Progress in Materials Science 103 (2019) 596–677
Contents lists available at ScienceDirect
Progress in Materials Science journal homepage: www.elsevier.com/locate/pmatsci
Cobalt oxide-based nanoarchitectures for electrochemical energy applications
T
Jun Mei, Ting Liao, Godwin A. Ayoko, John Bell, Ziqi Sun
⁎
School of Chemistry, Physics and Mechanical Engineering, Queensland University of Technology, 2 George Street, Brisbane, QLD 4000, Australia
ARTICLE INFO
ABSTRACT
Keywords: Cobalt oxides Co3O4 Graphene Batteries Electrocatalysis Supercapacitors
Cobalt oxide nanostructures have been considered as promising electrode materials for various electrochemical applications, especially for batteries, supercapacitors, and electrocatalysis, owing to their unparalleled advantages of high theoretical capacity, highly-active catalytic properties, and outstanding thermal/chemical stability. If hybridized with property-complementary nanomaterials, such as nanocarbon, CNTs, graphene, metal oxides/sulfides and conductive polymers, their electrochemical properties can be further enhanced in terms of specific reversible capacity/capacitance, rate capability, cycling stability, and catalytic activity. In this review, we first give a comprehensive overview on recent progress in both monolithic cobalt oxide nanostructures and their hybrid nanomaterials for batteries, supercapacitors, and electrocatalysis applications. Then, structure-property relationships of the cobalt oxide-based nanomaterials and current challenges in both nanoarchitectures design and their applications in electrochemical energy devices are proposed, and an outlook on future research of this family of
Abbreviations: 0D, zero-dimensional; 1D, one-dimensional; 2D, two-dimensional; 3D, three-dimensional; AAO, anodic aluminum oxide; AC, activated carbon; AEMFCs, anion exchange membrane fuel cells; AFM, atomic force microscopy; AIMD, ab initio molecular dynamics; APS, 3-aminopropyltrimethoxysilane; ASCs, asymmetric supercapacitors; BA, benzyl alcohol; BDHC, blood powder-derived heteroatom doped carbon; BFSTEM, bright-field scanning transmission electron microscope; BP, blood powder; BT, barberry tannin; CA, carbon aerogel; Ca-AF, calcium alginate fibers; CE, Coulombic efficiency; CNBs, cubic nanoboxes; CNTs, carbon nanotubes; COPs, covalent organic polymers; CP, carbon paper; CST, chitosan; CV, cyclic voltammetry; CVD, chemical vapor deposition; CWs, carbonized wing scales; DETA, diethylenetriamine; DMF, N, N-dimethylformamide; EDLCs, electric double-layer capacitors; EFM, electrostatic force microscopy; EIS, electrochemical impedance spectrum; EPD, electrophoretic deposition; f-CNTs, functionalized CNTs; FFT, fast Fourier transform; FTO, fluorine-doped tin oxide glass; GHCS, graphene-like holey Co3O4 nanosheets; GO, graphene oxide; HER, hydrogen evolution reaction; HR-SEM, high-resolution scanning electron microscope; HRTEM, highresolution transmission electron microscopy; HOR, hydrogen oxidation reaction; HTSs, hierarchical tubular structures; LDH, layered double hydroxide; LIBs, lithium-ion batteries; LSV, linear sweep voltammetry; MFCs, microbial fuel cells; moCNT, mildly oxidized CNTs; MOFs, metal-organic frameworks; MWCNTs, multiwall CNTs; NBAs, nanobelt arrays; NC, N-doped carbon; N-CN, nitrogen-doped carbon networks; NCNT, nitrogendoped CNT; NMEG, nitrogen modified microwave exfoliated graphite oxide; NPs, nanoparticles; NWAs, nanowire arrays; OER, oxygen evolution reaction; ORR, oxygen reduction reaction; ox-CNTs, oxygen-containing CNTs; PAH, poly(allylamine hydrochloride); PAN, poly(acrylonitrile); PANI, polyaniline; PBA, 1-pyrenebutyric acid; p-CNT, pristine CNTs; PEI, polyethyleneimine; Pind, polyindole; PPy, polypyrrole; PS, polystyrene; p-TSA, ptoluenesulfonic acid; PVD, physical vapor deposition; PVP, poly(vinylpyrrolidone); QDs, quantum dots; QTNBs, quadrate tubular nanoboxes; RRDE, rotating ring-disk electrode; rGO, reduced graphene oxide; RNBs, rectangular nanoboxes; RT-SOFCs, reduced-temperature solid oxide fuel cells; TCFP, Teflon-coated carbon fiber paper; TEM, transmission electron microscope; SACNT, super-aligned CNT array; SAED, selected area electronic diffraction; SAXS, synchrotron small angel X-ray scattering; SCF, supercritical fluid; SEI, solid-electrolyte interphase; SEM, scanning electron microscope; SIBs, sodium-ion batteries; S/N, signal-to-noise ratio; SSCs, symmetric supercapacitors; STEM, scanning transmission electron microscopy; SWCNTs, single-wall CNTs; TMOs, transition metal oxides; URFCs, unitized regenerative fuel cells; UV–Vis, ultraviolet–visible; VLS, vapor-liquidsolid; VSS, vapor-solid-solid; XAFS, X-ray absorption fine structure; XANES, X-ray absorption near-edge spectroscopy; XAS, X-ray absorption spectroscopy; XPS, X-ray photoelectron spectroscopy; XRD, X-ray diffraction; ZIF, zeolite imidazolate frameworks ⁎ Corresponding author. E-mail address:
[email protected] (Z. Sun). https://doi.org/10.1016/j.pmatsci.2019.03.001 Received 14 February 2018; Received in revised form 1 March 2019; Accepted 4 March 2019 Available online 05 March 2019 0079-6425/ © 2019 Elsevier Ltd. All rights reserved.
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
materials in electrochemical energy applications are brought forward. This understanding on the relationships of synthesis-nano/microstructure-property-performance of cobalt oxide-based nanomaterials is expected to lay a good foundation for pushing this promising class of materials to the practical application in energy conversion and storage devices and to provide a good reference for the readers in the fields of materials, chemistry, sustainable energy, and nanotechnology.
1. Introduction Cobalt oxides, possessing the merits of earth abundance, low cost, and good environmental compatibility, have been a focus in many research areas. There are four typical types of cobalt oxides, namely, cobalt(II) oxide (CoO), cobalt(III) oxide (Co2O3), cobalt (IV) oxide (CoO2), and cobalt(II,III) oxide (Co3O4) [1,2], in which Co3O4 and CoO are the most common ones, due to their excellent thermal stability and exceptional physical and chemical properties [3–6]. Fig. 1 presents the crystal structures of the low-valence cobalt oxide, CoO in a face-centered cubic structure, and the mixed-valence Co3O4 in a well-known spinel crystal structure. In the structure of Co3O4, the tetrahedral 8(a) sites and the octahedral 16(d) sites are occupied by Co(II) and Co(III) ions, respectively [7,8]. To date, cobalt oxides with different dimensionalities have been successfully synthesized and exploited for various applications, especially the electrochemical applications in renewable energy technologies, such as batteries, supercapacitors, and electrocatalytic reactions [9–11], primarily due to their high theoretical specific capacity/capacitance, high catalytic activity, and mechanical and chemical stability [12–21]. It is noteworthy that spinel Co3O4 demonstrates a remarkable theoretical capacity as high as 890 mA h g−1 for lithium-ion batteries (LIBs) and an excellent capacitance of over 3000 F g−1 for supercapacitors [22–24]. Moreover, it has been validated that the spinel cobalt oxide is one of the most promising candidates as efficient non-precious electrocatalysts in metal-oxygen batteries [25]. Nevertheless, bulk monolithic cobalt oxides still present some drawbacks in practical applications. As electrode materials, they often suffer from low rate performance and poor cycling stability, attributed to its inherently poor electronic conductivity, slow reaction kinetics, and severe volume expansion during the repeated ion uptake and removal processes of rechargeable batteries, which cause intensive local stress and eventually lead to electrode failure and electrolyte degradation [26]. On the other hand, some electrochemical energy devices involve more than one type of catalytic reactions, for example, unitized regenerative fuel cells (URFCs) includes both oxygen evolution reaction (OER) and oxygen reduction reaction (ORR), which demand high-performance catalysts that can be functionalized to different reactions simultaneously [27]. Most cobalt oxides, however, present excellent monofunctional catalytic performances but are limited for only one targeted reaction. For instance, the typical Co3O4 has good OER activity but inferior ORR kinetics [28]. If we want to make this family of materials capable of triggering two or more catalytic reactions simultaneously, elaborate design on their chemical states, surface properties, compositions, crystallinities, morphologies, dimensionalities, etc., has to be approached. One of the general strategies to address the above-mentioned issues is the fabrication of nanometer-sized materials with welldesigned structures, in which the reduced dimensionalities increases significantly the rate of ion insertion/removal and shortens dramatically the distance of ion and charge/carrier transport, the enhanced surface area permits high contact area with electrolyte and allows more active sites for chemical/electrochemical reactions, and the small particle size allows better accommodation of the strain associated with electrochemical reactions [29–32]. Some intrinsic issues of cobalt oxides, such as the low charge/ionic conductivity, however, cannot be conquered via this strategy. Recent research has shown that the formation of cobalt oxide-based hybrid structures by the introduction of one of two property-complementary nanomaterials, such as nano-carbon, graphene and CNTs, is another effective approach to resolve these critical issues. It is believed that the hybrid structures can take the full use of both the high capacity of cobalt oxides and the excellent cycle performance of additive materials [33]. Indeed, the addition of even a very small
Fig. 1. Crystal structures of (a) CoO and (b) Co3O4 (Navy blue, dark green and red balls indicate Co2+, Co3+, and O2− ions, respectively). 597
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
amount of carbon nanomaterials into Co3O4 can significantly enhance its electron transfer, accelerate the reaction kinetics, suppress the aggregation of metal oxide nanostructures, and accommodate the volume change of the electrodes, and thus improve the rate and cycling performance of the Co3O4-based anodes. It is argued that, however, the addition of large amount of carbon would sacrifice the capacity of Co3O4 due to the low capacity of carbon materials. Therefore, reliable fabrication of well-designed cobalt oxide-based nanomaterials is yet complex and remains a major challenge, and rational design of cobalt oxide-based nanomaterials for electrochemical energy devices is constantly being pursued. Owing to the promising potential of cobalt oxide-based materials, the research on this type of materials has been constantly increasing in recent years. To date, around 40,000 records have been found on topics related to cobalt oxides by searching the keywords of “Co3O4 OR CoO OR Co2O3 OR cobalt oxide” from the Web of Science™ database (by 23th, January 2018), among which 38,800 records are related to their electrochemical applications, when the keywords were changed “Co3O4 OR CoO OR Co2O3 OR cobalt oxide AND electrochemistry” in the same database. Relevant papers published in the most recent five years accounted for more than one-third of the overall records between 1900 and 2017. As shown in Fig. 2a, the steady increase in publication numbers every year also attests to the significance of this family of materials in the fields of materials and electrochemistry. In addition, among the three types of cobalt oxides, considerable progress has been made on the synthesis and electrochemical applications of Co3O4, which occupies a heavy proportion of the scientific papers published to date, particularly in the past five years. It is interesting that even though a large number of publications and very active researches about this family of materials, based on the knowledges of authors, currently only one review article was published on cobalt oxides, which is on the fabrication of hollow Co3O4 nanostructures appeared in 2012 [7]. Therefore, a comprehensive review on the fabrication, structure, and electrochemical applications of cobalt oxide-based nanomaterials is urgently desired to be a good reference for the readers in the fields of materials, chemistry, sustainable energy, and nanotechnology. This review intends to summarize the emerging studies related to the monolithic nanostructures and the hybrid materials of cobalt oxide semiconductors, especially for Co3O4 and Co3O4-based nanoarchitectures, in electrochemical energy-related applications, as shown in Fig. 2b. Firstly, we try to draw a general profile of cobalt oxide nanomaterials in the applications of different types of electrochemical energy devices together with a brief introduction on the energy storage mechanisms of this family of materials for batteries, supercapacitors, and electrocatalysis. Then, recent progress in the synthesis and electrochemical energy applications of cobalt oxides with different dimensionalities is explored and the effect of morphologies and structures of the nanomaterials on the electrochemical performances is reviewed. Subsequently, the design and construction of the cobalt oxide-based hybrids by combining with various property-complementary nanomaterials, which has been approved to be an important approach to enhance their electrochemical performances in energy-related applications, is comprehensively reviewed. In this section, the fabrication and the electrochemical performances of the cobalt oxide-based nanomaterials are summarized based on the selection of complimentary counterparts and then the dimensionality. Finally, current challenge of cobalt oxide-based materials in energy-related applications and some tangible strategies for improving the electrochemical performances of cobalt oxide-based nanomaterials are proposed and
Fig. 2. (a) Statistic analysis of the scientific publications on cobalt oxides from the Web of Science™ database (by 23th, January 2018). (b) An overview of the structure of this review on cobalt oxide-based nanoarchitectures for electrochemical energy applications. 598
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
outlined. Besides, the structure-property relationships of cobalt oxide-based nanomaterials and their impacts on their electrochemical performances are fully discussed, and at last a brief outlook is given to provide a guideline for the further research of this class of materials. 2. Electrochemical energy applications of cobalt oxide-based nanomaterials Electrochemical cells and systems play a key role in a wide range of critical technologies enabling for renewable energy conversion and storage, including the most established technologies such as rechargeable batteries, ion transport membranes and supercapacitors, and electrocatalytic reactors. Cobalt oxide-based materials are one family of the foremost electrode materials in these electrochemical energy devices. In this section, the application of cobalt oxide nanomaterials in those typical electrochemical energy conversion and storage devices together with their energy conversion and storage mechanisms will be summarized and outlined. 2.1. Rechargeable batteries Rechargeable batteries are one of the dominate power sources in a diverse range of applications, from electronic vehicles to microchips. Each battery is composed of a positive and a negative electrode, which are separated by solid/liquid electrolytes, enabling ion transfer between two electrodes [34–36]. When we evaluate the performance of rechargeable batteries, several factors are usually taken into concerns as summarized as follows. (i) Voltage: normally, it is difficult to achieve a thermodynamically equilibrium voltage for a battery, due to the reversibility deviation of the electrode processes. Instead, we often measure an open cell voltage (OCV). (ii) Capacity: the capacity of a battery is the amount of electrical charge which can be drawn from the battery at a specific voltage. This can be calculated based on the t equation: C = 0 I ·dt , where C, I, and t refer to the capacity (Ah), discharge current (A), and discharge time (h), respectively. (iii)
Fig. 3. (a) Voltage-composition profiles for various metal oxide electrodes for LIBs at a rate of C/5 in a voltage range of 0.01–3 V, and (b) capacity fading curves of metal oxide electrodes within 50 cycles (inset: rate capability of CoO electrode) [22] (Copyright 2000, Nature Publishing Group). (c) Potential-composition profiles of Co3O4 and the corresponding in-situ X-ray diffraction (XRD) patterns (x = 2.0) [23] (Copyright 2000, The Electrochemical Society). (d) Division of three distinct regions of a thick Co3O4 film deposited on the two-contact geometry transistor device used for in-situ monitoring the electric change during the insertion and deinsertion of lithium, (e) the origin of inferior conductivity of metallic Co phase owing to the formation of less conductive Li2O on the surface, and (f) the origin of the increase of resistance at the Co3O4 NPs/current collector interface after the lithiation of Co3O4 NPs owing to the formation of Li2O at the interface [69] (Copyright 2014, American Chemical Society). 599
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
t
Specific energy and energy density: the whole energy that a battery can deliver is represented by the equation: E = 0 U ·I ·dt , where E, U, I, and t represent the energy content (Wh), voltage (V), discharge current (A), and discharge time (h), respectively. The specific energy (Wh kg−1), or we called gravimetric density, refers to the energy content at a basis of the weight of a battery. The energy density (Wh L−1 or Wh cm−3), an important parameter for designing portable batteries, is the amount of energy stored in a unit volume of a battery. (iv) C-rate and cycle life: the C-rate specifies the speed a battery being charged or discharged relative to its maximum capacity, while the cycle life of a battery is the number of repeated charge/discharge processes before that the capacity decays under a specific percent (e.g. 80%). To date, a wide range of rechargeable batteries have been designed and investigated, such as metal ion (e.g. Li+, Na+, K+, Ca+, Mg2+, Al3+, etc.) batteries, metal-air/O2 batteries (e.g. Li-O2, Li-CO2, Na-O2, Zn-O2 battery, etc.), metal-sulfur batteries (e.g. Li-S, NaS, K-S battery, etc.), metal-selenium batteries (e.g. Li-Se, Na-Se battery, etc.), and some hybrid batteries (e.g. Mg-Li battery). In terms of cobalt oxide-based nanomaterials, the applications are mainly involved in lithium-ion batteries (LIBs), sodium-ion batteries (SIBs), and Li-O2/air batteries. 2.1.1. LIBs LIBs are a member of the family of rechargeable batteries, in which lithium ions move from the anode to the cathode during discharging and backwards when charging [37,38]. Since the first practical application of lithium batteries by Sony in 1991, LIBs have become one of the most popular rechargeable batteries owing to their many outstanding features, including superior energy density, no memory effect, low maintenance, and little self-discharge [39–43]. Since the function of LIBs is based on reversible shuttling of Li+ ions and differences in electrochemical potentials between the cathode and the anode, the inherent properties of the electrode materials are a crucial factor that largely determines the overall performance of batteries. Even to date, one of the most attractive research areas in LIBs is to design elaborate nanostructure of the electrode to solve the grand challenges in LIBs, such as the unsatisfied energy density, slow lithium ion and electron transport, and large volume change of electrode materials during cycling processes [44,45]. As one of the most attractive alternatives to commercial graphite, cobalt oxides have been widely studied as an anode material since their first use in LIBs was released in 2000 by Poizot et al. [22,46–61]. It revealed a conversion mechanism in Li storage for cobalt oxides and some other similar transition metal oxides (TMOs), such as FeO and NiO. This mechanism, involving the reduction and oxidation of metal nanoparticles (NPs) accompanied by the formation and decomposition of Li2O, differs from the previous Li insertion/deinsertion or Li-alloying storage mechanisms [22]. For example, the reversible electrochemical reaction of cobalt oxide electrodes in LIBs can be expressed as: CoO + 2Li ↔ Li2O + Co and Co3O4 + 8Li ↔ 4Li2O + 3Co, for CoO and Co3O4 electrodes, respectively. As shown in Fig. 3a, cobalt oxide electrode exhibits similar charging/discharging behaviors to FeO an NiO with a voltage window ranging from 0.01 to 3 V and a reversible capacities between 600 and 800 mA h g−1. The capacity retention capabilities of these metal oxide electrodes, however, are distinct from each other. Both CoO and Co3O4 electrodes demonstrated more stable capacities over cycling than these of other metal oxide electrodes (Fig. 3b), and a good rate capability with 85% of the remained capacity was found for CoO electrode even at a high rate of 2 C (inset in Fig. 3b). Furthermore, Co3O4, in theoretically, demonstrates a specific capacity as high as 890 mA h g−1 with eight lithium reacting with one cobalt oxide unit in each formula unit, which is almost three times higher than that of the commercially used graphite anode (less than 372 mA h g−1) [62–66]. Surprisingly, it is noteworthy that the experiments concluded that Co3O4 electrode can deliver an initial capacities of over 1200 mA h g−1 [67]. The reaction pathways and products of Co3O4 during charging and discharging are relatively complex. For example, a LixCo3O4 intercalated phase always forms during the reduction of Co3O4 phase [68], whose stability is highly associated with the specific discharge rate, the cycling temperature, and the structural parameters, such as crystallite size and specific surface area, of the Co3O4 material [23]. Fig. 3c shows the composition of Co3O4-based LIBs at different discharge current densities. It is clearly observed that a LixCo3O4 phase exists and remains stable at a relatively high rate but is decomposed into CoO phase at a low rate. Despite the fact that cobalt oxides often exhibit an excellent initial specific capacity as anode materials for LIBs, a dramatic initial capacity loss always appears, leading to a low reversible capacity after several cycles. To make this problem more explicit, Kim et al. systematically explored the origins of the irreversible capacity loss of cobalt oxide by probing the changes of the electronic and structural properties of hollow-structured Co3O4 NPs aroused by the conversion reactions related with the formation/decomposition of Co and Li2O [69]. In this method, the Co3O4 NPs were embedded into the channel of a two-contact transistor device with source and drain electrodes as current collectors to in-situ monitor the insertion and deinsertion of lithium during the lithiation/delithiation process [69]. By this means, the contributions from the Co3O4 film, the Co3O4/electrolyte interface, and the Co3O4/current collector interface can be separated to study the origin of the irreversible capacity loss of Co3O4 in the first cycle. As displayed in Fig. 3d, the thick Co3O4 NP film can be divided into three distinct regions. In “Region 1”, in which the active materials were directly in contact with liquid electrolyte, severe electrical and structural degradation occurred during lithiation/delithiation process, as confirmed by an obvious conductivity decrease monitored by the transistor device and further the structural change characterized by electrostatic force microscopy (EFM) and atomic force microscopy (AFM). In “Region 2”, the electrical and structural degradation was not as severe as those in the front region. In “Region 3”, dramatic drops of the conductivity of Co3O4 NPs and the voltage at the source/drain contacts were observed, which was not favorable for efficient electronic transport. Generally, the conversion reaction accompanying with the formation of metallic Co clusters increases the electrical conductivity in the first lithiation step. In this measurement, however, the increase of conductance was far lower than the production of Co metallic phase. The possible explanation is that the produced Co clusters might be covered by a more electrically resistive Li2O phase (Fig. 3e). In the same manner, the increase of the resistance between Co3O4 NPs and the current collector after the first lithiation step could also be attributed to the reduced physical contacts induced by the formation of Li2O at the interface (Fig. 3f). Therefore, they believed that the increased contact resistance 600
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
between the Co3O4 NPs and the current collector and in “Region 3” contributes to the irreversible capacity loss during the initial charging/discharging process for the cobalt oxide-based nanomaterials [69]. Besides the dramatic initial capacity loss, there are other drawbacks for cobalt oxides as electrode materials for LIBs, including low electrical conductivity, serious aggregation, and obvious volume expansion/shrink ratio during the intercalation and de-intercalation processes of Li+ ions, which result in considerable capacity loss at high rates or after long-term cycling. Moreover, the resource for cobalt minerals is relatively limited in most regions of the world. To make cobalt oxide nanomaterials more applicable and cost-effective, extensive studies on optimization of both the structures and the components of cobalt oxide-based composites have been proceeded to overcome the above-described shortcomings. The effective approaches include: (i) the incorporation of cobalt oxides into a conductive matrix, such as nanocarbon, CNTs, graphene, and organic polymers, can greatly improve the overall electrical conductivity and inhibit the initial capacity loss; (ii) the design of core-shell structures, such as core-shell nanowires and nanospheres with cobalt oxide nanostructures as cores, can effectively buffer the stress originated from the insertion and deinsertion of Li+ ions; (iii) the combination of cobalt oxides with other heterogeneous metal-based components, such as metal sulfides, metal alloys, metal NPs, and metal hydroxides, can exhibit a surprising synergetic effect to enhance both the reactive activity of the materials and the cycling stability of the LIBs; (iv) the construction of porous structures can offer more active sites for Li+ anchoring and more diffusion channels for electrolyte penetrating; (v) the introduction of heteroatoms into the cobalt oxide crystals can change the lattice parameters and lead to the formation of defects for better accommodating the Li+ insertion behaviors. It has been validated that these strategies are effective to improve the electrochemical performances of cobalt oxides in LIBs with enhanced electron transfer, accelerated reaction kinetics, suppressed self-aggregation, and accommodated volume change, and thus significantly enhance the overall battery performances of LIBs. 2.1.2. SIBs Owing to the abundant and low-cost sodium resources, SIBs have attracted considerable research attention as a potential alternative to LIBs [70–95]. SIBs share a similar working mechanism to LIBs, however, the Na+ (1.02 Å) has a larger ionic radius than Li+ (0.76 Å), leading to more sluggish diffusion kinetics and more significant volumetric changes during repeated charging/discharging cycles. Many electrode materials that show outstanding electrochemical performances for LIBs are proved to be unsuitable for SIBs. Hence, the current major research target for SIBs is still to explore suitable electrode materials.
Fig. 4. Electrochemical performances and Na+ storage mechanism of cobalt oxides. (a) transmission electron microscope (TEM) image of mesoporous Co3O4 (m-Co3O4) microspheres (inset: schematic illustration of electron and ion transport pathways upon discharging) [106], (b) schematic illustration of the sodiation/de-sodiation cycle of m-Co3O4 electrode [106] (Copyright 2015, Wiley-VCH), (c) rate capabilities of the atomically thin Co3O4 nanosheets grown on stainless steel mesh (ATCSM), the conventional Co3O4 nanostructures (CWM), and the Co3O4 electrode prepared by conventional casting method (ATCSC) (inset: capacity ratios vs. current densities), (d) scanning electron microscope (SEM) image of ATCSM [98] (Copyright 2016, IOP Publishing Ltd). 601
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
NiCo2O4 spinel oxide was the first reported metal oxide as anode electrode for SIBs in 2002 [96,97]. It was observed that this spinel showed an initial discharge capacity of ∼618 mA h g−1, which is lower than the theoretical value (890 mA h g−1) based on the reduction reaction to form four Na2O per formula unit (NiCo2O4 + 8Na ↔ Ni + 4Na2O + 2Co). Moreover, a reversible capacity of 200 mA h g−1 was obtained for the resultant NiCo2O4-based SIBs [96]. Inspired by this preliminary study on the spinel structures for SIBs and the outstanding electrochemical performances of cobalt oxide-based nanomaterials for LIBs, a variety of cobalt oxides and their hybrids (e.g. Co3O4/C, Co3O4/CNT, Co3O4/graphene, etc.) have been explored as potential electrode materials for SIBs [98–105]. Yang et al. synthesized highly-ordered mesoporous Co3O4 (m-Co3O4) microspheres (Fig. 4a) by using mesoporous silica (KIT-16) as sacrificial hard template as an electroactive anode material for SIBs [106]. The dual porosity mesoporous ordered structure with enhanced electrode-electrolyte interface facilitated mass transport and Na+ diffusion. Compared to the referential bulk Co3O4 (bCo3O4) and m-Co3O4 without fluoroethylene carbonate (FEC) in the electrolyte, the m-Co3O4 porous structures with FEC additives in organic electrolyte delivered an initial capacity of 782 mA h g−1 at a current density of 30 mA g−1, with 267 mA h g−1 remained at 2430 mA g−1 and 416 mA h g−1 remained at 90 mA g−1 after 100 cycles. Ex situ XRD patterns confirmed the occurrence of reversible conversion reaction during the uptake/extraction of sodium ions. As summarized in Fig. 4b, Na+ ions were inserted into m-Co3O4 by surface reconstruction reactions during the first discharge to form NaxCoyOz, which was then partially converted into CoxO and Na2O based on conversion reactions, accompanied by a surface amorphization process along with further sodiation. The subsequent charge process mainly contributed by de-intercalation reactions of Na+ [106]. Later, Dou et al. reported an atomically thin Co3O4 nanosheetcoated stainless steel mesh (Fig. 4d) with improved capacitive Na+ storage for SIBs, which delivered average capacities of 509.2 mA h g−1 for the initial 20 cycles (50 mA g−1) and 427.0 mA h g−1 at a high rate of 500 mA g−1 (Fig. 4c) [98]. The enhanced electrochemical performance is primarily ascribed to the unique atomic thickness of the obtained cobalt oxide nanosheets and the
Fig. 5. Cobalt oxide for the application in Li-O2/air batteries. (a) Schematic illustration of cobalt oxide nanosheets/carbon paper (Co3O4 NSs/CP) electrode for Li-air batteries [139] (Copyright 2013, Royal Society of Chemistry). (b) (top) schematic illustration of OER mechanism of Li2O2 on Co3O4 surface; (bottom) the variation of catalytic activity of transition-metal (TM)-doped Co3O4 (1 1 1) in OER as a function of ionization potentials of the doped metals [141] (Copyright 2015, American Chemical Society). (c) Schematic illustration of the reaction mechanism of the {1 1 2} faceted Co3O4 platelets enclosed by Co3+ and Co2+ sites for both ORR and OER, and (d) a comparison of cyclic behaviors of the {1 0 0} faceted Co3O4 cubes enclosed by only Co2+ sites and the {1 1 2} faceted Co3O4 platelets enclosed by Co3+ and Co2+ sites [145] (Copyright 2015, American Chemical Society). 602
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
direct growth technique on the current collector, which give rise to improved surface redox pseudocapacitance and interfacial double layer capacitance for Na+ storage behaviors. It has been demonstrated that the electrochemical performances of cobalt oxide-based nanomaterials for Na+ storage are much inferior to those for Li+ storage. To classify the distinct Li+ and Na+ storage performances, Xu et al. carried out a comparative study on the possible structural changes of porous cobalt oxide during lithiation and sodiation processes by in operando synchrotron small angel X-ray scattering (SAXS), X-ray diffraction (SXRD), and X-ray absorption fine structure (XAFS) techniques [107]. They found that during the sodiation process, the porous cobalt oxide performed less pore structure changes, local wall structure changes, oxidation state shifting, and crystal structure distortions than these in the lithiation process. Further ab initio molecular dynamics (AIMD) simulations indicated that cobalt oxide possessed a low intrinsic sodiation activity, which led to inferior Na+ storage properties compared to its Li+ storage performance, and thus had less structural and electronic disturbances [107]. Overall, the electrochemical performances of cobalt oxides for SIBs are much under satisfactory and the underlying reaction mechanisms for Na+ storage are still vague. To make cobalt oxide promising for SIBs, it is highly desired to conquer the intrinsic low sodiation activity and increase the reversible specific capacity. In spite of some attempts on cobalt oxide-based hybrid structures for SIBs have been carried out in recent years, very limited progress has been made and only slow improvement on the Na+ storage performances for cobalt oxide-based nanomaterials has been achieved. Therefore, the development of cobalt oxide-based nanomaterials in the application of SIBs is still at an infant stage, and more detailed studies are urgently needed. 2.1.3. Li-O2/air batteries Rechargeable Li-O2 batteries have been receiving a great deal of interest because their theoretical specific energy density far exceeds LIBs [108–111]. In this type of batteries, the basic operation mechanism critically depends on O2 being reduced at the porous cathode surface to O22−, which then combines with Li+ from the electrolyte to form Li2O2 during discharging (O2 + 2Li+ + 2e− ↔ Li2O2), while the reverse reaction during charging is the decomposition of the peroxide [112–114]. If ambient air is used as the oxygen sources for Li-air batteries, it is essential to use high-efficiency oxygen selective membranes to effectively separate O2 from the air and to avoid the negative influences derived from CO2, H2O, and other contaminants in the air [115–117]. Currently, some other challenges, such as the issues of stability, anti-pollution, oxygen solubility and diffusivity of the electrolytes, rational design of architecture, catalytic activity, etc. of the air cathodes, and dendrite, cycling efficiency, etc. of the lithium metal anodes, still need to be urgently addressed to achieve large-scale and practical applications of Li-O2/air batteries [118]. In particular, the sluggish oxygen kinetics for OER during the charge process involving in the decomposition of Li2O2, are one of the most critical issues, which greatly restrict the overall electrochemical performances of this family of batteries. Cobalt oxides have been extensively used in traditional heterogeneous catalysis, which is largely attributed to its appealing surface redox reactivity, high adsorption capacity, and high specific area achieved by optimized design [119]. In 2007, Débart et al. compared the catalytic activities of iron-, copper- and cobalt-based catalysts for rechargeable Li-O2 batteries, and found that Co3O4 exhibited the best compromise between the discharge capacity and the cycling capacity retention with an initial capacity of 2000 mA h·g−1 and a capacity retention of 6.5% per cycle together with the lowest charging voltage of 4 V [120]. It has also been verified that cobalt-based oxides exhibit a high affinity with Li2O2, an intermediate towards LiO2, and promote the electrochemical oxidation of Li2O2, which plays an important role in enhancing the cycling performance [121–123]. These results indicate that cobalt oxide-based nanomaterials have a great potential as efficient cathode catalysts for Li-O2 batteries. To date, cobalt oxides with different structures, including mesoporous NPs [124], nanowires [125], nanosheets [126], nanorods [127], flakes [128], spherical shell [129], mesoporous spheres [122], and their derivatives, such as Co3O4/C [130–132], Co-Mn-O [133], Co3O4/Pd [134], Co3O4/N-graphene [135], CoO/rGO [136], CoO/C [121], etc., have been investigated as promising electrocatalysts at the O2 electrodes of Li-O2/air batteries [137,138]. As shown in Fig. 5a, cobalt oxide nanosheets grown on carbon paper (Co3O4 NSs/CP) were reported as the free-standing cathode for Li-air batteries [139]. In this design, a combination of hierarchically porous and interconnected Co3O4 nanosheets with highly-conductive CP facilitated electron transfer throughout the cathode, and the porous structure of Co3O4 also offered sufficient diffusion channels and abundant active sites to achieve efficient mass transfer and superior catalytic activity. The catalytic mechanism of Co3O4 exposed with different crystal facets for OER in a Li-O2 battery has been studied by firstprinciples calculations based on an interfacial model of Li2O2/Co3O4/O2 [140]. As shown in Fig. 5b (top), the computational results indicated that the O-rich Co3O4 (1 1 1)C with a relatively low surface energy and high oxygen concentration presented an excellent catalytic activity in reducing the overpotential and the O2 desorption barrier, resulted by the electron transfer from Li2O2 to the underlying surfaces. In the meantime, the basic sites of Co3O4 (1 1 0)B surface induced the decomposition of Li2O2 into Li2O and the formation of dangling Co-O bonds, which led to a high charging voltage in following cycles. Furthermore, the calculations on transition-metal (TM)-doped Co3O4 (1 1 1) suggested that the Pd-doped Co3O4 (1 1 1) exhibited a comparable charging overpotential and a better catalytic activity in reducing O2 desorption energy (Fig. 5b, bottom) [140]. Later, the same group found that the oxygen desorption barriers manifested a linear correlation with surface acidity but the Li+ desorption energies presented a volcano-like relationship with surface acidity in OER [141]. It was inferred that this family of materials with an appropriate surface acidity can achieve higher catalytic activity, reduced charging voltage, and decreased activation barrier of the rate-determinant step. Based on this result, it has been validated that CoO has comparable activity with that of Co3O4 for OER [141]. Radin et al. also confirmed the Li-O2 electrode containing a trace of Co can significantly promote the charge transport in Li2O2 through first-principle calculations [142]. For example, if the doping level of Co in the Li-O2 batteries reached 13 ppm, only 10 mV of potential was needed to drive a current density of 1 μA cm−2 through a 100 nm thick film [142]. Besides the theoretical prediction on the effect of different exposed facets on the catalytic activity, the experimental investigations 603
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
have also demonstrated that the exposed surfaces of cobalt oxides play a critical role in the catalytic performances [29,119,143]. Su et al. explored the effects of the exposed crystal planes of single crystalline Co3O4 on their catalytic properties for Li-O2 batteries [144]. They verified that the correlation of the exposed surfaces of Co3O4 crystal on reducing charge/discharge overpotential is in the order of {1 0 0} < {1 1 0} < {1 1 2} < {1 1 1} [144]. Song et al. further studied the specific catalytic active sites of the {1 0 0} faceted Co3O4 cube enclosed by only Co2+ sites and the {1 1 2} faceted Co3O4 platelets enclosed by Co3+ and Co2+ sites for both ORR and OER in Li-O2 batteries (Fig. 5c) [145]. Electrochemical measurements revealed that the {1 1 2} faceted Co3O4 platelets delivered a slightly lower initial overpotential compared to the {1 0 0} faceted Co3O4 cubes electrode. The discharge capacity of the Co3O4 platelets remained stable over 45 cycles, while that of the Co3O4 cube was stable for only 20 cycles (Fig. 5d). Based on this result, the Co3+ sites can significantly promote the adsorption properties and enable high round-trip efficiency and good cyclic stability [145]. 2.1.4. Other batteries Apart from the above-mentioned batteries, cobalt oxide nanomaterials and their composites are also employed as electrode materials for other types of batteries, such as Zn-O2 batteries [146–153] and Li-S batteries [153,154]. In spite of the high energy density for Li-O2 batteries (3458 Wh kg−1), the metallic lithium used in this type of batteries is expensive and unsafe aroused by its inherently chemical activity with water or air. Zn-O2 batteries are more promising due to the high Zn reservation on earth and the relatively stable chemical properties of Zn metals compared with lithium. It has been calculated that the theoretical specific energy density for Zn-O2 batteries is about 1350 Wh kg−1 [155] Most reported Zn-O2 batteries are operated in aqueous electrolytes, and one of the major challenges for aqueous Zn-O2 batteries is unsatisfying catalytic activity for oxygen-related reactions near the cathode. Thereby, to achieve excellent reversible capacity for electrically rechargeable Zn-O2 battery, bifunctional electrocatalysts for catalyzing the evolution and reduction of oxygen are highly required. Some cobalt oxide-base composites have been proven to be efficient bifunctional catalysts for both OER and ORR, which will be systemically discussed in the following part on electrocatalysis. Li-S batteries are believed to be another promising next-generation high-energy density (2600 Wh kg−1 or 2800 Wh L−1) rechargeable batteries [156–160]. This is largely dependent on the ample sulfur resources in earth’s crust and the high theoretical capacity (1672 mA h g−1) calculated based on the conversion reaction of sulfur to Li2S [161–169]. Cobalt oxide-based nanomaterials have been explored for Li-S batteries with the functions of effective adsorption of the soluble lithium polysulfides during discharging and active catalytic transformation of insoluble Li2S2/Li2S into soluble polysulfides during charging [154]. Chang et al. designed an electrode with bottlebrush-like Co3O4 nanoneedle arrays on flexible carbon cloth fibers (CC@Co3O4) as a multifunctional “super-
Fig. 6. Study on the electrochemical reaction steps of cobalt oxide for the application in Li-S batteries. (a) The first galvanostatic cycle at 0.5 C, (b) Raman spectra in different voltages of Co3O4 nanowires on carbon cloth (CC@Co3O4) [154], (c) TEM images of CC@Co3O4 in fully discharged states at the 100th (c-1), 200th (c-2), 300th (c-3) and 500th (c-4) cycles at 0.5C, and (d) CC@Co3O4 in fully charged states at the 100th (d-1), 200th (d-2), 300th (d-3) and 500th (d-4) cycles at 0.5 C [154] (Copyright 2016, Royal Society of Chemistry). 604
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
reservoir” to prolong the cycle life of Li-S batteries [154]. During discharging, soluble lithium polysulfides (Li2Sn, 4 < n < 8) were effectively absorbed on polar Co3O4 nanoneedle arrays and then transformed into insoluble solid Li2S2/Li2S. When charging occurred, Co3O4 nanoneedles catalyzed the electrochemical transformation reactions of Li2S2/Li2S into soluble polysulfides (Fig. 6a and b). Transmission electron microscope (TEM) was employed to monitor the morphology of the “super-reservoir” within CC@Co3O4 over different cycles, as shown in Fig. 6c and d. After 100 cycles, the uniform layer of Li2S2/Li2S formed on the Co3O4 nanoneedle surfaces in the fully discharged state (Fig. 6c-1) nearly disappeared in the fully charged state (Fig. 6d-1). No obvious change in the morphology was observed with the cycling numbers prolonged to 200 (Fig. 6c-2 and d-2) and 300 cycles (Fig. 6c-3 and d-3). Even at the 500th cycle (Fig. 6c-4 and d-4), the reversible transformation of the formation and decomposition of Li2S2/Li2S in the fully discharged and charged states, respectively, were still clearly observed. As a result, the CC@Co3O4 delivered an initial capacity as high as 1231 mA h g−1 at 0.5 C and a low capacity decay of 0.049% per cycle at 2.0 C over 500 cycles [154]. Although some explorations in cobalt oxide-based nanomaterials for different types of batteries have been reported, the electrochemical performances in these batteries are still inferior to these of Li-based batteries. Hence, to fully understand the roles of cobalt oxides in non-Li batteries, more work is still needed to figure out the underlying electrochemical reactions and ion storage mechanisms with the assistance of advanced electrochemical monitoring and characterization techniques. With in-depth investigations on the battery application of cobalt oxides, novel types of battery s may also be innovated based on this family of promising nanomaterials. 2.2. Supercapacitors Supercapacitors or ultracapacitors, can provide high power density, long cycle life (> 100,000 cycles), and high dynamic of charge propagation compared to normal rechargeable batteries [170–176]. Unlike the traditional electric double-layer capacitor (EDLC) using carbon-based electrode materials with only double-layer capacitance, the supercapacitors using metal oxide-based electrode materials have additional pseudocapacitance derived from reversible Faradic redox reactions [177,178]. There are serval key parameters to evaluate the performance of supercapacitors. (i) Capacitance: the capacitance of a super1 1 1 capacitor can be expressed as the equation of C = C + C , where C, C+, and C− are the capacitances of the whole device, the + _ positive electrode, and the negative electrode, respectively. In order to evaluate specific electrode materials for supercapacitors, specific capacitance (F g−1) is usually used via dividing the overall electrode capacitance by the mass of the corresponding electrode materials. (ii) Energy density and power density: the energy density (E, Wh kg−1) is the energy stored in a supercapacitor per unit 1 mass and can be calculated according to the equation of E = 2 CV 2 (V stands for the voltage window). While the power density (P, W kg−1) is the energy delivered per unit time, which can be obtained by the equation of P = 4R V 2 (R is the equivalent inner resistance). It is easily concluded that high energy density and power density of a supercapacitor can be achieved by increasing the electrode 1
