [FeN2] moieties reduced the overpotential by ∼0.12 eV, contributing to the higher activity. The corresponding electrocatalytic parameters of these selected cases are listed in Table 1.
Fig. 11 Planar tetragonal [TMN2] configurations for ORR. (a) TEM image for the Co SAs/N–C (900) sample. The insets represent the HAADF image and the structural model for planar [FeN2]. The black, gray and white balls represent N, C, and Co atoms, respectively. (b) ORR polarization curves, (c) Tafel slopes and (d) Jk and E1/2 for Co–N–C materials and the commercial Pt/C electrocatalyst. (a–d) Reproduced with permission from ref. 101, Copyright 2017, John Wiley & Sons, Inc. |
Catalysts | Electrocatalysis | Electrolyte | Overpotential | Tafel slope (mV dec−1) | Ref. |
---|---|---|---|---|---|
Morphology design | |||||
Nano-FeS2-rGO | HER | 0.5 M H2SO4 | 139 mV @ 10 mA cm−2 | 66 | 33 |
Micro-FeS2-RGO | 260 mV @ 10 mA cm−2 | 124 | |||
Co9−xNixS8 nanocages | OER | 1.0 M NaOH | 364 mV @ 10 mA cm−2 | 74.7 | 34 |
CaMn2O4 nanorods | ORR | 0.1M KOH | 0.563 V vs. RHE@4.25 mA cm−2 | 64 | 40 |
NiMoP2 nanowires | HER | 0.5 M H2SO4 | 195 mV @ 100 mA cm−2 | 56 | 41 |
OER | 1.0 M KOH | 320 mV @ 10 mA cm−2 | 96.7 | ||
NiCo2Px nanoneedles | HER | 1.0 M KOH | 58 mV @ 10 mA cm−2 | 34.3 | 42 |
1.0 M PBS | 63 mV @ 10 mA cm−2 | 63.3 | |||
0.5 M H2SO4 | 104 mV @ 10 mA cm−2 | 32.5 | |||
NiFe LDH nanoplates | OER | 1.0 M KOH | 350 mV @ 10 mA cm−2 | 67 | 47 |
NiFe LDH nanosheets | 300 mV @ 10 mA cm−2 | 40 | |||
NiFe LDH/rGO superlattice | OER | 1.0 M KOH | 217 mV @ 10 mA cm−2 | 40 | 48 |
NiMn LDH nanosheets | OER | 1.0 M KOH | 360 mV @ 10 mA cm−2 | 65 | 50 |
NiMn LDH/rGO superlattice | 260 mV @ 10 mA cm−2 | 46 | |||
Hollow NiCo2O4 cuboids | HER | 1.0 M NaOH | 110 mV @ 10 mA cm−2 | 49.7 | 52 |
OER | 290 mV @ 10 mA cm−2 | 53 | |||
Hollow MoOx/Ni3S2 nanospheres | HER | 1.0 M KOH | 106 mV @ 10 mA cm−2 | 90 | 53 |
OER | 310 mV @ 100 mA cm−2 | 50 | |||
NiFe diselenide nanocages | OER | 1.0 M KOH | 240 mV @ 100 mA cm−2 | 24 | 54 |
Composition tuning | |||||
Ni0.75V0.25-LDH | OER | 1.0 M KOH | 350 mV @ 27.0 ± 1.6 mA cm−2 | 50 | 58 |
Ni0.75Fe0.25-LDH | 350 mV @ 11.7 ± 1.5 mA cm−2 | 64 | |||
Ni4/5Fe1/5 LDH nanosheets | OER | 1.0 M KOH | 280 mV @ 5 mA cm−2 | 82 | 48 |
Ni3/4Fe1/4 LDH nanosheets | 340 mV @ 5 mA cm−2 | 80 | |||
Ni2/3Fe1/3 LDH nanosheets | 370 mV @ 5 mA cm−2 | 76 | |||
P-Co3O4 | HER | 1.0 M KOH | 120 mV @ 10 mA cm−2 | 52 | 60 |
OER | 280 mV @ 10 mA cm−2 | 51.6 | |||
Co3Se4 | OER | 1.0 M KOH | 345 mV @ 10 mA cm−2 | 65 | 59 |
NiCo2Se4 | 290 mV @ 10 mA cm−2 | 53 | |||
Ni1.5Co1.5Se4 | 315 mV @ 10 mA cm−2 | 57.5 | |||
Ni2CoSe4 | 310 mV @ 10 mA cm−2 | 65 | |||
Ni3Se4 | 349 mV @ 10 mA cm−2 | 76 | |||
FeCo oxyhydroxide | OER | 1 M KOH | 215 ± 6 mV @ 10 mA cm−2 | — | 61 |
FeWCo oxyhydroxide | 191 ± 3 mV @ 10 mA cm−2 | — | |||
NiP oxyhydroxide | OER | CO2-saturated 0.5 M KHCO3 | 588 mV @ 10 mA cm−2 | — | 65 |
NiCoP oxyhydroxide | 547 mV @ 10 mA cm−2 | — | |||
NiCoFeP oxyhydroxide | 330 mV @ 10 mA cm−2 | — | |||
CoNi@graphene | HER | 0.1 M H2SO4 | 142 mV @ 10 mA cm−2 | 104 | 67 |
Ni-Al LDH/rGO composite | OER | 1.0 M KOH | 380 mV @ 10 mA cm−2 | 89 | 68 |
Ni/rGO composite | 330 mV @ 10 mA cm−2 | 68 | |||
Fe2P/NGO | HER | 0.5 M H2SO4 | 138 mV @ 10 mA cm−2 | 65 | 69 |
Ni(OH)2/MoS2 | HER | 1.0 M KOH | 80 mV @ 10 mA cm−2 | 60 | 70 |
Surface engineering | |||||
{111} facets-exposed Co3O4 | OER | 1.0 M KOH | 285 mV @ 10 mA cm−2 | 49 | 73 |
HER | 195 mV @ 10 mA cm−2 | 50 | |||
{100} facets-exposed Cu2O | ORR | 0.1 M KOH | E onset = 0.8 V | — | 75 |
{100} facets-exposed MnO | ORR | 0.1 M KOH | E 1/2 = 0.77 V vs. RHE, ΔE = 1.02 V | — | 77 |
O vacancies-rich CoO NDs | OER | 1.0 M KOH | 330 mV @ 10 mA cm−2 | 44 | 79 |
ORR | E 1/2 = 0.85 V, ΔE = 0.71 V | 47 | |||
Defected NiCo LDH | OER | 1.0 M KOH | 254 mV @ 10 mA cm−2 | 32 | 80 |
NiFe LDH/NiFe:Pi | OER | 1.0 M KOH | 290 mV @ 10 mA cm−2 | 38 | 87 |