Fig. 7. Effect of oxygen vacancies on the supercapacitor performances of cobalt oxide nanowires [222]. (a) Schematic illustration of the formation of oxygen vacancies in Co3O4 nanowires, (b) galvanostatic charge/discharge profiles of the reduced (red) and the pristine (black) Co3O4 nanowires at 2.0 A g−1, (c) cycling stability of the reduced and pristine Co3O4 nanowires at 2.0 A g−1, (d) partial charge density of the reduced Co3O4 nanowires, (e) the calculated conductivity of the reduced and pristine Co3O4 nanowires (Inset: an enlarged plot of the reduced Co3O4 nanowires) (Copyright 2014, Wiley-VCH). 605
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
capacitance and voltage and reducing the inner resistance. (iii) Rate capability and cycle life: similar to batteries, the rate capability and cycle life are two critical indexes to evaluate the performance of a supercapacitor. Among the reported pseudocapacitive type electrode materials (e.g. RuO2, Co3O4, MnO2, NiO, V2O5, SnO2, Fe2O3, etc.), Co3O4 is particularly attractive for supercapacitor application, due to its low cost, remarkable redox activity, and extremely high theoretical specific capacitance of around 3560 F g−1 [51,172,179–181]. Many work have reported on the synthesis of various cobalt oxide nanostructures for pseudocapacitor application, because the specific capacitance strongly depends on the morphologies and structures of electrode materials [182–199]. The reported capacitance values of the supercapacitors by using cobalt oxides as electrode, however, are typically lower than 1000 F g−1, and the capacity retention is also under satisfactory. This is largely due to limited ion diffusion pathways and poor electron transport for cobalt oxide semiconductors [200]. To address these issues existing in the pristine cobalt oxides, some cobalt oxide-based hybrids, such as cobalt oxide/carbon, cobalt oxide/CNTs, cobalt oxide/graphene, cobalt oxide/polymers, and cobalt oxide/other metal oxides, have developed to further enhance the capacitance [59,201–210]. The purposes for the introduction of other components to hybridize with cobalt oxides mainly include: (i) improving the specific capacitance, (ii) widening the operation voltage window, (iii) enhancing the reaction reversibility, (iv) boosting the overall electrochemical performances at high rates. Hence, relying on the exquisite design, cobalt oxide-based nanomaterials endowed with more exposed active sites, higher reaction activity, better electrical conductivity, and superior chemical stability have been synthesized to meet the requirements for high-performance supercapacitors. Hierarchically porous structures with the above-mentioned structural merits are one class of the most attractive and potential structures for the applications of catalysis and energy-related devices [211,212]. Kong et al. reported a unique combination of mesoporous Co3O4 nanoneedles with MnO2 nanosheets to form ordered 3D core-shell hierarchical porous nanowire arrays (NWAs) on nickel foam as superior anode for supercapacitors. This 3D hierarchical structure gave rise to outstanding capacitive properties with high specific capacitances of 932.8 F g−1 at a scan rate of 10 mV s−1 and 1693.2 F g−1 at a current density of 1.0 A g−1, and delivered a remarkable long-term cycling stability (only 10.2% capacitance loss after 5000 cycles at 2.0 A g−1) and a high energy density (66.2 Wh kg−1 at a power density of 0.25 kW kg−1), demonstrating a superior performance to that of individual Co3O4 nanoneedles [201]. Oxygen vacancies play a significant role in influencing the physical properties and chemical reactivity of metal oxides [213], such as TiO2 [214,215], WO3 [216], MoO3 [217–219], and MnO2 [220,221]. It has been proved that the properties of cobalt oxide-based metal oxides, such as the electronic structure, charge transport, and surface properties, are closely related to oxygen vacancies. Wang
Fig. 8. Phosphate ion functionalized cobalt oxide nanosheets for pseudocapacitors [223]. (a) Schematic illustration of the preparation of phosphate ion functionalized Co3O4 (PCO) nanosheets, (b) SEM image of PCO, (c) CV curves of the asymmetric supercapacitor (PCO//3DPG) device assembled by a PCO cathode and 3D porous graphene gel (3DPG) anode at different scan rates, (d) Ragone plots of various asymmetric supercapacitor devices, (e) schematic illustration of the surface reactivity and electrode kinetics on PCO surface (Copyright 2016, Wiley-VCH). 606
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
et al. applied NaBH4 to treat mesoporous Co3O4 nanowires to obtain reduced Co3O4 nanowires with oxygen vacancies for improving their catalytic and electrochemical performances (Fig. 7a) [222]. X-ray photoelectron spectroscopy (XPS) characterization confirmed that two new satellite peaks centered at 786.2 and 802.7 eV appeared in the Co 2p peaks of the reduced Co3O4, corresponding to a chemical state of Co2+, which suggest that the Co3+ ions were partly reduced into Co2+ ions accompanying with the formation of new oxygen vacancies. As electrode materials for supercapacitors, the galvanostatic charge/discharge profiles (Fig. 7b) indicated that the NaBH4-reduced Co3O4 nanowires delivered a high specific capacitance of 978 F g−1, more than three-fold enhancement than that of the pristine Co3O4 nanowires (288 F g−1). After 2000 charge/discharge cycles, the capacitance of the reduced Co3O4 nanowires could remain 883 F g−1 (Fig. 7c). The density functional theory (DFT) calculation revealed that a high electron delocalization occurred surrounding the positive charged oxygen vacancies (Vo2+) and the partial reduction of the neighboring Co3+ into Co2+ (Fig. 7d), resulting in much higher conductivity and electrocatalytic activity than that of the pristine Co3O4 [222]. Recently, phosphate ion functionalized Co3O4 (PCO) nanosheets with enhanced surface reactivity were fabricated by Zhai et al. as electrode materials for high-performance pseudocapacitors [223]. As displayed in Fig. 8a and b, the PCO was prepared via thermal treatment of the Co3O4 nanosheets grown on a carbon cloth in Ar atmosphere with the presence of NaH2PO2·H2O as phosphate ion sources for pseudocapacitors. The as-prepared PCO electrode delivered a large specific capacitance of 1716 F g−1 at 5 mV s−1 (∼4.5 A g−1) in 6.0 M KOH electrolyte with excellent capacitance retentions of 97% after 2000 cyclic voltammetry (CV) cycles and 85% after 10,000 cycles. Furthermore, the asymmetric supercapacitor (PCO//3DPG) device assembled by a PCO cathode and 3D porous graphene gel (3DPG) anode retained remarkable pseudocapacitive behavior with fast and reversible charge storage capability in the voltage window of 0–1.5 V (Fig. 8c), which achieved an energy density of 71.58 Wh kg−1 at an average power density of 1500 W kg−1 (Fig. 8d). This work demonstrated that the introduction of phosphate ion (H2PO4− and PO3−) onto Co3O4 nanosheets significantly enhances the surface reactivity and electrode kinetics to increase the active reaction sites and reduces the transfer resistance (Fig. 8e) [223], and thus indicated that heteroatom doping should be an effective way to enhance the electrochemical performances of cobalt oxide-based nanomaterials. In summary, even though significant advantages of low cost, excellent redox reactivity, and high theoretical specific capacitance endorse the promising application of cobalt oxides in supercapacitors, the conspicuous shortcomings of poor capacity retention and rate capacity hinder the practical utilization of this family of materials. Besides nanostructure engineering to enlarge the exposed surface area and the contact interfaces to electrolyte and to shorten the diffusion and mass transport length of ions and charges/ carriers, it has been demonstrated that the design of heterostructures, the introduction of vacancies and defects, and the implantation of heteroatoms are feasible strategies to further enhance the electrochemical performances of cobalt oxides for supercapacitor applications. 2.3. Electrocatalysis Electrocatalysis plays a key role in clean energy conversion and storage applications, such as water splitting, fuel cells, and metalO2 batteries. The primary goal for electrocatalysis process is to achieve high-efficiency conversion of common molecules in the atmosphere (e.g·H2O, O2, CO2, etc.) into highly value-added products (e.g. H2, CmHn, NH3, etc.) [224,225]. In the past decade, significant progress has been made in the exploration of some crucial electrochemical transformation reactions, especially those involving CO2, H2O, H2, and O2, such as hydrogen oxidation reaction (HOR), hydrogen evolution reaction (HER), CO2 reduction reaction (CRR), OER and ORR. Other emerging electrocatalytic reactions, such as H2O2 production and CO2/N2 reduction reactions, are also attracting more and more attentions [224]. Several factors in charge of the electrocatalysis performance are summarized as follows: (i) Onset potential: It is generally accepted that the potential value at a current density of 10 mA cm−2 (Ej = 10) can be used to compare the onset potentials of different catalysts. (ii) Overpotential: the overpotential in electrocatalysis means the extra energy over the thermodynamically expected energy needed to drive the electrocatalytic reactions, which is usually measured in the unit of mV. (iii) Tafel slope: Tafel slop is another crucial parameter to evaluate the activity of electrocatalysts, and it can be sued to evaluate the rate-determining steps of the electrocatalytic reactions, based on the equation of = b·log(j/ j0 ), where η, b, j, and j0 represent the overpotential, Tafel slope, current density, and exchange current density, respectively. (iv) Stability: in spite of many explorations have been performed to evaluate the stability of electrocatalysts, such as extended reaction time, varying reaction systems (e.g. acidic and alkaline electrolytes), and morphology and structure examination changes, until now, no well-defined parameter is utilized to judge the stability of a catalyst. (v) Turnover frequency (TOF) and mass activity: The TOF value can be calculated according to the following equation. TOF = (j ·S)/(4·F·n), where j, S, F and n refer to the current density, electrode area, Faraday constant (96,500 C mol−1) and the moles of the active materials, respectively. Besides, the mass activity is evaluated by j/m, in which j is measured current density and m is the mass loading of the working electrode. For electrocatalysis reactions, the selection of catalysts is very critical especially in practical applications, as we need to make a balance between cost and effectiveness. Great efforts have been made on the design of efficient but low-cost electrocatalysts. Cobalt oxides, as one of semiconductor-type catalysts with ample natural resources, are low cost and high catalytic efficiency, and expected to be one promising cost-efficient catalyst in electrocatalytic reactions to replace the low-abundant and noble metals (e.g. Pd, Ir, Pt, Au, etc.) and their alloys (e.g. Pt/Ir, Pt/Au, Pd/Au, etc.), if we can increase the active site numbers and enhance the activity of each site of the oxides. Cobalt oxide-based nanomaterials have been reported for various electrocatalytic reactions, including the widespread OER, ORR, HER, and CRR, as well as the traditional oxidation of methanol [226–230]. In this section, we mainly focus on the first four representative types of gaseous electrocatalytic reactions. Considering the fact that two or more reactions are often involved in the 607
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
practical applications, such as fuel cells, metal-O2 batteries, water splitting, and other related fields, bifunctional and multi-functional cobalt oxide-based catalysts are particularly stressed with a separated focus. 2.3.1. OER The reaction process of OER mainly includes a four proton-coupled electron transfer step and formation of oxygeneoxygen (OeO) bond. In acidic conditions, the reaction happens accompanied by the oxidation of water molecules to give one oxygen molecule and four protons (2H2O → 4H+ + O2 + 4e−), while the oxidation of hydroxyl groups to produce H2O and O2 (4OH− → 2H2O + O2 + 4e−) is involved in basic conditions [231–233]. Compared to RuO2 and IrO2, which are regarded as the state-of-the-art catalysts for OER with low overpotential and Tafel slope, the less active Co3O4 is still attractive due to its easy access in nature, low input-output ratio for industrialization, and good environmental compatibility [234–239]. In spite of the geometry-site-dependent performances are suggested for OER, the specific active sites of Co3O4 remain elusive. Some research indicated that Co3+ species on the surface of Co3O4 crystal should be active sites for OER catalysis [240–243], however, other studies revealed that the presence of Co2+ is also helpful for boosting OER performances, which are responsible for the formation of cobalt oxyhydroxide (CoOOH) [244–246]. Nevertheless, the sluggish kinetics and the relatively considerable overpotential are major challenges for the application of cobalt oxides in OER [247]. To meet the requirements of practical application of this family of promising materials, the electrocatalytic performances of cobalt oxides must be further enhanced by executing some effective strategies, such as the introduction of porous structures [185,248,249], the hybridization with carbon matrix [250–253], and the doping of some metal atoms [254–258], which help increase the contact interfaces, the electric conductivity of the system, and the number of the active sites of the materials. To increase the accessible surface area and the active sites on the surfaces of the electrocatalyst, 2D graphene-like holey Co3O4 nanosheets (GHCS) (inset in Fig. 9a and b) were fabricated by a bottom-up self-assembly approach and were utilized as an electrocatalyst for OER in 0.1 M KOH [248]. As shown in Fig. 9a, two peaks located at around 0.285 and 0.575 V in the anodic scan were assigned to the oxidation of Co2+ to Co3+ and Co3+ to Co4+, respectively. The linear sweep voltammetry (LSV) curves revealed that
Fig. 9. OER performance of 2D graphene-like holey cobalt oxide nanosheets [248]. (a) CV curves in 0.1 M KOH at a scan rate of 30 mV s−1, (b) linear sweep voltammetry (LSV) curves at 10 mV s−1, (c) Tafel plots, and (d) cycling stabilities of graphene-like holey Co3O4 nanosheets (GHCS) (Inset in a shows the schematic illustration of GHCS; inset in b shows SEM image; HCF represents holey Co3O4 nanoflowers; BCC represents bulk Co3O4 nanocubes) (Copyright 2016, Elsevier). 608
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
a low onset potential of 0.617 V (0.25 mA cm−2) was displayed for GHCS, which is almost comparable to commercial IrO2 (0.611 V) (Fig. 9b). Specifically, at a potential of 0.8 V, the obtained GHCS gave an OER current density of 12.26 mA cm−2, much higher than holey Co3O4 nanoflowers (HCF, 9.46 mA cm−2), bulk Co3O4 nanocubes (BCC, 2.45 mA cm−1), and referential IrO2 (3.42 mA cm−2). Moreover, the GHCS possessed a Tafel slope of 66 mV dec−1, which is lower than other referential samples (Fig. 9c). Long-term stability tests confirmed that the GHCS exhibited only 2.0% loss after 2000 cycles, greatly outperforming HCF (4.9%) and BCC (16.5%), respectively (Fig. 9d) [248]. Therefore, it is concluded that the design of graphene-like holey structures is very helpful to enhance the electrocatalytic performances of the materials, owing to significantly improved active surfaces and decreased energy barriers for mass conversion and transfer. The second strategy to enhance the electrocatalytic performance of cobalt oxides is to increase the intrinsically-inferior electrical conductivity. Ma et al. reported that carbon coating on the surface of cobalt oxide nanowires arrays could increase the electrical conductivity of each single wire, and thus significantly enhance their overall electrochemical properties. In this study, Co3O4-carbon porous NWAs (Co3O4C-NA) were prepared by carbonization of Co-based MOF precursor grown on Cu foil under an inert condition as free-standing and binder-free electrode for OER (Fig. 10a and b) [251]. When operated in alkaline solutions (0.1 M KOH), the electrode afforded a low onset potential of 1.47 V (Fig. 10c) and a stable current density of 10.0 mA cm−2 at 1.52 V for at least 30 h, accompanied by a high Faradaic efficiency of 99.3%. Moreover, the Co3O4C-NA hybrid electrode exhibited a Tafel slope value of 70 mV dec−1 (Fig. 10d), which is much lower than those of Co-based MOF (142 mV dec−1) and IrO2/C (97 mV dec−1). As displayed in the electrochemical impedance spectrum (EIS) in Fig. 10e, Co3O4C-NA exhibited a smaller contact and charge transfer impedance than the one without carbon coating. Furthermore, the porous structure of NWAs and the direct growth of active materials on Cu foil brought the electrode with enlarged active surface area and excellent structural stability for long-term cycles [251]. The significant enhancement of the electrocatalytic performances of the carbon-coated Co3O4 suggests the effectiveness of the strategy on increasing the electrical conductivity of the materials. 2.3.2. ORR Generally, there are two reaction pathways available for ORR. In acidic solutions, a four proton-electron transfer step with the reduction of O2 to H2O (O2 + 4H+ + 4e− → 2H2O) and a two-proton-electron one with reducing O2 to H2O2 (O2 + 2H+ + 2e− → H2O2) occur. The former is the preferred pathway for ORR and is also a desirable pathway for the operation of fuel cells. The latter, however, is detrimental to the ORR reaction efficiency albeit attractive for H2O2 production [259]. If operated in alkaline
Fig. 10. OER performance of carbon coated cobalt oxide porous nanowire arrays (NWAs). (a) Schematic illustration of the fabrication of Co3O4carbon porous NWAs (Co3O4C-NA), (b) SEM image of Co3O4C-NA [251], (c) polarization curves of cobalt oxides for OER (Inset: optical image of the generation of O2 bubbles on the electrode surface by using Co3O4C-NA as catalyst at 1.70 V), and (d) Tafel plots of the electrocatalytic reactions in an O2-saturated 0.1 M KOH solution at a scan rate of 0.5 mV s−1 [251], (e) electrochemical impedance spectrum (EIS) of the cobalt oxide electrode during reaction recorded at 1.60 V (Inset: the corresponding equivalent circuit diagram) [251] (Copyright 2014, American Chemical Society). 609
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
environment, O2 can be reduced by either the four-proton-electron pathway to form OH− (O2 + 2H2O + 4e− → 4OH−), or the two-proton-electron pathway to produce HO2− and then OH− (O2 + H2O + 2e− → HO2− + OH−; HO2− + H2O + 2e− → 3OH−) [260]. Pt metal has been considered as the best ORR catalyst. The limited resources and high cost, however, make the Pt-based catalysts less attractive for its commercial applications [261]. Cobalt oxides have been explored as an alternative ORR catalyst to replace the expensive Pt catalyst for fuel cells [262], such as microbial fuel cells (MFCs) [263], anion exchange membrane fuel cells (AEMFCs) [264], and the reduced-temperature solid oxide fuel cells (RT-SOFCs) [265]. It has also been proven that the active sites in Co3O4 for ORR are the surface Co2+ ions, and thus cobalt oxides exposed with Co2+ enriched {1 1 1} crystal planes show the best electrocatalytic activity compared to the cobalt oxides exposed with {1 1 0} and {1 0 0} faces nanorods (NR), nanocubes (NC) and nanooctahedrons (OC) with the different exposed nanocrystalline surfaces ({1 1 0}, {1 0 0}, and {1 1 1}), uniformly anchored on graphene sheets, which has allowed us to investigate the effects of the surface structure on the ORR activity. Results show that the catalytically active sites for ORR should be the surface Co2+ ions, whereas the surface Co3+ [266,267]. Unfortunately, as it appears in other electrocatalytic reactions, the intrinsically low electrical conductivity of cobalt oxides is a big barrier to achieve satisfying electrocatalytic performance in ORR. Hence, cobalt oxide-based nanomaterials with doping or the composites by hybridizing with highly conductive materials have been developed to address the critical issue to improve their electrochemical properties [28,268–276]. Among the cobalt oxide-based nanostructures, the formation of ternary mixed valence oxides with a spinel structure are an important class of materials to provide outstanding ORR catalytic activity in alkaline conditions. Liang et al. fabricated MnCo2O4/Ndoped graphene hybrids (MnCo2O4/N-rmGO) as high-performance ORR electrocatalysts (Fig. 11a), by taking advantage of the higher catalytic activity of MnCo2O4 than pristine Co3O4 and the strong coupling with N-doped graphene [268]. This hybrid showed more positive onset (0.95 V) and peak (0.88 V) potentials than these of the Co3O4/N-doped graphene (Co3O4/N-rmGO) hybrids (0.93 and 0.86 V, respectively), and presented only 20 mV of the half-wave potential difference to the Pt/C catalyst at the same mass loading (Fig. 11b). Moreover, rotating ring-disk electrode (RRDE) measurements in 1.0 M KOH solution confirmed that the yield of peroxide species was less than 10% for the hybrid, much lower than that of the corresponding physical mixture between MnCo2O4 NPs and Ndoped graphene nanosheets (about 15%). It was also found that Mn substitution mediates the NP size and phase, and provides further potential to enhance the catalytic performances [268]. Therefore, it is pretty effective to enhance the ORR performance of cobalt oxide to improve the electrical conductivity by incorporating with highly conductive species, such as nanocarbon and graphene. It is also demonstrated that doping is another useful way to further improve the catalytic activity of cobalt oxides. 2.3.3. HER HER is another important electrocatalytic reaction, which can generates H2 fuel via the reduction of water (2H+ + 2e− → H2 or 2H2O + 2e− → H2 + 2OH−). HER has also gained considerable attention primarily due to its potential applications in hydrogenbased fuel cells, alkaline water electrolysis, petroleum refining, NH3 synthesis, etc. [277–285]. The general catalytic mechanisms of HER in both acidic and alkaline media have been fully understood [286–288], however, the produce of other by-products is also possible in some cases when use some specific electrolytes [289]. In the electrocatalysts for HER, the well-known and best catalysts at present is still Pt and Pt-group metals because of their fast kinetics and nearly zero overpotential [290–292]. Cobalt oxides have been explored as potential catalysts for HER, but the activities in their pristine form are very poor [293]. It has been reported that the hybridization with highly conductive or chemically active components, such as nanocarbon, metal sulfides and metals, can greatly increase their electrochemical properties in HER application [293–296]. Yan et al. reported that an excellent HER activity could be achieved by the design of 3D core-shell Co/Co3O4 nanosheets, in which the metallic cobalt was core and the amorphous cobalt oxide was shell [293]. The 3D metallic framework was beneficial for
Fig. 11. ORR performance of MnCo2O4/N-doped graphene hybrids (MnCo2O4/N-rmGO) [268]. (a) Schematic illustration of ORR on MnCo2O4/NrmGO, (b) rotating-disk electrode (RDE) voltammograms in O2-saturated 1.0 M KOH at a sweep rate of 5 mV s−1 at 1600 rpm (Copyright 2012, American Chemical Society). 610
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
increasing electrical conductivity, and the thin amorphous Co3O4 surface layer (2–5 nm in thickness) enriched with hydroxyl groups could effectively improve the HER activity. This 3D Co/Co3O4 core-shell structure exhibited a very small onset potential of 30 mV and a current density of 10 mA cm−2 at an overpotential of ∼90 mV in 1.0 M KOH [293], providing a proof-of-concept on the ideas of the design of novel low-cost electrocatalysts toward high HER performances. 2.3.4. CRR To deal with the limited fossil feedstocks on earth and the continuously increasing CO2 emissions, the electroreduction of CO2 into fuels have become a very attractive topic. Depending on this reaction, a large quantity of useful products, such as CO, methane, formate, formic acid, methanol, formaldehyde, and C2+ hydrocarbons and oxygenates, can be formed via various multi-electron transfer pathways [297–312]. Electrocatalytic CO2 reduction mainly incorporates the following four steps: (i) adsorption of CO2 on active sites, (ii) activation of CO2 to CO2− or HCO2− or other intermediate, (iii) dissociation of CeO bonds accompanying with the transfer of protons and electrons, (iv) desorption of the products from active sites. In these steps, the activation step is the critical bottleneck in the reduction of CO2. Recent studies have indicated that the reduction reactions by using cobalt oxide and their derivations as electrocatalysts can be triggered at low overpotentials [313–315]. With a combination of theoretical and experimental study, Gao et al. found that oxygen-deficient cobalt oxide ultrathin nanosheets and partially oxidized atomic cobalt nanosheets had higher intrinsic activity and selectivity towards formate production, compared to the normal surface cobalt atoms on bulk samples [313,314]. Gao et al. compared the electrochemical performance of the Vo-rich Co3O4 nanosheets in CO2 reduction reaction (Fig. 12a). The presence of O(II) vacancy could facilitate CO2 adsorption and spontaneous HCOO− desorption, and thus significantly enhanced the catalytic performance of the Co3O4 nanosheets. The Vo-rich Co3O4 nanosheets exhibited a current density of 2.7 mA cm−2 at the potential of −0.87 V, almost two times larger than that of the Vopoor Co3O4 layers (Fig. 12b), and a decreased onset potential (0.78 vs. 0.81 V) as well as a lower Tafel slope (37 vs. 48 mV dec−1). At a given potential of −0.87 V, the Vo-rich Co3O4 nanosheets achieved a maximum faradaic efficiency of 87.6% after 4 h for producing formate, much higher than 67.3% of the Vo-poor layers (Fig. 12c), and presented negligible decay with a Faradaic efficiency of over 85% in 40 h (Fig. 12d). This study proved that the proton transfer in the CO2 reduction on Co3O4 surface is a rate-limiting step. The oxygen(II) vacancies in the surface layers decreased the activation barrier of the proton transfer from 0.51 to 0.40 eV and had the function of stabilizing the formate anion radical intermediate (HCOO−*) [313]. Partial oxidation of atomic cobalt layers or four-atom-thick nanosheets with coexistence of cobalt metal and cobalt oxide domains generated a current density of 10.59 mA cm−2 at −0.85 V (Fig. 13a) with the highest Faradaic efficiency (90.1%) at an overpotential to 0.24 V in the production of formate (Fig. 13b). Over 40 h, the partially oxidized layers did not show obvious decay in current density with approximately 90% formate selectivity [314]. The superior performance of the cobalt or cobalt oxide nanosheets can be attributed to their atomic thickness endowing Co3O4 with ample active sites for a large number of adsorption of CO2 and the enhanced electronic conductivity. Gao et al. also studied the effects of the thickness of nanosheets on the catalytic activity, based on two types of Co3O4 nanosheets in thicknesses of 1.72 and 3.51 nm, respectively [315]. It was found that the thinner the nanosheets, the better electrocatalytic performances were achieved. The Tafel slopes were 56 mV dec−1 for the nanosheets in 1.72 nm and 79 mV dec−1 for those in 3.51 nm, while it reached 97 mV dec−1 for bulk samples. The 1.72 nm Co3O4 layers delivered a current density of
Fig. 12. Effect of oxygen vacancies on CO2 electroreduction reaction of cobalt oxide layers [313]. (a) Schematic illustration of the formation of Vorich and Vo-poor Co3O4 layers, (b) LSV curves in CO2-saturated (solid line) and N2-saturated (dashed line) 0.1 M KHCO3 solutions, (c) Faradaic efficiencies of formate production, and (d) chronoamperometry results at −0.87 V of Vo-rich and Vo-poor Co3O4 layers (Copyright 2017, Nature Publishing Group). 611
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
Fig. 13. Electrochemical properties of CO2 reduction reaction of cobalt oxides. (a) LSV curves in a CO2-saturated (solid line) and N2-saturated (dashed line) in 0.1 M Na2SO4, and (b) the Faradaic efficiencies for formate using different Co-based catalysts at each given potential for 4 h [314] (Copyright 2016, Nature Publishing Group). (c) LSV curves in a CO2-saturated (solid line) and N2-saturated (dashed line) in 0.1 M KHCO3 solutions, and (d) the Faradaic efficiencies for formate using the different thickness of Co3O4 catalysts with at each given potential for 4 h [315] (Copyright 2016, Wiley-VCH).
0.68 mA cm−2 at −0.88 V, which was more than 1.5 and 20 times higher than these of Co3O4 layers 3.51 nm in thickness and the bulk counterpart, respectively (Fig. 13c), and the maximum faradaic efficiency of the 1.72 nm thick Co3O4 layers (over 60% for 20 h) was much higher than that of the 3.51 nm thick Co3O4 layers (51.2%) and the bulk counterpart (Fig. 13d) [315]. These investigations demonstrated that the manipulation of the chemical states of surface atoms and vacancies are very critical to increase the active sites and to lower the activation barrier of CO2 reduction reactions. 2.3.5. Bifunctional cobalt oxide-based catalysts In contrast to mono-functional catalysts which exhibit excellent activity towards one specific electrochemical reaction type, some catalysts are capable of catalyzing two different types of reactions simultaneously. The discovery of bifunctional catalysts is attracting considerable attention in the recent years due to the fact that there usually is more than one reaction in a certain practical application. For example, an entire water splitting process involves two half-reactions, HER and OER, for the evolution of hydrogen and oxygen, respectively [258,316–318]. In URFCs, the electrochemically conversion of H2 into electrical energy and the splitting of water into H2 and O2 occur simultaneously [319–322]. Compared to the mono-functional catalysts, the bi-functional catalysts are less common and more challenging to achieve high efficiencies at the same time for catalyzing two different reactions. As mentioned above, cobalt oxides in pristine state have good OER-favorable activity but inferior activities for ORR and HER. Hence, to make cobalt oxides applicable as bi-functional catalysts, one of the simplest approaches is to hybridize cobalt oxides with the materials with complementary functions. For example, the cobalt oxide-based bifunctional catalyst system can be easily constructed by integrating various cobalt oxide structures with functional carbonaceous materials, especially heteroatom-doped carbon, which has been considered as one of the most promising ORR-favorable catalysts [323]. Until now, many pioneering work has been reported using cobalt oxide-based nanomaterials as bifunctional catalysts for two pairs of reactions, HER/OER [294,295,324] and 612
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
OER/ORR [325,326], which are essential for the applications for water splitting, fuel cells, and metal-O2 batteries. Menezes et al. found that the Co3O4 nanochains exhibited superior catalytic performances toward OER than the solvothermally synthesized Co3O4 nanostructures and commercial Co3O4 and CoO in both alkaline (PH = 13) and neutral (PH = 7) solutions [325]. To enhance its ORR activity in alkaline media, the carbon-supported Co3O4 nanochains was fabricated by mixing with carbon black and it was observed that the resulting Co3O4/C hybrid outperformed some precious metal catalysts in terms of catalytic activity and chronoamperometric stability [325]. Besides carbon black, heteroatom-doped carbon was also introduced to couple with cobalt oxides improve the ORR and OER performances. Mao et al. reported a CoO NPs/N-doped graphene (N-CG-CoO) hybrid synthesized by the growth of CoO NPs on 3D hollow crumpled N-doped graphene as a bi-functional catalyst for both OER and ORR in alkaline solutions [326]. Compared to the CoO NPs/graphene (CG-CoO) hybrid without nitrogen doped, the N-CG-CoO hybrid exhibited enhanced ORR catalytic activity with a more positive onset potential of about 0.90 V, a more positive half-wave potential of 0.81 V at 1600 rpm, and a smaller Tafel slop of around 48 mV dec−1, etc. Moreover, the onset potential of N-CG-CoO was close to that of the commercial Pt/C (Fig. 14a). The durability of N-doped hybrid, with 12% decay after 20,000 s at 0.75 V, was superior to that of the Pt/ C (41%) in 1.0 M KOH electrolyte, as indicated in Fig. 14b. The OER study showed that the N-CG-CoO catalyst had a small overpotential of about 0.34 V (Fig. 14c) and a small Tafel slope of 71 mV dec−1 (Fig. 14d). These excellent catalytic properties suggested that the N-CG-CoO catalyst is a promising bifunctional electrocatalyst for water splitting and fuel cell applications [326]. In summary, many explorations on cobalt oxides and functional carbon hybrids has proven to be efficient bifunctional electrocatalysts, however, these performances are still far away from the requirements for the practical devices, such as metal-O2 batteries. Hence, further optimization in cobalt oxide/carbon hybrids is required in terms of the optimum ratio of the components, the sources of carbon, the dimensionality of cobalt oxides, the approaches of hybridization, and the avoidance of possible side reaction. Moreover, an in-depth exploration is needed for the formation of heterostructures with other components to simultaneously achieve high OER and ORR activities. 2.3.6. Multi-functional cobalt oxide-based catalysts Based on current progress, individual cobalt oxides can hardly be used for multi-functional electrocatalysts. Cobalt oxide-based hybrids, however, have great potential to achieve this goal via an ingenious combination of individual merits derived from each component to present an attractive synergetic effect. Nevertheless, it is challenging to design suitable structures with desired functions for simultaneously catalyzing different types of reactions. The reported multi-functional catalysts currently are largely focusing on the combination of cobalt oxide structures with functional graphene nanosheets [327,328].
Fig. 14. ORR and OER performances of of CoO NPs/N-doped graphene (N-CG-CoO) hybrid [326]. (a) ORR polarization curves in an O2-saturated 1.0 M KOH electrolyte at a sweep rate of 5 mV s−1 with a rotation speed of 1600 rpm, (b) chronoamperometric responses at 0.75 V in an O2saturated 1.0 M KOH electrolyte, (c) Oxygen evolution currents in an Ar-saturated 1.0 M KOH electrolyte at 20 mV s−1 (Inset: the oxidation peak of CoO), (d) Tafel plots of OER currents of N-CG-CoO (Copyright 2013, Royal Society of Chemistry). 613
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
Liu et al. systematically studied a cobalt/cobalt oxides/nitrogen-doped reduced graphene oxide (Co-CoO/N-rGO) hybrid composed of metallic cobalt, cobalt monoxide, and graphene as catalysts for OER, ORR, and HER [327]. As shown in Fig. 15a–c, Co/CoO nanorods with sub-micrometer lengths were anchored on N-rGO sheets. Among their synthesized cobalt-based materials, including CoO/N-rGO, Co-CoO/N-rGO, and Co-CoO + N-rGO, CoO/N-rGO showed the highest activity for OER (Fig. 15d) with much smaller charge transfer resistance (Rct) (∼20 Ω) than CoO/N-rGO (∼30 Ω), while Co-CoO/N-rGO exhibited the highest HER activity (Fig. 15e), superior to individual N-rGO, CoO/N-rGO hybrid, and Co-CoO + N-rGO physical mixture. As for ORR, as indicated in Fig. 15f, a higher activity was achieved by Co-CoO/N-rGO hybrid than CoO/N-rGO with an onset potential of ∼0.88 V and a halfwave potential of 0.78 V. Notably, similar Tafel slopes between Co-CoO/N-rGO (40 mV dec−1) and Pt/C (38 mV dec−1) were observed, especially in low current densities. Furthermore, stable current densities of −2.0 mA cm−2 at 0.75 V for ORR in 0.1 M KOH and −10 mA cm−2 at −0.15 V for HER in 1.0 M KOH were well-maintained in Co-CoO/N-rGO catalyst. Although the Co-CoO/N-rGO catalyst exhibited lower activity than Pt/C in the rotating-disk electrode (RDE) tests (Fig. 15f), it presented improved performance approaching to Pt/C + IrO2 cathode in the zinc-air battery tests [327]. Moreover, Co3O4 NPs anchored on N-doped rGO were also used as an effective multifunctional catalyst for H2O2 reduction, oxygen reduction and evolution reaction [328]. The resultant Co3O4/N-rGO composites demonstrated excellent electrocatalytic activities with a direct reduction to H2O2 at −0.6 V and a detection limit of 0.1 mM towards H2O2 sensor, a peak potential of −0.26 V (vs. Ag/AgCl) for ORR, and an onset potential of 1.54 V and a smaller Tafel slope (204 mV dec−1) compared with 20% Pt/C (308 mV dec−1) for OER [328]. 3. Fabrication and electrochemical performance of cobalt oxide nanomaterials In principle, nanomaterials represent the materials with at least one dimension less than 100 nm [329]. Nanomaterials can be generally classified into four categories based on their dimensionality. Materials with all three dimensions confined into 1–100 nm scales are considered as zero-dimensional (0D) nanomaterials, such as nano-sized particles, clusters, and quantum dots (QDs). Onedimensional (1D) nanomaterials possess two dimensions less than 100 nm and one ultra-long dimension along the radial direction, include nanorods, nanowires, nanoneedles, nanotubes, nanofibers, etc. Two-dimensional (2D) nanomaterials consist of two infinite dimensions in length and width and one nanoscale thickness dimension, and often present a paper-like morphology, involving nanosheets, nanodiscs, nanofoils and nanomeshes. One of the most typical examples of 2D nanomaterials is the isolated single-atomic carbon layer, graphene, which was first exfoliated from bulk graphite in 2004. Compared with other dimensional nanostructures, three-dimensional (3D) nanoarchitectures are assembled from different types of low-dimensional nanoarchitectures and their overall size in three dimensions is often beyond nanometer scales. The representative 3D nanomaterials include micro-/nanospheres, porous/ hollow structures, dendritic structures, hierarchical structures, heterogeneous networks, etc., assembled from 0D, 1D, 2D, or multiscale nanocomponents. In this section, we focus on the recent progress in the synthesis and fabrication of cobalt oxides in various dimensionalities,
Fig. 15. OER, HER, and ORR performances of cobalt oxide/graphene hybrids [327]. (a) SEM image, (b) low-magnification and (c) high-magnification TEM images of cobalt/cobalt oxides/nitrogen-doped reduced graphene oxide (Co-CoO/N-rGO) hybrid, (d) OER, (e) HER and (f) ORR polarization curves at a sweet rate of 10 mV s−1 with a rotating rate of 1600 rpm (Copyright 2015, Wiley-VCH). 614
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
including Co3O4, CoO, and mix-valenced CoxOy with different constitutional nanostructures, for typical electrochemical energy applications. 3.1. 0D cobalt oxide nanostructures 0D nanostructures, are usually defined as the overall size ranging from 1 to 100 nm in three dimensions, such as NPs and QDs. Owing to the unique confinement effect derived from electrons, excitons, holes, phonons, and/or plasmons in all dimensions, the physical and chemical properties of 0D nanomaterials usually are isotropic [330]. To date, cobalt oxide NPs have been synthesized via various methods, e.g. microwave irradiation [331], pulsed-laser ablation [332–334], mechanochemical reactions [335], laser ablation in liquid [336], bio-templating route [337], ionic liquid-assisted method [338], surfactant-assisted method [339], hydrothermal method [340], etc., for a wide range of applications, such as heterogeneous catalysts, batteries and supercapacitors [341–345]. As for the electrochemical reactions, the effect originating from the reduced sizes and the defined shapes of these NPs include increased surface area, more active crystal defects, and exposed highenergy crystal facets. The controlled synthesis of cobalt oxide NPs with tunable sizes and shapes thus has been attracting considerable interest. One of the simple strategies is to tune the size and shape by adjusting the synthesis parameters, such as the precursor concentration, the reaction temperature, and the reaction time [346–352]. For example, in 2005, the wurtzite-type hexagonal cobalt monoxide (CoO) nanocrystals with 11 ± 1.7 nm in width and 40 ± 7.3 nm in length were synthesized via thermal decomposition of cobalt(III) acetylacetonate precursor in oleylamine at 200 °C [347]. By change the molar ratio of the solutions and the reaction temperatures, hexagonal pyramid-shaped nanocrystals and cubic nanocrystals have been obtained [347]. It also has found that solid cubic CoO nanoparallelepipeds can be transformed into metallic cobalt hollow structures by thermolysis the solid cubic crystals in oleylamine at a temperature in 270–290 °C [348]. Later, Varón found that the presence of water in reaction solutions can facilitate the formation of a CoO hollow structure by a Kirkendal effect at the room temperature, otherwise Co-CoO core-shell NPs were obtained without water. These hollow CoO NPs, however, were only an intermediate state, which finally underwent disintegration to form minor CoO NPs of 2–3 nm in sizes [352]. As another important type of 0D nanostructures, QDs, have attracted much attention for their well-defined structure in terms of both size and shape, and thus exhibit extraordinary chemical and physical properties, such as the unique quantum confinement effect [353,354]. There are two basic features of this appealing material, e.g. small size and narrow size distribution. The size of QDs is generally not more than twice the exciton Bohr radius of the corresponding semiconductors in all three dimensions, and QDs are mostly in spherical or quasi-spherical morphologies with a diameter in the range of 2–20 nm. Hence, the accurate control of nucleation and growth steps during the synthesis is essential to produce well-defined QDs. In 2014, Zhang and his co-workers synthesized mono-dispersed Co3O4 QDs via a facile reverse micelle strategy assisted with
Fig. 16. Morphologies and performances of 0D cobalt oxide QDs. (a, b) TEM, magnified TEM (Inset: selected area electronic diffraction (SAED) pattern), and (c) high-resolution transmission electron microscopy (HRTEM) images of Co3O4 QDs [356] (Copyright 2014, Royal Society of Chemistry). (d) Comparison of OER activities in terms of overpotentials (η) at 10 mA cm−2 and Tafel slopes for cobalt-based catalysts [357] (Copyright 2017, American Chemical Society). 615
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
microwave-promoted solvothermal treatment [355]. In the synthesis, the introduction of microwaves contributed to accelerating the particle nucleation rate, by which the reaction time was easily controlled to effectively avoid the overgrowth of NPs. The specific size of Co3O4 QDs was around 3–4 nm, which is close to the exciton Bohr radius. Moreover, the ultraviolet–visible (UV–Vis) spectra revealed that the absorption edge of the obtained QDs exhibited a clear blue shift compared with that of the NPs synthesized by a solid-state reaction (Co3O4-SSR), which is consistence with the obvious change of exterior color from black (Co3O4-SSR) to yellow (Co3O4-QDs). Benefit from the confinement effect, the conduction band shifted to a higher level than the H+ reduction potential occurred, which makes these QDs attractive for overall water splitting under visible-light irradiation [355]. Subsequently, a one-step, surfactant-free, oil-bath-assisted method by using benzyl alcohol (BA) together with ammonium hydroxide has been employed to fabricate monodispersed, water-dispersible Co3O4 QDs (BA-Co3O4) with sizes of approximately 4.5 nm [356]. As shown in Fig. 16, the as-synthesized QDs presented uniform size (∼4.5 nm) and cube-like shape (Fig. 16a and b), and without obvious aggregation. Furthermore, these QDs showed well-defined Co3O4 spinel crystal structure as evidenced by the observed lattice spacing of 0.200 nm (Fig. 16c), which corresponds to the distance between the {4 0 0} planes. Without negative effect derived from agglomeration, the resultant product exhibited excellent visible light-driven oxygen evolution activities under mild pH conditions [356]. Quite recently, Zhang et al. reported the fabrication of Co3O4−δ QDs with a smaller crystallite size of about 2 nm through multicycle lithiation/ delithiation of mesoporous Co3O4 nanosheets [357]. Meanwhile, the concentration of oxygen vacancies and the population of Co2+ over Co3+ could be tuned. When applied for an OER electrocatalyst for water splitting, as shown in Fig. 16d, Co3O4−δ QDs prepared with 20 lithiation/delithiation cycles (Co3O4−δ QDs-20) exhibited an overpotential of only ∼270 mV at a current density of 10 mA cm−2, which is superior to many previously reported Co-based catalysts and the state-of-the-art IrO2, demonstrating the superior electrochemical reactivity of the 0D Co3O4 QDs [357]. Overall, 0D cobalt oxide nanostructures have been widely investigated for electrochemical applications. Two major issues, however, exist in the high-yield synthesis of highly homogeneous cobalt oxide QDs. The first challenge is the strong tendency of QDs to aggregate into larger aggregates during the drying process and electrode fabrication step. To address this problem, the strategy to hybridize with other supporting materials (e.g. graphene) is an effective solution, which will be discussed in the following sections. Based on this method, the homogeneous dispersion of cobalt oxide QDs can be achieved and the active surfaces and edges area can be maximally utilized. Another issue is that the controllable synthesis of uniform cobalt oxides NPs or QDs in high yields is still very difficult. To find simpler and more efficient preparation approaches, further efforts on the nucleation-and-growth mechanisms of QDs have to be further made.
Fig. 17. Morphologies and performances of 1D cobalt oxide nanostructures. SEM image of (a) nanorods [377] (Copyright 2005, Wiley-VCH), (b) needle-like nanotubes [379] (Copyright 2008, Wiley-VCH), (c, d) nanoneedles [380] (Copyright 2008, Royal Society of Chemistry). (e) Cycling stability for Li+ storage of three nanoneedle samples prepared by annealing the precursor β-Co(OH)2 nano-needles in air at 200 °C (A200), 300 °C (A300), and 400 °C (A400) at a current density of 150 mA g−1 [380] (Copyright 2008, Royal Society of Chemistry). 616
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
3.2. 1D cobalt oxide nanostructures The typical 1D nanostructures include nanowires, nanorods, nanofibers, nanoneedles, nanotubes, and nanobelts, and have presented great potential for the applications in electrochemical devices [354,358–374]. Among them, the aspect ratio (length/diameter) of the rod-like nanostructures is often less than 10, while for nanowire/fibers it is usually over 10 [375]. Versatile approaches have been developed to synthesize 1D nanostructures, such as template-assisted wet-chemical routes, chemical vapor deposition (CVD), physical vapor deposition (PVD), thermal decomposition, wet-chemical synthesis, and electrodeposition [177,376]. It is generally accepted that 1D nanostructures can provide a good platform to study the dependence of anisotropic electrical, thermal, and mechanical properties of nano-sized materials [375]. Li et al. fabricated various Co3O4 nanostructures, including nanotubes, nanorods and NPs [377]. Co3O4 nanotubes were fabricated by porous alumina template-assisted chemical decomposition of Co(NO3)2·6H2O precursor. Scanning electron microscope (SEM) confirmed that the length of Co3O4 nanotubes was between ca. 20 and 40 μm and the wall thickness was approximately 20–30 nm. The Co3O4 nanorods with uniform diameter in the range of 100–120 nm (Fig. 17a), were synthesized from inverse microemulsions [377,378]. The Co3O4 NPs with a typical size of around 100 nm were prepared by ball-milling and thermal decomposition of Co (NO3)2·6H2O. In terms of the electrochemical performance as anode materials for LIBs, the Co3O4 nanotube electrode exhibited an initial discharge capacity of 850 mA h g−1 with a wide discharge voltage window and a long plateau, which was higher than that of nanorod (815 mA h g−1) and NP (830 mA h g−1) electrodes. After 100 cycles at a current density of 50 mA g−1, the reversible capacities of these cobalt oxide nanostructures still maintained at 500 (nanotubes), 480 (nanorods), and 450 mA h g−1 (NPs), respectively, corresponding to approximately 58.8%, 57.8%, and 55.2% of the initial capacity [377]. The high performance of the cobalt oxide nanotubes was resulted by the decreased particle size, the increased proportion of the surface atoms, and the hollow structures, making the structural more active for lithium-involved reaction and easier for lithium ion diffusion through 1D nanostructures. Lou et al. reported a self-supported topotactic transformation method for the synthesis of needle-like Co3O4 nanotubes by transforming from β-Co(OH)2 nanorods. As shown in Fig. 17b, it can be clearly observed that these Co3O4 cylindrical nanotubes were constructed from building-blocks of less than 100 nm. These nanotubes delivered an ultrahigh initial discharge capacity of more than 2000 mA h g−1 [379]. Similarly, through controlled thermal oxidative decomposition and re-crystallization of β-Co(OH)2 nanoneedles, mesoporous single-crystalline Co3O4 nanoneedles (Fig. 17c and d) have also been synthesized [380]. When used as an electrode material for LIBs, these nanoneedles exhibited a high reversible capacity of 1079 mA h g−1 after 50 cycles at a current density of 150 mA h g−1 with a voltage window between 3 V and 10 mV (Fig. 17e), indicating their promising application in LIBs
Fig. 18. Morphologies and electrochemical performances of 1D cobalt oxide nanowires/nanobelts. SEM image of (a) Co3O4 nanowires on Ti foil [392] (Copyright 2014, Elsevier), (b) Co3O4 nanowires on Ni foam [394] (Copyright 2016, Elsevier), (c) Co3O4 nanobelt array on Ti foil [395] (Copyright 2010, American Chemical Society). (d) Cycling performance of LIBs based on Co3O4 nanowires on Ti foil at a current density of 500 mA g−1 [392] (Copyright 2014, Elsevier), and (e) Co3O4 nanobelt array on Ti foil at different rates [395] (Copyright 2010, American Chemical Society). (f) SEM image of Ni-doped Co3O4 nanowires arrays for electrocatalysis [397] (Copyright 2015, Royal Society of Chemistry). 617
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
[380]. Apart from the application in the common LIBs, 1D Co3O4 nanorods were also used as non-precious electrocatalysts for ORR in AEMFCs [264] and MFCs [263]. As a typical type of 1D nanostructures, nanowires have been widely explored for electronic, optoelectronic, sensors, and electrochemical device applications [365,381–391]. Nanowires are primarily grown via vapor-liquid-solid (VLS) growth, vapor-solidsolid (VSS) growth, wet-chemical synthesis, electrospinning, and electrochemical anodization, etc. [354]. It is worth noting that Co3O4 NWAs or nanobelt arrays (NBAs), which can be grown directly on some conductive current collectors, such as Ti foil (Fig. 18a and c) and Ni foam (Fig. 18b), are good candidates as binder-free electrodes for energy conversion and storage devices with enhanced storage kinetics and structural stability [392–399]. In this structure, each wire or belt is electrically connected to the current collector, enabling effective ion/charge transfer for ion insertion/extraction. Furthermore, the good contact between the active materials and the current collectors is of significance for providing better accommodation of the large volume changes during charging/ discharging cycles. These advantages make these arrays attractive for improving the rate capability and the cycling stability with a reduced overvoltage. For example, grass-like Co3O4 NWAs on Ti foil presented a remarkable Li+ capacity of 662 mA h g−1 even at a high current density of 5 A g−1 and maintained a stable capacity of 1031 mA h g−1 at a rate of 500 mA g−1 after 100 cycles (Fig. 18d) [392], and Co3O4 NBAs on Ti foil were capable of retaining the Li+ capacity of 770 mA h g−1 at 177 mA g−1 over 25 cycles and 330 mA h g−1 at 3350 mA g−1 beyond 30 cycles (Fig. 18e) [395]. Furthermore, via direct doping of Ni into Co3O4 NWAs (Fig. 18f) with the purpose to increase the electroactive sites and to enhance the overall conductivity of electrode, efficient ion/electron transport and low charge transfer resistance for electrocatalytic reactions have been achieved [397]. In short, the biggest advantage for 1D cobalt oxide nanostructures, especially the arrays grown in-situ on current collectors providing good binding force between the active materials and the conductive substrates, can greatly enhance the cycling stability and prolong the catalyst life. The possible influence of the structures, such as the length and diameter, the aspect ratio, the density, and the morphology of the wires/fibers/rods/needles/belts on their electrochemical properties, unfortunately, currently absent systematic investigation. Moreover, the serious expansion of 1D cobalt oxides during repeated cycling processes can destroy the original ordered structures and lead to a higher irreversible capacity/capacitance and a weaker catalytic activity. 3.3. 2D cobalt oxide nanostructures 2D nanostructures, or called ultrathin nanosheets, whose thickness can go downwards only a single or a few atomic layers, have high surface-to-volume ratios as well as unusual physical, chemical, or electronic properties compared to their bulk counterparts [33,354,400–405]. For example, some typical characteristics for graphene includes tunable band gap, high flexibility, an ambipolar electric field effect along with ballistic conductivity of charge carriers, a quantum Hall effect at room temperature, etc. [406]. These advantages make graphene has been widely applied in various fields, such as materials, chemistry, physics, biology, and medicine [407,408]. Furthermore, the research on graphene has also initiated a boom in the development of graphene-like nanomaterials
Fig. 19. Morphologies and performances of 2D cobalt oxide nanosheets. (a, b) SEM images of self-stacked Co3O4 nanosheets, (c) cycling stability of the self-stacked Co3O4 nanosheets compared to commercial Co3O4 products at a current density of 178 mA g−1 [421] (Copyright 2011, Royal Society of Chemistry). (d, e) SEM images of (d) hexagonal Co(OH)2 precursors and (e) hexagonal Co3O4 nanosheets, (f) rate capabilities of hexagonal Co3O4 nanosheets at various rates [422] (Copyright 2016, Wiley-VCH). 618
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
[409–411]. To date, the family of 2D nanomaterials encompasses a broad selection of various compositions including almost all the elements in the periodic table [412,413]. Three common approaches are employed for the fabrication of 2D metal oxide nanosheets, namely, CVD growth, mechanical/liquid exfoliation of layered host materials, and wet-chemical self-assembly routes [354,414,415]. We recently have comprehensively reviewed the applications of 2D metal oxide nanostructures in next-generation rechargeable batteries and proposed the strategies for how to further improve the storage performance of typical 2D nanomaterials [33,67,405], thus will not repeat here. For Co3O4 nanosheets, wet-chemical synthesis strategies are the dominant fabrication approaches [416–420]. In this section, three typical Co3O4 nanomaterials with 2D forms, namely, plate-like nanosheets, hexagonal nanosheets, and ultrathin nanosheets, are presented in detail. Wang et al. reported the synthesis of self-stacked Co3O4 nanosheets separated by carbon layers by a solvothermal approach in the
Fig. 20. Synthesis, characterization, and LIB applications of ultrathin 2D cobalt oxide nanosheets. (a) Schematic illustration of wet-chemical molecular self-assembly of metal oxide nanosheets, and (b) the corresponding SEM image of self-assembled ultrathin Co3O4 nanosheets [428] (Copyright 2014, Nature Publishing Group). (c) TEM image, (d) HRTEM image, and (e) the corresponding fast Fourier transform (FFT) pattern of atomically thick Co3O4 nanosheets fabricated by topochemical transformation [429] (Copyright 2013, Royal Society of Chemistry). (f) Schematic illustration of working mechanisms using nanosheets and their bulk counterpart as anode materials for LIBs [429] (Copyright 2013, Royal Society of Chemistry). 619
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
presence of poly(vinylpyrrolidone) (PVP) and water followed by a calcination process [421]. As shown in Fig. 19a, the as-obtained microstacks were built from vertically stacked nanoplates approximately 4 μm in diameter and about 40 nm thickness (Fig. 19b). Interestingly, they also found both PVP and water could affect the final morphology of the products. Flower-like particles and hollow spheres were obtained with the existence of only water and only PVP, respectively. The self-stacked nanosheets were used as electrode material for LIBs, and the discharge capacity reached 1644 mA h g−1 in the first cycle and maintained 1070 mA h g−1 after 50 cycles at a current density of 178 mA g−1 (Fig. 19c). Actually, the excellent electrochemical performances were largely ascribed to the carbon layers between the Co3O4 nanoplates, which provide enhanced electrical conductivity and act as buffer spaces between the Co3O4 sheets to accommodate the volume changes during charging/discharging [421]. Besides the self-stacked nanosheets, isolated ultrathin hexagonal Co3O4 nanosheets have also been synthesized via hydrothermal methods [422–427]. Ultrathin Co3O4 hexagonal nanosheets with exposed {1 1 1} facets, which are the reactive high-energy facets, were prepared via thermal treatment of hexagonal Co(OH)2 sheet-like precursors (Fig. 19d) synthesized by a general polyethyleneimine (PEI)-mediated hydrothermal reaction [422]. As indicated in Fig. 19e, the typical Co3O4 hexagonal nanosheets were approximately 100 nm in side length and 15 nm in thickness. It has also been found that pores from several to tens of nanometers simultaneously formed inside the Co3O4 hexagonal nanosheets, aroused by the dehydration of polymer precursors and the crystalline phase transformation during heat treatment. When tested as anode materials for LIBs, this unique Co3O4 nanostructure presented a remarkable capacity of 1007 and 858 mA h g−1 at 100 and 500 mA g−1, respectively (Fig. 19f), together with an excellent capacity retention of 96–99% [422]. Recently, Co3O4 nanosheets with the thickness at an atomic level were developed for electrochemical applications. Sun et al. proposed a generalized approach for molecular self-assembly synthesis of ultrathin 2D transition metal oxide nanosheets (e.g. TiO2, ZnO, Co3O4, WO3, Fe3O4, MnO2, etc.) by rationally employing lamellar reverse micelles, as shown in Fig. 20a [428]. Compared with the exfoliation and CVD methods, this route is suitable for metal oxides with non-layered host materials, and uniform products can be easily obtained in large quantities. Moreover, this process does not involve high temperature and/or pressures, so the cost is relatively low. Taking Co3O4 as an example, the thickness of the as-obtained Co3O4 ultrathin nanosheets was about 1.6 nm, corresponding to two unit-cell layers, but the lateral size can be extended to 10 μm (Fig. 20b). These nanosheets possess high surface area and chemical reactivity, which make them appealing in environment protection and energy conversion and storage applications [410,428]. Freestanding atomically thick Co3O4 nanosheets with exposed {1 1 1} facets have also been synthesized by topochemical transformation from few-layered α-Co(OH)2 precursors [429]. The resulting nanosheets were ultrathin with a thickness of only 1.5 nm and lateral sizes of about 400 nm (Fig. 20c). The crystal structure of spinel Co3O4 and the preferential crystal orientation of {1 1 1} facets were confirmed by a high-resolution transmission electron microscope (HRTEM) (Fig. 20d) and the corresponding fast Fourier
Fig. 21. Morphologies and performances of 3D cobalt oxide micro-/nano-spheres. SEM images of Co3O4 in morphology of (a) hollow-sphere monolayer arrays [453] (Copyright 2011, Royal Society of Chemistry), (b) single-shelled hollow spheres [456] (Copyright 2010, Wiley-VCH), (c) double-shelled hollow spheres [456] (Copyright 2010, Wiley-VCH), and (d) mesoporous hollow spheres [452] (Copyright 2014, Nature Publishing Group). (e) Comparison of rate capabilities for Co3O4 mesoporous hollow spheres before (blue squares) and after (red circles) lithiation-induced reactivation (scale bar, 100 nm), and (f) the capacity and coulombic efficiency (CE) at a rate of 5.62 C for the following 500 cycles after 1,000 cycles [452] (Copyright 2014, Nature Publishing Group). 620
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
transform (FFT) patterns (Fig. 20e). Compared with the bulk Co3O4, these atomically thin nanosheets are superior for the lithiation/ delithiation behaviors for LIBs (Fig. 20f). The synchrotron radiation X-ray absorption near-edge spectroscopy (XANES) has also identified that the proportion of Co2+ atoms is higher and the charge redistributed on the {1 1 1}-oriented surfaces, which facilitate the storage of lithium ions [429]. Besides the isolated nanomaterials, some highly-oriented 2D arrays directly grown on the substrates have also been fabricated. For example, highly-ordered 2D CoO nanosheets arrays vertically aligned to porous Ni foam substrate have been grown via a galvanostatic electrodeposition technique [430]. This structure was directly used as the anode for LIBs and exhibited excellent cyclability (retained 1000 mA h g−1 after 100 cycles at 1 A g−1) and rate capability (520 mA h g−1 at 10 A g−1) [430]. 2D cobalt oxide nanostructures with confined thickness and high surface-to-volume ratios have shown great potential as electrode materials for ion storage devices. Nevertheless, in practical applications, these ultrathin nanosheets are easily restacked into thick closely packed structures, which severely hamper the insertion and diffusion of the ions and the infiltration of electrolytes into the inside of active materials, especially at high current densities/sweeping rates. Therefore, the direct use of 2D cobalt oxide nanostructures for electrochemical devices is yet full of challenges. Some possible solutions are still under exploration from the perspective of material design and electrode assembly. 3.4. 3D cobalt oxide nanostructures Micro/nanospheres, solid/porous polyhedrons, core-shell structures, mesoporous structures, hollow structures, hierarchical/ heterogeneous structures assembled from low-dimensional nanoarchitectures, etc., with an overall size over 100 nm, are considered as 3D nanostructures. Compared with the lower dimensional nanostructures, 3D nanomaterials are by far possessing more advantageous for their high surface area and hierarchically porous structure, which lead to increased numbers of exposed active sites, good contact with electrolytes, and enhanced ions or electron transport and diffusion in the practical electrochemical applications. As for these 3D nanostructures assembled from 0D, 1D and/or 2D nanostructures, they not only possess the intrinsic properties of those constitutional low-dimensional structures, but also provide many advanced functions with improved performance for a wide variety of applications [182,431–443]. For example, through self-induced assembly of 2D ultrathin nanosheets into 3D hierarchical architectures via a solution-based reaction or a thermal treatment step, porous structures can form in the materials, and the self-aggregation between the 2D nanosheets can be effectively alleviated by this way [444–450]. Hence, 3D nanomaterials are promising for plenty of applications, such as sensors, solar energy harvesting devices, molecular separation membranes, nanoreactors, energy conversion and storage devices, pharmaceutical delivery, and advanced optical devices [435,443,451]. Some representative 3D sphere-like Co3O4 nanomaterials with designed pores or channels inside the shells, such as mesoporous, hollow, double-shelled, and multi-shelled micro-/nano-spheres/quasi-spheres, have been widely studied for electrochemical devices [452–463]. As we can see from Fig. 21a–d, the diameters of these spherical structures were in the range from several hundred nanometers to a few micrometers. These unique 3D spherical structures together with nanopores and nanochannels result in remarkable ion storage performance, due to the improved surface-to-volume ratio, the reduced mass/charge transport lengths, and the low-energy transport paths. Moreover, the hollow structures and the pores as well as channels can provide enough free space to mitigate the unfavorable volume changes caused by the repeated ion intercalation and de-intercalation processes. In addition, the limited outer surface brought by the relative large overall size is unfavorable for the formation of a solid-electrolyte interphase (SEI) layer, which is the main reason to consume lithium ions by forming irreversible compounds [464]. It is for sure that there are some drawbacks for those 3D nanostructures. For an instance, one of the biggest drawbacks for 3D hollow nanomaterials is poor mechanical stability. Degradation occurs when 3D nanomaterials suffer high mechanical stress originating from repetitive volume expansion/contraction cycles. Sun et al. reported that the mechanical degradation of the metal oxide hollow spheres during cycling could be deliberately controlled to hierarchically mesoporous structure together with the formation of a stable SEI layer by a highrate lithiation-induced reactivation process [452]. It is obviously demonstrated from Fig. 21e that Co3O4 mesoporous hollow spheres showed a significantly improved rate capability after lithiation-induced reactivation and an excellent cycling stability for lithium storage at a higher rate of 5.62 C (5.0 A g−1) for the subsequent 500 cycles with a stable reversible capacity of 335 mA h g−1 (Fig. 21f), which is almost 300% higher than these electrodes without reactivation. Moreover, the Co3O4 hollow mesospheres reactivated with a 400-cycle mixed charging and discharging rate scheme (MC-rate: a discharge rate of 2.81 C and a charge rate of 1.69 C in each cycle) exhibited a remarkable cycling stability with no obvious capacity fading after 7000 cycles at a high rate of 5.62 C [452]. The second typical type of 3D Co3O4 structures for electrochemical devices is cube-like/octahedron-like/dodecahedron-like architectures [465–471]. For understanding the structure-property relationships of the 3D Co3O4 polyhedrons, the relationship between the particle size and the electrochemical performance has been investigated. Xu et al. synthesized nano-sized (nO-Co3O4, 387 nm) and micron-sized (mO-Co3O4, 6.65 μm) solid Co3O4 octahedra as electrode materials for LIBs [467]. Electrochemical results showed that the nO-Co3O4 delivered a high charge capacity up to 955.5 mA h g−1 over 200 cycles with no obvious capacity decay at a current density of 0.1 A g−1 (ca. 0.11 C), while the mO-Co3O4 only maintained 288.5 mA h g−1 at the same conditions. This noticeable difference is ascribed to the size effect on volume expansion/extraction. The mO-Co3O4 with large particle sizes suffers larger volume changes during repeated cycling, leading to serious pulverization of electrode materials and subsequent appearance of an electric isolation layer. Interestingly, the nO-Co3O4 maintains the initial octahedral shape after over 300 cycles. On the contrary, the mOCo3O4 broke into small particles with an average size of 1–2 μm and lost the octahedral shape after only 42 cycles [467]. Morphology is another crucial parameter affecting electrochemical properties. Chen et al. explored cobalt-based nanostructures, including cubes, discs, and flowers, for LIBs and concluded that cube-like structure gave the lowest capacity while nanoflowers 621
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
exhibited the best performance [468]. Similarly, Song et al. synthesized three types of Co3O4 hollow nanoarchitectures, including quadrate tubular nanoboxes (QTNBs), rectangular nanoboxes (RNBs), and cubic nanoboxes (CNBs), via controlled mild annealing of cobalt coordination polymer nano-solids [472]. Electrochemical measurements demonstrated that QTNBs exhibited outstanding properties, such as a superior capacity of 1200 mA h g−1 at 0.2 A g−1 and a remarkable retention of 625 mA h g−1 at a high current density of 10 A g−1. CNBs, however, delivered the worst performance, such as higher irreversible capacity (344 mA h g−1) and higher charge transfer resistance (Rct) (∼250 Ω) than that of QTNBs (222 mA h g−1, ∼130 Ω) and RNBs (252 mA h g−1, ∼230 Ω), respectively [472]. To improve the performance of Co3O4 cubes or boxes, heterostructure and hierarchical structure were introduced. Huang et al. produced regular Co3O4 microcubes with 2.37 μm in length on average, assembled by many irregular Co3O4 NPs (20–200 nm in diameter and 30–40 nm in thickness, Fig. 22a) [473]. Remarkably, these mono-dispersed micro-/nanostructured cubes displayed discharge capacities as high as 1298 mA h g−1 at 0.1 C and 1041 mA h g−1 at 1 C, respectively [473]. Afterwards, hierarchically porous Co3O4 nanoboxes with well-defined interior voids and functional shells (Fig. 22b) were synthesized via an ion
Fig. 22. Morphologies and performances of 3D cobalt oxide polyhedrons. SEM images of Co3O4 in morphology of (a) mesoporous cubes [473] (Copyright 2014, American Chemical Society), (b) hierarchically porous boxes [474] (Copyright 2016, Royal Society of Chemistry), (c) hollow octahedra [475] (Copyright 2009, American Chemical Society), (d) solid cubes [476], (e) solid octahedra [476], (f) truncated octahedra [476] (Copyright 2012, Wiley-VCH). (g) Discharging and charging curves of Co3O4 octahedra at a current density of 1000 mA g−1, and (h) comparison of cycling stability of Co3O4 cubes, octahedra and truncated octahedra [476] (Copyright 2012, Wiley-VCH). 622
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
exchange reaction using Prussian blue analogue as precursor at a low temperature (60 °C) [474]. An extremely large surface area of about 272.5 m2 g−1 was achieved and it provided more active sites to promote the ORR and OER as a Li-O2 battery cathode [474]. Apart from cubes, single-crystalline Co3O4 octahedral hollow cages (Fig. 22c) with tunable surface aperture were also synthesized through a carbon-assisted carbothermal approach and applied for Li+ storage [475]. The hollow quasi-octahedra presented the highest initial discharge capacity (1470.2 mA g h−1) and the best cycling performance (maintained 670 mA g h−1 after 50 cycles at 178 mA g−1) among the fabricated different morphologies of quasi-octahedra, hollow octahedral, and hollow spheres [475]. In addition to the effect of particle size and morphology, it should be noted that exposed crystal facets also have a significant influence on the electrochemical performance. Xiao et al. synthesized three well-defined Co3O4 morphologies with different exposed crystal planes via a simple control of the ratio of Co(NO3)2 6H2O and NaOH [476]. Co3O4 cubes (Fig. 22d) with a particle size of about 500 nm and six exposed {0 0 1} planes, Co3O4 octahedra (Fig. 22e) with a size of 600–700 nm and eight exposed {1 1 1} planes, and Co3O4 truncated octahedral (Fig. 22f) with a size of around 600 nm and eight exposed {1 1 1} and six {0 0 1} planes. Based on those samples, the effect of exposed crystal planes on the electrochemical properties such as lithium ion storage has been systematically investigated. It has confirmed that the performances of these three types of Co3O4 nanostructures ranked as “octahedra > truncated octahedra > cubes” and the Co3O4 octahedra with exposed {1 1 1} planes delivered both the highest charge/discharge specific capacity and the best rate capability (Fig. 22g and h). In fact, the electrochemical performances of the cobalt oxides with different exposed facets depend largely on the redox reaction of Com+/Co0. Surface atomic configurations of Co3O4 unit cell (Fig. 23a) shows that the {1 1 1} plane have more Co2+ (3.75 Co2+) than the {0 0 1} plane (2 Co2+), leading to a faster reaction rate [476]. Subsequently, they found that the Co3O4 octahedra with exposed {1 1 1} planes showed superior electrocatalytic performance for Li-O2 battery than the Co3O4 cubes with exposed {0 0 1} facets (Fig. 23b and c). Based on theoretical calculations and experimental results, the facet-dependent electrocatalytic mechanism of the Co3O4 nano-polyhedrons with different crystal planes have been proposed, as shown in Fig. 23d and e. Compared to the Co3O4 with exposed {0 0 1} planes, Co3O4 with {1 1 1} surfaces has more active Co2+ sites exposed on the surface, which have a stronger absorption ability for Li2O2 in ORR and a lower activation barrier for O2 desorption in OER [477]. A similar study on the facet-dependent activity and stability of Co3O4 nanostructures towards OER has also been reported thereafter [478]. 3D hierarchical architectures, constructed by 1D, 2D, and 3D building blocks, are another important class of nanomaterials with exceptional physical and chemical properties for energy conversion and storage devices [443,479–484]. Generally, wet-chemical synthesis is the most common strategy for the preparation of these types of 3D nanostructures, including one-step self-assembly and two-step self-assembly approaches [443,485,486]. In particular, the versatile two-step synthetic route allows an effective combination with a variety of other techniques (e.g. lithography, hydrolysis, carbonization, etc.) to produce multifunctional nanostructures [487]. In 2012, Wang et al. reported the growth of symmetric Co3O4 hexapods assembled from numerous porous nanorods on copper foil by a facile hydrothermal method followed by a thermal annealing step in nitrogen environment (Fig. 24a) [488]. In this synthesis, NH4F plays a key role for the formation of nanorod-assembled Co3O4 hexapods. Without the addition of NH4F, only bunch-like products were obtained. The specific surface area of Co3O4 hexapods was 134.84 m2 g−1. Most importantly, this nanostructure grown directly on a conductive copper substrate could be used as binder-free anodes for LIBs, which exhibited a high reversible capacity of 800 mA h g−1 and 440 mA h g−1 at current densities of 100 and 500 mA g−1, respectively, after 40 cycles. Moreover, the lithium storage performance has been further enhanced via carbon coating and reached an improved reversible capacity of 1001 mA h g−1 after 40 cycles [488]. Flower-like Co3O4 nanomaterials have also been fabricated through a solvothermal reaction together with a subsequent postcalcination process [489,490]. As indicated in Fig. 24b, Co3O4 nanoflowers grown on Ni foam were consisted of many 2D nanosheets with jagged edges. These nanosheets, with a thickness of between 60 and 110 nm, possessed smooth surface [489]. Afterwards, 3D
Fig. 23. Effect of exposed crystal planes of cobalt oxide structures on electrocatalytic performances. (a) schematic illustration of atomic configurations for Co3O4 structures with different exposed planes [477], (b, c) cycling stability of the Li-O2 batteries based on the cathode catalysts of Co3O4 (b) cubes and (c) octahedrons at a current density of 100 mA h·g−1 with a limited capacity of 1000 mA h·g−1 [477], and (d, e) schematic illustration of facet-dependent electrocatalytic mechanisms of ORR and OER for Co3O4 (d) cubes and (e) octahedrons [477] (Copyright 2015, American Chemical Society). 623
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
Fig. 24. Morphologies and performances of 3D dendritic cobalt oxide microspheres. SEM images of Co3O4 in morphology of (a) hexapods [488] (Copyright 2012, Royal Society of Chemistry), (b) nanoflowers [489] (Copyright 2011, Elsevier), (c) red-headed calliandra-like nanoflowers [490] (Copyright 2015, American Chemical Society), (d) 3D pompon-like porous spheres [492]. (e) Comparison on rate capability of 3D pompon-like Co3O4 spheres, Co3O4 nanowires (NWs), and Co3O4 nanoparticles (NPs) at various current densities [492], (f) schematic illustration of possible Li+ transfer pathways for the pompon-like Co3O4 anode [492] (Copyright 2014, Royal Society of Chemistry).
red-headed calliandra-like (Fig. 24c) and dandelion-flower-like hierarchical mesoporous Co3O4 were prepared by calcining the solvothermal synthesized cobalt carbonate hydroxide precursors at 400 °C in air, in which the morphology of the precursors was control by adjusting the proportion of solvent, the amount of urea, and the reaction time [490]. When tested as anode materials for LIBs, the dandelion-flower-like cobalt oxide showed a high initial specific capacity of 1298 mA h g−1 and a stable reversible capacity of 1204 mA h g−1 after 20 cycles, while those of the red-headed calliandra-like Co3O4 exhibited capacities of 1340 and 833 mA h g−1, respectively, at a current density of 100 mA g−1 with a potential window ranging from 0.01 to 3.0 V [490]. Other hierarchical structures, such as star-like, pompon-like, shale-like, and conch-like structures, have been reported for energy storage applications [491–494]. In terms of ion storage, these hierarchical structures have advantages for achieving high-efficiency ion transport and accommodating volume strain during intercalation and de-intercalation processes. For example, the star-like Co3O4 hierarchical structure, synthesized through thermal decomposition of self-assembled Co(OH)F precursors., delivered an initial specific charge capacity of 1036 mA h g−1 at a current density of 50 mA g−1 as anode materials for LIBs [491]; the pompon-like Co3O4 hierarchical porous spheres, as shown in Fig. 24d, synthesized via a combination of hydrothermal method and calcination treatment, presented considerable capacities of about 730 and 650 mA h g−1 at high current densities of 400 and 1200 mA g−1, respectively (Fig. 24e) [492]. This superior electrochemical performance is believed to be resulted from the unique hierarchical structures, the short ion transport lengths, and the multiple transport directions, as illustrated in Fig. 24f. Besides, Co3O4 hexagonal plates assembled by Co3O4 nanocubes [495] and Co3O4 thin films formed from Co3O4 nanoflakes [496] were demonstrated with enhanced electrochemical properties. Recently, a general method for producing 3D mesoporous Co3O4 networks from ultrathin-nanosheet as anode materials for LIBs has been proposed by Zhu et al. [497]. During the fabrication, 3D nitrogen-doped carbon networks (N-CN) (Fig. 25a) were first synthesized via a salt template-assisted method. Cobalt ions were then adsorbed onto the N-CN via a dipping process in cobalt nitrate solution. After the removal of carbon template by heating at 500 °C in air, 3D mesoporous Co3O4 networks (3D-MN Co3O4) were finally obtained, as shown in Fig. 25b and c. The obtained interconnected 3D-MN Co3O4 structures exhibited a large surface area as high as 106.6 m2 g−1 and a pore size distribution ranging from 2 to 10 nm. The presence of mesoporous architecture, which provides increased charge transport pathways (Fig. 25e), demonstrated a superior performance as an anode material for LIB with a high capacity of 1033 mA h g−1 at the rate of 0.1 A g−1 and long-life stability (700 cycles at 5 A g−1) in the potential range from 5 mV to 3 V (Fig. 25d). Moreover, this strategy has been shown to be effective for the synthesis of other 3D mesoporous TMOs, such as Fe2O3, ZnO, Mn3O4, NiCo2O4, and CoFe2O4 [497]. To sum up, 3D cobalt oxides have much complex morphologies and porous structures compared to other types of structures. Generally speaking, the 3D cobalt oxide nanostructures with the existence of porous architectures, such as those with hollow structures and hierarchical structures, demonstrate better electrochemical performances than those solid particles. More attractively, 624
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
Fig. 25. Synthesis, characterization, and application of 3D cobalt oxide networks (3D-MN Co3O4) [497]. (a) Schematic illustration of a general synthesis route for 3D cobalt oxide network, (b) the corresponding SEM image, (c) TEM image, and (d) cycling stability of Li+ storage at a current density of 5 A g−1 [497], (e) schematic illustration of electron/ion transport pathways of 3D mesoporous Co3O4 networks (Copyright 2017, WileyVCH).
the 3D nanostructures assembled from low dimensional constituents are promising for practical applications, because they not only can take the full advantage of the merits of low dimensional structures, but also possess extra benefits from their 3D configurations, such as preventing the aggregation of the low-dimensional structures, improving the contact area with electrolytes, and accommodating the volume expansion/contraction during cycles. It is worth noting that the adhesion force between these 3D nanostructures and current collectors during the preparation of electrodes for batteries is always weak, which result in the delamination of active materials and rapid fading of the electrochemical performance during cycling. Therefore, further optimization on the electrode preparation of these 3D nanostructures is needed. 4. Fabrication and electrochemical performance of cobalt oxide-based nanocomposites To conquer the intrinsic issues of cobalt oxides, such as the low electric conductivity and the significant volume variation during charging/discharging, the construction of cobalt oxide-based nanocomposites is one of the most feasible strategies, owing to its appealing advantages, such as simple for fabrication, effective for enhancing the performance, and generalized for most other metal oxides. Via this method, a wide range of cobalt oxide nanocomposites has been produced with the configuration of 0D-1D, 0D-2D, 1D2D, and 2D-2D, etc., by using the techniques of anchoring, depositing, encapsulation, and coating, etc. In this section, we aim to summarize the recent achievements in the rational design of cobalt oxide-based nanocomposites for electrochemical energy applications. 4.1. Cobalt oxide/carbon nanocomposites Hybridizing carbon with cobalt oxide nanomaterials is a simple and effective way to overcome the intrinsic low conductivity of 625
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
cobalt oxides to achieve rapid ion insertion/disinsertion and promoted catalytic activity [498–511]. Two-step strategy is often employed for the fabrication of metal oxide nanostructures with complex configurations [354,432–434,484,485,487]. This strategy has also been used for the fabrication of cobalt oxide-based nanocomposites, in which a phase transformation process from the assynthesized cobalt-containing solid precursors (e.g. cobalt hydroxide, cobalt nitrate hydroxide, etc.) to the final pure cobalt oxides is usually performed by thermal treatment in air or in solvents at elevated temperatures [512]. In this section, we classified the cobalt oxides/carbon composites into six typical types based on the dimensionality of cobalt oxide nanostructures, namely 0D cobalt oxide/carbon hybrids, 1D cobalt oxide/carbon hybrids, 2D cobalt oxide/carbon hybrids, 3D cobalt oxide/carbon hybrids, cobalt oxide/heteroatom-doped carbon hybrids, and cobalt oxide/carbon/others hybrids. It should be noted what the carbon refers here is the amorphous carbon coatings or nanocarbon in the forms of fiber-like, sphere-like, or other morphologies, but except for CNTs and graphene, which will be separately summarized in the following sections. 4.1.1. 0D cobalt oxide/carbon hybrids The combination of 0D cobalt oxide NPs with carbon can be mainly classified into two main strategies, namely, carbon surface supported 0D cobalt oxides, and carbon encapsulated 0D cobalt oxides. For the carbon surface supported 0D cobalt oxides, the cobalt oxide NPs can be dispersed, anchored, or deposited on the carbon surfaces. This hybrid is easily to be fabricated by some simple techniques, such as direct mixing of cobalt oxide NPs and carbon in a solution. A uniform dispersion of NPs on the carbon surface with a strong binding force, however, is still a big challenge. In the case of carbon encapsulated 0D cobalt oxides, the cobalt oxide NPs are encapsulated or surrounded by carbon layers, or the NPs are embedded into a carbon matrix. There are some advantages in the carbon encapsulated 0D cobalt oxide hybrids. First, the carbon layers or matrix can help avoid the aggregation of cobalt oxide NPs. Second, the carbon coating on the surfaces of NPs can act as excellent conductive networks for the overall nanocomposites. Moreover, the carbon layers or matrix can contribute to ease the volume expansion of 0D cobalt oxide NPs during the repeated ion insertion/ extraction cycles. For the carbon supported 0D cobalt oxide, the most common configuration is the 0D NPs anchoring on carbon surfaces. Depends on the adsorption of 0D NPs on the outer surface or the inner surface of the carbon supporter, some different morphologies of the hybrids have been reported [513–519]. For example, as shown in Fig. 26a, cobalt oxide NPs (20–40 nm in size) were anchored on the pore walls of amorphous carbon [514]; while as shown in Fig. 26b, cobalt oxide NPs (< 10 nm in size) were adsorbed on the outer surface of carbon spheres (∼100–300 nm in size) [515]. It is clearly observed that the electrochemical performances of the cobalt oxide/carbon hybrids are superior to these of the individual cobalt oxides and carbon when these were used as electrode materials for LIBs [514] or cathode catalyst for Li-O2 batteries (Fig. 26c) [515]. Hao et al. reported a facile hydrothermal and sol-gel polymerization route for the large-scale fabrication of interconnected hybrids, where well-defined Co3O4 NPs (∼5 nm) were anchored on
Fig. 26. Morphologies and performances of 0D cobalt oxide NPs absorbed on carbon matrix hybrid materials. (a) TEM image of the hybrid material of 0D Co3O4 NPs anchored on the inner surface of porous carbon (Co3O4/PC) [514] (Copyright 2015, Elsevier). (b) The hybrid material of 0D Co3O4 adsorbed on the outer surfaces of carbon spheres [515], (c) cycling performance of the Li-O2 batteries with different air electrodes under a limited capacity of 1000 mA h gelectrode−1 and a stable charge voltage of 4.35 V [515] (Copyright 2013, Elsevier). (d) Cycling stability of individual carbon aerogel (CA), Co3O4, and Co3O4/CA hybrids at a current density of 50 mA g−1 [520], (e) schematic illustration of Li insertion/extraction process of the hybrid material of 0D Co3O4 NPs interconnected with CA hybrid nanospheres [520] (Copyright 2013, American Chemical Society). 626
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
the surfaces of carbon aerogel (CA) nanospheres to form interconnected networks [520]. Such an interconnected Co3O4/CA hybrid materials presented some unbeatable merits, such as large surface area, good conductivity, fast diffusion channels contributed by their mesoporous structure, mechanical flexibility, chemical stability, and short transport length of Li+ within CA matrix. The incorporation of Co3O4 NPs into the interconnected CA effectively reduced the irreversible reaction between Li and carbon matrix, and thus greatly increased the reversible specific capacity and the initial Coulombic efficiency (CE). The Co3O4/CA hybrid material with 25 wt% Co3O4 loading amount displayed an excellent storage performance, which a CE up to 99.5% was retained, and a high discharge capacity of 779 mA h g−1 after 50 cycles, corresponding to 10.1 and 1.6 times larger than that of Co3O4 (73 mA h g−1) and CA (478 mA h g−1), respectively, was observed (Fig. 26d) [520]. As illustrated in Fig. 26e, the improved Li+ storage performance of the mesoporous Co3O4/CA hierarchical hybrids are mainly due to an intimate contact and a synergistic effect between the Co3O4 NPs and CA matrices. Wang et al. reported the fabrication of hybrid architectures through anchoring of 0D Co3O4 NPs around 5 nm in diameter on 3D arrays of carbon nanosheets (Co3O4/CNS) [521]. The combination of the nanoscale 0D cobalt oxide NPs and the 3D interconnected sheet-like carbon provided some advantages for improving the electrochemical properties, such as the high surface area and ample porosity of conductive carbon nanosheets ensuring fast electrolyte transport. When used as anode materials for LIBs, high lithium storage over 1200 mA h g−1 was recorded, together with outstanding rate capability (e.g. ∼390 mA h g−1 at 10 A g−1). Furthermore, the Co3O4/CNS hybrid material displayed a reversible capacity of approximately 970 mA h g−1 at current density of 1 A g−1 after 500 cycles. The greatly enhanced capacity and initial CE were attributed to the synergistic effect between the nanosized cobalt oxide and the sheet-like interconnected carbon nanosheets of the hybrid material [521]. Cobalt oxide NPs with a size ranging 10–50 nm derived from cobalt-containing zeolite imidazolate frameworks (ZIF) have also deposited on a electrospinning fiber-like carbon matrix (Co3O4/CNF) for non-aqueous Li-air batteries [522]. As a self-standing, binder-free electrode, the Co3O4/CNF composite exhibited a high initial discharge capacity over 760 mA h g−1, improved cycling properties, and a lower charge overpotential at different current densities [522]. These were mainly attributed to enhanced catalytic activity of Co3O4 NPs and stable contact between the homogeneously distributed Co3O4 NPs and the carbon nanofibers. Encapsulation of cobalt oxide NPs into carbon matrix is often achieved via confined growth of NPs inside carbon tubes [523], spheres [524], and fibers [525–527]. Leng et al. reported a type of porous graphic carbon (PGC) encapsulated carbon-coated Co3O4 core-shell NPs (Co3O4@C@PGC) hybrid material. In the synthesis, Co3O4@C core-shell nanostructures, in which Co3O4 NPs were homogeneously encapsulated by a thin carbon shell, were further homogeneously embedded into porous graphitic carbon nanosheets to form the final Co3O4@C@PGC hybrid material [528]. As illustrated in Fig. 27a, when the hybrid is used as electrode for LIBs, the unique structure is helpful to reduce the pulverization and aggregation of cobalt oxide NPs. Furthermore, owing to the presence of abundant micropores and mesopores in PGC nanosheets, the stress originated from the volume changes of cobalt oxides it during the insertion and extraction process of Li+ ions can be easily accommodated. As a result, the resulting Co3O4@C@PGC hybrid electrodes exhibited a much larger reversible capacity at various rates compared with other Co3O4-based electrodes (Fig. 27b) [528]. Song et al. synthesized porous carbon/Co3O4 composites by heat treatment of metal-organic frameworks (MOFs) as cathode materials in rechargeable Li-O2 batteries [529]. During the thermal decomposition of cobalt-containing MOFs (Co-MOFs), an interconnected porous structure with uniform distributed Co3O4 NPs in the carbonaceous matrix was obtained (Fig. 27c). The as-synthesized porous carbon/ Co3O4 hybrids exhibited superior electrochemical performance, and a high initial capacity of around 9850 mA h g−1 at a current density of 100 mA g−1 was achieved (Fig. 27d) [529].