Metal coordination symmetry/geometry control | |||||
NaCo4(PO4)3 | OER | 0.05 M PBS | 373 mV @ 1 mA cm−2 | 121 | 96 |
FeN4/GN | ORR | 1.0 M NaOH | E 1/2 = 0.91 V vs. RHE | — | 100 |
[CoN2] + [CoN4]/N–C | ORR | 0.1 M KOH | E 1/2 = 0.881 V vs. RHE | 75 | 101 |
FeN2/NOMC | ORR | 0.1 M KOH | E 1/2 = 0.863 V vs. RHE | 58 | 102 |
In addition, the activity of TM–N–O coordination configurations has also been investigated. Yang et al. analyzed the ORR performance of planar [MnNxO4−x] (x = 1, 2 or 3) coordination sites theoretically.103 Typically, for planar tetragonal Mn moieties, various Mn–N–O configurations exist, i.e., [MnN1O3], [MnN2O2] and [MnN3O1]. DFT calculations showed that as a merit of the proper downshift of the d-band center and the position of the first peak close to the Fermi level in [MnN1O3] geometries, desorption and formation of the intermediates were more favorable, resulting in the best ORR kinetics for [MnN1O3] among all the planar Mn–N–O configurations. However, until now, only the material with mixed [MnNxO4−x] (x = 1, 2 or 3) configurations has been prepared. More work needs to be performed for the synthesis of electrocatalysts containing a single type of TM–N–O structure, e.g., [MnN1O3], to further verify the theoretical analysis and explore the experimental reactivity.
The ultimate goal of developing effective electrocatalysts based on TM elements is to construct stable and high-efficiency energy conversion systems and devices. Hence, further work may be directed towards the following aspects. (i) More efforts may need to be focused on identifying the most ideal nanostructures for durable and efficient electrocatalysis. So far various electrocatalysts based on 3d transition metals and related nanoarchitectures have been constructed. It is necessary to choose the best structure on account of the cost, maximum utilization of metal sites, activity, the feasibility of large-scale fabrication and applications, etc. (ii) Composition, valence and component tuning should be further carried out to achieve the optimal activity. (iii) More favorable metal coordination geometries are expected due to their fascinating adsorption/desorption abilities. In addition, for an in-depth understanding of the reaction mechanism, it is urgent to introduce some in situ and/or ex situ techniques to characterize the real-time reaction states and changes of typical electrocatalysts during electrocatalysis. The advances in the combination of theoretical calculations and experimental approaches are deemed valuable in exploiting low-cost and high-performance electrocatalysts.
Many investigation cases have verified the high activity of nanostructured electrocatalysts based on 3d TM elements, providing a great potential for the gradual replacement of precious metal-based electrocatalysts. To meet the industrial use, more endeavors are expected for non-precious 3d TM-based materials. In particular, it is urgent to improve the structural stability and electrochemical durability of the electrocatalysts. Current proton exchange membrane (such as Nafion N115 membrane) fuel cells can work only in acidic environments.104 Thus, it is rather difficult to completely replace precious metal-based materials due to the low durability of TM-based electrocatalysts. To make full and practical use of the earth-abundant and cheap 3d TM sources, developing alkaline condition-related electrocatalytic devices or systems is prospective and should be greatly appreciated. For instance, due to the high activity, reversibility and hydroxyl-transfer characters,45,48,105 TM hydroxides can be regarded as a kind of important multifunctional material for the future electrolyzers. In this regard, designing and developing economical electrocatalysts for alkaline systems are significantly promising for sustainable energy conversion.
This journal is © The Royal Society of Chemistry 2019 |