Fig. 27. Hybrid materials of 0D cobalt oxide NPs encapsulated into a carbon matrix. (a) Schematic illustration of the Li+ extraction/insertion behaviors of 0D Co3O4@C core-shell NPs encapsulated into porous graphitic carbon nanosheets (Co3O4@C@PGC), (b) comparison of specific capacity for different Co3O4-based electrodes at various rates [528] (Copyright 2015, Nature Publishing Group). (c) Schematic illustration of the fabrication and (d) initial discharge/charge profile of MOF-derived carbon/Co3O4 composites [529] (Copyright 2017, Elsevier). 627
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
Recently, biomass-derived carbon materials have attracted an increasing attention for their low cost and abundance, which have also been used as the conductive matrix to hybridize with cobalt oxide NPs for renewable energy applications. Li et al. prepared a Co3O4/carbon fiber hybrid structure (Co3O4-CF) by pyrolysis of wet-spun cobalt alginate fibers, in which Co3O4 NPs (∼30–50 nm) were embedded in a micro-sized carbonaceous fibrous matrix (∼5–8 μm) [530]. In the synthesis, cobalt alginate fibers (Co-AF) were first obtained by a metal ion exchange of calcium alginate fibers (Ca-AF), and then the Co-AF was carbonized into final Co3O4-CF. When used for LIBs, the Co3O4-CF hybrid exhibited a reversible capacity up to 780 mA h g−1 at the current density of 89 mA g−1 after 100 cycles [530]. Similarly, they also prepared nano/micro hierarchical Co3O4/cellulose fibers composites (Co3O4/CF), where Co3O4 NPs (∼50–100 nm) were adsorbed on the surfaces of carbonized cellulose microfibers (CFs). This Co3O4/CF hybrid structure showed a reversible capacity of 730 mA h g−1 under the same test conditions with the Co3O4-CF [531]. The excellent electrochemical performances of these biomass derived Co3O4/C hybrid nanocomposites were mainly attributed to the confinement of the biomass substrates to retard the growth of Co3O4 NPs, and the bi-functionality of the carbon fiber matrix, which served simultaneously as scaffold and conducting network and helped improve the overall conductivity and release the mechanical strain induced during the long-time charging/discharging processes. Furthermore, the porous structure of the hybrid materials also greatly enhances the contact between electrolytes and active materials and the effective diffusion of electrolyte through it. A peapod-like cobalt oxide/carbon nanostructure, where the cobalt oxide NPs encapsulated uniformly in a tube-like carbon matrix, as shown in Fig. 28a, were synthesized by Wang et al. and used as electrode materials for batteries [532]. In the synthesis of the peapod-like Co3O4@carbon hierarchical nanostructure, cobalt carbonate hydroxide (Co(CO3)0.5(OH) 0.11H2O) nanobelts were used as templates to be coated with polymeric layers by a hydrothermal reaction, and then were calcined in Ar atmosphere at 700 °C for 200 min, followed with annealing in air at 250 °C for 200 min to obtain the final product. Scanning transmission electron microscopy (STEM) image (Fig. 28a) indicated that Co3O4 NPs 20–30 nm in diameter were encapsulated in tubular carbon fibers with regular intervals. When tested as anode materials for LIBs, the peapod-like nanostructure demonstrated a high specific capacity of about 1000 mA h g−1 at the rate of 100 mA g−1 and wonderful cycling stability with 80% retention when cycled back from very high rate of 1 A g−1 [532]. Later, Gu et al. reported mesoporous peapod-like Co3O4@carbon nanotube arrays (Co3O4@CNT) through a controllable nano-casting process by using an ordered mesoporous carbon (CMK-5) as substrates [533]. SEM (Fig. 28b) and TEM (Fig. 28c) images revealed that the product had a rod-like morphology with a particle size around 0.8–1.0 μm and no cobalt oxide NPs were found on the external surface of the carbon substrate. To analyze the distribution of NPs in the carbon matrix, high-resolution SEM (HR-SEM) and bright-field scanning transmission electron microscope (BF-STEM) images analysis confirmed that Co3O4 NPs were uniformly embedded into the tubular mesopores of the carbon matrix. Furthermore, it was clearly demonstrated in the TEM image that Co3O4 NPs were highly dispersed and homogeneously confined in the mesoporous channels with uniform sizes of 4–5 nm and distances of approximately 3 nm. They also found that the wall thickness of the tube-like carbon in the products could be easily tuned by adjusting the initial loading precursor amount in the SBA-15 template. By varying the thickness of the carbon layer, the sizes of Co3O4 NPs could also be controlled in the range of 3–7 nm and their loading amount could be varied from 45 to 70 wt%. This
Fig. 28. Morphologies and performances of peapod-like cobalt oxide/carbon hybrid nanostructures. (a) Scanning transmission electron microscopy (STEM) image of peapod-like Co3O4@carbon composite [532] (Copyright 2010, American Chemical Society). (b) SEM and (c) TEM images of mesoporous peapod-like Co3O4@carbon nanotube arrays (Co3O4@CNT) [533], (d) Discharge/charge curves, (e) cycling stability at 0.1 A g−1, and (f) rate capability of the peapod-like Co3O4@carbon nanotube arrays (inset in d: schematic illustration of the peapod-like Co3O4@carbon nanotube arrays) [533] (Copyright 2015, Wiley-VCH). 628
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
unique mesoporous structure facilitates the insertion/extraction behaviors of lithium ion for LIBs. The Co3O4@CNT mesoporous peapod-like structure exhibited a high specific capacity of 780 mA h g−1 (Fig. 28d) or volumetric capacity of about 1370 mA h cm−3 at a current density of 100 mA g−1, an outstanding cycling performance (700 mA h g−1 over 60 cycles at 0.1 A g−1) (Fig. 28e), and an excellent rate capacity (453 and 408 mA h g−1 at 1.0 and 5.0 A g−1, respectively) (Fig. 28f) [533]. According to the above selected examples, we can know that 0D cobalt oxides dispersed onto or encapsulated into a carbon matrix can effectively enhance the overall conductivity and thus improve the electrochemical properties. However, there are several major shortcomings of the physical dispersion of cobalt oxide NPs on carbon surfaces, including weak binding force between cobalt oxide NPs and carbon matrix, and poor buffer capacity accommodating volume changes. The encapsulation of these NPs into carbon can greatly suppress the serious aggregation of cobalt oxide NPs into bulks, but the slight aggregation inside the carbon may still exist. Meanwhile, confining cobalt oxide NPs within carbon matrix results in indirect contact between active materials and electrolytes because the electrolyte ions need to penetrate through the outer carbon layers. Hence, the design of cobalt oxide NPs uniformly encapsulated within thin carbon layers to form peapod-like analogous structures is a potential solution for improving the electrochemical performance of 0D cobalt oxide NPs, however, more further work on enhancing the diffusion of electrolyte through the carbon matrix, by decreasing the thickness of carbon layers, for example, is needed. 4.1.2. 1D cobalt oxide/carbon hybrids Currently, there are two main strategies used for the synthesis of 1D cobalt oxide/carbon hybrids. In the first strategy, 1D cobalt oxide nanostructures are firstly prepared and a following carbon coating is achieved via physical sputtering or chemical deposition. Via this method, a carbon layer can be coated onto the cobalt oxide nanowires/fibers to produce core-shell structures. The outer carbon layer in the structure is an effective way to overcome the low conductivity, weak reaction kinetics, and obvious volume expansion of cobalt oxides during electrochemical applications. It has reported that this carbon coated 1D cobalt oxide hybrids demonstrated an excellent capacity of more than 1000 mA h g−1 for LIBs and a remarkable cycling stability up to 8000 cycles for supercapacitors [534,535]. However, the homogeneous coating and the accurate control on the thickness of carbon layer on 1D cobalt oxide nanostructures remain challenging. The second strategy on preparing 1D carbon@cobalt oxide composites is the template-assisted route, in which cobalt precursors/cobalt oxides are coated on fiber-like carbonaceous template and followed with a thermal treatment process. Using this strategy, porous tubular structures can be easily obtained. The presence of cavities and porous structures within 1D nanostructures offer more available space to accommodate the mechanical strain aroused by ion insertion/ deinsertion, and is helpful to shorten ion diffusion distance, promote electrochemical reaction kinetics, and improve electrochemical properties. Co3O4-C core-shell NWAs were fabricated by combining hydrothermal synthesis and direct current magnetron sputtering by Chen et al. [535]. As shown in Fig. 29a and b, compared with Co3O4 NWAs, the surface of the carbon-coated nanowires became a little coarser. TEM analysis revealed that the core nanowires had many nanopores with diameters in the range from 2 to 4 nm, and the amorphous carbon layer was homogeneously coated on the porous nanowire surface, with a thickness of 18 nm (Fig. 29c). This array delivered an initial discharge capacity as high as 1330.8 mA h g−1 at 0.5 C and maintained a high reversible capacity of
Fig. 29. Morphologies and performances of cobalt oxide NWAs and cobalt oxide/carbon core-shell NWAs. SEM images of (a) Co3O4 NWAs and (b) Co3O4-C core-shell NWAs [535], (c) TEM image of a Co3O4-C core-shell NWA [535] (Copyright 2012, Royal Society of Chemistry). (d) Cycling stability at 200 mA g−1 [537], and (e) rate capability of carbon-doped Co3O4 hollow nanofibers (C-doped Co3O4 HNFs) and undoped Co3O4 HNFs [537] (Copyright 2016, Wiley-VCH). 629
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
989.0 mA h g−1 over 50 cycles, much higher than the unmodified Co3O4 nanowire array (490.5 mA h g−1) [535]. Later, similar structures have also been reported for asymmetric supercapacitors (ASCs) with cobalt oxide nanowires as positive electrode and activated carbon (AC) as negative electrode [534]. These arrays exhibited a high specific capacity of 116 mA h g−1 at 4 A g−1 and a remarkable cycle life with around 92% retention after 8000 cycles [534]. Tan et al. [536] and Yan et al. [537] fabricated carbon-doped fiber-like Co3O4 nanomaterials by using human hair and poly (acrylonitrile) nanofibers (PAN NFs) as templates, respectively [536,537]. In the work of Tian et al., by using human hairs as templates, 1D hierarchical porous Co3O4 nanofibers with {2 2 0} exposed facets, 20–30 nm in width and 3–5 μm in length on carbon matrix (H2@Co3O4), were prepared by solvothermal synthesis and followed calcination. In this structure, well crystallized Co3O4 particles 8–12 nm in diameter, were closely aggregated together inside the nanofibers [536]. Electrochemical tests of H2@Co3O4 for LIBs displayed a high initial discharge capacity of 1368 mA h g−1 and maintained 916 mA h g−1 after 100 cycles at the current density of 0.1 A g−1. Also, 659 and 573 mA h g−1 could be still achieved at high current densities of 1 and 2 A g−1, respectively [536]. In the study of Yan et al., carbon-doped Co3O4 hollow nanofibers (C-doped Co3O4 HNFs) were synthesized by combining electrospinning technique and hydrothermal method, in which PAN NFs acted as both templates and carbon source [537]. This unique configuration possessed many outstanding advantages, such as more Co2+ ions for faster Co2+/Co0 redox reaction, high surface-to-volume ratio, and a large number of oxygen vacancies, plenty of free space for accommodating volumetric expansion, and short diffusion pathway for electron or Li+ ion, which resulted in a high reversible capacity of 1121 mA h g−1 over 100 cycles at the current density of 200 mA g−1, much higher than that of the undoped Co3O4 HNFs (663 mA h g−1) (Fig. 29d). Moreover, reversible specific capacities ranged from 1173 to 607 mA h g−1 have been achieved, which could be recovered to around 1280 mA g h−1 with the current density increased from 50 to 3000 mA g−1 for every 10 successive cycles and subsequently back to 50 mA g−1 for 20 cycles (Fig. 29e) [537]. Currently, there are two main strategies used for the synthesis of 1D cobalt oxide/carbon hybrids. In the first strategy, 1D cobalt oxide nanostructures are firstly prepared and the following carbon coating is achieved via physical sputtering or chemical deposition. Via this method, a carbon layer can be coated onto the cobalt oxide nanowires/fibers to produce core-shell structures. The outer carbon layer in the structure is an effective way to overcome the low conductivity, weak reaction kinetics, and obvious volume expansion of cobalt oxides during electrochemical applications. It has reported that this carbon coated 1D cobalt oxide hybrids demonstrated an excellent capacity of more than 1000 mA h g−1 for LIBs and a remarkable cycling stability up to 8000 cycles for supercapacitors. However, the homogeneous coating and the accurate control on the thickness of carbon layer on 1D cobalt oxide nanostructures remain challenging. The second strategy on preparing 1D carbon@cobalt oxide composites is the template-assisted route, in which cobalt precursors/cobalt oxides are coated on fiber-like carbonaceous template and followed with a thermal treatment process. Using this strategy, porous tubular structures can be easily obtained. The presence of cavities and porous structures
Fig. 30. Morphology and performances of 2D cobalt oxide/carbon hybrids. (a–c) FESEM images of Co3O4/C nanoplates [538] (Copyright 2013, Royal Society of Chemistry). (d) SEM images of carbon-coated Co3O4 nanosheets on carbon paper (CP) (Co3O4@C/CP) [250], (e) polarization curves in 0.5 M H2SO4 electrolytes with a scan rate of 5 mV s−1 [250], (f) galvanostatic tests for OER at a current density of 100 mA cm−2 with Co3O4 and RuO2-based catalysts [250] (Copyright 2016, Elsevier). 630
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
within 1D nanostructures offer more available space to accommodate the mechanical strain aroused by ion insertion/deinsertion, and are helpful to shorten ion diffusion distance, promote electrochemical reaction kinetics, and improve electrochemical properties. 4.1.3. 2D cobalt oxide/carbon hybrids Coating carbon on 2D cobalt oxide nanosheets has attracted much attention for electrochemical devices for the purpose of taking the full advantages of the merits and overcoming the shortcomings of the 2D sheet-like nanostructures. The presence of carbon coating not only strengthens the structure stability, but also prevents the nanosheets from agglomeration during long-time catalytic cycles or repeated charging/discharging processes. Most importantly, the wider interspaces between the carbon coated nanosheets allow good electrolyte penetration and offer elastic buffering spaces to accommodate possible volume changes, which is very helpful to enhance the electrochemical performance of the materials. Sun et al. fabricated hierarchically structured Co3O4/C nanoplates on copper foil from nanosheet-assembled cobalt(II)-cobalt(III) layered double hydroxide (CoII-CoIII-LDH) precursors, which were synthesized from oil-droplet templates [538]. As shown in Fig. 30a–c, the surface of smooth Co3O4 nanoplates, with a thickness of about 30 nm, was covered by a thin amorphous carbon layer derived from the carbonation of the surfactant. When employed as binder-free electrodes for LIBs, this Co3O4/C composite exhibited initial discharge and reversible charge capacities of 1254 mA h g−1 and 1035 mA h g−1, respectively, corresponding to an initial CE as high as about 85%, as well as a remarkable reversible capacity of 1079 mA h g−1 over 50 cycles at a current density of 0.1 A g−1, which were better than Co3O4 NPs. Even at high rates, such as 3.0, 5.6 and 7.0 A g−1, the electrode still maintained specific capacities of 837, 732 and 540 mA h g−1, respectively [538]. Moreover, Ni foam has also been used as conductive substrate for the growth of Co3O4/C nanosheet arrays as free-standing electrode materials with improved rate capability, cycling stability, and ion storage performance [539,540]. During electrocatalytic reactions, apart from the activity of catalysts, the interface between catalysts and substrates is also of significance and the selection of both high-efficient catalysts and suitable substrates is crucial. Poor adhesion between the active material and the substrate usually causes the delamination of targeted catalysts, especially in corrosive solutions. Yang et al. grew cobalt oxide nanosheets on CP though a combination of electrodeposition and two-step calcination process [250]. In this experiment, the first electrodeposition of Co-species on pre-treated CP was carried in a standard three-electrode system at the current density of 10 mA cm−2 with 0.1 M Co(NO3)2 solution as electrolyte. It should be noted that a hydrophilic treatment of CP with an acid was essential before coating Co-species. After being heated in air over night, the Co-species/CP sample was treated at 350 °C first under oxygen-deficient condition for one hour, where the electroplated Co-species were converted to CoO (CoO/CP), and then under ambient condition for five hours further to make CoO nanosheets transform into Co3O4 nanosheets (Co3O4/CP). Moreover, carboncoated Co3O4 nanosheets on CP (Co3O4@C/CP) (Fig. 30d) were obtained without changes on morphology, if the Co-species/CP immersed in a glucose solution before the second treatment step. Further characterization by HRTEM revealed that the thickness of the glucose-derived carbon layer was 3.6 ± 0.5 nm. When used as catalyst for OER, the Co3O4@C/CP provided low over-potentials
Fig. 31. Morphologies and performances of 3D cobalt oxide/carbon spheres. (a, b) TEM images of (a) hollow Co3O4-C spheres [541], and (b) coreshell Co3O4-C spheres [541], (c) the discharge/charge curves in the first cycle [541], and (d) cycling performances at a current rate of 178 mA g−1 of the Co3O4-C core-shell and hollow spheres (insets show the structural changes during the insertion/extraction of Li+) [541] (Copyright 2011, Royal Society of Chemistry). (e) TEM image of Co3O4/carbon (C@Co3O4) spheres [542] (Copyright 2011, Elsevier). (f) TEM image of Co3O4/C hollow quasi-nanosphere [543] (Copyright 2012 Wiley-VCH). 631
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
of 370 mV in acid (0.5 M H2SO4) solution (Fig. 30e) and 310 mV in alkaline (1.0 M KOH) solution to achieve a current density of 10 mA cm−2. Furthermore, these electrodes maintained stable as high as 86.8 h (Fig. 30f) and 413.8 h (at 100 mA cm−2) in acid and alkaline conditions, respectively [250]. Generally, 2D cobalt oxide/carbon hybrids can be synthesized by coating carbon on cobalt oxide nanosheets or growing cobalt oxide nanosheets on carbon matrix. In-situ growth of cobalt oxide nanosheets on carbon-based substrates as binder-free electrode for electrocatalysis can take the unique merits of 2D nanostructures and provide enhanced binding force between the active materials and the current collectors. 4.1.4. 3D cobalt oxide/carbon hybrids Co3O4/carbon spheres, the most typical example of 3D hybrid structures, have always been the focus of research, due to the attractive synergetic effects of carbon coating and cobalt oxide nanomaterials. A uniformly distributed carbon layer can greatly enhance the surface conductivity, decrease the overall electrical resistance, improve the electrical contact over the electrode, and from uniform and thin SEI films. In particularly, for hollow spheres, the volume change during cycles can be partly buffered by the void spaces and the porous structures, which also can provide large contact areas between the active materials and the electrolyte. In electrocatalysis applications, the introduction of a carbon layer into the 3D cobalt oxide nanostructures is helpful to reduce overpotential, allow better reaction kinetics, and thus accelerate ion and electron transport rates. As shown in Fig. 31a, Co3O4-C hybrid hollow spheres, as reported by Zhong et al. by using a ligand exchange etching method, possessed a diameter of around 0.5 μm and the thickness of the shells was estimated to be about 50–100 nm [541]. If the reaction time of hydrothermal synthesis decreased from 2 h to 16 h, core-shell Co3O4-C spheres (Fig. 31b) were obtained. As anode materials for LIBs, both the core-shell and hollow Co3O4-C spheres exhibited higher initial discharge capacities (1002 mA h g−1 and 1485 mA h g−1, respectively) than the theoretical values of Co3O4 (890 mA h g−1) (Fig. 31c). Moreover, as shown in Fig. 31d, the reversible specific capacity of the Co3O4-C core-shell spheres reached 825 mA h g−1 at a rate of 178 mA g−1 over 40 cycles, higher than that of hollow spheres (670 mA h g−1) [541]. The superior performances for the core-shell structure to hollow structure are largely attributed to the higher void-space utilizing rate and the larger contact surface with electrolyte. Later, Jayaprakash et al. synthesized Co3O4/carbon (C@Co3O4) nanostructured spheres via high-temperature carbonization of the glucose coated cobalt oxide
Fig. 32. Synthesis, morphology, and performances of cobalt oxide/N-doped carbon (NC) hybrids. (a) Schematic illustration of the synthesis of interconnected NC framework with Co/Co3O4 NPs (Co/Co3O4/C-N) hybrid nanostructure, and (b) the corresponding STEM image [273] (Copyright 2014, Elsevier). (c) Schematic illustration of the synthetic procedure of Co3O4/N-doped porous carbon (Co3O4/N-PC) hybrid nanostructure, and (d) the corresponding TEM images [551], (e) rate capability of Co3O4/N-PC electrode for Li+ storage [551], (f) a comparison on Tafel plots of the individual Co3O4 and N-PC, Co3O4/N-PC hybrids, and Ir/C at 5 mV s−1 in 0.1 M KOH solution [551]. (Copyright 2014, Elsevier). 632
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
particles. As presented in Fig. 31e, a 22 nm-thick carbon layer was coated uniformly on the surface of Co3O4 particles approximately 400–500 nm in diameter. Electrochemical measurement confirmed that the Co3O4-C spheres exhibited a high lithium capacity of 567 mA h g−1 after 107 cycles at a current density of 440 mA g−1 [542]. Liu et al. prepared carbon-coated Co3O4 hollow quasi-nanospheres (Co3O4/C) by an in-situ carbon-coating and Ostwald ripening solvothermal process [543]. The products had a uniform quasi-spherical morphology of porous carbon shells coated on Co3O4 hollow particles (Fig. 31f). The diameter of the hybrid structures was around 250 nm and the thickness of the uniform carbon layer was about 20 nm. These carbon-coated Co3O4/C quasi-nanospheres exhibited a stable reversible capacity of about 1150 mA h g−1 over 20 cycles as anode of LIBs, whereas the bare Co3O4 hollow NPs showed only 70% retention of the initial capacity after 20 cycles. Furthermore, the rate capability of Co3O4/C remarkably maintained 700 and 150 mA h g−1 at 2 C and even 15 C, respectively [543]. Besides the hollow spherical morphologies, hierarchical and mesoporous cobalt oxide/carbon composites have also been reported as electrode materials for high-performance energy storage devices [503,544]. Lan et al. prepared mesoporous CoO nanocubes on natural rose-derived 3D porous carbon skeleton for the electrode of pseudocapacitor [545]. The obtained hybrid exhibited high capacitances of 1672 F g−1 at 1 A g−1 and 521 F g−1 at 40 A g−1. Notably, about 82% of the capacitance was maintained after 3000 cycles at 5 A g−1, and only 40% capacitance loss was observed after 1500 cycles at a relatively high rate of 10 A g−1 [545]. Wu et al. adopted an interface-modulated calcination strategy to obtain hierarchical yolk-shell Co3O4/C dodecahedrons derived from ZIF-67 frameworks [100]. An interface separation between the ZIF-67 core and the carbon-based shell occurred during the pyrolysis process. As an anode material for LIBs, the yolk-shell Co3O4/C dodecahedrons exhibited a high initial specific capacity (1209 mA h·g−1 at 200 mA·g−1) and remarkable cycling stability (1100 mA h·g−1 after 120 cycles). This type of composites was also explored for sodium ion storage and it delivered an excellent rate capability with average discharge capacities of 307 and 269 mA h·g−1 at 1.0 and 2.0 A·g−1, respectively [100]. 4.1.5. Cobalt oxide/heteroatom-doped carbon hybrids To further improve the electrochemical performance, some heteroatoms, especially nitrogen atoms, have been introduced into carbon skeletons during the hybridizing process to form cobalt oxides/N-doped carbon (NC) nanostructures for electrochemical devices [546–550]. The implantation of a large number of nitrogen atoms and N-containing functional groups can offer abundant both binding sites to anchor with cobalt oxide nanomaterials and active sites to improve ion storage performances or catalytic activities for HER, OER, and ORR. Wu et al. combined an interconnected NC framework with Co/Co3O4 NPs to form a Co/Co3O4/C-N nanostructure by using chitosan (CST) as carbon and nitrogen sources [273]. As demonstrated in Fig. 32a, Co/CST hydrogel was formed via a gelification
Fig. 33. MOF-derived cobalt oxide/NC hybrid for supercapacitors. (a) Schematic illustration of the preparation process of nanoporous carbon and Co3O4 from ZIF-67 precursor [552], (b) schematic illustration of asymmetric supercapacitors (ASCs) with Co3O4 as the positive electrode and carbon as the negative electrode [552], (c) Ragone plots of SSCs (carbon//carbon and Co3O4//Co3O4) and ASCs (Co3O4//carbon) [552], (d) comparison of the ASC cell (Co3O4//carbon) with some previously reported ACS cells [552] (Copyright 2015, American Chemical Society). 633
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
step, which was achieved by dissolving Co(NO3)2 6H2O and CST in acetic acid solution (5%), and then treated under ammonia environment. After filtration and drying processes, final product of Co/Co3O4/C-N hybrid was obtained after annealing the aerogel in N2 flow at 750 °C. The corresponding SEM image and STEM (Fig. 32b) image revealed the production of an interconnected carbon framework anchored with NPs in an average size of about 45 nm. The presence of nitrogen was confirmed by elemental mapping analysis. HRTEM images suggested the coexistence of Co and Co3O4 NPs according to the lattice fringes of Co {1 1 1} and Co3O4 {3 1 1} planes with d-spacing of 2.17 Å and 2.56 Å, respectively. Also, the product manifested a high BET surface area of 320.5 m2 g−1. In an ORR application, this hybrid nanomaterial showed remarkable catalytic activity (i.e.. an onset potential of 64.0 mV and an estimated electron transfer number of 3.4–3.9) and stability (i.e. only 18.0 mV shift for half-wave potential after 5000 continuous cycles), as well as excellent tolerance to methanol poisoning effects in an alkaline media [273]. Later, Hou et al. reported the synthesis of Co3O4/N-doped porous carbon (Co3O4/N-PC) hybrid with dodecahedrons structure via thermal transformation of cobaltbased ZIF-67 precursors [551], as illustrated in Fig. 32c. Fig. 32d shows the morphology of Co3O4 NPs with a size of 15–30 nm embedded within N-PC networks. XRD pattern and HRTEM images with the corresponding selected area electronic diffraction (SAED) pattern confirmed the presence of graphitic carbon and cubic Co3O4 NPs in the structure. Benefiting from the unique hybridized structure, compared with pure Co3O4 nanomaterial, the Co3O4/N-PC nanomaterial presented superior Li+ storage performance, including a large discharge capacity of 1730 mA h g−1 at 100 mA g−1, excellent rate capability of 560 mA h g−1 at a current density of 10 A g−1 (Fig. 32e), and remarkable cycling stability with 91.7% capacity retention (1117 mA h g−1) of the second cycle after 110 cycles. Moreover, this hybrid also showed enhanced catalytic activity for OER for water splitting, such as a low onset potential of 1.52 V, high current density, low Tafel slope of 72 mV/decade (Fig. 32f), and only 4.4% decay after 6000 s at 1.7 V [551]. With the successful synthesis of Co3O4/NC hybrid nanomaterials, this type of nanomaterials has attracted strong interest in the application as electrode materials for electrochemical capacitors. Salunkhe et al. synthesized either nanoporous carbon or nanoporous Co3O4 from a single MOF (ZIF-67) by optimizing annealing conditions [552]. Actually, MOFs, with ordered metal cluster coordinated by organic linkers, are suitable and promising candidates to produce uniform and small-sized metal oxide nanostructures. MOF-derived Co3O4 materials, including plate-like/rod-like structure [553], hollow parallelepipeds [554], and mesoporous structure [555], were utilized as electrode materials for lithium ion storage in previous reports. Meanwhile, MOF-derived carbon has also caused much attention in energy and environment-related applications [556,557]. In Salunkhe’s work, as illustrated in Fig. 33a, the carbonization of ZIF-67 at 800 °C in nitrogen produced porous carbon after purified by a followed acid treatment to remove the remained Co NPs, while nanoporous Co3O4 polyhedra can be obtained from the same ZIF-67 by heating in nitrogen at 500 °C for 30 min and then 350 °C in air. The chemical composition of the nanoporous carbon was dominant C together with 1.4 at% Co, 2.7 at% and 5.1 at% N, while the nanoporous Co3O4 was 1.7 at% C and 4 at% N, besides 36.9 at% Co and 57.4 at% O. The specific surface areas of the resulted ZIF-derived nanoporous carbon and Co3O4 were 350 and 148 m2 g−1, respectively. When utilized as electrode materials for supercapacitor, the MOF-derived nanoporous carbon presented a high capacitance value of 272 F g−1, and the Co3O4 showed a high value of 504 F g−1 at a scan rate of 5 mV s−1. Subsequently, symmetric supercapacitors (SSCs) and ASCs were also fabricated to evaluate their electrochemical properties with the as-synthesized carbon and Co3O4 as electrodes [552]. They found that the enhanced capacitance performance for the ASCs (Co3O4//carbon) has been achieved, compared with the SSCs based on carbon (carbon//carbon) and Co3O4 materials (Co3O4//Co3O4). As demonstrated in Fig. 33b, the Co3O4 positive electrode of the ASCs
Fig. 34. Morphologies and performances of the carbonized butterfly wing templates (CWs, hierarchical porous NC frameworks) and the corresponding cobalt oxide/CWs hybrids. (a-d) SEM images of carbonized wing scales with (a) type-I and (b) type-II wing scale structures, and the corresponding Co3O4 nanostructure on CWs composites (CWs-Co3O4) from the (c) type-I and (d) type-II scale frameworks, (e) CV and (f) galvanostatic discharge curves of the CWs-Co3O4 hybrids [558] (Copyright 2015, Elsevier). 634
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
offered easy access to a large amount of ions to penetrate into the interior of the porous material. On the other side, nanoporous carbon could play an effective role by providing rapid charge transfer process. Thus, this designed ASC with an optimal mass loading could be operated within a wide potential window between 0 and 1.6 V, leading to an excellent specific energy of 36 Wh kg−1. Compared with the carbon//carbon SSC cell, the Co3O4//Co3O4 SSC cell (Fig. 33c), and the previously reported ASC cells (Fig. 33d), this ASC (Co3O4//carbon) exhibited superior rate capability with the highest specific power of 8000 W kg−1 at a specific energy of 15 Wh kg−1, accompanied by a long-time stability for up to 2000 cycles [552]. Later, multi-channeled hierarchical N-doped porous carbon incorporated Co3O4 nanopillar arrays made from nature Morpho butterfly wings were also employed as binder-free electrodes for high-performance supercapacitors (Fig. 34) [558]. In the experiments, two types of wing scales of natural Morpho butterfly, the basal scale (type I, Fig. 34a) and the cover scale (type II, Fig. 34b), were first carbonized into hierarchical porous NC nanostructures, then the corresponding Co3O4 on carbonized wing scales (CWs) composites (CWs-Co3O4) were synthesized by direct deposition of Co3O4 NPs on the hierarchical porous carbonized wing templates in aqueous Co(Ac)2 solution. As shown in Fig. 34, the carbonized type-I CWs presented a dorsal surface consisting of parallel longitudinal ridge networks with rectangular windows in the interior part (Fig. 34a), and the carbonized type II structure illustrated a scale surface consisting of ridge lamellae with serrated edges (Fig. 34b). After the deposition of Co3O4 nanomaterials, a uniform and close-packed layer of Co3O4 NPs with a diameter of about 20 nm have been coated on the type-I CWs skeleton (Fig. 34c). On the typeII CWs, however, Co3O4 nanopillar arrays with a diameter of about 20 nm and a height of about 150 nm were deposited (Fig. 34d). These 3D hierarchically CWs-Co3O4 composites have demonstrated improved capacitance when used as the electrodes of supercapacitors. Stable redox peaks at different scan rates (Fig. 34e), have been identified. A maximum specific capacity of 978.9 F g−1 at the rate of 0.5 A g−1 (Fig. 34f), 94.5% of capacity retention after 2000 cycles, and a maximum energy density of 99.11 Wh kg−1 have been achieved [558]. Jin et al. have explored cobalt-cobalt oxide/NC hybrids (CoOx@CN) composed of Co0, CoO, and Co3O4 as bifunctional electrocatalysts for simultaneously generating H2 and O2 via both HER and OER in water splitting [295]. The CoOx@CN hybrid in this work was synthesized from inexpensive starting materials (e.g. melamine, CoNO3·6H2O, and D(+)-glucosamine hydrochloride, etc.) and a simple thermal treatment technique (reaction conditions: 800 °C in N2). The formed CoOx NPs had an average size of 12.3 nm and were uniformly dispersed on CN sheets. In electrocatalysis applications, the CoOx@CN hybrid exhibited a small onset potential of
Fig. 35. (a) Schematic illustration of the synthesis of hybrid material of CoO nanocrystals spatially confined in holey N-doped carbon nanowires (CoO/NCWs) [559], (b) HAADF-STEM image and the corresponding elemental mapping of CoO/NCWs [559], (c) the kinetic current (JK) and the electron transfer number (n) at 0.7 V of Co-based and Pt-based catalysts [559] (Copyright 2015, Royal Society of Chemistry). (d) LSV curves of Co3O4 decorated blood powder-derived heteroatom doped porous carbon (BDHC) hybrid (Co3O4@BDHC2, Co3O4: 27 wt%), Pt/C, IrO2 in 0.1 M KOH at a sweep rate of 10 mV s−1 (inset shows the formation of Co3O4@BDHC hybrid nanomaterials) [564] (Copyright 2014, Wiley-VCH). 635
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
85 mV and a low charge-transfer resistance of 41 Ω for HER, and a low overpotential of 0.26 V for a current density of 10 mA cm−2 for OER. Most importantly, as an electrolysis catalyst for water splitting, a current density of 20 mA cm−2 at 1.55 V was achieved with CoOx@CN as cathode and anode catalysts [295]. The remarkable catalytic performance could be possibly attributed by a unique synergistic effect from the metal Co and cobalt oxides, the electron-rich nitrogen, and the highly conductive and high large surface area carbon matrix. CoO nanocrystals spatially confined in holey NC nanowires (CoO/NCWs) fabricated via CVD and colloidal assembly to achieve enlarged interfacial area for improved electrochemical performance was reported by Xu et al. [559]. As shown in Fig. 35a, holey NC nanowires (NCWs) were first synthesized from Fe-Co metal oxides decorated anodic aluminum oxide (AAO) template via pyrimidine vapor deposition and followed HF etching, and then CoO/NCWs hybrid was obtained via colloidal assembly of CoOx precursors performed by a solvent refluxing method and mild heat treatment under Ar atmosphere. As demonstrated in the STEM-HAADF image and the corresponding element mapping illustrated in Fig. 35b, ultrafine CoO NPs with a mean size of 5 nm were uniformly deposited on the 1D NCWs. The rough surfaces, enriched nanoholes, and defects of the holey NCWs could be favorable for the immobilization of cobalt oxide NPs with uniform distribution. The as-obtained catalyst LAO manifested an enlarged electrochemically accessible interfacial area for the stabilization of CoO at a low valence [559]. As a result, the CoO/NCW electrode delivered a remarkable catalytic activity and superior stability for ORR with an electron-transfer number (∼4.0 V) close to that of Pt/C catalyst (Fig. 35c). Furthermore, cobalt oxide NPs coupled with O-and NC nanoweb (Co3O4/ON-CNW) have also proved to be a high-efficient ORR catalyst for hybrid Li-air batteries [560]. Recently, CoO/NC hybrid (NC-CoO/C) was synthesized by embedded CoO NPs into NC layers via an one-step thermal conversion reaction [561]. In this fabrication, polypyrrole-coated Co3O4 NPs supported on carbon were heat-treated in Ar at 900 °C to make the transfer of polypyrrole (PPy) into NC shell and Co3O4 phase into CoO. TEM characterization confirmed the successful transfer of Co3O4 into cubic CoO. With this strategy, high-contact areas between the CoO NPs and the NC layers were realized, and the carbon layer also effectively prevented the aggregation of CoO NPs. This NC-CoO/C catalyst provided a dual-site synergistic activity towards ORR, in which the NC could initiate reaction by reducing oxygen to the peroxide ion (HO2−) and further reduced to hydroxide ions (OH−) at metal oxide sites [562]. Moreover, the introduction of electronegative nitrogen species into carbon induced an electron-deficient environment for cobalt atoms and lead to enhanced OER kinetics [561]. Similarly, Huang et al. reported an effective cathodic catalyst (CoO@N-AC) for ORR in MFCs, where cubic CoO nanosheets exposed with {2 0 0} plane were in-situ grown on nitrogen-doped AC [563]. Strikingly, this nanocatalyst possessed a large surface area of 1577.2 m2 g−1 and a low total resistance of 9.26 Ω. When utilized as air-cathode catalysts for MFCs, a maximum power density of 1650.1 ± 36.2 mW m−2 has been achieved [563]. Besides the most used N-doping, other heteroatoms were also implanted into the carbon matrix to enhance the performance of cobalt oxides/doped carbon hybrid nanomaterials. Zhang et al. have synthesized foam-like porous nitrogen, phosphorus, and sulfur ternary-doped doped carbon/Co3O4 NPs hybrid material (inset in Fig. 35d) [564]. Low-cost commercial CaCO3 NPs with 23 nm in diameter were applied as template and activating agent and blood powders (BP) were used as a carbon source. In this synthesis, Co3O4 NPs were grown on the blood powder-derived heteroatom doped carbon (BDHC) by a thermal deposition technique. XPS results revealed that the atomic percentages of carbon, oxygen, cobalt, nitrogen, phosphorous, and sulfur in the final product were estimated as 66%, 20%, 5%, 3.4%, 3.4%, 1%, respectively. The Co3O4 NPs with an average particle size of 3 ± 1 nm were highly crystallized and well-dispersed in the amorphous carbon matrix, leading to a highest surface area of 1288 m2 g−1. As an efficient bifunctional electrochemical catalyst, the blood powder-derived heteroatom doped porous carbon and Co3O4 hybrid (Co3O4/BDHC) catalysts delivered an outstanding activity and performance towards both OER and ORR in alkaline conditions, which showed a characteristic overpotential of only 380 mV (10 mA cm−2) for OER and a half wave potential of 0.83 V (vs RHE) for ORR. It was concluded that the heterojunction could facilitate electron transfer by a reduced accumulation of electron density within Co3O4 NPs, and suggested that this type of hybrid nanomaterials can give comparable performances to commercial IrO2 catalyst for OER and Pt/C catalyst for ORR (Fig. 35d) [564]. In general, heteroatoms-doped carbon matrix has been widely used to hybridize with cobalt oxides with different dimensionalities for electrochemical applications. Among them, nitrogen atoms are the most common dopant, which is usually introduced into the carbon matrix synchronously in the carbonization process by using a raw material containing both carbon and nitrogen elements. With the introduction of heterogeneous atoms, the carbon skeletons usually become much more affinity to cobalt oxide nanostructures and help to form uniform and strongly anchored cobalt oxide nanostructures. The doped carbon and cobalt oxide hybrids have been verified to exhibit better electrochemical properties than these undoped composites. Some problems, such as the accurate management of doping amounts, the control of doping states, and the specific doping sites, still remains out of control. Most importantly, a quantitative structure-property relationship between the doped structures and the electrochemical properties are still very vague and more scientific research on this relationship is urgently demanded. 4.1.6. Cobalt oxide/carbon/others hybrids In order to meet the requirement of different electrochemical reactions for various applications, a third heterogeneous component or more is often introduced into the cobalt oxide/carbon systems to further optimize some properties, such as mechanical, electrical, physical, and chemical properties [565–572]. For example, the thickness of the carbon layer on Co3O4 NP surface has great influence on the overall electrochemical performance for LIBs. A thicker carbon layer benefits the cycling stability of materials. At the same time, however, it is also a barrier to increase the Li+ diffusion distance, which leads to lower discharge capacity and poor rate capacity. A thinner carbon layer, as a contradiction, will be easily destroyed during the repeated lithiation processes and thus loss its functionality. With the purpose to increase its rigidness and robustness of the carbon coating, Wang et al. proposed to incorporate an analogue of Al oxide, COAl, into 636
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
carbon coated Co3O4 nanostructures to form a sandwich-structured composite nanofiber bundle (C@Co3O4@COAl) by using natural collagen fibers (CF) as template, and barberry tannin (BT) as carbon source and bridge reagent to link with COAl, as shown in Fig. 36a [573]. After calcination, the inner CF@Co2+ was converted into core-shell structured C@Co3O4 nanofiber bundle, and simultaneously the outer AlOC complex was transformed into COAl covered on C@Co3O4. SEM image of C@Co3O4@COAl (Fig. 36b) and the corresponding element mapping of Al, Co, C, and O indicated that the final nanofiber bundle exhibited a typical sandwich structure. Apart from the benefits from a continuous conductive pathway provided by the inner carbon nanofiber, the hybrid coating with metal oxides and carbon plays an important role to buffer the volumetric change of Co3O4 nanomaterials and to facilitate fast electron and Li+ transports. As a consequence, the as-obtained composites exhibited significantly improved rate capability (778 and 229 mA h g−1 at the rates of 0.2 and 15.0 A g−1, respectively) and cycling stability (4000 cycles at 5 A g−1) (Fig. 36c), superior to the referred sample with double carbon layers (C@Co3O4@C) [573]. 4.2. Cobalt oxide/CNTs nanocomposites CNTs, including single-wall CNTs (SWCNTs), or multiwall CNTs (MWCNTs), feature their inherent extraordinary electrical and mechanical properties [574–577]. The walls of CNTs, however, are often chemically inert and highly hydrophobic, which bring negative effects on the uniform growth and distribution of metal oxide NPs on their surface. Moreover, the inert feature of the walls and surfaces for CNTs usually results in weak interaction between CNTs and decoration NPs. Some functionalization strategies, such as covalent attachment, noncovalent adsorption/wrapping, and the endohedral filling, are often undertaken to improve the chemical reactivity of CNTs [578,579]. For example, pre-oxidation of CNTs is an effective approach to provide strong coupling with metal oxide NPs. For the case of cobalt oxide nanostructures hybridizing with CNTs, the 0D cobalt oxide NPs can be either deposited on the surface or encapsulated in the inner space of the tube-like structure of CNTs, and 2D/3D cobalt oxide nanostructures can be interconnected or networked by CNT fibers. Besides the two-phase cobalt oxide/CNTs systems, other components can also be introduced to form ternary or multi-species composites. All these hybrid cobalt oxides/CNTs composites have been validated to be very effective to improve the electrochemical properties of this family of materials. 4.2.1. 0D cobalt oxide/CNTs hybrids To date, the assembly of 0D cobalt oxide NPs with CNTs to form 0D-1D hybrids has been paid much attention especially for their applications in electrochemical devices, which can be further classified into two subtypes, namely, cobalt oxide NPs coated on the surface of CNTs and cobalt oxide NPs encapsulated inside CNTs.
Fig. 36. Synthesis, morphology, and performance of sandwich-structured cobalt oxide-based composite nanofiber bundle (C@Co3O4@COAl) [573]. (a) Schematic illustration of the preparation of C@Co3O4@COAl, (b) SEM image of an individual C@Co3O4@COAl fiber, (c) cycling stability of C@Co3O4@COAl at a current density of 5.0 A g−1 (Copyright 2016, Elsevier). 637
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
4.2.1.1. 0D cobalt oxide NPs coated on CNTs. The most common configuration of the 0D cobalt oxide/CNTs system is the decoration of cobalt oxide NPs on the surface of CNTs. Owing to the highly hydrophobic property of the pristine nanotubes, the 0D cobalt oxide/ CNTs composites are often fabricated by solid ball-milling, solution-based mixing under strong sonication/magnetic stirring, or supercritical CO2 system, etc. [580–583]. In 2007, a general self-assembly approach was proposed by Li et al. and six different kinds of CNTs-based nanocomposites, including Co3O4/CNTs, TiO2/CNTs, Au/CNTs, TiO2/Co3O4/CNTs, Au/TiO2/CNTs, and Co/CoO/ Co3O4/CNTs, were successfully synthesized [584]. Among them, the Co3O4/CNTs was synthesized by directly mixing Co3O4 NPs with CNTs suspension in maleic acid-ethanol solution. This structure of Co3O4 NPs (4–5 nm in size) on CNTs (Fig. 37a and b) were stable and maintained largely intact even after sonicating treatment [584]. Super-aligned CNT array (SACNT), a special type of vertically aligned CNT arrays possessing the capability of being transformed into continuous films and yarns [585–589], were explored as binder-free current collector for in-situ growth of Co3O4 NPs [590]. In the synthetic procedure, SACNTs were first assembled into continuous CNT films as a flexible and highly conductive scaffold to host Co3O4 NPs (Fig. 37c), and then, Co3O4 NPs were grown on the SACNTs by the pyrolysis of Co(NO3)2 absorbed on the SACNTs. After synthesis, as shown in Fig. 37d, the ordered structure of the film was largely retained and no obvious aggregation of Co3O4 NPs was observed. The resulting Co3O4/SACNT composite anodes for LIBs exhibited an initial discharge capacity of 1200 mA h g−1 (Fig. 37e), a reversible capacity of 910 mA h g−1 and a CE of 97% after 50 cycles (Fig. 37f), and outstanding rate capability of about 820 mA h g−1 at 1 C [590]. Even though the successful loading of cobalt oxide NPs on the surfaces of CNTs, the chemically inert and hydrophobic walls of the pristine CNTs (P-CNTs) result in a weak binding between them, if the CNTs and the Co3O4 NPs are directly mixed without prior functionalization. Surface modification is often used to promote the chemical activity and the wettability. Zhuo et al. compared the electrochemical properties of two types of CNTs, p-CNTs and functionalized CNTs (f-CNTs), with cobalt oxide particles coating [591]. The p-CNTs were purified with nitric acid in an ultrasonic bath and showed a smooth surface with a diameter of 40–60 nm and a length of several micrometers. The f-CNTs presented a rough surface with the existence of holes and functional groups obtained through a steaming treatment in acid vapor [592], and the length reduced to hundreds of nanometers. Then, Co3O4 were thermally decomposed on these CNTs. As shown in Fig. 38a and d, Co3O4 with a diameter of about 150–300 nm were anchored uniformly and strongly on the external surfaces along the axis of the f-CNTs. In contrast, in the Co3O4-p-CNT composite, Co3O4 particles with a larger diameter of 300–500 nm did not present close contact with the p-CNTs, but were dispersed freely (Fig. 38c and d). It is clear that the f-CNTs that have ample functional groups and lateral defects on the walls offer more nucleation sites for the growth of Co3O4, leading to better dispersing and anchoring of Co3O4 as well as a higher mass loading of 87%. When used as anode materials for lithium ion storage devices, the Co3O4-f-CNT composites delivered high discharge capacities (e.g. 719 mA h g−1 at the 2nd cycle) and excellent cycling performance (776 mA h g−1 after 100 cycle) at a current density of 200 mA g−1. Even at a high current density of 1
Fig. 37. Morphology and performance of cobalt oxide NPs/CNT hybrids. (a, b) TEM images of Co3O4/CNTs composite [584] (Copyright 2007, American Chemical Society), (c) SEM images of a pristine super-aligned CNT film (SACNT) [590], (d) SEM image of a Co3O4/SACNT composite electrode [590], (e) galvanostatic charge/discharge curves, and (f) cycling performance of Co3O4/SACNT composite at 0.1 C [590] (Copyright 2013, Royal Society of Chemistry). 638
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
Fig. 38. Morphology and performance of cobalt oxide NPs/CNT hybrids. (a) SEM and (b) TEM images of Co3O4/functionalized CNTs hybrid (Co3O4f-CNT) [591], (c) SEM and (d) TEM images of Co3O4/pristine CNTs hybrid (Co3O4-p-CNT) [591], (e) a comparison on the rate capability of Co3O4-fCNT and Co3O4-p-CNT hybrid electrodes [591] (Copyright 2012, Royal Society of Chemistry). (f) TEM of the beaded necklace-like Co3O4/MWCNTs structures [596] (Copyright 2005, Wiley-VCH). (g, h) SEM images of Co3O4-CNT heterostructure with beaded “necklace” morphology [597] (Copyright 2013, Royal Society of Chemistry).
A g−1, the capacity of Co3O4-f-CNT still maintained about 600 mA h g−1, higher than that of Co3O4-p-CNT composites (350 mA h g−1) under the same testing conditions (Fig. 38e) [591]. Other synthetic approaches, such as electrophoretic deposition (EPD) and chemical co-precipitation, were also exploited to grow cobalt oxide NPs anchored on f-CNTs [593–595]. It is noteworthy that the necklace-like cobalt oxide NPs/CNTs structures have been studied for LIBs. Fu et al. fabricated an ordered beaded necklace-like Co3O4/MWCNTs architecture, in which spherical Co3O4 nanocrystallines with a diameter around 90 nm were
Fig. 39. Morphology, characterization and electrocatalytic performance of cobalt oxide NPs on CNTs hybrid nanomaterials. (a) SEM image, (b) TEM image (Inset: electron diffraction pattern), and (c) HRTEM image of CoO/nitrogen-doped CNT (NCNT) hybrid nanostructure (CoO/NCNT) [274]. (d) Oxygen reduction polarization curves (Inset: a scheme of CoO/NCNT for ORR catalysis) [274] (Copyright 2012, American Chemical Society). (e) Oxygen activities of the Co3O4-based and noble metal-based catalysts for bifunctional OER and ORR catalysis [601] (Copyright 2016, The Electrochemical Society). 639
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
connected stepwise by MWCNTs to form a beaded “necklace” (Fig. 38f) via the simple treatment of cobalt precursor and MWCNTs under mixed supercritical fluid (SCF) containing CO2 and ethanol in a high-pressure vessel [596]. Xu et al. also reported a Co3O4-CNT heterostructure with a similar bead-on-string architecture by using alkylcarboxyl group-decorated CNTs [597]. The introduction of carboxyl groups onto the CNTs surface makes the CNTs easily dispersed in water before the in-situ growth of Co3O4 spheres. As shown in Fig. 38g and h, The Co3O4 spheres with an average size of ca. 100 nm were threaded with CNTs to form a bead-on-string architecture. Owing to the fact that the CNTs passed through the inner Co3O4 spheres to provide intimate contact and a continuous electron transport channel, the necklace-like nanoarchitectures exhibited superior mechanical stability and electrochemical performance. As anode materials for LIB applications, the composites delivered a high initial discharge capacity of 1248 mA h g−1 and high capacity retention of 758 mA h g−1 after 30 cycles. Even at a current density as high as 6000 mA g−1, this composites offered a high capacity up to ∼400 mA h g−1, while the referenced material, aggregated Co3O4 spheres, showed no performance at such high rates [597]. Apart from the lithium ion storage, this type of cobalt oxide NPs on CNTs hybrids has also been used as an electrocatalyst for both OER and ORR [598–600]. Liang et al. developed CoO/CNT covalent hybrid as highly efficient ORR catalyst by directly growing CoO nanocrystals on mildly KMnO4-oxidized MWCNTs (moCNT), including a mild solution-phase synthesis followed by annealing at 400 °C in an NH3 atmosphere [274]. The coated CoO NPs on CNTs had a mean size of around 5 nm (Fig. 39a and b) and a cubic CoO rock-salt structure (Inset in Fig. 39b and c). The resultant CoO/nitrogen-doped CNT (NCNT) hybrid exhibited much higher ORR current density than that of CoO, NCNT and their physical mixture loaded on Teflon-coated carbon fiber paper (TCFP) (Fig. 39d). Furthermore, the CoO/CNTs hybrid was found to be advantageous to the one hybridized with graphene because of more active sites and much rapid charge transport provided by CNTs. Electrochemical and XANES spectroscopy measurements confirmed the strong coupling between NPs and CNTs, which favor high-efficient charge transport in ORR catalysis. The oxidation degree of CNTs was of significance to the activity of the hybrids, likely due to the balance between the electrical conductivity and the density of functional groups on CNTs. Compared with thermal annealing in Ar, the reduction in NH3 environment gave rise to improved electrochemical properties. It is interesting that this strongly coupled CoO/CNTs hybrid even presented high ORR activity and stability under a corrosive alkaline condition (10 M NaOH) [274]. Further research on the influence of the length of CNTs on the catalytic activity of Co3O4/CNTs has been carried out by So et al. [601]. Three types of CNTs (CM150, CM250, and CM280) with different lengths (30, 70, and 200 μm) were hybridized with Co3O4 and tested for both OER and ORR performances. The results indicated that the Co3O4/ CM150 CNTs with the highest surface area of 170.46 m2 g−1 delivered the highest OER activity, while the CM280 CNTs with the longest length coupled with Co3O4 presented the highest ORR performances. It has demonstrated that, to achieve high-efficient bifunctional performance, the oxygen activity of Co3O4/CM150 CNTs can be further enhanced by using oxygen-containing CNTs (oxCNTs) (Fig. 39e), for example, by chemical functionalization in acidic KMnO4 solution [601].
Fig. 40. Synthesis, morphology, and performance of 1D cobalt oxide/CNTs hybrid nanostructure. (a) Schematic illustration of the formation of hierarchical CNT/Co3O4 hybrid microtubes [606], (b) the corresponding SEM image of CNT/Co3O4 hybrid microtubes [606], (c) cycling performance and CE of CNT/Co3O4 hybrid microtubes [606] (Copyright 2016, Wiley-VCH). (d) TEM image of the Co3O4@MWCNT hybrid, [604], (e) schematic illustration of the synergetic effect of Co3O4 nanorod and MWCNT for water oxidation [604] (Copyright 2014, Royal Society of Chemistry).
640
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
4.2.1.2. 0D cobalt oxide NPs encapsulated inside CNTs. The second type of 0D cobalt oxide/CNTs hybrid materials is to encapsulate/ embed/fill cobalt oxide NPs into CNTs. Park et al. fabricated vertically aligned mesoporous carbon nanotubes filled with Co3O4 NPs (Co3O4/MCT) on a copper current collector through a dual template method, where an AAO membrane was applied as hard template and a block copolymer surfactant F127 was employed as soft template [602]. Co3O4 NPs (10–20 in diameter) were dispersed uniformly on the inside surface of MCT. As an anode material for LIBs, this electrode displayed a high reversible capacity of 627 mA h g−1 over 50 discharge/charge cycles [602]. Overall, for the 0D cobalt/CNT hybrid nanostructures, the deposition of cobalt oxide NPs on the surfaces of CNTs is easy to be achieved, but a uniform dispersion is still very difficult due to the inactive CNT walls. The functionalization of CNTs before the deposition of cobalt oxide nanostructures provides a good solution to address this challenge. Surface modification changes the surface groups, surface roughness, and the wettability of CNTs and provides much higher affinity to the precursors of cobalt oxides, which leads to much stronger contact between cobalt oxide NPs and CNTs and thus superior electrochemical performances. The fabrication of cobalt oxide NPs embedded or encapsulated inside CNTs is more challenging and further study on this interesting type of hybrids is still on the way. 4.2.2. 1D cobalt oxide/CNTs hybrids 1D cobalt oxide nanostructures, with a high aspect ratio, are also integrated with CNTs to improve their electrochemical performances [603–605]. Hierarchical tubular structures (HTSs) composed of Co3O4 hollow NPs and CNTs were designed and fabricated via a multi-step route, as illustrated in Fig. 40a [606]. This hybrid nanostructure, different to the above-mentioned configuration of Co3O4 NPs on the surface of CNT, consist of Co3O4 tubes and vertically aligned CNTs from the walls (Fig. 40b). In this work, electrospun PAN-cobalt acetate (Co(Ac)2) composite nanofibers were selected as the templates to provide cobalt source for the following growth of ZIF-67. Then, owing to a strong coordination of 2-methylimidazole to Co2+ within the fibers, uniform shells of ZIF-67 nanocrystals grew on these nanofibers. After being dispersed in N, N-dimethylformamide (DMF) for removing PAN-Co(Ac)2 core, the obtained ZIF-67 tubulars were further converted into CNT/Co-carbon hybrids by thermal treated in Ar/H2. Finally, Co NPs were oxidized into Co3O4 hollow NPs via annealing in air to generate the hierarchical CNT/Co3O4 microtube hybrid nanostructures. The inner diameters of the Co3O4 hollow NPs and the CNTs were approximately 10 ± 5 nm and 3 ± 2 nm, respectively. These HTSs constructed by hollow NPs offered ample active interfacial sites, and effectively alleviated volume changes during electrochemical reactions. As a result, the as-prepared hierarchical CNT/Co3O4 microtubes delivered a high initial discharge capacity of 1840 mA h g−1 at 0.1 A g−1, an exceptional rate capability (515 mA h g−1 at 6.0 A g−1) and long cycle life over 200 cycles (782 mA h g−1 at 1.0 A g−1) (Fig. 40c) as an anode material for LIBs [606]. Zhang et al. synthesized Co3O4 nanorods anchored on MWCNTs (Co3O4@MWCNT) via a diethylenetriamine (DETA)-assisted
Fig. 41. Synthesis of 2D cobalt oxide nanoplates/2D CNT fishing-net-like layers for LIBs. (a) Schematic illustration of the preparation of layer-bylayer 2D Co3O4 nanoplates/2D CNT nets structure [611], (b) TEM image of the 2D-2D hybrid structure [611], (c) discharge profiles and (d) discharge capacity at different current densities of the 2D Co3O4 nanoplate/2D CNT hybrid [611] (Copyright 2015, American Chemical Society). 641
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
solution-phase synthesis method [604]. It is presented in Fig. 40d and e that Co3O4 nanorods were intimately wrapped on the CNTs to form a 1D/1D heterostructure, which is favorable for electron transport and surface electrochemical reactions. In this 1D/1D heterostructure, the Co3O4 nanorods exposed with {1 1 0} facets. As an efficient catalyst for OER, this hybrid exhibited an onset potential of 474 mV (vs. Ag/AgCl) and an overpotential of 310 mV to achieve a current density of 10 mA cm−2 in alkaline electrolytes for water oxidation [604]. In summary, 1D cobalt oxide/CNTs hybrid nanostructures are advantageous for providing high surface-to-volume ratio, remarkable conductivity, and excellent surface activities. The electrochemical properties, however, are quite unsatisfying. The possible reasons include the long-distance electron transport in the 1D geometry and the non-uniform and non-continuous hybridization forms of 1D-1D configuration. It is still relatively difficult to achieve the uniform growth of 1D cobalt oxide nanowires or nanorods on the surfaces of CNTs. Furthermore, the related exploration of 1D/1D hybrid structures on electrochemical energy applications remains very limited. Most reported 1D/1D heterostructures are mainly related to photocatalysis, largely attributed to the significantly enhanced light absorption and scattering contributed by the high length-to-diameter ratio [607–610]. Hence, much more investigated are needed to make this type of materials appealing in electrochemical energy related applications. 4.2.3. 2D cobalt oxide/CNTs hybrids Compared to 1D cobalt oxides, 2D cobalt oxide nanostructures are much easier to form hybrid structures with 1D CNTs. It has been reported that Co3O4 nanoplates coupled with capillary-like MWCNT nets to form a 3D layer-by-layer structure demonstrated both high-rate and long-term cycling stability, when it was employed as LIB anode material [611]. As shown in Fig. 41a, 2D hexagonal α-Co(OH)2 nanoplates and 2D CNT fishing-net-like layer were transferred layer-by-layer to copper foil, and then α-Co(OH)2 nanoplates were converted into Co3O4 nanoplates by thermal treatment at 300 °C for 2 h in air. TEM characterization revealed that the MWCNT nets enveloped the surfaces of each α-Co(OH)2 nanoplate (Fig. 41b). Electrochemical results confirmed that this 2D Co3O4 nanoplates/2D CNT network multilayer structure presented an initial discharge capacity of about 1550 mA h g−1. The discharge capacity still remained at 710 mA h g−1 at a current density as high as 50 A g−1 (Fig. 41c), which was fully recovered when the rate was returned to 2 A g−1 after 75 cycles. In this work, differential capacities of the hybrid at current densities from 0.1 to 50 A g−1 were also examined, as shown in Fig. 41d. Two broad peaks in the 1.5–1.0 V and 1.0–0.5 V represented the intercalation of Li+ ions into the Co3O4 to form a LixCo3O4 intermediate phase, and the full reduction of the intermediate phase to cobalt metal, respectively. It was observed that both broad peaks shifted to the negative direction and the intensity of the peaks associated with the conversion reaction gradually decreased with the increasing current density. These results indicated that the capacity loss at high rates was primarily due to the slow kinetics of the conversion reaction accompanying with a phase transformation process. Meanwhile, no obvious capacity was lost aroused by the formation of the intermediate phase, which was ascribed to the more stable phase state of LixCo3O4 even at high rates [611]. Besides, the fully recovered capacity of the Co3O4 nanoplates/CNT network multilayer structure proved that Co3O4 nanoplates were in good electrical contact with the electrically conductive CNT nets, contributing to the stable long-term cycling performances [611].
Fig. 42. Morphology and performance of MOF-derived cobalt oxide polyhedral hybridized with CNTs nanostructures (MWCNTs/Co3O4) [619]. (a) SEM and (b) TEM images of MWCNTs/ZIF-67, (c) SEM and (d) TEM images of MWCNTs/Co3O4, (e) cycling stability (100 mA g−1) and (f) rate capability of the MWCNTs/Co3O4 (Copyright 2015, American Chemical Society). 642
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
4.2.4. 3D cobalt oxide/CNTs hybrids 3D cobalt oxide nanostructures, such as hollow, flake-like, or sphere-like structures with an overall size over 100 nm, have been widely studied to hybridize with CNTs to improve the electrochemical properties of cobalt oxides [612–616]. Some advantages can be brought by 3D nanostructures for the applications in electrochemical devices. For example, hollow structures, which usually associate with low mass density, high porosity, large reaction surface area, and good strain accommodation, high specific surface areas, have been demonstrated to be effective to accommodate the large volume change during cycling, provide extra space for ion storage, facilitate lithium diffusion, and allow better reaction kinetics when this type of nanostructures are used in rechargeable batteries [617,618]. Venugopal et al. reported a template-free soft chemical method to fabricate hollow mesoporous Co3O4 and Co3O4-CNT composites by using acid-treated f-CNTs [612]. It was observed that each Co3O4 hollow particle had a double-layered nanowall 50–100 nm in thickness, and the f-CNTs were penetrated inside the hollow particle. When applied as negative electrodes for LIBs, the Co3O4-CNT hybrid materials showed an initial discharge specific capacity of 1420 mA h g−1, higher than that of Co3O4 (1283 mA h g−1) and the physically mixed Co3O4 + CNT (1380 mA h g−1). What is more, the Co3O4-CNT composites presented superior rate capability with capacity retention at 620 mA h g−1 even at a high rate of 1000 mA g−1 [612]. The excellent performances of the hollow Co3O4/f-CNT hybrid material resulted from the unique hollow and mesoporous structure of Co3O4 NPs and the hybridization with CNTs, which is helpful to mitigate the mechanical stress and enhance the kinetics of the electrochemical reactions. 3D Co3O4 polyhedra/CNTs hybrids have been reported as anode materials for ion storage [619,620]. Huang et al. synthesized MOF-derived Co3O4 polyhedra/MWCNTs hybrid composites by thermal treatment of the in-situ synthesized MWCNTs/ZIF-67 [619]. The MWCNTs/ZIF-67 composite, synthesized by co-precipitation of cobalt ions and 2-methylimidazolate together with the presence of MWCNTs, composed of uniform MWCNT-inserted polyhedral with smooth surfaces and a size of about 500 nm (Fig. 42a and b). After calcination at 400 °C, the MWCNTs/ZIF-67 nanocomposites converted to MWCNTs/MOF-Co3O4 hybrid nanocomposites with well-maintained size distribution and morphology. The most significant difference between the ZIF-67 and the Co3O4 polyhedra is the formation of mesoporous structure, owing to the release of CO2 and H2O from the ZIF-67 polyhedra during calcination (Fig. 42c and d). TEM characterization further confirmed the existence of hierarchical pores, and the intertwined and inserted MWCNTs within the Co3O4 polyhedra. As anode materials for LIBs, the MWCNTs/MOF-Co3O4 composites displayed excellent cycling stability (813 mA h g−1 at 100 mA g−1 after 100 cycles) (Fig. 42e) and good rate capability (514 mA h g−1 at 1000 mA g−1) (Fig. 42f). It is worth noting that this protocol can also be used to prepare MWCNTs/ZnCo2O4 composites by using hetero-bimetallic MWCNTs/ZIF67 (Zn, Co) as precursors [619].
Fig. 43. In-situ electrochemical XAS analyses at an open circuit potential (OCV) and (a) OER potential of 1.8 V and (b) ORR potential of 0.6 V, (c) the effect of NP size on OER and ORR performances [623] (Copyright 2016, American Chemical Society). (d, e) TEM images of MOF-derived Co3O4 nanocrystals embedded in N-doped mesoporous carbon and MWCNT hybrid structures (Co3O4@C-MWCNTs) [624] (Copyright 2015, Royal Society of Chemistry). (f) Schematic illustration of assembled two-electrode rechargeable Zn-air battery [625], (g) charge and discharge polarization curves of the assembled Zn-air battery based on Co-embedded CNT/porous carbon (Co-CNT/PC) and Pt/C + RuO2 mixture cathode catalysts [625] (Copyright 2016, Royal Society of Chemistry). 643
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
4.2.5. Cobalt oxide/CNTs/others hybrids Heterogeneous cobalt and its oxide nanocomposites, such as Co/CoO and Co/Co3O4, have been hybridized with CNTs to promote the electrocatalytic efficiency for oxygen catalysis or Li-air batteries [621,622]. In the design of cobalt oxide-based catalysts, it was found that the nanoscale particle size and the valence states of cobalt and cobalt oxide significantly influence the electrocatalytic reactions. Seo et al. studied the size-dependent catalytic activity of bifunctional cobalt oxide (CoOx) NP/CNTs catalysts for both OER and ORR [623]. In-situ X-ray absorption spectroscopy (XAS) revealed that the majority of the CoOx NPs during both OER (Fig. 43a) and ORR (Fig. 43b) consisted of Co3O4 and CoOOH phases regardless of their particle sizes. As shown in Fig. 43c, the OER activity of the CoOx/CNT catalysts increased as the decrease of the CoOx NP size, which is correlated to the increased amount of Co(III) species and the larger surface area of CoOx NPs, whereas the ORR activity of the CoOx/CNT catalysts was nearly independent on the CoOx NP size [623]. Cobalt oxide/carbon/CNTs hybrid is another important ternary composite for electrocatalysis. MOF-derived Co3O4 nanocrystals (10–25 nm in diameter) in-situ embedded in N-doped mesoporous graphitic carbon layer and MWCNT hybrid structures (Co3O4@CMWCNTs) (Fig. 43d and e) were synthesized by a facile carbonization and a subsequent oxidation of cobalt-based MOF precursor, ZIF-9, with the presence of MWCNTs [624]. The as-obtained hybrid material showed remarkable electrochemical performances for OER and ORR. As OER catalysts, an onset potential of 1.50 V and an overpotential of 320 mV achieved for a stable current density of 10 mA cm−2 for at least 25 h. Compared to the commercial 20 wt% Pt/C catalyst, the hybrid exhibited superior stability for ORR [624]. Later, Co-embedded CNT/porous carbon (Co-CNT/PC) was synthesized by pyrolysis of ZIF-67 encapsulated Co3O4 NPs, and used as bifunctional electrocatalysts towards ORR and OER in rechargeable Zn-air batteries (Fig. 43f), which showed a similar performance to the assembled battery using the state-of-the-art Pt/C and RuO2 mixture as oxygen cathode electrocatalysts (Fig. 43g) [625]. Besides the above-mentioned nanocarbon, other active components, such as MnO2 and polydopamine, were also introduced into the cobalt-CNTs hybrids as a ternary component for improving the electrochemical performances of the materials [626,627]. In spite of increasing reports on the cobalt oxide/CNT hybrid for electrochemical applications, the fabrication of uniform composites consisting of cobalt oxide and CNTs is a grand challenge, largely owing to the slender tubular structure and the intrinsic chemically inert and highly hydrophobic surfaces of CNTs. Moreover, the high cost of CNTs is still a major barrier for the applications
Fig. 44. Synthesis and performance of 0D cobalt oxide/graphene hybrids. (a) Schematic illustration of the fabrication of 0D Co3O4/graphene composite [648], (b) the corresponding TEM image [648], (c) charging/discharging curves of the 0D Co3O4/graphene composite at a current density of 50 mA g−1 [648], (d) comparison of the cycling stability of Co3O4, graphene, and the Co3O4/graphene composite [648] (Copyright 2010, American Chemical Society). (e) AFM image of Co3O4 NPs/graphene composite with smaller Co3O4 NPs [649] (Copyright 2010, Elsevier). (f) Low magnification and (g) high magnification SEM images of CoO QD/graphene (CQD/GN) composites [655] (Copyright 2012, American Chemical Society). 644
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
of CNT-based materials. 4.3. Cobalt oxide/graphene hybrids Graphene, an emerging 2D nanomaterial, have attracted considerable attention for the potential applications in energy conversion and storage devices. Graphene possesses some distinct advantages, including striking electrical conductivity, ultrahigh surface area, together with excellent mechanical/chemical stability. The salient properties of graphene offer promising potentials to create incredible synergic effects to the materials, when it is used as supportive materials to combine with other substances or materials. By integrating with cobalt oxide nanostructures, the graphene-metal oxide composites have presented enhanced performances as electrocatalysts or electrode materials, where graphene mainly acts as conductive supports to reduce the inner resistance. Various architectures composed of graphene nanosheets and cobalt oxide nanomaterials with various morphologies and dimensions have been synthesized in the past decades, such as 0D-2D hybrids, 1D-2D hybrids, 2D-2D hybrids, and 3D-2D hybrids, in which cobalt oxide nanostructures can be deposited/anchored/grown on the surface of graphene nanosheets, or encapsulated/ wrapped by the sheet-like graphene, or assembled into layered/sandwich-like structures in a layer-on-layer or sheet-on-sheet manner. In the 0D-2D configuration, NPs-on-graphene (0D-2D) can effectively improve the dispersion degree of the cobalt oxide NPs and avoid serious aggregation. The nanosheets-encapsulated/wrapped-NPs (2D-0D) configuration is beneficial to accommodate volume changes of cobalt oxide nanostructures during repeated cycles. The nanosheets-on-graphene (2D-2D) stacks can take the full advantage of the merits of geometrical compatibility to enhance the interface force between cobalt oxide layers and graphene layers. 4.3.1. 0D cobalt oxide/graphene hybrids The integration of 0D cobalt oxide NPs with graphene has been widely studied and can be easily achieved via directly mixing dispersants of Co3O4 NPs and graphene/graphene oxide (GO)/reduced graphene oxide (rGO), or in-situ growth of cobalt oxide nanomaterials on graphene surface [628–647]. Wu et al. fabricated Co3O4 NPs/graphene composite via thermal treatment of the Co(OH)2/graphene precursor at 450 °C, as illustrated in Fig. 44a [648]. The resulting Co3O4 NPs, with a size of around 10–30 nm, were homogeneously anchored on the surfaces of graphene nanosheets (Fig. 44b). Even after a long time of sonication, the Co3O4 NPs were still strongly anchored on the graphene surface with a high density, indicating a strong interaction between them. As a result, the obtained nanocomposites exhibited an initial discharge capacity of 1097 mA h g−1 (Fig. 44c) at 50 mA g−1, a reversible capacity of 935 mA h g−1 after 30 cycles at 50 mA g−1 (Fig. 44d), and capacity retention of 484 mA h g−1 after 40 cycles at 500 mA g−1 [648]. Later, Co3O4/graphene hybrid composed of smaller Co3O4 particle sizes (∼5 nm) (Fig. 44e) was reported by Kim et al. for highly reversible anode for LIBs [649]. This anode delivered a high reversible capacity of over 800 mA h g−1 at a current density of 200 mA g−1 in the voltage window between 3.0 and 0.001 V. Remarkably, more than 550 mA h g−1 was retained even at a high rate of 1000 mA g−1 [649]. 3D Porous graphene assembled from 2D graphene sheets can provide a larger surface area with abundant pores and more exposed active sites for electrochemical reactions. Hence, cobalt oxide species deposited on the defective surfaces of the 3D graphene, or cobalt oxide NPs confined inside 3D porous structure, will be helpful for the access of the electrolyte ions, and thus greatly reduce the irreversible capacity and improve the CE as electrode materials or electrocatalysts for electrochemical applications [105,650–653]. For example, Choi et al. fabricated 3D hierarchical porous Co3O4/rGO composite films by the deposition of Co3O4 NPs (5.87 ± 2.1 nm in diameter) onto the surfaces of the 3D porous graphene [651]. The porous structure of graphene was controlled by using polystyrene (PS) spheres as structural guiding templates. When the porous Co3O4/rGO heterogeneous films were directly applied as anode materials without binders and current collectors, they exhibited a good cycling performance with 90.6% retention during 50 cycles. Moreover, an excellent rate capability of 71% retention was maintained at the current rate as high as 1000 mA g−1, which was much better than these of physically mixed Co3O4/rGO (48%) and barely Co3O4 (10%) [651]. Besides the composition with Co3O4 NPs, 3D hollow crumpled graphene were also employed to hybridize with CoO NPs to form CoO/graphene hybrid for the applications of electrochemical reactions, which presented excellent catalytic activity and durability to become promising nonprecious metal-based bi-functional catalysts for both ORR and OER [326,654]. Apart from NPs, cobalt oxide QDs were also hybridized with graphene to form QDs/graphene composites by using Co4(CO)12 as precursor [655]. With the assistance of sonication at ambient temperature, the precursor clusters on graphene nanosheets were decomposed into metallic Co clusters, which were subsequently oxidized into CoO by solution-dissolved O2. As illustrated in Fig. 44f and g, CoO QDs with small sizes of 3–8 nm were firmly anchored and uniformly distributed on the graphene nanosheets. Such CoO QD/graphene nanosheet (CQD/GN) composites with metal oxide QDs on conductive graphene support are advantageous to inhibiting aggregation, buffering volume change of cobalt oxides during cycling, and offering lots of active sites and short ion-diffusion pathways. As expected, as an anode material for LIBs, the composites delivered an outstanding reversible lithium storage capacity of 1592 mA h g−1 at a current density of 50 mA g−1 after 50 cycles, and excellent rate capability (e.g. 1008 mA h g−1 at 1000 mA g−1) [655]. Owing to the excellent performances of 0D cobalt oxide NPs/graphene hybrid nanostructures, extensive research has been reported to fabricate this type of nanocomposites for the applications in electrochemical devices. In most cases, to improve and distribution of cobalt oxide NPs on graphene surfaces, chemically active GO is often employed initially to absorb or couple with cobalt-based precursors, and then to be reduced into rGO with the assistance of thermal treatment and/or reducing agents. After the reduction of GO, however, some functional groups may still remain and some defects may form on the surface, which can result in decreased conductivity compared to the pristine graphene. Therefore, the use of oxidized graphene and the selection of reduction methods are important to achieve a uniform deposition of cobalt oxide nanostructures on the surfaces of graphene but without much 645
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
loss of its conductivity. 4.3.2. 1D cobalt oxide/graphene hybrids 1D cobalt oxide nanostructures, including needle-like, wire-like, and fiber-like structures, anchored on 2D graphene nanosheets, or grown on 3D graphene aerogel/foams have been investigated as electrode materials for batteries, capacitors, and catalysts for oxygen electrocatalysis [656–664]. Ryu et al. proposed a bifunctional composite catalyst composed of 1D Co3O4 nanofibers immobilized on 2D nonoxidized graphene nanoflakes (Co3O4 NF/GNF) for an oxygen electrode in Li-O2 batteries [660]. Fig. 45a showed the whole fabrication procedure. First, 1D polycrystalline Co3O4 nanofibers were prepared via an electrospinning strategy to obtain 1D fiber-like network. Then, the Co3O4 NFs/GNFs composites (Fig. 45b) with NFs assembled on both sides of GNFs were synthesized by simply mixing Co3O4 nanofibers with graphene dispersion, where the nonoxidized graphene nanosheets were noncovalently functionalized by 1-pyrenebutyric acid (PBA). When employed as a bifunctional catalyst for the oxygen electrode of Li-O2 batteries, the composite delivered an excellent initial discharge capacity as high as 10,500 mA h g−1 and maintained a limited capacity of 1000 mA h g−1 for 80 cycles (Fig. 45c). The remarkable electrochemical properties were mainly attributed to the improved electrocatalytic activity for both OER and ORR, the fast electron transport, and the easy oxygen diffusion of Co3O4 NF/GNF composites [660]. Besides the layer-by-layer assembly of 1D Co3O4 nanofibers with graphene, vertically aligned CoO nanowires directly grown on 3D graphene networks via a wet chemistry process, as shown in Fig. 45d, were reported for lithium storage without extra addition of any binders or conductive additives, which exhibited a high capacity of 857 mA h g−1, an excellent rate capability, and good cycle performance [661]. Co3O4 nanowire/3D graphene composite was synthesized as free-standing electrode for supercapacitor and for enzymeless electrochemical detection of glucose [662]. In the synthesis, Co3O4 nanowires were in-situ hydrothermally grown on the CVD-
Fig. 45. Synthesis, morphology, and performance of 1D cobalt oxide/graphene hybrids. (a) Schematic illustration of the synthesis of Co3O4 nanofibers networks assembled on both sides of nonoxidized graphene nanoflakes (Co3O4 NF/GNF) [660], (b) TEM image of Co3O4 NF/GNF composite [660], (c) a comparison on cycling performance for Co3O4 NPs, Co3O4 NFs, Co3O4 NF/RGO hybrid, and Co3O4 NFs/GNF hybrid composite at a limited capacity of 1000 mA h g−1 at a current density of 200 mA g−1 [660] (Copyright 2013, American Chemical Society). (d) FESEM image of vertical 1D CoO nanowire grown on 3D graphene network [661] (Copyright 2014, IOP Publishing). (e) SEM images of 1D Co3O4 nanowire grown on 3D graphene form [662], (f) CV curves of 3D graphene/Co3O4 composite electrode at different scan rates [662], (c) galvanostatic charge and discharge curves at different current densities [662], (e) charge and discharge profile of the 3D graphene/Co3O4 composite electrode at 10.0 A g−1 [662] (Copyright 2012, American Chemical Society). 646
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
prepared graphene foam. As shown in Fig. 45e, Co3O4 nanowire networks fully covered the graphene skeleton with uniform distribution. Each nanowire was around 200–300 nm in diameter and several micrometers in length. As a free-standing electrode for supercapacitors, the obtained composites were capable of delivering a pseudocapacitive behavior (Fig. 45f) with a specific capacitance of 768 F g−1 at a current density of 10 A g−1 (Fig. 45g), and a long cycling time of over 25,000 s (Fig. 45h). Moreover, 1D Co3O4/3D graphene had great potential for electrochemical sensing. For example, this hybrid nanostructure has been used for nonenzymatic detection of glucose and presented an extraordinary sensitivity of 3.39 mA mM−1 cm−2, and a detection limit lower than 25 nM with a signal-to-noise ratio (S/N) of 8.5 [662]. In general, directly mixing 1D cobalt oxides with graphene benefits to handicap the self-agglomeration of graphene sheets. The binding between the graphene and the cobalt oxide nanostructures, however, is relatively weak through physical contact, and cobalt oxide nanowires are easy to aggregate into random bundles between the graphene layers. 1D cobalt oxide nanostructures grown on graphene foams via a much stronger chemical contact can effectively prevent the restack of graphene, enlarge the specific surface area, and enhance the overall conductivity of the composites. More promisingly, these hybrids can be applied as free-standing electrode for electrochemical devices without any additional binders or current collectors. 4.3.3. 2D cobalt oxide/graphene hybrids The hybridization of 2D cobalt oxide nanomaterials with graphene can be divided into three types, namely, horizontal growth on graphene, vertical growth on graphene, and layer-by-layer sandwiched 2D-2D nanostructures [33,665–667]. Among them, the sheeton-sheet 2D-2D structure is appealing for electrochemical applications due to its enhanced binding force between layers and improved structural stability without serious aggregation of graphene [668–670]. Free-standing 2D Co3O4/graphene hybrid films were synthesized by vacuum filtration followed by a thermal treatment (Fig. 46a), in which the sheet-like Co3O4 (∼20 nm in thickness) and graphene were assembled into a lamellar hierarchical structure, as shown in Fig. 46b, by means of electrostatic interactions [671]. This layered morphology with strong interfacial interactions significantly promotes the interfacial electron and Li+ transport. As binder-free and free-standing electrodes for LIBs, the hybrid films delivered a high specific capacity of ∼1400 mA h g−1 at a current density of 100 mA g−1, improved rate capability, and superior cyclic stability with 1200 mA h g−1 at the 100th cycle at 200 mA g−1 [671].
Fig. 46. Synthesis, morphology, and performance of 2D cobalt oxide/graphene hybrids. (a) Schematic illustration of the fabrication process of Co3O4/graphene nanosheets hybrid films, (b) the corresponding SEM image [671] (Copyright 2013, Royal Society of Chemistry). (c, d) STEM images of atomically thin mesoporous 2D Co3O4 nanosheets/graphene (ATMCNs-GE) composite [672], (e) a comparison on cycling performances of ATMCNs, graphene and ATMCNs-GE at a rate of 2.25 C [672] (Copyright 2016, Wiley-VCH). 647
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
Dou et al. fabricated atomically thin mesoporous Co3O4 nanosheets/graphene (ATMCNs-GE) composite under stirring (Fig. 46c and d) [672]. Benefiting from the unique structure of the atomic-level thickness and mesoporous structure of the Co3O4 nanosheets and the conductive graphene, this hybrid exhibited extraordinary electrochemical performance, which delivered high discharge capacities of 2014.7 and 1134.4 mA h g−1 at 0.11 and 2.25 C, respectively, and an outstanding capacity retention of 92.1% at 2.25 C after 2000 cycles (Fig. 46e). These performances significantly outperformed those of the bare ATMCNs and the previously reported Co3O4/C composites [672]. Recently, 2D-2D Co3O4/rGO composites consisting of ultrathin 2D Co3O4 nanosheets and rGO synthesized through a one-pot hydrothermal strategy have been reported as highly-efficient OER catalyst [246]. The obtained 2D-2D hybrid exhibited a low overpotential of 290 mV at 10 mA cm−2 with a small Tafel slope of 68 mA dec−1 in 1.0 M KOH, much smaller than that of bare Co3O4 catalyst. In this work, 0D carbon spheres and 1D CNTs were also employed to replace 2D rGO to be hybridized with Co3O4 nanosheets to form 2D Co3O4/0D carbon spheres and 2D Co3O4/1D CNTs hybrids. Interestingly, they found that the overpotentials decreased with the increased dimensionality numbers of the carbon nanostructures and they suggested that the electrochemical activity of the Co3O4/nanocarbon composites is very closely related to the surface area of the carbon substrates [246]. Overall, the layer-by-layer stacked 2D cobalt oxides and graphene nanosheets 2D-2D hybrid structures have the best geometric compatibility and usually provide superior structural stability for electrochemical applications. In this type of 2D-2D structures, the agglomeration of both cobalt oxide nanosheets and graphene can be largely inhibited. In the fabrication, particularly for wetchemistry synthesis of this type of hybrid nanostructures, we need to pay close attention to the surface potential of both metal oxide and graphene nanosheets. 4.3.4. 3D cobalt oxide/graphene hybrids Various 3D cobalt oxide nanostructures (e.g. spheres, cubes, polyhedrons, etc.) have been reported to be coupled with graphene nanosheets by two primary strategies, namely, anchored on graphene and or encapsulated by graphene [673–678]. The “anchoredon-graphene” strategy is very simple for synthesis and has the potential for large-scale production. One of the biggest drawbacks for this strategy, however, is that there is a weak interface between the metal oxide nanostructures and graphene, which usually leads to strong aggregation of metal oxide nanostructures during electrochemical reactions and then loss of the synergic effect of the hybrids. On the contrary, 3D cobalt oxide/graphene nanocomposites fabricated via the “encapsulated-by-graphene” strategy possess enhanced binding force between the cobalt oxide nanostructures and the graphene. According to the reported studies on the application of the “encapsulated-by-graphene” type nanocomposites, the conductive graphene coating on Co3O4 nanomaterials provides much improved charge and ion transfer and avoids the formation of unstable SEI. Furthermore, the void space between the metal oxide nanomaterials and graphene layer contributes to buffer the volume expansion during long-time cycling life. Graphene-encapsulated 3D cobalt oxide nanostructures are often achieved by self-assembly of the negatively charged GO and the positively charged oxide nanostructures. In 2010, Yang et al. synthesized graphene-encapsulated solid Co3O4 nanospheres (NSs) (GECo3O4) with high electrochemical activity by employing the electrostatic interactions of differently charged Co3O4 nanospheres and graphene [675]. In this experiment, Co3O4 NSs were first modified with 3-aminopropyltrimethoxysilane (APS) to render a positively
Fig. 47. Synthesis, morphologies, and performances of graphene-encapsulated cobalt oxide nanostructures. (a) Schematic illustration of the fabrication of graphene-encapsulated mesoporous Co3O4 microspheres (G-Co3O4) [677], (b) the corresponding FESEM image of G-Co3O4 (Inset: FESEM image of graphene) [677] (Copyright 2012, Royal Society of Chemistry). (c) SEM image of graphene-wrapped mesoporous Co3O4 hollow spheres core-shell (Co3O4@G) nanostructures [676], (d) cycling performance and CE of Co3O4@G electrode at 1.0 A g−1 (inset shows the fabrication process of the core-shell composites) [676] (Copyright 2014, American Chemical Society). 648
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
charged surface, and then were encapsulated by negatively charged GO. After chemical reduction of GO, the GE-Co3O4 composites were finally obtained. Electrochemical measurements indicated that the composites exhibited excellent cycle performance with a reversible capacity of ∼1100 mA h g−1 in the first 10 cycles, and 1000 mA h g−1 after 130 cycles [675]. Besides the solid Co3O4 nanospheres, mesoporous Co3O4 spheres featuring high surface area and highly-ordered inner channels were used to be encapsulated by graphene nanosheets to produce Co3O4/graphene hybrid (G-Co3O4) (Fig. 47a) [677]. SEM images revealed that the mesoporous Co3O4 spheres with a diameter of about 700 nm and micropores of 7.2 nm were fully wrapped by graphene shells (Fig. 47b). The obtained composites as anode materials for LIBs exhibited an initial discharge capacity as high as 1533 mA h g−1 at a current density of 100 mA g−1 and remained at a reversible capacity up to 820 mA h g−1 at 200 mA g−1 [677]. Later, a similar principle has also applied to fabricate graphene-wrapped mesoporous Co3O4 hollow spheres to form core-shell (Co3O4@G) nanostructures by using poly(allylamine hydrochloride) (PAH) as the surface modifier of Co3O4 spheres [676]. As shown in Fig. 47c, Co3O4 hollow spheres with a diameter of ∼270 nm were covered by graphene nanosheets. As anode materials for LIBs, this core-shell structure delivered a high reversible capacity of 1076 mA h g−1 over 10 cycles at a current density of 0.1 A g−1 and displayed an excellent cyclic stability (over 600 mA h g−1 after 500 cycles at a high rate of 1.0 A g−1) (Fig. 47d) [676]. Likewise, graphene-encapsulated porous Co3O4 cubes (Co3O4@G) were also synthesized by using uniform Co3[Co(CN)6]2 cubes as the precursor [679]. This structure was expected to be advantageous to fast transport of Li+/electron and alleviation of the volume change during long-term cycles, and thereby a remarkable cycling and rate performances. When evaluated as anode materials, this cubic Co3O4@G exhibited a reversible discharge capacity of 980 mA h g−1 after 80 cycles at 200 mA g−1 and 500 mA h g−1 at high rate of 2 A g−1 [679]. Generally, the encapsulation of 3D cobalt oxide structures by graphene is usually achieved by electrostatic attraction between cobalt oxides and graphene with different surface charges. However, it is unlikely to ensure the cobalt oxides to be completely covered by the graphene nanosheets. The encapsulation of metal oxide by graphene can give rise to superior electric conductivity on the one hand. On the other hand, the diffusion of electrolyte and the transport of ions through the graphene shells would be largely retarded. Further design on the shell structure of this type of hybrid nanostructure is still demanded to reach expected electrochemical performances. 4.3.5. Cobalt oxide/heteroatom-doped graphene hybrids Similar to the heteroatom-doped nanocarbon and CNTs, some heteroatoms have also introduced into the graphene skeleton with the purpose to increase its chemical reactivity to modify its electronic structure, or to enhance its chemical affinity to metal oxides [272,678,680–682]. Hence, compared to the cobalt oxide/pristine graphene composites, the hybrids of cobalt oxide nanostructures with heteroatom-doped graphene have many incomparable advantages for their potential applications in electrochemical devices. For example, the introduction of nitrogen atoms in graphene can offer stronger coupling between cobalt oxides and graphene and be favorable for the nucleation of cobalt oxide at the N-doped sites during synthesis. Furthermore, Co-Nx species in which cobalt is coordinated to or strongly interacts with nitrogen are believed to be potential active sites for ORR [683–685]. Liang et al. reported Co3O4 nanocrystals grown on N-doped RGO nanosheets as a high-performance bi-functional catalyst for ORR
Fig. 48. Morphology and performance of cobalt oxide NPs on N-doped graphene [28]. (a) SEM, (b) TEM, and (c) HRTEM images of the Co3O4 NPs/ nitrogen-doped reduced mildly oxidized GO hybrid (Co3O4/N-rmGO), (d) oxygen electrode activities of Co3O4, Co3O4/N-rmGO and Pt/C catalysts in 0.1 M KOH (catalyst loading ∼0:24 mg cm−2), (e) oxygen evolution currents and (f) Tafel plots of Co3O4, Co3O4/rmGO and Co3O4/N-rmGO (catalyst loading: ∼1 mg cm−2) in 1 M KOH [28] (Copyright 2011, Nature Publishing Group). 649
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
and OER [28]. Based on their synthesis route, well-crystallization Co3O4 NPs were grown on graphene sheets by hydrolysis and oxidation of cobalt acetate. When NH4OH was added into this system, the Co3O4 NPs/N-doped reduced moGO hybrid (Co3O4/NrmGO) was obtained. SEM (Fig. 48a) and TEM (Fig. 48b and c) images revealed that well-crystallized Co3O4 NPs (4–8 nm) in spinel structure were anchored on graphene nanosheets. When used as catalyst for electrocatalysis, the Co3O4/N-rmGO hybrid showed remarkable ORR activities than Co3O4 or GO alone in alkaline solutions. As shown in Fig. 48d, the Co3O4/N-rmGO electrode affords higher OER currents compared with Co3O4 nanocrystals and Pt/C. Moreover, when the catalyst loading on the substrate was increased into ∼1 mg cm−2, a small overpotential of 0:31 V affording a current density of 10 mA cm−2 (Fig. 48e) and a small Tafel slope of 67 mV dec−1 (Fig. 48f) were achieved. Besides, it is obviously observed that the Co3O4/N-rmGO hybrid delivered only slightly higher OER activity than the Co3O4/rmGO hybrid, indicating that the N-doped graphene has no effect on OER activity [28]. The integration of nitrogen-doped graphene with cobalt oxide nanomaterials also delivers more appealing electrochemical properties for rechargeable batteries. Lai et al. synthesized Co3O4 NPs grown on nitrogen modified microwave exfoliated graphite oxide (Co3O4/NMEG) and Co3O4 on thermally rGO (Co3O4/tRG-O) with Co3O4 loading ratio of 10–70% [686]. Electrochemical measurements revealed that the 70% Co3O4/NMEG composite delivered an initial reversible capacity of 799 mA h g−1 and retained 910 mA h g−1 after 100 cycles, and the 70% Co3O4/tRG-O had a reversible capacity of 750 ± 20 mA h g−1 but the irreversible capacity loss during the first cycle was 700 ± 20 mA h g−1. More interestingly, Co3O4/NMEG composites with different Co3O4 loading weight ratios from 10% to 70% showed outstanding cycling behaviors with little capacity decay over 100 cycles at a current density of 100 mA g−1. The advantages brought by N-doping becomes much significant at high charge/discharge rate. The enhancement of the electrochemical performances of the Co3O4/NMEG composites was supposed to be from the nitrogen functional groups in NMEG, especially the pyridinic (C]N) and the pyrrolic (C^N) N, which are advantageous for the growth of Co3O4 nanomaterials, the reduced oxygen content of graphene, finally the amelioration of the first cycle efficiency [686]. Then, Yao et al. incorporated 3D scaffold of graphene aerogel with synchronous chemical-doping of nitrogen atoms and 1D mesoporous Co3O4 nanowires to produce N-doped graphene/Co3O4 nanowires composites (NGA-Co3O4) [687]. As illustrated in Fig. 49a and b, the Co (CO3)0.5(OH)·0.11H2O precursor was first synthesized within a self-assembled 3D graphene network, and then transformed into Co3O4 nanowires after cryodesiccation and thermal decomposition. For LIB application, a remarkably high capacity of 1229 mA h g−1 was achieved at a current density of 100 mA g−1 over 100 cycles, which was four folds higher than that of Co3O4 nanowires (314 mA h g−1) and NGA (419 mA h g−1) (Fig. 49c). Moreover, 812 mA h g−1 was still retained after 230 cycles at a rate of 1000 mA g−1 (Fig. 49d) [687]. Other heteroatoms, such as sulfur [688] and boron [689], have also been incorporated into graphene framework for enhancing its electrochemical performances. Recently, CoOx NPs with rich oxygen vacancies strongly coupled with B, N-decorated graphene (CoOx NPs/BNG) via Co-N-C bridging bonds have been reported by Tong et al. [689]. This synthesis was achieved by in-situ NH3 treatment of a Co-Bi/G precursor (Fig. 50a). TEM (Fig. 50b and c) characterizations showed that small CoOx NPs grown uniformly on the graphene sheets, and the existence of CoO was confirmed by the exposed {2 0 0} lattice planes. In the material, abundant oxygen vacancies, Co-N and C-B-N bonds bonds, and strong Co-N-C bridging bonds formed at the pyridinic-N and pyrrolic-N sites were identified by XPS and XANES. It has been reported that N atoms can induce positive polarization of C atoms, while B atoms with a low electronegativity can be positively polarized with the existence of N atoms [690], and these positively polarized C-B and N-C bonds favorer the capture of O2 molecules during ORR. As an efficient bifunctional electrocatalyst for OER/ORR, the CoOx NPs/BNG hybrid
Fig. 49. Synthesis, morphology, and performance of 1D cobalt oxide nanowires on N-doped graphene aerogel [687]. (a) Schematic illustration of the fabrication of nitrogen-doped 3D graphene aerogel/1D Co3O4 nanowires (NGA-Co3O4) composites [687], (b) SEM image of the NGA-Co3O4 hybrids, (c) cycling performance of Co3O4 nanowires, NGA, and NGA-Co3O4 hybrids at a current density of 100 mA g−1, (d) cycling stability of NGACo3O4 hybrid electrode at a high rate of 1000 mA g−1 (Copyright 2016, Wiley-VCH). 650
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
Fig. 50. Preparation, morphology, and performance of cobalt oxide/B, N-doped graphene hybrids [689]. (a) Schematic illustration of the preparation of CoOx NPs/B, N-decorated graphene composites (CoOx NPs/BNG), (b) TEM and (c) HRTEM images of CoOx NPs/BNG hybrids, (d) polarization curves in the both ORR and OER regions in 0.1 M KOH solution (Copyright 2017, Wiley-VCH).
was highly efficient for OER with a low overpotential (295 mV at 10 mA cm−1) and Tafel slope (57 mV dec−1), and also active for ORR with a positive half-wave potential (805 mV) and high limiting current density (∼5.67 mV cm−2 at 0.2 V) in alkaline medium, which is accompanied by a low potential gap of 0.72 V (Fig. 50d) [689]. In spite of some major progress have been achieved in the doped graphene/cobalt oxide composites, similar to the heteroatomdoped nanocarbon/cobalt oxide hybrids, further effort on low-cost and large-scale preparation with controllable doping levels is still urgently needed to further enhance their electrochemical performances.
Fig. 51. Morphology of Co3O4-Mn3O4/GO hybrid and the electrocatalytic pathways [700]. (a) TEM and (b, c) STEM dark field images of Co3O4Mn3O4/GO hybrid, (d) STEM-EDS elemental mapping (green and red points show Co and Mn elements, respectively) of Co3O4-Mn3O4/GO composite, schematic illustration of (e) the interfacial structures and (f) the ORR pathway catalyzed by Co3O4-Mn3O4/GO composite (Copyright 2016, Elsevier). 651
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
4.3.6. Cobalt oxide/graphene/others hybrids Carbon (black) [691–694], CNTs [695–697], polymers (e.g. PPy, etc.) [698], and other metal oxides (e.g. MnO2, Mn3O4, CoFe2O4, etc.) [271,699–701] have also hybridized with cobalt oxide nanostructures and graphene networks to further enhance the ion storage performance and catalytic activity of the materials, and to trigger more appealing synergistic effects to meet the requirements for the practical application of cobalt oxide based nanomaterials. Dai et al. synthesized 3D stacked-up Co3O4-Mn3O4/GO ternary composites via coherent nucleation and growth of Mn3O4 on GO supported Co3O4 based on a two-step aqueous synthesis [700]. It can be easily found from Fig. 51a that these oxide NPs, with a mean size of about 13.5 nm (Fig. 51b), were uniformly dispersed on the graphene surfaces. Furthermore, as shown in Fig. 51c and d, the smaller Mn3O4 nanocrystals had been successfully dispersed onto the surfaces of Co3O4 NPs to form an oxide-on-oxide nanostructure on the surface of graphene. In an alkaline environment, compared to individual Mn3O4/GO and Co3O4/GO catalysts, the Co3O4Mn3O4/GO composite exhibited much better electrocatalytic activity (e.g. a half-wave potential of −0.2235 V, etc.) towards ORR. This might be largely attributed to the enhanced covalent electron transfer at the interfaces between these two metal oxides and the GO matrix, resulting from their interphase ligands and synergetic effects (Fig. 51e). During ORR, as depicted in Fig. 51f, the Co3O4 phase on the graphene serves as the main catalyst for catalyzing ORR with the formation of a large amount of OH− via nearly fourelectron pathway and a small amount of H2O2 intermediates via two-electron pathway. Subsequently, with the presence of Mn3O4 phase on the surface of Co3O4 phase, the resultant H2O2 undergoes a chemical disproportionation reaction to produce OH− and O2 specie, which is repeatedly utilized as reactants for the further ORR. The oxide-on-oxide structure in this hybrid brings about a small diffusion distance, which is beneficial to promote the decomposition of H2O2 and the further utilization of O2. Moreover, defects and slight lattice strain were found in the spinel Mn3O4 phase of the composite, which also contribute to the decomposition of H2O2 [700]. Recently, flexible films/papers composed of 1D CNTs and 2D graphene have been paid much attention for energy conversion and storage devices [702–707]. This class of unique combination of 1D/2D hybrid film can effectively prevent the self-stack tendency of graphene nanosheets and possess excellent electronic conductivity and electrochemical performances. What is more, these flexible and conductive films are outstanding substrates as free-standing electrodes for sustainable applications when depositing with active materials [270,696,708–710]. Yuan et al. reported a facile and efficient route to fabricate flexible and free-standing Co3O4/rGO/ CNTs hybrid paper, in which Co3O4 monolayer microsphere arrays were directly grown on the rGO/CNTs film by a simple one-step hydrothermal deposition [696]. As displayed in Fig. 52a, the hybrid paper possessed a layered structure with an average thickness of about 50 μm. As showed in Fig. 52b, the hybrid paper demonstrated good flexibility. The enlarged SEM images (Fig. 52c) revealed that interconnected Co3O4 microsphere array grew on the surface of the rGO/CNTs film with 1 μm in thickness. Electrochemical evaluation for its application as electrochemical capacitors demonstrated that this flexible paper electrode delivered specific capacitance of 378 and 297 F g−1 at 2 and 8 A g−1, respectively in 3.0 M KOH solution with a voltage window between −0.2 to 0.45 V
Fig. 52. Morphology of cobalt oxide/graphene/CNTs hybrid paper. (a) FE-SEM image of Co3O4/rGO/CNTs hybrid paper (inset shows the enlarged part) - [696], (b) photograph showing the flexibility of the hybrid paper [696], (c) the enlarged cross-sectional view of the hybrid paper [696] (Copyright 2012, Wiley-VCH). (d) SEM image of the cross-section of N-doped graphene/CNT/Co3O4 (NG/CNT/Co3O4) paper (inset shows a photograph of the flexible paper) [270] (Copyright 2014, Royal Society of Chemistry). 652
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
[696]. In 2014, flexible N-doped graphene/CNT/Co3O4 (NG/CNT/Co3O4) paper (Fig. 52d) was synthesized in ammonia solution with in-situ formation of Co3O4 NPs, reduction of GO into rGO, and doping of nitrogen species [270]. This paper has a thickness of around 4 μm, with CNTs and Co3O4 NPs (∼25 nm in size) existed between the graphene nanosheets. As working electrode for ORR, it delivered an enhanced catalytic performance with high current density and positive onset potential. Furthermore, this as-prepared paper exhibited good cycling stability and tolerance to methanol poisoning [270]. 4.4. Cobalt oxide/metal hybrids The implantation of heterogeneous transition metal atoms/ions/NPs, such as Fe, Ni, Ru, Pd and Pt into cobalt oxide nanostructures, greatly promotes the electrochemical performances of cobalt oxide-based nanomaterials, particularly for electrocatalytic reactions in related fuel cells and metal-air batteries [269,711–719]. Cao et al. reported a binder/carbon-free air electrode with tips-bundled Pt/Co3O4 nanowires grown directly on Ni foam substrate via a hydrothermal route [718]. As shown in Fig. 53a-c, the obtained Pt/Co3O4 nanowires with sharp tips were uniformly arranged on the Ni foam substrate, in which Pt nanocrystals (< 5 nm in size) were homogeneously distributed within Co3O4 nanowires. Interestingly, if without the presence of Pt nanocrystals, only a random arrangement of Co3O4 nanowires was found on the Ni foam substrate, indicating the presence of Pt plays a crucial role on the formation of the ordered tips-bundled structure. Consequently, a LiO2 battery by using this tips-bundled Pt/Co3O4 nanostructures as catalysts (Fig. 53d) showed a stabilized discharge plateau (> 2.50 V) (Fig. 53e) and low charge end voltages (< 4.0 V) up to 55 cycles. In contrast, the cycling of the battery by only using Co3O4 as catalysts maintained only 10 cycles with high charge terminal voltages (> 4.0 V) and low discharge end voltages (< 2.5 V). Further characterization of these electrodes after discharge confirmed that Pt also affected the crystallization habit of the discharge product Li2O2. With Pt nanocrystals, a fluffy and thin Li2O2 layer was uniformly deposited only on the periphery of Pt/Co3O4 nanowires (Fig. 53d), which makes the Li2O2 nanocrystals easy to decompose at low potentials with reduced side reactions. Otherwise, large-sized flower-like particles and plate aggregates of Li2O2 randomly distributed on the bare Co3O4 nanowires and resulted in inferior electrochemical performances [718]. 4.5. Cobalt oxide/metal oxides heterogeneous hybrids The combination of different metal oxide nanostructures into heterogeneous structures has demonstrated to be an effective way to design novel electronic correlation materials and catalytic materials owing to its extraordinary interfacial coupling effects, regulated charge/carrier transport behaviors, synergetic catalytic activities, etc. Lots of cobalt oxide/metal oxide heterogeneous hybrids composed of cobalt oxides and other metal oxides, including Fe3O4/Co3O4 [720,721], Co3O4/RuO2 [722–724], Co3O4/ZnO [725], MnO2/Co3O4 [726–731], Mn3O4/Co3O4 [732], NiO/Co3O4 [324,733–735], Co3O4/CeO2 [736], NiCo2O4/Co3O4 [737], Co3O4/ NiMoO4 [738], Co3O4/CoMoO4 [739], ZnO/NiO/Co3O4 [740], etc., have been designed and synthesized in various architectures, such as core-shell, shuttle-like, flake-like, NPs-on-fiber, NPs-on-wire, rod-on-sheet, horn-on-sheet structures, etc. [741–743]. These hybrids have been extensively studied for the potential applications in batteries, capacitors, and electrocatalysis. Xia et al. reported a two-step solution-based approach to synthesize TMO core/shell nanoarrays on a variety of substrates including fluorine-doped tin oxide glass (FTO), nickel foam, and nickel foil [733]. In this synthesis, the Co3O4 or ZnO NWAs, synthesized by hydrothermal reactions in the first step, served as the backbone for the following chemical bath deposition of NiO nanoflakes. The morphologies of Co3O4/NiO core/shell structures with Co3O4 core nanowire and NiO shell nanoflake were confirmed by SEM and TEM. As shown in Fig. 54a–c, the highly porous Co3O4 nanowires had an average diameter of 70 nm and the thickness of
Fig. 53. Morphology and performance of tips-bundled Pt/Co3O4 nanostructures for Li-O2 battery [718]. (a) schematic illustration of the Pt/Co3O4 cathode architecture, (b) SEM and (c) TEM images of the tips-bundled Pt/Co3O4 nanostructures on Ni foam substrates, (d) schematic illustration of the Pt/Co3O4 cathode architecture and the working mechanism of Li-O2 battery, (e) voltage profiles of Li-O2 batteries at 100 mA g−1 with Pt/Co3O4 cathode (Copyright 2015, American Chemical Society). 653
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
Fig. 54. Morphology and performance of Co3O4/NiO core/shell NWAs for supercapacitor [733]. (a, b) SEM and (c) TEM images of Co3O4/NiO core/ shell NWAs on fluorine-doped tin oxide glass (FTO) substrate, (d) a comparison of specific capacitances of different electrodes at various current densities, (e) a comparison on cycling performances of different electrodes at 2 A g−1 (Copyright 2012, American Chemical Society).
NiO nanoflakes was about 10 nm. Benefited from the porous wire core and the flake shell, the Co3O4/NiO NWAs exhibited a quite large surface area as high as 415 m2 g. As electrode material for supercapacitor, the Co3O4/NiO NWAs on nickel foam delivered excellent rate capability (e.g. 452 F g−1 at 2 A g−1 and 384 F g−1 at 40 A g−1) (Fig. 54d), and remarkable cycling stability with an areal capacitance of 2.56 F cm−2 after 6000 cycles, much higher than those of the individual Co3O4 NWAs (1.36 F cm−2) and NiO nanoflake arrays (0.16 F cm−2) (Fig. 54e) [733]. Kim et al. designed a sectionalized MnO2-Co3O4 electrode for metal-air batteries via an agarose gel-mediated strategy, in which the selective electrodeposition of MnO2 and Co3O4 were achieved on 3D porous nickel foam [726]. At the first stage, semi-flexible hydrogel was produced by mixing molten agarose and cobalt precursor ions in an aqueous solution. After an electric current was applied to the gel attached to the nickel foam, cobalt ions were reduced to Co(OH)2, which was further converted into Co3O4 by heat treatment. Finally, MnO2 was incorporated into the opposite side of the substrate by using a similar procedure to form the final sectionalized MnO2-Co3O4 electrode on Ni foam (MnO2-Co3O4/NiF), as illustrated in Fig. 55a. This hybrid structure effectively enhanced catalytic activities for both ORR and OER. The sectionalized electrode showed outstanding cycling performance and longterm stability (e.g. 400 h of operation time) in rechargeable Zn-air batteries (Fig. 55b) [726].
Fig. 55. (a) Schematic illustration of sectionalized bifunctional MnO2-Co3O4 on Ni foam electrode (MnO2-Co3O4/NiF) [726], (b) cycling stability of the assembled rechargeable Zn-air battery using MnO2-Co3O4/NiF electrode at a current density of 1 mA cm−2 [726] (Copyright 2016, Elsevier). 654
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
4.6. Cobalt oxide/polymers hybrids Two types of polymers, including porous network polymers (e.g. covalent organic polymers (COPs), etc.) and conductive polymers (e.g. PPy, polyaniline (PANI), polyindole (Pind), etc.), are commonly used to hybridize with cobalt oxide nanostructures to improve its electric conductivity, enhance adhering force, and facilitate rapid charge transfer for electrochemical applications [200,744–750]. The storage mechanism of the conductive polymers depends on the redox reaction process. During the oxidation step, ions are transferred to the backbone of the polymer, which will be released again into the electrolyte when the reduction reaction occurs. Hence, the combination of cobalt oxides with conductive polymers can greatly promote the reaction rate and enhance the ionstorage performances. Zhou et al. developed a novel supercapacitor electrode composed of well-aligned CoO nanowire arrays grown vertically on 3D nickel foam in the presence of conductive PPy to boost the pseudocapacitive performance [200]. As illustrated in Fig. 56a, CoO NWAs were first grown on commercial nickel foam via a combination of hydrothermal and post-annealing steps, and PPy was subsequently immobilized onto these arrays based on a chemical polymerization reaction by using ammonium persulphate as the oxidant and ptoluenesulfonic acid (p-TSA) as the doping agent. After the integration of PPy, the CoO arrays remained their original ordered structures (Fig. 56b). Further TEM analysis confirmed that each nanowire surface was coated by a thin PPy layer with a nanosized thickness and some PPy was attached firmly to the nanowires in particulate forms. This electrode architecture could take full advantage of the high electrochemical activities derived from both CoO and PPy, the high electronic conductivity of PPy polymer, and the short ion diffusion pathway within these ordered mesoporous nanowires. These structural advantages as well as the synergy effects of the hybrid electrode for systematic supercapacitor led to a high specific capacitance of 2223 F g−1 at 1.0 mA cm−1, good rate capability (∼647 F g−1 at 50 mA cm−1), and excellent cycling stability (∼99.8% capacitance retention after 2000 cycles). Furthermore, an aqueous asymmetric supercapacitor device was assembled by using the CoO@PPy as positive electrode and AC film on nickel foam as negative electrode, in 3 M NaOH electrolyte, as depicted in Fig. 56c. The cell voltage was as large as 1.8 V and the device delivered a high energy density of ∼43.5 Wh kg−1 at 87.5 W kg−1, high power density of ∼5500 W kg−1 at 11.8 Wh kg−1 (Fig. 56d), and outstanding cycling life up to ∼20,000 times. This asymmetric supercapacitor thus had the potential to bridge the performance gap between thin-film Li-ion batteries and EDLCs (Fig. 56d and e) [200]. The integration of organic conductive polymers and cobalt oxide nanostructures to form cobalt oxide/polymers hybrids is an effective strategy to contribute to a wide voltage window and a remarkable storage capacity for energy storage devices, especially for the widely studied supercapacitors. Moreover, with the assistance of these polymers, the resulting hybrids feature adjustable redox activities because the functionalization of these organic polymers is easy to be realized via different organic chemical modification techniques. Furthermore, the redox reactions for these conductive polymers occur in the entire polymer skeleton instead of preferentially on the surface for inorganic metal oxides such as cobalt oxides. In this regard, a uniform distribution of metal oxide component in cobalt oxide/polymers hybrids, such as the above-mentioned PPy-coated CoO nanowires, can bring forward an integral conductive network over the whole hybrid frameworks. Another advantage for the introduction of organic polymers is that no phase change behavior for the polymers exists during the charging/discharging processes, which theoretically makes the redox reactions reversible and thus leads to an excellent cycling stability. In practical application, however, these polymers may mechanically
Fig. 56. Preparation, characterization, and electrochemical performances of cobalt oxide/polymer (CoO/PPy) hybrid [200]. (a) Schematic illustration of the synthesis of 3D hybrid nanowire electrode, (b) SEM image of the CoO/PPy hybrid, (c) schematic illustration of the configuration of CoO/PPy and carbon asymmetric supercapacitor, (d) Ragone plot of the assembled supercapacitor device and the referenced EDLC and Li-ion battery, (e) a comparison on volumetric energy and powder densities of the asymmetric supercapacitor (Copyright 2013, American Chemical Society). 655
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
degraded aroused by swelling and shrinking, and some side reactions may also appear under different types of electrolyte systems, resulting in a fast capacity decay over cycling. Considering these side effects, optimized design for enhancing the structural stability is still needed to achieve excellent electrochemical properties for this type of materials. 4.7. Other cobalt oxide-based hybrids Other cobalt oxide-based hybrids, including ion-doped cobalt oxides [751–753], cobalt oxides/metal sulfides/selenides [296,754,755], cobalt oxides/CxNy [294,756], cobalt oxides/metal hydroxides [757–759], and some complex cobalt oxide-based hybrids with three or more components [760–763], have been reported as electrode materials or catalysts for electrochemical applications [712,764–768]. Recently, Zhou et al. developed a MOF template-directed fabrication of hierarchical Co3O4@X (X = Co3O4, CoS, C, and CoP) nanostructures, which were derived from cobalt carbonatehydroxide@ZIF-67 (CCH@ZIF-67) [769]. As shown in Fig. 57a, in-situ growth of ZIF-67 on well-aligned CCH nanorods supported on Ni foil was first achieved by hydrothermal synthesis, and then various products (e.g. Co3O4@Co3O4, Co3O4@CoS, Co3O4@C, or Co3O4@CoP) were obtained by different post-treatment methods, including oxidation, sulfurization, carbonization, or phosphorization. This unique hierarchical structure gave rise to ample exposed active sites, shortened ion diffusion pathway, and enhanced conductivity. As electrocatalysts for OER, these derivatives revealed high efficiency catalytic properties, which were much superior to cobalt-based catalysts and the Ir/C catalyst [769]. In particular, Co3O4@CoP exhibited the highest catalytic capability among the Co3O4@X nanostructures with an overpotential of 0.238 V at 10 mA cm−2 (Fig. 57b), and a current density of 45 mA cm−2 at the overpotential of 0.300 V. Moreover, these hybrids also demonstrated improved catalytic performances to other small molecules (e.g., glycerol, methanol, or ethanol) towards electro-oxidation reaction [769]. Cobalt oxide-based composites with other heterogeneous components have been fully explored with the purpose to achieve the best combination of both enhanced structural stability and improved electrochemical properties. Compared with the cobalt oxidebased hybrids with carbon, CNTs, and graphene, the properties are less attractive for these types of nanomaterials. For example, the integration of emerging active metal sulfides with cobalt oxides can result in enhanced activities of the related redox reactions for energy conversion and storage devices, however, the conductivity is still difficult to be alleviated. Moreover, multiple components of the active materials can bring about more complex reactions and the chance of the appearance of some possible side reactions is greatly increased. Moreover, how to identify the specific contribution of each component for the overall electrochemical properties is still a technical issue. Therefore, it is highly demanded that in-depth studies on the underlying reaction mechanisms in the hybrids or nanocomposites with multiple components should be paid more attention. 5. Conclusion and outlook Rapid progress has been made in the research of cobalt oxides and their hybrids for electrochemical applications, particularly in the recent five years. Cobalt oxide nanostructures in different dimensionalities have been fabricated for various electrochemical energy applications, including rechargeable batteries, supercapacitors, and electrocatalysis, for their appealing features of low cost and abundance in earth, high theoretical capacity/capacitance and electrochemical activity, in electrochemical energy storage, and excellent chemical and mechanical stability, etc., compared to other metal oxides. It is out of question that cobalt oxides will be one of the most promising materials as an alternative electrode material for next-generation electrochemical energy devices. The electrochemical performances of cobalt oxides in their pristine forms, as we reviewed in this article, however, are still far from their practical applications, owing to some intrinsic defects, such as poor electrical conductivity, slow reaction kinetics, several morphology and volume changes during electrochemical reactions, etc., which usually result in fast fading in electrochemical performance, poor structural stability and cycling performance, and slow activation process. To address these issues, cobalt oxide-based hybrids with unique merits of each individual component and synergetic effects can
Fig. 57. (a) Schematic illustration of the fabrication of hierarchically Co3O4@X (X = Co3O4, CoS, C, and CoP) structures [769], (b) LSV curves of the Co3O4@X structures [769] (Copyright 2017, Wiley-VCH). 656
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
effectively improve the electronic conductivity, enhance the reaction kinetics, and buffer the volume changes during repeated charging and discharging processes. Compared with the pristine cobalt oxide nanostructures, these hybrids exhibit improved rate capability and cycling stability, decreased overpotential, and overall promoted electrochemical properties for energy-related devices. Despite the fact that much progress has been achieved on the investigation of various types of cobalt oxide-based composites, some challenges still exist in their practical electrochemical energy applications In this section, we will give a detailed analysis on the current challenges of the development of cobalt oxide-based materials for electrochemical energy applications. Then, some effective strategies for enhancing the electrochemical properties of cobalt oxidebased nanomaterials will be summarized. To make a good understanding, we also emphasis the structure-property relationships between the cobalt oxide-based structures and the electrochemical properties in energy applications. Finally, an outlook will be given to offer some reference to the further research on this topic. 5.1. Current challenges In spite of applausive progress on this amazing family of materials has been achieved in recent years, some challenges still need to be addressed to meet the high-performance electrochemical applications. As summarized in this review, cobalt oxide-based nanomaterials share some common challenges to other metal oxide-based materials in electrochemical energy applications, such as that the serious agglomeration of the low-dimensional cobalt oxide nanostructures will lead to the loss of effective active sites and accelerate the pulverization or crack of electrode during cycling processes. Below are some typical challenges in the application of cobalt oxide nanomaterials. The first challenge is from the synthesis of the property-on-demand cobalt oxide nanostructures. Even though lots of work on the synthesis and fabrication of cobalt oxide nanostructures have been reported, it is still a grand challenge to achieve a balance between the homogeneity and quality of the morphology/size and the productivity of the nanomaterials. Moreover, the design of some complex nanostructures possessing some extraordinary properties to achieve higher performances, such as 3D multiscale-ordered nanostructures, 3D hierarchical networks and porous structures, highly-dispersed 3D nanocomposites and nanohybrids, and the materials with unusual crystal structures or exposed facets, always brings extra difficulties in materials synthesis. Usually, the finer the nanostructures, the smaller the productivity is. Therefore, the large-scale production of property-on-demand nanostructures with highly-controlled morphology and narrow-size distribution is the next target for possible commercial application of this promising class of materials. The second challenge is from the optimization of cobalt oxide-based nanocomposites or nanohybrids. As reviewed in this paper, a wide range of cobalt oxide-based composites have been reported to date, however, it is still difficult to find an optimum design, including dimensionality match, component ratios, and configurations, etc., for this type of materials. First, as for the dimensionality match, cobalt oxide nanostructures with different dimensionalities (0D, 1D, 2D, and 3D) can be hybridized with the other heterogeneous components with different dimensional structures as well (0D, 1D, 2D, and 3D). There are theoretically sixteen types of combinations for binary cobalt oxide-based hybrids, and even more combined forms for ternary hybrids. Based on the recent progress, it is obviously found that each type has their own advantages and disadvantages. For example, (i) 0D/0D or 1D/1D heterogeneous hybridization can produce good interfacial contacts, however, the self-aggregation issue is the most serious; (ii) 2D/2D heterogeneous hybridization have the best geometric compatibility and can effective prevent the self-aggregation of 2D nanosheets, however, the precise control on the layer-by-layer assembly is relatively difficult; (iii) 0D/2D, 1D/2D, or 3D/2D heterogeneous hybridization can achieve an homogeneous dispersion of different dimensional nanostructures on 2D conductive substrates, however, the binding force between these nanostructures and 2D sheet-like substrates is often weak. Second, in terms of components, quite a few selections (e.g. carbon, CNTs, graphene, metal oxides, polymers, metal NPs, etc.) are available to hybridize with cobalt oxide nanostructures. People have been persistently trying to find whether there is the most suitable heterogeneous component to achieve the most promising cobalt oxide-based material for electrochemical applications, but so far no great progress is approached. Finally, the formation of a suitable configuration of the nanocomposites to give suitable mass transport channels and exposed active sites, such as core-shell, layer-on-layer, and hierarchical porous structures, is also recognized as another grand challenge. The currently designed configurations for cobalt oxide-based nanomaterials all present their own pros and cons. For example, the coating of carbon species onto cobalt oxides or the encapsulation of cobalt oxide into carbon matrices, which can effectively enhance the electrical conductivity and accommodate the volume expansion/extraction of cobalt oxide nanostructures, however, significantly resist the infiltration the wetting of electrolytes inside the cobalt oxide nanostructures. Therefore, more systematically investigations are needed to achieve an optimum design for cobalt oxide-based hybrids to meet the various requirements to achieve outstanding electrochemical properties for energy-related devices. The third challenge is from the side reactions in the complex practical conditions of the electrochemical energy devices. Much theoretical effort is still urgently needed to reveal the overall reaction pathways of cobalt oxides for ion storage and electrocatalysis. In the practical devices, the electrochemical reactions are much more complicated and can be affected by various factors, including the influences from the electrodes, electrolytes, and separators, and some side reactions. For example, the electrochemical reaction pathways in the electrolytes with different components, PH values, additives, etc. are strongly varied. It is therefore highly recommended that the electrochemical performances of the materials for practical applications should be evaluated in entire devices instead of those exanimated in half cells with sole working electrode in laboratory. Furthermore, the formation of cobalt oxide-based hybrids will make the reactions more complex. Further studies focusing on the actual mechanisms in practical devices are highly demanded to get a clear idea on some specific contributions from the constituent components and the interfaces of the electrochemical energy devices. 657
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
Besides the above major challenges, there are some other concerns in regards of the practical application of this family of materials, such as abundance, toxicity, and possible environment issues. The abundance of cobalt in the crust ranks around 32nd among all elements, which is very close to lithium, yttrium, neodymium, etc. The overall reservation of cobalt is therefore actually not too rare in the Earth's crust. The distribution, however, is quite uneven. Over half of the world’s cobalt reserves are in Democratic Republic of Congo, which greatly impede the development of cobalt-related industries in other countries. In recent years, driven by the increasing demands in electrochemical energy devices, the price of cobalt in the international metal markets maintains an upward trend. The introduction of cheap metals or metal oxides to form hybrid nanostructures with even superior electrochemical performance should be an effective way to reduce the consumption of cobalt. On the other hands, cobalt oxide and their composites have some negative effects on humans and environment. IARC has classified cobalt and cobalt compounds as Group 2B carcinogens: possibly carcinogenic to humans. ACGIH has classified cobalt and cobalt compounds as Group A3 carcinogens: proven carcinogenic to animals. It has been reported that the maximum tolerated dose (MTD) of CoO is higher than 2000 mg kg−1, which can be classified as a low toxic chemical substance. A high concentration of cobalt oxides, however, has the risk to cause an oxidative stress in human cells, inflammatory response, and even DNA damage, as evidenced by biological toxicity tests. Therefore, during the utilization of cobalt oxide-based materials, the disposal and recycling of cobalt-containing waste should be considered properly to avoid any potential biological and environmental risks. 5.2. Strategies for improving electrochemical properties For addressing the encountered challenges of cobalt oxide nanomaterials for electrochemical energy applications, we intend to propose some effective strategies based on the recent publications. The first strategy is nanostructure engineering, by which at least one of the dimensionalities is less than 100 nm, to enlarge the exposed surfaces, increase the mass transport paths, and maximize the active sites. In this way, the active surface of the oxides, the contact interfaces in the devices, and the mass diffusion and charge/carrier transport via the thinnest dimension can be dramatically improved, and thus improve the performance of the materials. Via proper nanostructure design, it has been found that the volume changes, restacking or agglomerations, and structural stability can be accommodated during electrochemical applications, and thus prolong the lifetime and provide better cycling performances. Some high energy facets with superior chemical and catalytic activities that cannot be obtained on common bulk materials have also been achieved by nanostructure engineering, which are also a novel pathway to enhance the electrochemical performance of cobalt oxides. Moreover, nanostructure engineering increases the chemical and structural affinity of cobalt oxides to combine with other active materials to form nanocomposites or nanohybrids, which usually exhibited a synergetic effect and present superior performance to their physical mixtures. The second strategy is to address the biggest shortcoming of cobalt oxides, the low electrical conductivity, by incorporating with some highly conductive and/or chemical active components to form nanocomposites or nanohybrids. Via the hybridization with amorphous carbon, CNTs, graphene, metal NPs, conductive polymers, and heterogeneous doping, and multiple components. As demonstrated in the examples summarized in this review, the electrical properties of the cobalt oxide nanocomposites and nanohybrids presented significant improvement, which facilitate charge/carrier transport through the materials, and as a result, considerably enhance the specific reversible capacity/capacitance, rate capability, cycling stability, and catalytic activity in the electrochemical energy applications. The introduction of the secondary active materials has also the effect to separate some easily agglomerated cobalt oxide nanostructures and increase the dispersibility of both the oxide nanostructures and the active materials, and thus accommodate the volume variation and boost the structural stability during operation, which can dramatically contribute to the rate and cycling performances of the energy devices. The nanocomposites and nanohybrids have provided better catalytic activities for their increased active sites and internal stresses. Owing to that the interfacial coupling and the contact between cobalt oxides and the secondary active materials are very critical for the electrochemical performance of the nanocomposites, rational design of the morphology, the interfacial chemical bonds, and the functional groups on their surfaces is an important method to further enhance the performance of the materials. 5.3. Structure-property relationships of cobalt oxide nanomaterials Structure-property relationships are a central theme for material research. Different types of structures may bring about some distinctive properties, such as the differences between graphite and graphene. Hence, understanding on the structure-property relationships is critical to get better use of materials. In this review, it is concluded that some factors, such as crystal structure, dimensionality, and composition, performed significant impact on the electrochemical properties of cobalt oxides and cobalt oxidebased hybrid materials. In this section, a summary of three primary categories of structure-property relationships is given in terms of the effects of crystal structure, dimensionality, and composition of cobalt oxide-based nanomaterials on their electrochemical properties. 5.3.1. Relationship between crystal structure and electrochemical performances of cobalt oxides Co3O4 has been verified as a promising candidate as high-performance electrode material or active electrocatalysts for electrochemical energy applications. It is demonstrated that the appealing electrochemical properties are highly related to the unique crystal structure of Co3O4, which results in distinct mechanisms or pathways for electrochemical reactions. In a typical Co3O4 spinel structure, there are 64 tetrahedral sites, one-eighth (8a) of which are occupied by the Co2+ cations, and 32 octahedral sites, one-half (16d) of which are occupied by the Co3+ cations, accompanied by the cubic-close-packed O2− anions (32e). In the interstitial space, 658
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
there are 56 empty tetrahedral sites (8b and 48f) and 16 empty octahedral sites (16c). It should be noted that every tetrahedron shares faces with octahedron in the cubic close packing group. As electrode materials for rechargeable batteries, taking the common Li+ storage as example, there are two competing reaction pathways. Either the initial reduction of Co3O4 into CoO/Li2O composite intermediate or the formation of LixCo3O4 intermediate has been observed before the full decomposition of both intermediates into the final Co/Li2O composites upon further reduction, which is quite different from the storage mechanisms for other metal oxides, such as the insertion/deinsertion and the alloying mechanisms. As catalyst for electrochemical reactions, in spite of some controversial conclusions on the specific active sites of Co3O4 for different types of electrocatalytic reactions, the mixed-valence Co3O4 spinel has demonstrated superior electrochemical properties compared with other single-valence metal oxides. It has revealed that the catalytic activity of Co3O4 strongly depends on the exposed facets in an order of {1 0 0} < {1 1 0} < {1 1 2} < {1 1 1} and the {1 1 2} facets enclosed by Co3+ and Co2+ sites presented much superior activity to the {1 0 0} facets enclosed by only Co2+ sites for both ORR and OER. Moreover, it has been demonstrated that the electrochemical properties can be further enhanced by targeted metal substitution of Co2+ with other guest cations, such as Ni and Mn. 5.3.2. Relationship of dimensionality-electrochemical properties of cobalt oxides Based on the review on cobalt oxides with different dimensionalities, it can be concluded that the dimensionality has distinct influences on their electrochemical properties for energy-related devices. A summary is given as follows on the pros and cons of different dimensionality on the electrochemical properties. (i) 0D nanostructures feature unique confinement effects in three directions, but they are easy to restack into large aggregates or bulks during the sample drying process and the electrode assembly step, which greatly reduces the effective active sites and compromises the merits of 0D nanostructures. (ii) 1D nanostructures, especially these arrays grown on the conductive substrates, can offer efficient transport pathways as binder-free electrode materials. The structural deformation, however, usually occurs easily during the cycling processes. (iii) The highly exposed surface area of 2D nanostructures makes them appealing for electrochemical reactions, whereas the strong restacking towards thick layered structures is a major barrier. Moreover, 2D nanosheets with large lateral size are not favorite for the diffusion and mass transport through their thickness dimension and lower the activity of the materials. (v) 3D structures with hierarchical and porous morphologies are attractive, due to the excellent structural stability, high surface area for the good contact with electrolytes, and outstanding buffering capacity for volume changes during the long-time cycles. Nevertheless, the synthetic procedures are always complex and the interfaces between the active materials and the current collectors need further optimization. 5.3.3. Relationship of composition-electrochemical properties of cobalt oxide-based hybrids Compared with cobalt oxide nanostructures, cobalt oxide-based nanocomposites with heterogeneous structure-complementary components have been verified as excellent electrode materials with enhanced electrochemical properties. By the introduction of different types of components, some issues for the pristine cobalt oxide nanostructures can be effective addressed. A summary on the advantages and disadvantages of each cobalt oxide-based category is given as follows: (i) cobalt oxide/carbon hybrids are the most widely studied hybrids. This type of hybridization is easily achievable for its versatile cobalt precursors and carbon sources. An optimum quantitative ratio of the cobalt oxides and the carbon species, however, has not confirmed yet. (ii) Cobalt oxide/CNTs hybrids, especially the common 0D cobalt oxides hybridized with 1D CNTs, can possess enhanced overall electrical conductivity and partly prevent the self-aggregation of cobalt oxide NPs. However, the chemically inert feature of the p-CNTs is a major barrier for a uniform dispersion of cobalt oxide NPs on the CNTs. (iii) Cobalt oxide/graphene hybrids have received a great deal of attention in recent years, and the heteroatom-doped graphene has been confirmed as an excellent conductive candidate to improve the electrochemical properties of cobalt oxide nanostructures with different dimensionalities. However, the current high-cost preparation technique for graphene will greatly hinder the practical application of cobalt oxide/graphene hybrids. (iv) Cobalt oxide/metal hybrids can take the full use of the remarkable electrochemical activities of metal NPs. The inferior cycling stability and the possible use of noble metals, unfortunately, make this type of hybrids less attractive. (v) Cobalt oxide/metal oxides heterogeneous hybrids usually exhibit some synergetic effects. This type of hybrids, however, share the common disadvantage of low electrical conductivity of metal oxides. (vi) Cobalt oxide/conductive polymers hybrids can bring forward more active redox reactions and good conductive networks. The mechanical degradation of polymers is always an issue and often leads to a quick decay of electrochemical performances upon repeated cycling. 5.4. Future outlooks Great progress has been achieved for the design and synthesis of cobalt oxides as well as their hybrids for electrochemical energy devices, especially for rechargeable batteries, supercapacitors, and electrocatalysis. In spite of some major challenges, this family of metal oxides is still unbeatable for the applications in electrochemical energy devices for the salient chemical and physical properties. It is believed that, with further study on the underlying electrochemical reaction mechanisms, some effective solutions to address the current challenges will be found and the way towards practical application of this family of materials will be paved together with the innovation of novel electrochemical techniques and advanced energy devices. Meanwhile, theoretical calculations and simulations will also assist the design of novel cobalt oxide-based materials to meet the practical requirements for electrochemical applications. The reveal of the structure-property relationships of cobalt oxide-based materials for electrochemical applications will significantly promote the development of material science, nanoscience and nanotechnology, and energy-related technology. 659
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
Acknowledgements This work was partly supported by an Australian Research Council (ARC) Future Fellowship project (FT180100387) for ZS, an ARC Future Fellowship project (FT160100281) for TL, and an ARC Discovery Project (DP160102627). References [1] [2] [3] [4] [5] [6] [7] [8] [9] [10] [11] [12] [13] [14] [15] [16] [17] [18] [19] [20] [21] [22] [23] [24] [25] [26] [27] [28] [29] [30] [31] [32] [33] [34] [35] [36] [37] [38] [39] [40] [41] [42] [43] [44] [45] [46] [47]
Gulino A, Dapporto P, Rossi P, Fragala I. A novel self-generating liquid MOCVD precursor for Co3O4 thin films. Chem Mater 2003;15:3748–52. Choi J, Ch Y. Study of the nonstoichiometric compositions of cobaltous oxide. Inorg Chem 1974;13:1720–4. Chen M, Hallstedt B, Gauckler LJ. Thermodynamic assessment of the Co-O system. J Phase Equilibr 2003;24:212–27. Barakat NAM, Khil MS, Sheikh FA, Kim HY. Synthesis and optical properties of two cobalt oxies (CoO and Co3O4) nanofibers produced by electrospinning process. J Phys Chem C 2008;112:12225–33. Kuboon S, Hu YH. Study of NiO−CoO and Co3O4−Ni3O4 solid solutions in multiphase Ni−Co−O systems. Ind Eng Chem Res 2011;50:2015–20. Varghese B, Hoong TC, Yanwu Z, Reddy MV, Chowdari BVR, Wee ATS, et al. CO3O4 nanostructures with different morphologies and their field-emission properties. Adv Funct Mater 2007;17:1932–9. Wang X, Tian W, Zhai T, Zhi C, Bando Y, Golberg D. Cobalt(II, III) oxide hollow structures: fabrication, properties and applications. J Mater Chem 2012;22:23310–26. Smith WL, Hobson AD. The structure of cobalt oxide, Co3O4. Acta Crystallogr Sect B Struct Crystallogr Cryst Chem 1973;29:362–3. He Q, Cairns EJ. Review—recent progress in electrocatalysts for oxygen reduction suitable for alkaline anion exchange membrane fuel cells. J Electrochem Soc 2015;162:F1504–39. Lee J, Jeong B, Ocon JD. Oxygen electrocatalysis in chemical energy conversion and storage technologies. Curr Appl Phys 2013;13:309–21. Raj CR, Samanta A, Noh SH, Mondal S, Okajima T, Ohsaka T. Emerging new generation electrocatalysts for the oxygen reduction reaction. J Mater Chem A 2016;4:11156–78. Tahir M, Pan L, Idrees F, Zhang X, Wang L, Zou JJ, et al. Electrocatalytic oxygen evolution reaction for energy conversion and storage: A comprehensive review. Nano Energy 2017;37:136–57. Jiao Y, Zheng Y, Jaroniec M, Qiao SZ. Design of electrocatalysts for oxygen- and hydrogen-involving energy conversion reactions. Chem Soc Rev 2015;44:2060–86. Suen N-T, Hung S-F, Quan Q, Zhang N, Xu Y-J, Chen HM. Electrocatalysis for the oxygen evolution reaction: recent development and future perspectives. Chem Soc Rev 2017;46:337–65. Reier T, Nong HN, Teschner D, Schlögl R, Strasser P. Electrocatalytic oxygen evolution reaction in acidic environments – Reaction mechanisms and catalysts. Adv Energy Mater 2017;7:1601275. Hong WT, Risch M, Stoerzinger KA, Grimaud A, Suntivich J, Shao-Horn Y. Toward the rational design of non-precious transition metal oxides for oxygen electrocatalysis. Energy Environ Sci 2015;8:1404–27. Zhu YP, Guo C, Zheng Y, Qiao S-Z. Surface and interface engineering of noble-metal-free electrocatalysts for efficient energy conversion processes. Acc Chem Res 2017;50:915–23. Zhang L, Xiao J, Wang H, Shao M. Carbon-based electrocatalysts for hydrogen and oxygen evolution reactions. ACS Catal 2017;7:7855–65. Shao M, Chang Q, Dodelet JP, Chenitz R. Recent advances in electrocatalysts for oxygen reduction reaction. Chem Rev 2016;116:3594–657. Wu G, Zelenay P. Nanostructured nonprecious metal catalysts for oxygen reduction reaction. Acc Chem Res 2013;46:1878–89. Zhu C, Dong S. Recent progress in graphene-based nanomaterials as advanced electrocatalysts towards oxygen reduction reaction. Nanoscale 2013;5:1753–67. Poizot P, Laruelle S, Grugeon S, Dupont L, Tarascon J-M. Nano-sized transition-metal oxides as negative-electrode materials for lithium-ion batteries. Nature 2000;407:496–9. Larcher D, Sudant G, Leriche J, Chabre Y, Tarascon J. The electrochemical reduction of Co3O4 in a lithium cell. J Electrochem Soc 2002;149:A234–41. Rakhi RB, Chen W, Cha D, Alshareef HN. Substrate dependent self-organization of mesoporous cobalt oxide nanowires with remarkable pseudocapacitance. Nano Lett 2012;12:2559–67. Wang Z-L, Xu D, Xu J-J, Zhang X-B. Oxygen electrocatalysts in metal–air batteries: from aqueous to nonaqueous electrolytes. Chem Soc Rev 2014;43:7746–86. Qiu HJ, Liu L, Mu YP, Zhang HJ, Wang Y. Designed synthesis of cobalt-oxide-based nanomaterials for superior electrochemical energy storage devices. Nano Res 2015;8:321–39. Pettersson J, Ramsey B, Harrison D. A review of the latest developments in electrodes for unitised regenerative polymer electrolyte fuel cells. J Power Sources 2006;157:28–34. Liang Y, Li Y, Wang H, Zhou J, Wang J, Regier T, et al. Co3O4 nanocrystals on graphene as a synergistic catalyst for oxygen reduction reaction. Nat Mater 2011;10:780–6. Sun H, Ang HM, Tadé MO, Wang S. Co3O4 nanocrystals with predominantly exposed facets: synthesis, environmental and energy applications. J Mater Chem A 2013;1:14427–42. Zhang W, Zeng Y, Xiao N, Hng HH, Yan Q. One-step electrochemical preparation of graphene-based heterostructures for Li storage. J Mater Chem 2012;22:8455–61. Xie X, Shen W. Morphology control of cobalt oxide nanocrystals for promoting their catalytic performance. Nanoscale 2009;1:50–60. Grassian VH. When size really matters: size-dependent properties and surface chemistry of metal and metal oxide nanoparticles in gas and liquid Phase environments. J Phys Chem C 2008;112:18303–13. Mei J, Zhang Y, Liao T, Sun Z, Dou SX. Strategies for improving the lithium-storage performance of 2D nanomaterials. Natl Sci Rev 2018;5:389–416. Tarascon JM, Armand M. Issues and challenges facing rechargeable lithium batteries. Nature 2001;414:359–67. Armand M, Tarascon J-M. Building better batteries. Nature 2008;451:652–7. Winter M, Brodd RJ. What are batteries, fuel cells, and supercapacitors? Chem Rev 2004;104:4245–69. Goodenough JB, Park K. The Li-ion rechargeable battery: a perspective. J Am Chem Soc 2013;135:1167–76. Zhou G, Li F, Cheng H-M. Progress in flexible lithium batteries and future prospects. Energy Environ Sci 2014;7:1307–38. Cabana J, Monconduit L, Larcher D, Palacín MR. Beyond intercalation-based Li-ion batteries: the state of the art and challenges of electrode materials reacting through conversion reactions. Adv Mater 2010;22:E170–92. Reddy MV, Subba Rao GV, Chowdari BV. Metal oxides and oxysalts as anode materials for Li ion batteries. Chem Rev 2013;113:5364–457. Wang X, Weng Q, Yang Y, Bando Y, Golberg D. Hybrid two-dimensional materials in rechargeable battery applications and their microscopic mechanisms. Chem Soc Rev 2016;45:4042–73. Zhao Y, Li X, Yan B, Xiong D, Li D, Lawes S, et al. Recent developments and understanding of novel mixed transition-metal oxides as anodes in lithium ion batteries. Adv Energy Mater 2016;6:1502175. Cheng F, Tao Z, Liang J, Chen J. Template-directed materials for rechargeable lithium-ion batteries. Chem Mater 2008;20:667–81. Sun Y, Liu N, Cui Y. Promises and challenges of nanomaterials for lithium-based rechargeable batteries. Nat Energy 2016;1:16071. Chen G, Yan L, Luo H, Guo S. Nanoscale engineering of heterostructured anode materials for boosting lithium-ion storage. Adv Mater 2016;28:7580–602. Wang Y, Wang B, Xiao F, Huang Z, Wang Y, Richardson C, et al. Facile synthesis of nanocage Co3O4 for advanced lithium-ion batteries. J Power Sources 2015;298:203–8. Liu HC, Yen SK. Characterization of electrolytic Co3O4 thin films as anodes for lithium-ion batteries. J Power Sources 2007;166:478–84.
660
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
[48] Do JS, Weng CH. Preparation and characterization of CoO used as anodic material of lithium battery. J Power Sources 2005;146:482–6. [49] Zhang XX, Xie QS, Yue GH, Zhang Y, Zhang XQ, Lu AL, et al. A novel hierarchical network-like Co3O4 anode material for lithium batteries. Electrochim Acta 2013;111:746–54. [50] Wang Y, Zhang HJ, Wei J, Wong CC, Lin J, Borgna A. Crystal-match guided formation of single-crystal tricobalt tetraoxygen nanomesh as superior anode for electrochemical energy storage. Energy Environ Sci 2011;4:1845–54. [51] Lou X. Growth of ultrathin mesoporous Co3O4 nanosheet arrays on Ni foam for high-performance electrochemical capacitors. Energy Environ Sci 2012;5:7883–7. [52] Yao W, Dai Q, Liu Y, Zhang Q, Zhong S, Yan Z. Microwave-assisted synthesis of Co3O4 sheets for reversible Li-storage: regulation of structure and performance. ChemElectroChem 2017;4:1236–42. [53] Wen W, Wu J-M, Cao M-H. Facile synthesis of a mesoporous Co3O4 network for Li-storage via thermal decomposition of an amorphous metal complex. Nanoscale 2014;6:12476–81. [54] Bian S-W, Zhu L. Template-free synthesis of mesoporous Co3O4 with controlled morphologies for lithium ion batteries. RSC Adv 2013;3:4212–5. [55] Liu D, Wang X, Wang X, Tian W, Bando Y, Golberg D. Co3O4 nanocages with highly exposed 110 facets for high-performance lithium storage. Sci Rep 2013;3:2543. [56] Zhao G, Xu Z, Zhang L, Sun K. Hierarchical porous Co3O4 films with size-adjustable pores as Li ion battery anodes with excellent rate performances. Electrochim Acta 2013;114:251–8. [57] Wen W, Wu J-M, Tu J-P. A novel solution combustion synthesis of cobalt oxide nanoparticles as negative-electrode materials for lithium ion batteries. J Alloys Compd 2012;513:592–6. [58] Reddy MV, Beichen Z, Loh KP, Chowdari BVR. Facile synthesis of Co3O4 by molten salt method and its Li-storage performance. CrystEngComm 2013;15:3568–74. [59] Wang D, Yu Y, He H, Wang J, Zhou W, Abruna H. Template-free synthesis of hollow-structured Co3O4 nanoparticles as high-performance anodes for lithium-ion batteries. ACS Nano 2015;9:1775–81. [60] Xiao Y, Hu C, Cao M. High lithium storage capacity and rate capability achieved by mesoporous Co3O4 hierarchical nanobundles. J Power Sources 2014;247:49–56. [61] Do JS, Weng CH. Electrochemical and charge/discharge properties of the synthesized cobalt oxide as anode material in Li-ion batteries. J Power Sources 2006;159:323–7. [62] Binotto G, Larcher D, Prakash AS, Urbina RH, Hegde MS, Tarascon J-M. Synthesis, characterization, and Li-electrochemical performance of highly porous Co3O4 powders. Chem Mater 2007;19:3032–40. [63] Shaju KM, Jiao F, Débart A, Bruce PG. Mesoporous and nanowire Co3O4 as negative electrodes for rechargeable lithium batteries. Phys Chem Chem Phys 2007;9:1837–42. [64] Kang YM, Song MS, Kim JH, Kim HS, Park MS, Lee JY, et al. A study on the charge-discharge mechanism of Co3O4 as an anode for the Li ion secondary battery. Electrochim Acta 2005;50:3667–73. [65] Wang GX, Chen Y, Konstantinov K, Lindsay M, Liu HK, Dou SX. Investigation of cobalt oxides as anode materials for Li-ion batteries. J Power Sources 2002;109:142–7. [66] Wang GX, Chen Y, Konstantinov K, Yao J, Ahn JH, Liu HK, et al. Nanosize cobalt oxides as anode materials for lithium-ion batteries. J Alloys Compd 2002;340:L5–10. [67] Mei J, Liao T, Sun Z. Two-dimensional metal oxide nanosheets for rechargeable batteries. J Energy Chem 2018;27:117–27. [68] Goodenough JB. Lithium insertion into Co3O4: a preliminary investigation. Solid State Ionics 1985;17:175–81. [69] Kim Y, Lee JH, Cho S, Kwon Y, In I, Lee J, et al. Additive-free hollow-structured Co3O4 nanoparticle Li-ion battery: the origins of irreversible capacity loss. ACS Nano 2014;8:6701–12. [70] Hwang J-Y, Myung S-T, Sun Y-K. Sodium-ion batteries: present and future. Chem Soc Rev 2017;46:3529–614. [71] Chen S, Wu C, Shen L, Zhu C, Huang Y, Xi K, et al. Challenges and perspectives for NASICON-type electrode materials for advanced sodium-ion batteries. Adv Mater 2017;1700431. [72] Wang LP, Yu L, Wang X, Srinivasan M, Xu ZJ. Recent developments in electrode materials for sodium-ion batteries. J Mater Chem A 2015;3:9353–78. [73] Kundu D, Talaie E, Duffort V, Nazar LF. The emerging chemistry of sodium ion batteries for electrochemical energy storage. Angew Chem Int Edit 2015;54:3431–48. [74] Han MH, Gonzalo E, Singh G, Rojo T. A comprehensive review of sodium layered oxides: powerful cathodes for Na-ion batteries. Energy Environ Sci 2015;8:81–102. [75] Pan H, Hu Y-S, Chen L. Room-temperature stationary sodium-ion batteries for large-scale electric energy storage. Energy Environ Sci 2013;6:2338–60. [76] Sawicki M, Shaw LL. Advances and challenges of sodium ion batteries as post lithium ion batteries. RSC Adv 2015;5:53129–54. [77] Balogun MS, Luo Y, Qiu W, Liu P, Tong Y. A review of carbon materials and their composites with alloy metals for sodium ion battery anodes. Carbon 2016;98:162–78. [78] Yabuuchi N, Kubota K, Dahbi M, Komaba S. Research development on sodium-ion batteries. Chem Rev 2014;114:11636–82. [79] Nithya C, Gopukumar S. Sodium ion batteries: a newer electrochemical storage. Wiley Interdiscip Rev Energy Environ 2015;4:253–78. [80] Ponrouch A, Monti D, Boschin A, Steen B, Johansson P, Palacín MR. Non-aqueous electrolytes for sodium-ion batteries. J Mater Chem A 2015;3:22–42. [81] Xiang X, Zhang K, Chen J. Recent advances and prospects of cathode materials for sodium-ion batteries. Adv Mater 2015;27:5343–64. [82] Ortiz-Vitoriano N, Drewett NE, Gonzalo E, Rojo T. High performance manganese-based layered oxide cathodes: overcoming the challenges of sodium ion batteries. Energy Environ Sci 2017;10:1051–74. [83] Luo W, Shen F, Bommier C, Zhu H, Ji X, Hu L. Na-ion battery anodes: materials and electrochemistry. Acc Chem Res 2016;49:231–40. [84] Cheng F, Liang J, Tao Z, Chen J. Functional materials for rechargeable batteries. Adv Mater 2011;23:1695–715. [85] Guo S, Yi J, Sun Y, Zhou H. Recent advances in titanium-based electrode materials for stationary sodium-ion batteries. Energy Environ Sci 2016;9:2978–3006. [86] Fang C, Huang Y, Zhang W, Han J, Deng Z, Cao Y, et al. Routes to high energy cathodes of sodium-ion batteries. Adv Energy Mater 2016;6:1501727. [87] Xiao Y, Lee SH, Sun Y-K. The application of metal sulfides in sodium ion batteries. Adv Energy Mater 2017;7:1601329. [88] Meng X. Atomic-scale surface modifications and novel electrode designs for high-performance sodium-ion batteries via atomic layer deposition. J Mater Chem A 2017;5:10127–49. [89] Slater MD, Kim D, Lee E, Johnson CS. Sodium-ion batteries. Adv Funct Mater 2013;23:947–58. [90] Che H, Chen S, Xie Y, Wang H, Amine K, Liao X-ZX, et al. Electrolyte design strategies and research progress for room-temperature sodium-ion batteries. Energy Environ Sci 2017;10:1075–101. [91] Zhu Y, Choi SH, Fan X, Shin J, Ma Z, Zachariah MR, et al. Recent progress on spray pyrolysis for high performance electrode materials in lithium and sodium rechargeable batteries. Adv Energy Mater 2017;7:1601578. [92] Ren W, Zhu Z, An Q, Mai L. Emerging prototype sodium-ion full cells with nanostructured electrode materials. Small 2017;13:1604181. [93] Kim SW, Seo DH, Ma X, Ceder G, Kang K. Electrode materials for rechargeable sodium-ion batteries: potential alternatives to current lithium-ion batteries. Adv Energy Mater 2012;2:710–21. [94] Kang W, Wang Y, Xu J. Recent progress of layered metal dichalcogenide nanostructures as electrodes for high-performance sodium-ion batteries. J Mater Chem A 2017;5:7667–90. [95] Palomares V, Serras P, Villaluenga I, Hueso KB, Carretero-Gonzalez J, Rojo T. Na-ion batteries, recent advances and present challenges to become low cost energy storage systems. Energy Environ Sci 2012;5:5884–901. [96] Alcántara R, Jaraba M, Lavela P, Tirado JL. NiCo2O4 spinel: first report on a transition metal oxide for the negative electrode of sodium-ion batteries. Chem Mater 2002;14:2847–8.
661
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
[97] Chadwick AV, Savin SLP, Fiddy S, Alcántara R, Lisbona DF, Lavela P, et al. Formation and oxidation of nanosized metal particles by electrochemical reaction of Li and Na with NiCo2O4: X-ray absorption spectroscopic study. J Phys Chem C 2007;111:4636–42. [98] Dou Y, Wang Y, Tian D, Xu J, Zhang Z, Liu Q, et al. Atomically thin Co3O4 nanosheet-coated stainless steel mesh with enhanced capacitive Na+ storage for highperformance sodium-ion batteries. 2D Mater 2017;4. [99] Wen JW, Zhang DW, Zang Y, Sun X, Cheng B, Ding CX, et al. Li and Na storage behavior of bowl-like hollow Co3O4 microspheres as an anode material for lithium-ion and sodium-ion batteries. Electrochim Acta 2014;132:193–9. [100] Wu Y, Meng J, Li Q, Niu C, Wang X, Yang W, et al. Interface-modulated fabrication of hierarchical yolk–shell Co3O4/C dodecahedrons as stable anodes for lithium and sodium storage. Nano Res 2017;10:2364–76. [101] Longoni G, Fiore M, Kim JH, Jung YH, Kim DK, Mari CM, et al. Co3O4 negative electrode material for rechargeable sodium ion batteries: an investigation of conversion reaction mechanism and morphology-performances correlations. J Power Sources 2016;332:42–50. [102] Zhou Y, Sun W, Rui X, Zhou Y, Ng WJ, Yan Q, et al. Biochemistry-derived porous carbon-encapsulated metal oxide nanocrystals for enhanced sodium storage. Nano Energy 2016;21:71–9. [103] Rahman MM, Glushenkov AM, Ramireddy T, Chen Y. Electrochemical investigation of sodium reactivity with nanostructured Co3O4 for sodium-ion batteries. Chem Commun 2014;50:5057–60. [104] Li Q, Wu J, Xu J, Dravid VP. Synergistic sodiation of cobalt oxide nanoparticles and conductive carbon nanotubes (CNTs) for sodium-ion batteries. J Mater Chem A 2016;4:8669–75. [105] Liu Y, Cheng Z, Sun H, Arandiyan H, Li J, Ahmad M. Mesoporous Co3O4 sheets/3D graphene networks nanohybrids for high-performance sodium-ion battery anode. J Power Sources 2015;273:878–84. [106] Yang J, Zhou T, Zhu R, Chen X, Guo Z, Fan J, et al. Highly ordered dual porosity mesoporous cobalt oxide for sodium-ion batteries. Adv Mater Interfaces 2016;3:1500464. [107] Xu GL, Sheng T, Chong L, Ma T, Sun CJ, Zuo X, et al. Insights into the distinct lithiation/sodiation of porous cobalt oxide by in operando synchrotron X-ray techniques and Ab initio molecular dynamics simulations. Nano Lett 2017;17:953–62. [108] Peng Z, Freunberger SA, Chen Y, Bruce PG. A reversible and higher-rate Li-O2 battery. Science 2012;337:563–6. [109] Ottakam Thotiyl MM, Freunberger SA, Peng Z, Chen Y, Liu Z, Bruce PG. A stable cathode for the aprotic Li-O2 battery. Nat Mater 2013;12:1050–6. [110] Chen Y, Freunberger SA, Peng Z, Fontaine O, Bruce PG. Charging a Li-O2 battery using a redox mediator. Nat Chem 2013;5:489–94. [111] Luntz AC, Mccloskey BD, Dupré N, Cuisinier M, Guyomard D. Nonaqueous Li-air batteries: a status report. Chem Rev 2014;114:11721–50. [112] Bruce PG, Freunberger SA, Hardwick LJ, Tarascon JM. Li–O2 and Li–S batteries with high energy storage. Nat Mater 2012;11:19–29. [113] Lu J, Li L, Park J-B, Sun Y-K, Wu F, Amine K. Aprotic and aqueous Li-O2 batteries. Chem Rev 2014;114:5611–40. [114] Black R, Adams B, Nazar LF. Non-aqueous and hybrid Li-O2 batteries. Adv Energy Mater 2012;2:801–15. [115] Girishkumar G, McCloskey B, Luntz AC, Swanson S, Wilcke W. Lithium-air battery: promise and challenges. J Phys Chem Lett 2010;1:2193–203. [116] Christensen J, Albertus P, Sanchez-Carrera RS, Lohmann T, Kozinsky B, Liedtke R, et al. A critical review of Li/air batteries. J Electrochem Soc 2012;159:R1–30. [117] Kraytsberg A, Ein-Eli Y. Review on Li-air batteries – Opportunities, limitations and perspective. J Power Sources 2011;196:886–93. [118] Shao Y, Ding F, Xiao J, Zhang J, Xu W, Park S, et al. Making Li-air batteries rechargeable: material challenges. Adv Funct Mater 2013;23:987–1004. [119] Xie X, Li Y, Liu Z, Haruta M, Shen W. Low-temperature oxidation of CO catalysed by Co3O4 nanorods. Nature 2009;458:746–9. [120] Débart A, Bao J, Armstrong G, Bruce PG. An O2 cathode for rechargeable lithium batteries: the effect of a catalyst. J Power Sources 2007;174:1177–82. [121] Gao R, Li Z, Zhang X, Zhang J, Hu Z, Liu X. Carbon-dotted defective CoO with oxygen vacancies: a synergetic design of bifunctional cathode catalyst for Li-O2 batteries. ACS Catal 2016;6:400–6. [122] Shang C, Dong S, Hu P, Guan J, Xiao D, Chen X, et al. Compatible interface design of CoO-based Li-O2 battery cathodes with long-cycling stability. Sci Rep 2015;5:8335. [123] Yao KPC, Risch M, Sayed SY, Lee YL, Harding JR, Grimaud A, et al. Solid-state activation of Li2O2 oxidation kinetics and implications for Li-O2 batteries. Energy Environ Sci 2015;8:2417–26. [124] Zeng J, Francia C, Amici J, Bodoardo S, Penazzi N. Mesoporous Co3O4 nanocrystals as an effective electro-catalyst for highly reversible Li-O2 batteries. J Power Sources 2014;272:1003–9. [125] Liu Q, Jiang Y, Xu J, Xu D, Chang Z, Yin Y, et al. Hierarchical Co3O4 porous nanowires as an efficient bifunctional cathode catalyst for long life Li-O2 batteries. Nano Res 2015;8:576–83. [126] Shen C, Wen ZY, Wang F, Huang X, Rui K, Wu XW. Reduced free-standing Co3O4@Ni cathode for lithium-oxygen batteries with enhanced electrochemical performance. RSC Adv 2016;6:16263–7. [127] Cui Y, Wen Z, Liu Y. A free-standing-type design for cathodes of rechargeable Li-O2 batteries. Energy Environ Sci 2011;4:4727–34. [128] Zhao G, Xu Z, Sun K. Hierarchical porous Co3O4 films as cathode catalysts of rechargeable Li-O2 batteries. J Mater Chem A 2013;1:12862–7. [129] Wang F, Wen Z, Shen C, Rui K, Wu X, Chen C. Open mesoporous spherical shell structured Co3O4 with highly efficient catalytic performance in Li-O2 batteries. J Mater Chem A 2015;3:7600–6. [130] Lee CK, Park YJ. Carbon and binder-free air electrodes composed of Co3O4 nanofibers for Li-air batteries with enhanced cyclic performance. Nanoscale Res Lett 2015;10:319. [131] Wang S, Sha Y, Zhu Y, Xu X, Shao Z. Modified template synthesis and electrochemical performance of a Co3O4/mesoporous cathode for lithium-oxygen batteries. J Mater Chem A 2015;3:16132–41. [132] Kim DS, Park YJ. Ketjen black/Co3O4 nanocomposite prepared using polydopamine pre-coating layer as a reaction agent: effective catalyst for air electrodes of Li/air batteries. J Alloys Compd 2013;575:319–25. [133] Zhang J, Wang L, Xu L, Ge X, Zhao X, Lai M, et al. Porous cobalt-manganese oxide nanocubes derived from metal organic frameworks as a cathode catalyst for rechargeable Li-O2 batteries. Nanoscale 2015;7:720–6. [134] Cao C, Yan Y, Zhang H, Xie J, Zhang S, Pan B, et al. Controlled growth of Li2O2 by cocatalysis of mobile Pd and Co3O4 nanowire arrays for high-performance LiO2 batteries. ACS Appl Mater Interfaces 2016;8:31653–60. [135] Zhang Z, Chen Y, Bao J, Xie Z, Wei J, Zhou Z. Co3O4 hollow nanoparticles and Co organic complexes highly dispersed on N-doped graphene: an efficient cathode catalyst for Li-O2 batteries. Part Part Syst Charact 2015;32:680–5. [136] Yuan M, Lin L, Yang Y, Nan C, Ma S, Sun G, et al. In-situ growth of ultrathin cobalt monoxide nanocrystals on reduced graphene oxide substrates: an efficient electrocatalyst for aprotic Li-O2 batteries. Nanotechnology 2017;28. [137] Khajehbashi SMB, Li J, Wang M, Xu L, Zhao K, Wei Q, et al. A crystalline/amorphous cobalt(II, III) oxide hybrid electrocatalyst for lithium-air batteries. Energy Technol 2017;5:568–79. [138] Zeng J, Francia C, Amici J, Bodoardo S, Penazzi N. A highly reversible Li-O2 battery utilizing a mixed electrolyte and a cathode incorporating Co3O4. RSC Adv 2015;5:83056–64. [139] Liu Q, Xu J, Chang Z, Zhang X. Direct electrodeposition of cobalt oxide nanosheets on carbon paper as free-standing cathode for Li-O2 battery. J Mater Chem A 2014;2:6081–5. [140] Zhu J, Ren X, Liu J, Zhang W, Wen Z. Unraveling the catalytic mechanism of Co3O4 for the oxygen evolution reaction in a Li-O2 battery. ACS Catal 2015;5:73–81. [141] Zhu J, Wang F, Wang B, Wang Y, Liu J, Zhang W, et al. Surface acidity as descriptor of catalytic activity for oxygen evolution reaction in Li-O2 battery. J Am Chem Soc 2015;137:13572–9. [142] Radin MD, Monroe CW, Siegel DJ. How dopants can enhance charge transport in Li2O2. Chem Mater 2015;27:839–47. [143] Hu L, Peng Q, Li Y. Selective synthesis of Co3O4 nanocrystal with different shape and crystal plane effect on catalytic property for methane combustion. J Am Chem Soc 2008;130:16136–7. [144] Su D, Dou S, Wang G. Single crystalline Co3O4 nanocrystals exposed with different crystal planes for Li-O2 batteries. Sci Rep 2014;4:5767.
662
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
[145] Song K, Cho E, Kang YM. Morphology and active-Site engineering for stable round-trip efficiency Li-O2 batteries: a search for the most active catalytic site in Co3O4. ACS Catal 2015;5:5116–22. [146] Li Y, Gong M, Liang Y, Feng J, Kim J-E, Wang H, et al. Advanced zinc-air batteries based on high-performance hybrid electrocatalysts. Nat Commun 2013;4:1805. [147] Li L, Liu C, He G, Fan D, Manthiram A. Hierarchical pore-in-pore and wire-in-wire catalysts for rechargeable Zn– and Li–air batteries with ultra-long cycle life and high cell efficiency. Energy Environ Sci 2015;8:3274–82. [148] Lee DU, Choi JY, Feng K, Park HW, Chen Z. Advanced extremely durable 3D bifunctional air electrodes for rechargeable zinc-air batteries. Adv Energy Mater 2014;4:1301389. [149] Li B, Ge X, Goh FWT, Hor TSA, Geng D, Du G, et al. Co3O4 nanoparticles decorated carbon nanofiber mat as binder-free air-cathode for high performance rechargeable zinc-air batteries. Nanoscale 2015;7:1830–8. [150] Lin C, Shinde SS, Jiang Z, Song X, Sun Y, Guo L, et al. In situ directional formation of Co@CoOx-embedded 1D carbon nanotubes as an efficient oxygen electrocatalyst for ultra-high rate Zn–air batteries. J Mater Chem A 2017;5:13994–4002. [151] Hao Y, Xu Y, Han N, Liu J, Sun X. Boosting the bifunctional electrocatalytic oxygen activities of CoOx nanoarrays with a porous N-doped carbon coating and their application in Zn–air batteries. J Mater Chem A 2017;5:17804–10. [152] Song Z, Han X, Deng Y, Zhao N, Hu W, Zhong C. Clarifying the controversial catalytic performance of Co(OH)2 and Co3O4 for oxygen reduction/evolution reactions toward efficient Zn-air batteries. ACS Appl Mater Interfaces 2017;9:22694–703. [153] He Y, Zhang J, He G, Han X, Zheng X, Zhong C, et al. Ultrathin Co3O4 nanofilm as an efficient bifunctional catalyst for oxygen evolution and reduction reaction in rechargeable zinc-air batteries. Nanoscale 2017;9:8623–30. [154] Chang Z, Dou H, Ding B, Wang J, Wang Y, Hao X, et al. Co3O4 nanoneedle arrays as a multifunctional “super-reservoir” electrode for long cycle life Li–S batteries. J Mater Chem A 2017;5:250–7. [155] Fu J, Cano ZP, Park MG, Yu A, Fowler M, Chen Z. Electrically rechargeable zinc–air batteries: progress, challenges, and perspectives. Adv Mater 2017;29:1604685. [156] Huang C, Xiao J, Shao Y, Zheng J, Bennett WD, Lu D, et al. Manipulating surface reactions in lithium-sulphur batteries using hybrid anode structures. Nat Commun 2014;5:3015. [157] Liu X, Huang J-Q, Zhang Q, Mai L. Nanostructured metal oxides and sulfides for lithium-sulfur batteries. Adv Mater 2017;29:1601759. [158] Croy JR, Abouimrane A, Zhang Z. Next-generation lithium-ion batteries: the promise of near-term advancements. MRS Bull 2014;39:407–15. [159] Yang Y, Zheng G, Cui Y. Nanostructured sulfur cathodes. Chem Soc Rev 2013;42:3018–32. [160] Wang D-W, Zeng Q, Zhou G, Yin L, Li F, Cheng H-M, et al. Carbon-sulfur composites for Li-S batteries: status and prospects. J Mater Chem A 2013;1:9382–94. [161] Evers S, Nazar LF. New approaches for high energy density lithium-sulfur battery cathodes. Acc Chem Res 2013;46:1135–43. [162] Cao R, Xu W, Lv D, Xiao J, Zhang JG. Anodes for rechargeable lithium-sulfur batteries. Adv Energy Mater 2015;5:1402273. [163] Ma L, Hendrickson KE, Wei S, Archer LA. Nanomaterials: science and applications in the lithium-sulfur battery. Nano Today 2015;10:315–38. [164] Wild M, O’Neill L, Zhang T, Purkayastha R, Minton G, Marinescu M, et al. Lithium sulfur batteries, a mechanistic review. Energy Environ Sci 2015;8:3477–94. [165] Manthiram A, Fu Y, Su Y. Challenges and prospects of lithium- sulfur batteries. Acc Chem Res 2013;46:1125–34. [166] Rosenman A, Markevich E, Salitra G, Aurbach D, Garsuch A, Chesneau FF. Review on Li-sulfur battery systems: an integral perspective. Adv Energy Mater 2015;5:1500212. [167] Manthiram A, Fu Y, Chung S, Zu C, Su Y. Rechargeable lithium-sulfur batteries. Chem Rev 2014;114:11751–87. [168] Yin YX, Xin S, Guo YG, Wan LJ. Lithium-sulfur batteries: electrochemistry, materials, and prospects. Angew Chem Int Edit 2013;52:13186–200. [169] Liang J, Sun ZH, Li F, Cheng HM. Carbon materials for Li-S batteries: functional evolution and performance improvement. Energy Storage Mater 2016;2:76–106. [170] Miller JR, Simon P. Electrochemical capacitors for energy management. Sci Mag 2008;321:651–2. [171] Snook GA, Kao P, Best AS. Conducting-polymer-based supercapacitor devices and electrodes. J Power Sources 2011;196:1–12. [172] Zhi M, Xiang C, Li J, Li M, Wu N. Nanostructured carbon-metal oxide composite electrodes for supercapacitors: a review. Nanoscale 2013;5:72–88. [173] González A, Goikolea E, Barrena JA, Mysyk R. Review on supercapacitors: technologies and materials. Renew Sustain Energy Rev 2016;58:1189–206. [174] Chabot V, Higgins D, Yu A, Xiao X, Chen Z, Zhang J. A review of graphene and graphene oxide sponge: material synthesis and applications to energy and the environment. Energy Environ Sci 2014;7:1564–96. [175] Bin Tan Y, Lee J-M. Graphene for supercapacitor applications. J Mater Chem A 2013;1:14814–43. [176] Wang G, Zhang L, Zhang J. A review of electrode materials for electrochemical supercapacitors. Chem Soc Rev 2012;41:797–828. [177] Yu Z, Tetard L, Zhai L, Thomas J. Supercapacitor electrode materials: nanostructures from 0 to 3 dimensions. Energy Environ Sci 2015;8:702–30. [178] Liu C, Yu Z, Neff D, Zhamu A, Jang BZ. Graphene-based supercapacitor with an ultrahigh energy density. Nano Lett 2010;10:4863–8. [179] Mei J, Zhang L. Facile and economic synthesis of nitrogen doped graphene/manganese dioxide composites in aqueous solution for energy storage devices. Mater Lett 2015;143:163–6. [180] Mei J, Zhang L. Anchoring high-dispersed MnO2 nanowires on nitrogen doped graphene as electrode materials for supercapacitors. Electrochim Acta 2015;173:338–44. [181] Mei J, Zhang L. Novel MnOOH-graphene nanocomposites: preparation, characterization and electrochemical properties for supercapacitors. J Solid State Chem 2015;221:178–83. [182] Yuan YF, Xia XH, Wu JB, Huang XH, Pei YB, Yang JL, et al. Hierarchically porous Co3O4 film with mesoporous walls prepared via liquid crystalline template for supercapacitor application. Electrochem Commun 2011;13:1123–6. [183] Xiao Y, Liu S, Li F, Zhang A, Zhao J, Fang S, et al. 3D hierarchical Co3O4 twin-spheres with an urchin-like structure: large-scale synthesis, multistep-splitting growth, and electrochemical pseudocapacitors. Adv Funct Mater 2012;22:4052–9. [184] Wang X, Yao S, Wu X, Shi Z, Sun H, Que R. High gas-sensor and supercapacitor performance of porous Co3O4 ultrathin nanosheets. RSC Adv 2015;5:17938–44. [185] Chen S, Zhao Y, Sun B, Ao Z, Xie X, Wei Y, et al. Microwave-assisted synthesis of mesoporous Co3O4 nanoflakes for applications in lithium ion batteries and oxygen evolution reactions. ACS Appl Mater Interfaces 2015;7:3306–13. [186] Kung CW, Chen HW, Lin CY, Vittal R, Ho KC. Synthesis of Co3O4 nanosheets via electrodeposition followed by ozone treatment and their application to highperformance supercapacitors. J Power Sources 2012;214:91–9. [187] Vezvaie M, Kalisvaart P, Fritzsche H, Tun Z, Mitlin D. The penetration depth of chemical reactions in a thin-film Co3O4 supercapacitor electrode. J Electrochem Soc 2014;161:A798–802. [188] Godillot G, Guerlou-Demourgues L, Taberna P-L, Simon P, Delmas C. Original conductive nano-Co3O4 investigated as electrode material for hybrid supercapacitors. Electrochem Solid-State Lett 2011;14:A139–42. [189] Du W, Liu R, Jiang Y, Lu Q, Fan Y, Gao F. Facile synthesis of hollow Co3O4 boxes for high capacity supercapacitor. J Power Sources 2013;227:101–5. [190] Meng F, Fang Z, Li Z, Xu W, Wang M, Liu Y, et al. Porous Co3O4 materials prepared by solid-state thermolysis of a novel Co-MOF crystal and their superior energy storage performances for supercapacitors. J Mater Chem A 2013;1:7235–41. [191] Fan Y, Shao G, Ma Z, Wang G, Shao H, Yan S. Ultrathin nanoflakes assembled 3D hierarchical mesoporous Co3O4 nanoparticles for high-rate pseudocapacitors. Part Part Syst Charact 2014;31:1079–83. [192] Wang X, Sumboja A, Khoo E, Yan C, Lee PS. Cryogel synthesis of hierarchical interconnected macro-/mesoporous Co3O4 with superb electrochemical energy storage. J Phys Chem C 2012;116:4930–5. [193] Feng C, Zhang J, He Y, Zhong C, Hu W, Liu L, et al. Sub-3 nm Co3O4 nano films with enhanced supercapacitor properties. ACS Nano 2015;9:1730–9. [194] Wang Y, Lei Y, Li J, Gu L, Yuan H, Xiao D. Synthesis of 3D-nanonet hollow structured Co3O4 for high capacity supercapacitor. ACS Appl Mater Interfaces 2014;6:6739–47. [195] Xie L, Li K, Sun G, Hu Z, Lv C, Wang J, et al. Preparation and electrochemical performance of the layered cobalt oxide (Co3O4) as supercapacitor electrode
663
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
material. J Solid State Electrochem 2013;17:55–61. [196] Aghazadeh M. Electrochemical preparation and properties of nanostructured Co3O4 as supercapacitor material. J Appl Electrochem 2012;42:89–94. [197] Wang H, Zhang L, Tan X, Holt CMB, Zahiri B, Olsen BC, et al. Supercapacitive properties of hydrothermally synthesized Co3O4 nanostructures. J Phys Chem C 2011;115:17599–605. [198] Yuan C, Yang L, Hou L, Shen L, Zhang F, Lia D, et al. Large-scale Co3O4 nanoparticles growing on nickel sheets via a one-step strategy and their ultra-highly reversible redox reaction toward supercapacitors. J Mater Chem 2011;21:18183–5. [199] Tummala R, Guduru RK, Mohanty PS. Nanostructured Co3O4 electrodes for supercapacitor applications from plasma spray technique. J Power Sources 2012;209:44–51. [200] Zhou C, Zhang Y, Li Y, Liu J. Construction of high-capacitance 3D CoO@polypyrrole nanowire array electrode for aqueous asymmetric supercapacitor. Nano Lett 2013;13:2078–85. [201] Kong D, Luo J, Wang Y, Ren W, Yu T, Luo Y, et al. Three-dimensional Co3O4@MnO2 hierarchical nanoneedle arrays: morphology control and electrochemical energy storage. Adv Funct Mater 2014;24:3815–26. [202] Zhang G, Wang T, Yu X, Zhang H, Duan H, Lu B. Nanoforest of hierarchical Co3O4@NiCo2O4 nanowire arrays for high-performance supercapacitors. Nano Energy 2013;2:586–94. [203] Lang J, Yan X, Xue Q. Facile preparation and electrochemical characterization of cobalt oxide/multi-walled carbon nanotube composites for supercapacitors. J Power Sources 2011;196:7841–6. [204] Madhu R, Veeramani V, Chen SM, Manikandan A, Lo AY, Chueh YL. Honeycomb-like porous carbon-cobalt oxide nanocomposite for high-performance enzymeless glucose sensor and supercapacitor applications. ACS Appl Mater Interfaces 2015;7:15812–20. [205] Wang X, Li M, Chang Z, Yang Y, Wu Y, Liu X. Co3O4@MWCNT nanocable as cathode with superior electrochemical performance for supercapacitors. ACS Appl Mater Interfaces 2015;7:2280–5. [206] Liao Q, Li N, Jin S, Yang G, Wang C. All-solid-state symmetric supercapacitor based on Co3O4 nanoparticles on vertically aligned graphene. ACS Nano 2015;9:5310–7. [207] Wang N, Liu Q, Kang D, Gu J, Zhang W, Zhang D. Facile self-cross-linking synthesis of 3D nanoporous Co3O4/carbon hybrid electrode materials for supercapacitors. ACS Appl Mater Interfaces 2016;8:16035–44. [208] Zhang J, Fu J, Zhang J, Ma H, He Y, Li F, et al. Co@Co3O4 core-shell three-dimensional nano-network for high-performance electrochemical energy storage. Small 2014;10:2618–24. [209] Zhou W, Liu J, Chen T, Tan KS, Jia X, Luo Z, et al. Fabrication of Co3O4-reduced graphene oxide scrolls for high-performance supercapacitor electrodes. Phys Chem Chem Phys 2011;13:14462–5. [210] Ma L, Zhou H, Shen X, Chen Q, Zhu G, Ji Z. Facile synthesis of Co3O4 porous nanosheets/reduced graphene oxide composites and their excellent supercapacitor performance. RSC Adv 2014;4:53180–7. [211] Parlett CMA, Wilson K, Lee AF. Hierarchical porous materials: catalytic applications. Chem Soc Rev 2013;42:3876–93. [212] Cai X, Lai L, Lin J, Shen Z. Recent advances in air electrodes for Zn-air batteries: electrocatalysis and structural design. Mater Horiz 2017;4:945–76. [213] Wang G, Yang Y, Han D, Li Y. Oxygen defective metal oxides for energy conversion and storage. Nano Today 2017;13:23–39. [214] Wang G, Wang H, Ling Y, Tang Y, Yang X, Fitzmorris RC, et al. Hydrogen-treated TiO2 nanowire arrays for photoelectrochemical water splitting. Nano Lett 2011;11:3026–33. [215] Chen X, Liu L, Yu PY, Mao SS. Increasing solar absorption for photocatalysis with black hydrogenated titanium dioxide nanocrystals. Science 2011;331:746–50. [216] Wang G, Ling Y, Wang H, Yang X, Wang C, Zhang JZ, et al. Hydrogen-treated WO3 nanoflakes show enhanced photostability. Energy Environ Sci 2012;5:6180–7. [217] Kim H-S, Cook JB, Lin H, Ko JS, Tolbert SH, Ozolins V, et al. Oxygen vacancies enhance pseudocapacitive charge storage properties of MoO3−x. Nat Mater 2017;16:454–60. [218] Vasilopoulou M, Douvas AM, Georgiadou DG, Palilis LC, Kennou S, Sygellou L, et al. The influence of hydrogenation and oxygen vacancies on molybdenum oxides work function and gap states for application in organic optoelectronics. J Am Chem Soc 2012;134:16178–87. [219] Hanson ED, Lajaunie L, Hao S, Myers BD, Shi F, Murthy AA, et al. Systematic study of oxygen vacancy tunable transport properties of few-layer MoO3-x enabled by vapor-based synthesis. Adv Funct Mater 2017;27:1605380. [220] Zhu G, Zhu J, Jiang W, Zhang Z, Wang J, Zhu Y, et al. Surface oxygen vacancy induced α-MnO2 nanofiber for highly efficient ozone elimination. Appl Catal B Environ 2017;209:729–37. [221] Tompsett DA, Parker SC, Islam MS. Rutile (β-)MnO2 surfaces and vacancy formation for high electrochemical and catalytic performance. J Am Chem Soc 2014;136:1418–26. [222] Wang Y, Zhou T, Jiang K, Da P, Peng Z, Tang J, et al. Reduced mesoporous Co3O4 nanowires as efficient water oxidation electrocatalysts and supercapacitor electrodes. Adv Energy Mater 2014;4:1400696. [223] Zhai T, Wan L, Sun S, Chen Q, Sun J, Xia Q, et al. Phosphate ion functionalized Co3O4 ultrathin nanosheets with greatly improved surface reactivity for high performance pseudocapacitors. Adv Mater 2017;29:1604167. [224] Seh ZW, Kibsgaard J, Dickens CF, Chorkendorff I, Nørskov JK, Jaramillo TF. Combining theory and experiment in electrocatalysis: insights into materials design. Science 2017;355:eaad4998. [225] Stamenkovic VR, Strmcnik D, Lopes PP, Markovic NM. Energy and fuels from electrochemical interfaces. Nat Mater 2017;16:57–69. [226] Wu JB, Li ZG, Huang XH, Lin Y. Porous Co3O4/NiO core/shell nanowire array with enhanced catalytic activity for methanol electro-oxidation. J Power Sources 2013;224:1–5. [227] Sun S, Xu ZJ. Composition dependence of methanol oxidation activity in nickel-cobalt hydroxides and oxides: an optimization toward highly active electrodes. Electrochim Acta 2015;165:56–66. [228] Zafeiratos S, Dintzer T, Teschner D, Blume R, Hävecker M, Knop-Gericke A, et al. Methanol oxidation over model cobalt catalysts: influence of the cobalt oxidation state on the reactivity. J Catal 2010;269:309–17. [229] Jia W, Guo M, Zheng Z, Yu T, Rodriguez EG, Wang Y, et al. Electrocatalytic oxidation and reduction of H2O2 on vertically aligned Co3O4 nanowalls electrode: toward H2O2 detection. J Electroanal Chem 2009;625:27–32. [230] Shahid MM, Pandikumar A, Golsheikh AM, Huang NM, Lim HN. Enhanced electrocatalytic performance of cobalt oxide nanocubes incorporating reduced graphene oxide as a modified platinum electrode for methanol oxidation. RSC Adv 2014;4:62793–801. [231] Birss VI, Damjanovic A, Hudson PG. Oxygen evolution at platinum electrodes in alkaline solutions. J Electrochem Soc 1986;133:1621–5. [232] Conway BE, Liu TC. Characterization of electrocatalysis in the oxygen evolution reaction at platinum by evaluation of behavior of surface intermediate states at the oxide film. Langmuir 1990;6:268–76. [233] Fabbri E, Habereder A, Waltar K, Kötz R, Schmidt TJ, Kotz R, et al. Developments and perspectives of oxide-based catalysts for the oxygen evolution reaction. Catal Sci Technol 2014;4:3800–21. [234] Koza JA, He Z, Miller AS, Switzer JA. Electrodeposition of crystalline Co3O4-A catalyst for the oxygen evolution reaction. Chem Mater 2012;24:3567–73. [235] Yao L, Zhong H, Deng C, Li X, Zhang H. Template-assisted synthesis of hierarchically porous Co3O4 with enhanced oxygen evolution activity. J Energy Chem 2016;25:153–7. [236] Chen P, Xu K, Zhou T, Tong Y, Wu J, Cheng H, et al. Strong-coupled cobalt borate nanosheets/graphene hybrid as electrocatalyst for water oxidation under both alkaline and neutral conditions. Angew Chem 2016;128:2534–8. [237] Zhang L, Li H, Li K, Li L, Wei J, Feng L, et al. Morphology-controlled fabrication of Co3O4 nanostructures and their comparative catalytic activity for oxygen evolution reaction. J Alloys Compd 2016;680:146–54. [238] Gong M, Dai H. A mini review of NiFe-based materials as highly active oxygen evolution reaction electrocatalysts. Nano Res 2015;8:23–39. [239] Jeon HS, Jee MS, Kim H, Ahn SJ, Hwang YJ, Min BK. Simple chemical solution deposition of Co3O4 thin film electrocatalyst for oxygen evolution reaction. ACS
664
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
Appl Mater Interfaces 2015;7:24550–5. [240] Bajdich M, García-Mota M, Vojvodic A, Nørskov JK, Bell AT. Theoretical investigation of the activity of cobalt oxides for the electrochemical oxidation of water. J Am Chem Soc 2013;135:13521–30. [241] Zhang Y, Ding F, Deng C, Zhen S, Li X, Xue Y, et al. Crystal plane-dependent electrocatalytic activity of Co3O4 toward oxygen evolution reaction. Catal Commun 2015;67:78–82. [242] Hutchings GS, Zhang Y, Li J, Yonemoto BT, Zhou X, Zhu K, et al. In situ formation of cobalt oxide nanocubanes as efficient oxygen evolution catalysts. J Am Chem Soc 2015;137:4223–9. [243] Kim TW, Woo MA, Regis M, Choi K-S. Electrochemical synthesis of spinel type ZnCo2O4 electrodes for use as oxygen evolution reaction catalysts. J Phys Chem Lett 2014;5:2370–4. [244] Wang HY, Hung SF, Chen HY, Chan TS, Chen HM, Liu B. In operando identification of geometrical-site-dependent water oxidation activity of spinel Co3O4. J Am Chem Soc 2016;138:36–9. [245] Xu L, Jiang Q, Xiao Z, Li X, Huo J, Wang S, et al. Plasma-engraved Co3O4 nanosheets with oxygen vacancies and high surface area for the oxygen evolution reaction. Angew Chem Int Edit 2016;55:5277–81. [246] Chen Y, Hu J, Diao H, Luo W, Song YF. Facile preparation of ultrathin Co3O4/nanocarbon composites with greatly improved surface activity as a highly efficient oxygen evolution reaction catalyst. Chem Eur J 2017;23:4010–6. [247] Suntivich J, May KJ, Gasteiger HA, Goodenough JB, Shao-Horn Y. A perovskite oxide optimized for oxygen evolution catalysis from molecular orbital principles. Science 2011;334:1383–5. [248] Dou Y, Liao T, Ma Z, Tian D, Liu Q, Xiao F, et al. Graphene-like holey Co3O4 nanosheets as a highly efficient catalyst for oxygen evolution reaction. Nano Energy 2016;30:267–75. [249] Zhou X, Xia Z, Tian Z, Ma Y, Qu Y. Ultrathin porous Co3O4 nanoplates as highly efficient oxygen evolution catalysts. J Mater Chem A 2015;3:8107–14. [250] Yang X, Li H, Lu AY, Min S, Idriss Z, Hedhili MN, et al. Highly acid-durable carbon coated Co3O4 nanoarrays as efficient oxygen evolution electrocatalysts. Nano Energy 2016;25:42–50. [251] Ma TY, Dai S, Jaroniec M, Qiao SZ. Metal-organic framework-derived hybrid Co3O4-carbon porous nanowire arrays as reversible oxygen evolution electrodes. J Am Chem Soc 2014;136:13925–31. [252] Bayatsarmadi B, Zheng Y, Casari CSS, Russo V, Qiao S. Pulsed laser deposition of porous N-carbon supported cobalt (oxide) thin films for highly efficient oxygen evolution. Chem Commun 2016;52:11947–50. [253] Jadhav AR, Bandal HA, Tamboli AH, Kim H. Environment friendly hydrothermal synthesis of carbon–Co3O4 nanorods composite as an efficient catalyst for oxygen evolution reaction. J Energy Chem 2017;26:695–702. [254] Li Y, Hasin P, Wu Y. NixCo3-xO4 nanowire arrays for electrocatalytic oxygen evolution. Adv Mater 2010;22:1926–9. [255] Xiao C, Lu X, Zhao C. Unusual synergistic effects upon incorporation of Fe and/or Ni into mesoporous Co3O4 for enhanced oxygen evolution. Chem Commun 2014;50:10122–5. [256] Yang Y, Fei H, Ruan G, Xiang C, Tour JM. Efficient electrocatalytic oxygen evolution on amorphous nickel-cobalt binary oxide nanoporous layers. ACS Nano 2014;8:9518–23. [257] Liu X, Chang Z, Luo L, Xu T, Lei X, Liu J, et al. Hierarchical ZnxCo3-xO4 nanoarrays with high activity for electrocatalytic oxygen evolution. Chem Mater 2014;26:1889–95. [258] McCrory CCL, Jung SH, Peters JC, Jaramillo TF. Benchmarking heterogeneous electrocatalysts for the oxygen evolution reaction. J Am Chem Soc 2013;135:16977–87. [259] Ben Liew K, Daud WRW, Ghasemi M, Leong JX, Su Lim S, Ismail M. Non-Pt catalyst as oxygen reduction reaction in microbial fuel cells: a review. Int J Hydrogen Energy 2014;39:4870–83. [260] Guo S, Zhang S, Sun S. Tuning nanoparticle catalysis for the oxygen reduction reaction. Angew Chem Int Edit 2013;52:8526–44. [261] Dai L, Xue Y, Qu L, Choi HJ, Baek JB. Metal-free catalysts for oxygen reduction reaction. Chem Rev 2015;115:4823–92. [262] Wang B. Recent development of non-platinum catalysts for oxygen reduction reaction. J Power Sources 2005;152:1–15. [263] Kumar R, Singh L, Zularisam AW, Hai FI. Potential of porous Co3O4 nanorods as cathode catalyst for oxygen reduction reaction in microbial fuel cells. Bioresource Technol 2016;220:537–42. [264] Xu J, Gao P, Zhao TS. Non-precious Co3O4 nano-rod electrocatalyst for oxygen reduction reaction in anion-exchange membrane fuel cells. Energy Environ Sci 2012;5:5333–9. [265] Ren Y, Cheng Y, Gorte RJ, Huang K. Toward stabilizing Co3O4 nanoparticles as an oxygen reduction reaction catalyst for intermediate-temperature SOFCs. J Electrochem Soc 2017;164:F3001–7. [266] Liu H, Long W, Song W, Liu J, Wang F. Tuning the electronic bandgap: an efficient way to improve the electrocatalytic activity of carbon-supported Co3O4 nanocrystals for oxygen reduction reactions. Chem Eur J 2017;23:2599–609. [267] Xiao J, Kuang Q, Yang S, Xiao F, Wang S, Guo L. Surface structure dependent electrocatalytic activity of Co3O4 anchored on graphene sheets toward oxygen reduction reaction. Sci Rep 2013;3:2300. [268] Liang Y, Wang H, Zhou J, Li Y, Wang J, Regier T, et al. Covalent hybrid of spinel manganese−cobalt oxide and graphene as advanced oxygen reduction electrocatalysts. J Am Chem Soc 2012;134:3517–23. [269] Indra A, Menezes PW, Sahraie NR, Bergmann A, Das C, Tallarida M, et al. Unification of catalytic water oxidation and oxygen reduction reactions: amorphous beat crystalline cobalt iron oxides. J Am Chem Soc 2014;136:17530–6. [270] Li S-S, Cong H-P, Wang P, Yu S-H. Flexible nitrogen-doped graphene/carbon nanotube/Co3O4 paper and its oxygen reduction activity. Nanoscale 2014;6:7534–41. [271] Huo R, Jiang W-J, Xu S, Zhang F, Hu J-S, Bing Y, et al. Co/CoO/CoFe2O4/G nanocomposites derived from layered double hydroxides towards mass production of efficient Pt-free electrocatalysts for oxygen reduction reaction. Nanoscale 2014;6:203–6. [272] Odedairo T, Yan X, Ma J, Jiao Y, Yao X, Du A, et al. Nanosheets Co3O4 interleaved with graphene for highly efficient oxygen reduction. ACS Appl Mater Interfaces 2015;7:21373–80. [273] Wu ZY, Chen P, Wu QS, Yang LF, Pan Z, Wang Q. Co/Co3O4/C-N, a novel nanostructure and excellent catalytic system for the oxygen reduction reaction. Nano Energy 2014;8:118–25. [274] Liang Y, Wang H, Diao P, Chang W, Hong G, Li Y, et al. Oxygen reduction electrocatalyst based on strongly coupled cobalt oxide nanocrystals and carbon nanotubes. J Am Chem Soc 2012;134:15849–57. [275] Malkhandi S, Trinh P, Manohar AK, Jayachandrababu KC, Kindler A, Surya Prakash GK, et al. Electrocatalytic activity of transition metal oxide-carbon composites for oxygen reduction in alkaline batteries and fuel cells. J Electrochem Soc 2013;160:F943–52. [276] Li T, Liu J, Jin X, Wang F, Song Y. Composition-dependent electro-catalytic activities of covalent carbon-LaMnO3 hybrids as synergistic catalysts for oxygen reduction reaction. Electrochim Acta 2016;198:115–26. [277] Popczun EJ, Read CG, Roske CW, Lewis NS, Schaak RE. Highly active electrocatalysis of the hydrogen evolution reaction by cobalt phosphide nanoparticles. Angew Chem 2014;126:5531–4. [278] Safizadeh F, Ghali E, Houlachi G. Electrocatalysis developments for hydrogen evolution reaction in alkaline solutions-A review. Int J Hydrogen Energy 2015;40:256–74. [279] Yan Y, Xia B, Xu Z, Wang X. Recent development of molybdenum sulfides as advanced electrocatalysts for hydrogen evolution reaction. ACS Catal 2014;4:1693–705. [280] Zheng Y, Jiao Y, Jaroniec M, Qiao SZ. Advancing the electrochemistry of the hydrogen- Evolution reaction through combining experiment and theory. Angew Chem Int Edit 2015;54:52–65. [281] Chen WF, Sasaki K, Ma C, Frenkel AI, Marinkovic N, Muckerman JT, et al. Hydrogen-evolution catalysts based on non-noble metal nickel-molybdenum nitride
665
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
nanosheets. Angew Chem Int Edit 2012;51:6131–5. [282] Zou X, Zhang Y. Noble metal-free hydrogen evolution catalysts for water splitting. Chem Soc Rev 2015;44:5148–80. [283] Wang HZ, Leung DYC, Leung MKH, Ni M. A review on hydrogen production using aluminum and aluminum alloys. Renew Sustain Energy Rev 2009;13:845–53. [284] Vesborg PCK, Seger B, Chorkendorff I. Recent development in hydrogen evolution reaction catalysts and their practical implementation. J Phys Chem Lett 2015;6:951–7. [285] Chen W-F, Muckerman JT, Fujita E. Recent developments in transition metal carbides and nitrides as hydrogen evolution electrocatalysts. Chem Commun 2013;49:8896–909. [286] Fournier J, Brossard L, TiIqui J-Y, Côté R, Dodelet J-P, Guay D, et al. Hydrogen evolution reaction in alkaline solution. J Electrochem Soc 1996;143:919–26. [287] Danilovic N, Subbaraman R, Strmcnik D, Stamenkovic V, Markovic N. Electrocatalysis of the HER in acid and alkaline media. J Serbian Chem Soc 2013;78:2007–15. [288] Sheng W, Gasteiger HA, Shao-Horn Y. Hydrogen oxidation and evolution reaction kinetics on platinum: acid vs alkaline electrolytes. J Electrochem Soc 2010;157:B1529–36. [289] Gong M, Wang DY, Chen CC, Hwang BJ, Dai H. A mini review on nickel-based electrocatalysts for alkaline hydrogen evolution reaction. Nano Res 2016;9:28–46. [290] Zeng M, Li Y. Recent advances in heterogeneous electrocatalysts for the hydrogen evolution reaction. J Mater Chem A 2015;3:14942–62. [291] Shi Y, Zhang B. Recent advances in transition metal phosphide nanomaterials: synthesis and applications in hydrogen evolution reaction. Chem Soc Rev 2016;45:1529–41. [292] Xiao P, Chen W, Wang X. A review of phosphide-based materials for electrocatalytic hydrogen evolution. Adv Energy Mater 2015;5:1500985. [293] Yan X, Tian L, He M, Chen X. Three-dimensional crystalline/amorphous Co/Co3O4 core/shell nanosheets as efficient electrocatalysts for the hydrogen evolution reaction. Nano Lett 2015;15:6015–21. [294] Tahir M, Mahmood N, Zhang X, Mahmood T, Butt FK, Aslam I, et al. Bifunctional catalysts of Co3O4@GCN tubular nanostructured (TNS) hybrids for oxygen and hydrogen evolution reactions. Nano Res 2015;8:3725–36. [295] Jin H, Wang J, Su D, Wei Z, Pang Z, Wang Y. In situ cobalt-cobalt oxide/N-doped carbon hybrids as superior bifunctional electrocatalysts for hydrogen and oxygen evolution. J Am Chem Soc 2015;137:2688–94. [296] Li T, Niu K, Yang M, Shrestha NK, Gao Z, Song YY. Ultrathin CoS2 shells anchored on Co3O4 nanoneedles for efficient hydrogen evolution electrocatalysis. J Power Sources 2017;356:89–96. [297] Amatore C, Saveant JM. Mechanism and kinetic characteristics of the electrochemical reduction of carbon dioxide in media of low proton availability. J Am Chem Soc 1981;103:5021–3. [298] Azuma M, Hashimoto K, Hiramoto M, Watanabe M, Sakata T. Electrochemical reduction of carbon dioxide on various metal electrodes in low-temperature aqueous KHCO3 media. J Electrochem Soc 1990;137:1772–8. [299] Peterson A, Nørskov J. Activity descriptors for CO2 electroreduction to methane on transition metal catalysts. J Phys Chem Lett 2012;3:251–8. [300] Costentin C, Robert M, Savéant J-M. Catalysis of the electrochemical reduction of carbon dioxide. Chem Soc Rev 2013;42:2423–36. [301] Lu Q, Rosen J, Zhou Y, Hutchings GS, Kimmel YC, Chen JG, et al. A selective and efficient electrocatalyst for carbon dioxide reduction. Nat Commun 2014;5:3242. [302] Zhang S, Kang P, Meyer TJ. Nanostructured tin catalysts for selective electrochemical reduction of carbon dioxide to formate. J Am Chem Soc 2014;136:1734–7. [303] Kumar B, Asadi M, Pisasale D, Sinha-Ray S, Rosen BA, Haasch R, et al. Renewable and metal-free carbon nanofibre catalysts for carbon dioxide reduction. Nat Commun 2013;4:2819. [304] Kim D, Resasco J, Yu Y, Asiri AM, Yang P. Synergistic geometric and electronic effects for electrochemical reduction of carbon dioxide using gold-copper bimetallic nanoparticles. Nat Commun 2014;5:4948. [305] Hori Y, Takahashi R, Yoshinami Y, Murata A. Electrochemical reduction of CO at a copper electrode. J Phys Chem B 1997;101:7075–81. [306] Hori Y, Murata A, Takahashi R. Formation of hydrocarbons in the electrochemical reduction of carbon dioxide at a copper electrode in aqueous solution. J Chem Soc Faraday Trans 1989;1(85):2309–26. [307] Hara K, Kudo A, Sakata T. Electrochemical reduction of carbon dioxide under high pressure on various electrodes in an aqueous electrolyte. J Electroanal Chem 1995;391:141–7. [308] Frese KW, Leach S. Electrochemical reduction of carbon dioxide to methane, methanol, and CO on Ru electrodes. J Electrochem Soc 1985;132:259–60. [309] Ikeda S, Takagi T, Ito K. Selective formation of formic acid, oxalic acid, and carbon monoxide by electrochemical reduction of carbon dioxide. Bull Chem Soc Jpn 1987;60:2517–22. [310] Rosen BA, Salehi-khojin A, Thorson MR, Zhu W, Whipple DT, Kenis PJA, et al. Ionic liquid-mediated selective conversion of CO2 to CO at low overpotentials. Science 2011;334:643–4. [311] Kuhl KP, Hatsukade T, Cave ER, Abram DN, Kibsgaard J, Jaramillo TF. Electrocatalytic conversion of carbon dioxide to methane and methanol on transition metal surfaces. J Am Chem Soc 2014;136:14107–13. [312] Kuhl KP, Cave E, Abram DN, Jaramillo TF. New insights into the electrochemical reduction of carbon dioxide on metallic copper surfaces. Energy Environ Sci 2012;5:7050–9. [313] Gao S, Sun Z, Liu W, Jiao X, Zu X, Hu Q, et al. Atomic layer confined vacancies for atomic-level insights into carbon dioxide electroreduction. Nat Commun 2017;8:14503. [314] Gao S, Lin Y, Jiao X, Sun Y, Luo Q, Zhang W, et al. Partially oxidized atomic cobalt layers for carbon dioxide electroreduction to liquid fuel. Nature 2016;529:68–71. [315] Gao S, Jiao X, Sun Z, Zhang W, Sun Y, Wang C, et al. Ultrathin Co3O4 layers realizing optimized CO2 electroreduction to formate. Angew Chem Int Edit 2016;55:698–702. [316] Chemelewski WD, Lee H, Lin J-F, Bard AJ, Mullins CB. Amorphous FeOOH oxygen evolution reaction catalyst for photoelectrochemical water splitting. J Am Chem Soc 2014;136:2843–50. [317] Feng LL, Yu G, Wu Y, Li GD, Li H, Sun Y, et al. High-index faceted Ni3S2 nanosheet arrays as highly active and ultrastable electrocatalysts for water splitting. J Am Chem Soc 2015;137:14023–6. [318] Gerken JB, Mcalpin JG, Chen JYC, Rigsby ML, Casey WH, Britt RD, et al. Electrochemical water oxidation with cobalt-based electrocatalysts from pH 0–14: the thermodynamic basis for catalyst structure, stability, and activity. J Am Chem Soc 2011;133:14431–42. [319] Zhang Y, Zhang H, Ma Y, Cheng J, Zhong H, Song S, et al. A novel bifunctional electrocatalyst for unitized regenerative fuel cell. J Power Sources 2010;195:142–5. [320] Jörissen L. Bifunctional oxygen/air electrodes. J Power Sources 2006;155:23–32. [321] Altmann S, Kaz T, Friedrich KA. Bifunctional electrodes for unitised regenerative fuel cells. Electrochim Acta 2011;56:4287–93. [322] Chen G, Bare SR, Mallouk TE. Development of supported bifunctional electrocatalysts for unitized regenerative fuel cells. J Electrochem Soc 2002;149:A1092–9. [323] Wang D, Su D. Heterogeneous nanocarbon materials for oxygen reduction reaction. Energy Environ Sci 2014;7:576–91. [324] Liu H, Xia G, Zhang R, Jiang P, Chen J, Chen Q. MOF-derived RuO2/Co3O4 heterojunctions as highly efficient bifunctional electrocatalysts for HER and OER in alkaline solutions. RSC Adv 2017;7:3686–94. [325] Menezes PW, Indra A, González-Flores D, Sahraie NR, Zaharieva I, Schwarze M, et al. High-performance oxygen redox catalysis with multifunctional cobalt oxide nanochains: morphology-dependent activity. ACS Catal 2015;5:2017–27. [326] Mao S, Wen Z, Huang T, Hou Y, Chen J. High-performance bi-functional electrocatalysts of 3D crumpled graphene-cobalt oxide nanohybrids for oxygen reduction and evolution reactions. Energy Environ Sci 2014;7:609–16.
666
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
[327] Liu X, Liu W, Ko M, Park M, Kim MG, Oh P, et al. Metal (Ni, Co)-metal oxides/graphene nanocomposites as multifunctional electrocatalysts. Adv Funct Mater 2015;25:5799–808. [328] Zhang T, He C, Sun F, Ding Y, Wang M, Peng L, et al. Co3O4 nanoparticles anchored on nitrogen-doped reduced graphene oxide as a multifunctional catalyst for H2O2 reduction, oxygen reduction and evolution reaction. Sci Rep 2017;7:43638. [329] Buzea C, Pacheco II, Robbie K. Nanomaterials and nanoparticles: sources and toxicity. Biointerphases 2007;2:MR17-MR71. [330] Sajanlal PR, Sreeprasad TS, Samal AK, Pradeep T. Anisotropic nanomaterials: structure, growth, assembly, and functions. Nano Rev 2011;2:5883. [331] Vijayakumar S, Kiruthika Ponnalagi A, Nagamuthu S, Muralidharan G. Microwave assisted synthesis of Co3O4 nanoparticles for high-performance supercapacitors. Electrochim Acta 2013;106:500–5. [332] Zhou Y, Dong CK, Han LL, Yang J, Du XW. Top-down preparation of active cobalt oxide catalyst. ACS Catal 2016;6:6699–703. [333] Koshizaki N, Narazaki A, Sasaki T. Size distribution and growth mechanism of Co3O4 nanoparticles fabricated by pulsed laser deposition. Scripta Mater 2001;44:1925–8. [334] Blakemore JD, Gray HB, Winkler JR, Müller AM. Co3O4 nanoparticle water-oxidation catalysts made by pulsed-laser ablation in liquids. ACS Catal 2013;3:2497–500. [335] García-Pacheco G, Cabañas-Moreno JG, Yee-Madeira H, Cruz-Gandarilla F. Co3O4 nanoparticles produced by mechanochemical reactions. Nanotechnology 2006;17:2528–35. [336] Liu XY, Gao YQ, Yang GW. A flexible, transparent and super-long-life supercapacitor based on ultrafine Co3O4 nanocrystal electrodes. Nanoscale 2016;8:4227–35. [337] Kim J-W, Choi SH, Lillehei PT, Chu S-H, King GC, Watt GD. Cobalt oxide hollow nanoparticles derived by bio-templating. Chem Commun 2005:4101–3. [338] Zou D, Xu C, Luo H, Wang L, Ying T. Synthesis of Co3O4 nanoparticles via an ionic liquid-assisted methodology at room temperature. Mater Lett 2008;62:1976–8. [339] Wang Y, Shi JC, Cao JL, Sun G, Zhang ZY. Synthesis of Co3O4 nanoparticles via the CTAB-assisted method. Mater Lett 2011;65:222–4. [340] Nassar MY. Size-controlled synthesis of CoCO3 and Co3O4 nanoparticles by free-surfactant hydrothermal method. Mater Lett 2013;94:112–5. [341] Liu Y-R, Han G-Q, Li X, Dong B, Shang X, Hu W-H, et al. A facile synthesis of reduced Co3O4 nanoparticles with enhanced electrocatalytic activity for oxygen evolution. Int J Hydrogen Energy 2016;41:12976–82. [342] Zhu J, Kailasam K, Fischer A, Thomas A. Supported cobalt oxide nanoparticles as catalyst for aerobic oxidation of alcohols in liquid phase. ACS Catal 2011;1:342–7. [343] Deng J, Kang L, Bai G, Li Y, Li P, Liu X, et al. Solution combustion synthesis of cobalt oxides (Co3O4 and Co3O4/CoO) nanoparticles as supercapacitor electrode materials. Electrochim Acta 2014;132:127–35. [344] Fan S, Liu X, Li Y, Yan E, Wang C, Liu J, et al. Non-aqueous synthesis of crystalline Co3O4 nanoparticles for lithium-ion batteries. Mater Lett 2013;91:291–3. [345] Dong Y, He K, Yin L, Zhang A. A facile route to controlled synthesis of Co3O4 nanoparticles and their environmental catalytic properties. Nanotechnology 2007;18. [346] Iablokov V, Beaumont SK, Alayoglu S, Pushkarev VV, Specht C, Gao J, et al. Size-controlled model Co nanoparticle catalysts for CO2 hydrogenation: synthesis, characterization, and catalytic reactions. Nano Lett 2012;12:3091–6. [347] Seo WS, Shim JH, Oh SJ, Lee EK, Hur NH, Park JT. Phase- and size-controlled synthesis of hexagonal and cubic CoO nanocrystals. J Am Chem Soc 2005;127:6188–9. [348] Nam KM, Shim JH, Ki H, Il Choi S, Lee G, Jang JK, et al. Single-crystalline hollow face-centered-cubic cobalt nanoparticles from solid face-centered-cubic cobalt oxide nanoparticles. Angew Chem Int Edit 2008;47:9504–8. [349] Iablokov V, Barbosa R, Pollefeyt G, Van Driessche I, Chenakin S, Kruse N. Catalytic CO oxidation over well-defined cobalt oxide nanoparticles: size-reactivity correlation. ACS Catal 2015;5:5714–8. [350] Bartůněk V, Huber Š, Sedmidubský D, Sofer Z, Šimek P, Jankovský O. CoO and Co3O4 nanoparticles with a tunable particle size. Ceram Int 2014;40:12591–5. [351] Grzelczak M, Zhang J, Pfrommer J, Hartmann J, Driess M, Antonietti M, et al. Electro-and photochemical water oxidation on ligand-free Co3O4 nanoparticles with tunable sizes. ACS Catal 2013;3:383–8. [352] Varón M, Ojea-Jimenez I, Arbiol J, Balcells L, Martínez B, Puntes VF. Spontaneous formation of hollow cobalt oxide nanoparticles by the Kirkendall effect at room temperature at the water-air interface. Nanoscale 2013;5:2429–36. [353] Xue J, Ma W, Wang L, Cui H. Surfactant-free large scale synthesis of Co3O4 quantum dots at room temperature. Adv Powder Technol 2016;27:2019–24. [354] Sun Z, Liao T, Kou L. Strategies for designing metal oxide nanostructures. Sci China Mater 2017;60:1–24. [355] Zhang N, Shi JW, Mao SS, Guo LJ. Co3O4 quantum dots: reverse micelle synthesis and visible-light-driven photocatalytic overall water splitting. Chem Commun 2014;50:2002–4. [356] Shi N, Cheng W, Zhou H, Fan T, Niederberger M. Facile synthesis of monodisperse Co3O4 quantum dots with efficient oxygen evolution activity. Chem Commun 2015;51:1338–40. [357] Zhang G, Yang J, Wang H, Chen H, Yang J, Pan F. Co3O4−δ quantum dots as a highly efficient oxygen evolution reaction catalyst for water splitting. ACS Appl Mater Interfaces 2017;9:16159–67. [358] Su B, Wu Y, Jiang L. The art of aligning one-dimensional (1D) nanostructures. Chem Soc Rev 2012;41:7832–56. [359] Yan D, Zhang H, Chen L, Zhu G, Li S, Xu H, et al. Biomorphic synthesis of mesoporous Co3O4 microtubules and their pseudocapacitive performance. ACS Appl Mater Interfaces 2014;6:15632–7. [360] Kumar M, Subramania A, Balakrishnan K. Preparation of electrospun Co3O4 nanofibers as electrode material for high performance asymmetric supercapacitors. Electrochim Acta 2014;149:152–8. [361] Huang H, Zhu W, Tao X, Xia Y, Yu Z, Fang J, et al. Nanocrystal-constructed mesoporous single-crystalline Co3O4 nanobelts with superior rate capability for advanced lithium-ion batteries. ACS Appl Mater Interfaces 2012;4:5974–80. [362] Yao M, Hu Z, Xu Z, Liu Y. Template synthesis of 1D hierarchical hollow Co3O4 nanotubes as high performance supercapacitor materials. J Alloys Compd 2015;644:721–8. [363] Zhu T, Chen JS, Lou XW. Shape-controlled synthesis of porous Co3O4 nanostructures for application in supercapacitors. J Mater Chem 2010;20:7015–20. [364] Liu W, Li X, Zhu M, He X. High-performance all-solid state asymmetric supercapacitor based on Co3O4 nanowires and carbon aerogel. J Power Sources 2015;282:179–86. [365] Cheng H, Lu ZG, Deng JQ, Chung CY, Zhang K, Li YY. A facile method to improve the high rate capability of Co3O4 nanowire array electrodes. Nano Res 2010;3:895–901. [366] Li Y, Tan B, Wu Y. Freestanding mesoporous quasi-single-crystalline Co3O4 nanowire arrays. J Am Chem Soc 2006;128:14258–9. [367] Chen M, Xia X, Yin J, Chen Q. Construction of Co3O4 nanotubes as high-performance anode material for lithium ion batteries. Electrochim Acta 2015;160:15–21. [368] Yang Z, Wang S, Liu Y, Lei X. Cobalt oxide microtubes with balsam pear-shaped outer surfaces as anode material for lithium ion batteries. Ionics 2015;21:2423–30. [369] Zheng F, Shi K, Xu S, Liang X, Chen Y, Zhang Y. Facile fabrication of highly porous Co3O4 nanobelts as anode materials for lithium-ion batteries. RSC Adv 2016;6:9640–6. [370] Wang D, Wang Q, Wang T. Morphology-controllable synthesis of cobalt oxalates and their conversion to mesoporous Co3O4 nanostructures for application in supercapacitors. Inorg Chem 2011;50:6482–92. [371] Cui L, Li J, Zhang XG. Preparation and properties of Co3O4 nanorods as supercapacitor material. J Appl Electrochem 2009;39:1871–6. [372] Xu J, Gao L, Cao J, Wang W, Chen Z. Preparation and electrochemical capacitance of cobalt oxide (Co3O4) nanotubes as supercapacitor material. Electrochim Acta 2010;56:732–6. [373] Meher SK, Rao GR. Effect of microwave on the nanowire morphology, optical, magnetic, and pseudocapacitance behavior of Co3O4. J Phys Chem C
667
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
2011;115:25543–56. [374] Meng T, Xu QQ, Wang ZH, Li YT, Gao ZM, Xing XY, et al. Co3O4 nanorods with self-assembled nanoparticles in queue for supercapacitor. Electrochim Acta 2015;180:104–11. [375] Chen J, Wiley BJ, Xia Y. One-dimensional nanostructures of metals: large-scale synthesis and some potential applications. Langmuir 2007;23:4120–9. [376] Sun Z, Kim JH, Liao T, Zhao Y, Bijarbooneh F, Malgras V, et al. Continually adjustable oriented 1D TiO2 nanostructure arrays with controlled growth of morphology and their application in dye-sensitized solar cells. CrystEngComm 2012;14:5472–8. [377] Li WY, Xu LN, Chen J. Co3O4 nanomaterials in lithium-ion batteries and gas sensors. Adv Funct Mater 2005;15:851–7. [378] Liu Y, Wang G, Xu C, Wang W. Fabrication of Co3O4 nanorods by calcination of precursor powders prepared in a novel inverse microemulsion. Chem Commun 2002:1486–7. [379] Lou XW, Deng D, Lee JY, Feng J, Archer LA. Self-supported formation of needlelike Co3O4 nanotubes and their application as lithium-ion battery electrodes. Adv Mater 2008;20:258–62. [380] Lou XW, Deng D, Lee JY, Archer LA. Thermal formation of mesoporous single-crystal Co3O4 nano-needles and their lithium storage properties. J Mater Chem 2008;18:4397–401. [381] Law M, Greene LE, Johnson JC, Saykally R, Yang P. Nanowire dye-sensitized solar cells. Nat Mater 2005;4:455–9. [382] Yu Z, Cheng Z, Tai Z, Wang X, Subramaniyam CM, Fang C, et al. Tuning the morphology of Co3O4 on Ni foam for supercapacitor application. RSC Adv 2016;6:45783–90. [383] Thelander C, Agarwal P, Brongersma S, Eymery J, Feiner LF, Forchel A, et al. Nanowire-based one-dimensional electronics. Mater Today 2006;9:28–35. [384] Li Y, Qian F, Xiang J, Lieber CM. Nanowire electronic and optoelectronic devices. Mater Today 2006;9:18–27. [385] Wanekaya AK, Chen W, Myung NV, Mulchandani A. Nanowire-based electrochemical biosensors. Electroanalysis 2006;18:533–50. [386] Patolsky F, Lieber CM. Nanowire nanosensors. Mater Today 2005;8:20–8. [387] Weber WM, Heinzig A, Trommer J, Martin D, Grube M, Mikolajick T. Reconfigurable nanowire electronics-a review. Solid State Electron 2014;102:12–24. [388] Li Y, Tan B, Wu Y. Mesoporous Co3O4 nanowire arrays for lithium ion batteries with high capacity and rate capability. Nano Lett 2008;8:265–70. [389] Zheng J, Zhang B. Facile chemical bath deposition of Co3O4 nanowires on nickel foam directly as conductive agent- and binder-free anode for lithium ion batteries. Ceram Int 2014;40:11377–80. [390] Li CC, Yin XM, Chen LB, Li QH, Wang TH. High capacity and excellent cycling stability of branched cobalt oxide nanowires as Li-insertion materials. Appl Phys Lett 2010;97. [391] Zhang C, Xie L, Song W, Wang J, Sun G, Li K. Electrochemical performance of asymmetric supercapacitor based on Co3O4/AC materials. J Electroanal Chem 2013;706:1–6. [392] Zhan L, Wang S, Ding LX, Li Z, Wang H. Grass-like Co3O4 nanowire arrays anode with high rate capability and excellent cycling stability for lithium-ion batteries. Electrochim Acta 2014;135:35–41. [393] Xia X, Tu J, Mai Y, Wang X, Gu C, Zhao X. Self-supported hydrothermal synthesized hollow Co3O4 nanowire arrays with high supercapacitor capacitance. J Mater Chem 2011;21:9319–25. [394] Zhang Y, Wu Y, Chu Y, Li L, Yu Q, Zhu Y, et al. Self-assembled Co3O4 nanostructure with controllable morphology towards high performance anode for lithium ion batteries. Electrochim Acta 2016;188:909–16. [395] Wang Y, Xia H, Lu L, Lin J. Excellent performance in lithium-ion battery anodes: rational synthesis of Co(CO3)0.5(OH) 0.11H2O nanobelt array and its conversion into mesoporous and single-crystal Co3O4. ACS Nano 2010;4:1425–32. [396] Lee H, Kim Y-J, Lee DJ, Song J, Lee YM, Kim H-T, et al. Directly grown Co3O4 nanowire arrays on Ni-foam: structural effects of carbon-free and binder-free cathodes for lithium–oxygen batteries. J Mater Chem A 2014;2:11891–8. [397] Hong S, Fan J. Efficient oxygen reduction reaction using mesoporous Ni-doped Co3O4 nanowire array electrocatalysts. J Mater Chem A 2015;3:18372–9. [398] Xia X, Tu J, Zhang Y, Mai Y, Wang X, Gu C, et al. Freestanding Co3O4 nanowire array for high performance supercapacitors. RSC Adv 2012;2:1835–41. [399] Zhang X, Zhao Y, Xu C. Surfactant dependent self-organization of Co3O4 nanowires on Ni foam for high performance supercapacitors: from nanowire microspheres to nanowire paddy fields. Nanoscale 2014;6:3638–46. [400] Peng X, Peng L, Wu C, Xie Y. Two dimensional nanomaterials for flexible supercapacitors. Chem Soc Rev 2014;43:3303–23. [401] Koski KJ, Cui Y. The new skinny in two-dimensional nanomaterials. ACS Nano 2013;7:3739–43. [402] Zhang X, Xie Y. Recent advances in free-standing two-dimensional crystals with atomic thickness: design, assembly and transfer strategies. Chem Soc Rev 2013;42:8187–99. [403] Chimene D, Alge DL, Gaharwar AK. Two-dimensional nanomaterials for biomedical applications: emerging trends and future prospects. Adv Mater 2015;27:7261–84. [404] Tan C, Zhang H. Two-dimensional transition metal dichalcogenide nanosheet-based composites. Chem Soc Rev 2015;44:2713–31. [405] Mei J, Liao T, Kou L, Sun Z. Two-dimensional metal oxide nanomaterials for next-generation rechargeable batteries. Adv Mater 2017;29:1700176. [406] Rao CNR, Sood AK, Subrahmanyam KS, Govindaraj A. Graphene: the new two-dimensional nanomaterial. Angew Chem Int Edit 2009;48:7752–77. [407] Xu M, Liang T, Shi M, Chen H. Graphene-like two-dimensional materials. Chem Rev 2013;113:3766–98. [408] Liu B, Zhang JG, Shen G. Pursuing two-dimensional nanomaterials for flexible lithium-ion batteries. Nano Today 2016;11:82–97. [409] Sheng L, Liao T, Kou L, Sun Z. Single-crystalline ultrathin 2D TiO2 nanosheets: a bridge towards superior photovoltaic devices. Mater Today Energy 2017;3:32–9. [410] Dou Y, Tian D, Sun Z, Liu Q, Zhang N, Kim JH, et al. Fish gill inspired crossflow for efficient and continuous collection of spilled oil. ACS Nano 2017;11:2477–85. [411] Liao T, Sun Z, Dou SX. Theoretically manipulating quantum dots on two-dimensional TiO2 monolayer for effective visible light absorption. ACS Appl Mater Interfaces 2017;9:8255–62. [412] Miró P, Audiffred M, Heine T. An atlas of two-dimensional materials. Chem Soc Rev 2014;43:6537–54. [413] Tan C, Cao X, Wu X-J, He Q, Yang J, Zhang X, et al. Recent advances in ultrathin two-dimensional nanomaterials. Chem Rev 2017;117:6225–331. [414] Zhang H. Ultrathin two-dimensional nanomaterials. ACS Nano 2015;9:9451–69. [415] Tan C, Zhang H. Wet-chemical synthesis and applications of non-layer structured two-dimensional nanomaterials. Nat Commun 2015;6:7873. [416] Li Z, Yu X-Y, Paik U. Facile preparation of porous Co3O4 nanosheets for high-performance lithium ion batteries and oxygen evolution reaction. J Power Sources 2016;310:41–6. [417] Eom W, Kim A, Park H, Kim H, Han TH. Graphene-mimicking 2D porous Co3O4 nanofoils for lithium battery applications. Adv Funct Mater 2016;26:7605–13. [418] Meher SK, Rao GR. Ultralayered Co3O4 for high-performance supercapacitor applications. J Phys Chem C 2011;115:15646–54. [419] Yang Q, Lu Z, Sun X, Liu J. Ultrathin Co3O4 nanosheet arrays with high supercapacitive performance. Sci Rep 2013;3:3537. [420] ten Elshof JE, Yuan H, Gonzalez Rodriguez P. Two-dimensional metal oxide and metal hydroxide nanosheets: synthesis, controlled assembly and applications in energy conversion and storage. Adv Energy Mater 2016;6:1600355. [421] Wang X, Guan H, Chen S, Li H, Zhai T, Tang D, et al. Self-stacked Co3O4 nanosheets for high-performance lithium ion batteries. Chem Commun 2011;47:12280–2. [422] Wang B, Lu X-Y, Tang Y, Ben W. General polyethyleneimine-mediated synthesis of ultrathin hexagonal Co3O4 nanosheets with reactive facets for lithium-ion batteries. ChemElectroChem 2016;3:55–65. [423] He M, Zhang P, Xu S, Yan X. Morphology engineering of Co3O4 nanoarrays as free-standing catalysts for lithium-oxygen batteries. ACS Appl Mater Interfaces 2016;8:23713–20. [424] Pan A, Wang Y, Xu W, Nie Z, Liang S, Nie Z, et al. High-performance anode based on porous Co3O4 nanodiscs. J Power Sources 2014;255:125–9. [425] Zhan F, Geng B, Guo Y. Porous Co3O4 nanosheets with extraordinarily high discharge capacity for lithium batteries. Chem Eur J 2009;15:6169–74. [426] Wang B, Lu X-Y, Tang Y. Synthesis of snowflake-shaped Co3O4 with a high aspect ratio as a high capacity anode material for lithium ion batteries. J Mater Chem
668
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
A 2015;3:9689–99. [427] Deori K, Ujjain SK, Sharma RK, Deka S. Morphology controlled synthesis of nanoporous Co3O4 nanostructures and their charge storage characteristics in supercapacitors. ACS Appl Mater Interfaces 2013;5:10665–72. [428] Sun Z, Liao T, Dou Y, Hwang SM, Park M-S, Jiang L, et al. Generalized self-assembly of scalable two-dimensional transition metal oxide nanosheets. Nat Commun 2014;5:3813. [429] Zhu J, Bai L, Sun Y, Zhang X, Li Q, Cao B, et al. Topochemical transformation route to atomically thick Co3O4 nanosheets realizing enhanced lithium storage performance. Nanoscale 2013;5:5241–6. [430] Li D, Ding L-X, Wang S, Cai D, Wang H. Ultrathin and highly-ordered CoO nanosheet arrays for lithium-ion batteries with high cycle stability and rate capability. J Mater Chem A 2014;2:5625–30. [431] Polshettiwar V, Baruwati B, Varma RS. Self-assembly of metal oxides into three-dimensional nanostructures: synthesis and application in catalysis. ACS Nano 2009;3:728–36. [432] Sun Z, Liao T, Sheng L, Kim JH, Dou SX, Bell J. Fly compound-eye inspired inorganic nanostructures with extraordinary visible-light responses. Mater Today Chem 2016;1–2:84–9. [433] Sun Z, Liao T, Sheng L, Kou L, Kim JH, Dou SX. Deliberate design of TiO2 nanostructures towards superior photovoltaic cells. Chem Eur J 2016;22:11357–64. [434] Liu Q, Dou Y, Ruan B, Sun Z, Chou S-L, Dou SX. Carbon-coated hierarchical SnO2 hollow spheres for lithium ion batteries. Chem Eur J 2016;22:5853–7. [435] Sun Z, Kim JH, Zhao Y, Attard D, Dou SX, Ho J, et al. Morphology-controllable 1D–3D nanostructured TiO2 bilayer photoanodes for dye-sensitized solar cells. Chem Commun 2013;49:966–8. [436] Jadhav HS, Rai AK, Lee JY, Kim J, Park CJ. Enhanced electrochemical performance of flower-like Co3O4 as an anode material for high performance lithium-ion batteries. Electrochim Acta 2014;146:270–7. [437] Xiong S, Yuan C, Zhang X, Xi B, Qian Y. Controllable synthesis of mesoporous Co3O4 nanostructures with tunable morphology for application in supercapacitors. Chem Eur J 2009;15:5320–6. [438] Pang H, Gao F, Chen Q, Liu R, Lu Q. Dendrite-like Co3O4 nanostructure and its applications in sensors, supercapacitors and catalysis. Dalt Trans 2012;41:5862–8. [439] Yang W, Gao Z, Ma J, Wang J, Wang B, Liu L. Effects of solvent on the morphology of nanostructured Co3O4 and its application for high-performance supercapacitors. Electrochim Acta 2013;112:378–85. [440] Kwak JH, Lee YW, Bang JH. Supercapacitor electrode with an ultrahigh Co3O4 loading for a high areal capacitance. Mater Lett 2013;110:237–40. [441] Luo F, Li J, Lei Y, Yang W, Yuan H, Xiao D. Three-dimensional enoki mushroom-like Co3O4 hierarchitectures constructed by one-dimension nanowires for highperformance supercapacitors. Electrochim Acta 2014;135:495–502. [442] Yang Q, Lu ZY, Chang Z, Zhu W, Sun JQ, Liu JF, et al. Hierarchical Co3O4 nanosheet@nanowire arrays with enhanced pseudocapacitive performance. RSC Adv 2012;2:1663–8. [443] Sun Z, Kim JH, Zhao Y, Bijarbooneh F, Malgras V, Lee Y, et al. Rational design of 3D dendritic TiO2 nanostructures with favorable architectures. J Am Chem Soc 2011;133:19314–7. [444] Cao X, Yin Z, Zhang H. Three-dimensional graphene materials: preparation, structures and application in supercapacitors. Energy Environ Sci 2014;7:1850–65. [445] Fattakhova-rohl D, Zaleska A, Bein T. Three-dimensional titanium dioxide nanomaterials. Chem Rev 2014;114:9487–558. [446] Wu Z, Sun Y, Tan Y, Yang S. Three-dimensional graphene-based macro- and mesoporous frameworks for high-performance electrochemical capacitive energy storage. J Am Chem Soc 2012;134:19532–5. [447] Che H, Liu A, Liang S, Zhang X, Mu J, Bai Y, et al. Superlattices and microstructures facile synthesis of three-dimensional hierarchical Co3O4 peony-like microspheres and their lithium storage performance. Superlattices Microstruct 2015;83:538–48. [448] Xiao A, Zhou S, Zuo C, Zhuan Y, Ding X. Controllable synthesis of mesoporous Co3O4 nanoflake array and its application for supercapacitor. Mater Res Bull 2014;60:674–8. [449] Liu R, Jiang Z, Liu Q, Zhu X, Chen W. Cu2+ ions inducing the growth of porous Co3O4 nanospheres as high-capacity supercapacitors. CrystEngComm 2015;17:4449–54. [450] Cheng JP, Chen X, Wu J-S, Liu F, Zhang XB, Dravid VP. Porous cobalt oxides with tunable hierarchical morphologies for supercapacitor electrodes. CrystEngComm 2012;14:6702–9. [451] Van Gough D, Juhl AT, Braun PV. Programming structure into 3D nanomaterials. Mater Today 2009;12:28–35. [452] Sun H, Xin G, Hu T, Yu M, Shao D, Sun X, et al. High-rate lithiation-induced reactivation of mesoporous hollow spheres for long-lived lithium-ion batteries. Nat Commun 2014;5:4526. [453] Xia XH, Tu JP, Wang XL, Gu CD, Zhao XB. Mesoporous Co3O4 monolayer hollow-sphere array as electrochemical pseudocapacitor material. Chem Commun 2011;47:5786–8. [454] Wang L, Liu X, Wang X, Yang X, Lu L. Preparation and electrochemical properties of mesoporous Co3O4 crater-like microspheres as supercapacitor electrode materials. Curr Appl Phys 2010;10:1422–6. [455] Wang J, Yang N, Tang H, Dong Z, Jin Q, Yang M, et al. Accurate control of multishelled Co3O4 hollow microspheres as high-performance anode materials in lithium-ion batteries. Angew Chem 2013;125:6545–8. [456] Wang X, Wu XL, Guo YG, Zhong Y, Cao X, Ma Y, et al. Synthesis and lithium storage properties of Co3O4 nanosheet-assembled multishelled hollow spheres. Adv Funct Mater 2010;20:1680–6. [457] Yan N, Hu L, Li Y, Wang Y, Zhong H, Hu X, et al. Co3O4 nanocages for high-performance anode material in lithium-ion batteries. J Phys Chem C 2012;116:7227–35. [458] Ma J, Manthiram A. Precursor-directed formation of hollow Co3O4 nanospheres exhibiting superior lithium storage properties. RSC Adv 2012;2:3187. [459] Hu L, Yan N, Chen Q, Zhang P, Zhong H, Zheng X, et al. Fabrication based on the Kirkendall effect of Co3O4 porous nanocages with extraordinarily high capacity for lithium storage. Chem Eur J 2012;18:8971–7. [460] Duan BR, Cao Q. Hierarchically porous Co3O4 film prepared by hydrothermal synthesis method based on colloidal crystal template for supercapacitor application. Electrochim Acta 2012;64:154–61. [461] Shim H-W, Lim A-H, Kim J-C, Jang E, Seo S-D, Lee G-H, et al. Scalable one-pot bacteria-templating synthesis route toward hierarchical, porous-Co3O4 superstructures for supercapacitor electrodes. Sci Rep 2013;3:2325. [462] Zhou X, Shen X, Xia Z, Zhang Z, Li J, Ma Y, et al. Hollow fluffy Co3O4 cages as efficient electroactive materials for supercapacitors and oxygen evolution reaction. ACS Appl Mater Interfaces 2015;7:20322–31. [463] Du H, Jiao L, Wang Q, Yang J, Guo L, Si Y, et al. Facile carbonaceous microsphere templated synthesis of Co3O4 hollow spheres and their electrochemical performance in supercapacitors. Nano Res 2013;6:87–98. [464] Wu H, Chan G, Choi JW, Ryu I, Yao Y, McDowell MT, et al. Stable cycling of double-walled silicon nanotube battery anodes through solid–electrolyte interphase control. Nat Nanotechnol 2012;7:310–5. [465] Guo J, Chen L, Zhang X, Jiang B, Ma L. Sol-gel synthesis of mesoporous Co3O4 octahedra toward high-performance anodes for lithium-ion batteries. Electrochim Acta 2014;129:410–5. [466] Lee HT, Kwon S, Youn CM, Choi T, Lee JH. Topochemical reaction of exfoliated layered cobalt(II) hydroxide for the synthesis of ultrapure Co3O4 as an oxygen reduction catalyst. Eur J Inorg Chem 2017;2017:2184–9. [467] Xu GL, Li JT, Huang L, Lin W, Sun SG. Synthesis of Co3O4 nano-octahedra enclosed by 111 facets and their excellent lithium storage properties as anode material of lithium ion batteries. Nano Energy 2013;2:394–402. [468] Chen JS, Zhu T, Hu QH, Gao J, Su F, Qiao SZ, et al. Shape-controlled synthesis of cobalt-based nanocubes, nanodiscs, and nanoflowers and their comparative lithium-storage properties. ACS Appl Mater Interfaces 2010;2:3628–35. [469] Liu X, Long Q, Jiang C, Zhan B, Li C, Liu S, et al. Facile and green synthesis of mesoporous Co3O4 nanocubes and their applications for supercapacitors.
669
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
Nanoscale 2013;5:6525–9. [470] Cao Y, Yuan F, Yao M, Bang JH, Lee J-H. A new synthetic route to hollow Co3O4 octahedra for supercapacitor applications. CrystEngComm 2014;16:826–33. [471] Zhang Y-Z, Wang Y, Xie Y-L, Cheng T, Lai W-Y, Pang H, et al. Porous hollow Co3O4 with rhombic dodecahedral structures for high-performance supercapacitors. Nanoscale 2014;6:14354–9. [472] Song H, Shen L, Wang C. Template-free method towards quadrate Co3O4 nanoboxes from cobalt coordination polymer nano-solids for high performance lithium ion battery anodes. J Mater Chem A 2014;2:20597–604. [473] Huang G, Xu S, Lu S, Li L, Sun H. Micro-/nanostructured Co3O4 anode with enhanced rate capability for lithium-ion batteries. ACS Appl Mater Interfaces 2014;6:7236–43. [474] Zhang J, Lyu Z, Zhang F, Wang L, Xiao P, Yuan K, et al. Facile synthesis of hierarchical porous Co3O4 nanoboxes as efficient cathode catalysts for Li–O2 batteries. J Mater Chem A 2016;4:6350–6. [475] Wang X, Yu L, Wu X-L, Yuan F, Guo Y-G, Ma Y, et al. Synthesis of single-crystalline Co3O4 octahedral cages with tunable surface aperture and their lithium storage properties. J Phys Chem C 2009;113:15553–8. [476] Xiao X, Liu X, Zhao H, Chen D, Liu F, Xiang J, et al. Facile shape control of Co3O4 and the effect of the crystal plane on electrochemical performance. Adv Mater 2012;24:5762–6. [477] Gao R, Zhu J, Xiao X, Hu Z, Liu J, Liu X. Facet-dependent electrocatalytic performance of Co3O4 for rechargeable Li-O2 battery. J Phys Chem C 2015;119:4516–23. [478] Chen Z, Kronawitter CX, Koel BE. Facet-dependent activity and stability of Co3O4 nanocrystals towards the oxygen evolution reaction. Phys Chem Chem Phys 2015;17:29387–93. [479] Ding YL, Wen Y, Wu C, Van Aken PA, Maier J, Yu Y. 3D V6O13 nanotextiles assembled from interconnected nanogrooves as cathode materials for high-energy lithium ion batteries. Nano Lett 2015;15:1388–94. [480] Niu C, Liu X, Meng J, Xu L, Yan M, Wang X, et al. Three dimensional V2O5/NaV6O15 hierarchical heterostructures: controlled synthesis and synergistic effect investigated by in situ X-ray diffraction. Nano Energy 2016;27:147–56. [481] Jung SM, Jung HY, Dresselhaus MS, Jung YJ, Kong J. A facile route for 3D aerogels from nanostructured 1D and 2D materials. Sci Rep 2012;2:849. [482] Alenezi MR, Henley SJ, Emerson NG, Silva SRP. From 1D and 2D ZnO nanostructures to 3D hierarchical structures with enhanced gas sensing properties. Nanoscale 2014;6:235–47. [483] Suber L, Plunkett WR. Formation mechanism of silver nanoparticle 1D microstructures and their hierarchical assembly into 3D superstructures. Nanoscale 2010;2:128–33. [484] Sun Z, Liao T, Liu K, Jiang L, Kim JH, Dou SX. Fly-eye inspired superhydrophobic anti-fogging inorganic nanostructures. Small 2014;10:3001–6. [485] Sun ZQ, Liao T, Kim JGH, Liu KS, Jiang L, Kim JGH, et al. Architecture designed ZnO hollow microspheres with wide-range visible-light photoresponses. J Mater Chem C 2013;1:6924–9. [486] Sun Z, Liao T, Liu K, Jiang L, Kim JH, Dou SX. Robust superhydrophobicity of hierarchical ZnO hollow microspheres fabricated by two-step self-assembly. Nano Res 2013;6:726–35. [487] Liu Q, Sun Z, Dou Y, Kim JH, Dou SX. Two-step self-assembly of hierarchically-ordered nanostructures. J Mater Chem A 2015;3:11688–99. [488] Wang L, Liu B, Ran S, Huang H, Wang X, Liang B, et al. Nanorod-assembled Co3O4 hexapods with enhanced electrochemical performance for lithium-ion batteries. J Mater Chem 2012;22:23541–6. [489] Qing X, Liu S, Huang K, Lv K, Yang Y, Lu Z, et al. Facile synthesis of Co3O4 nanoflowers grown on Ni foam with superior electrochemical performance. Electrochim Acta 2011;56:4985–91. [490] Liu P, Hao Q, Xia X, Lu L, Lei W, Wang X. 3D hierarchical mesoporous flowerlike cobalt oxide nanomaterials: controllable synthesis and electrochemical properties. J Phys Chem C 2015;119:8537–46. [491] Li L, Seng KH, Chen Z, Guo Z, Liu HK. Self-assembly of hierarchical star-like Co3O4 micro/nanostructures and their application in lithium ion batteries. Nanoscale 2013;5:1922–8. [492] Hao W, Chen S, Cai Y, Zhang L, Li Z, Zhang S. Three-dimensional hierarchical pompon-like Co3O4 porous spheres for high-performance lithium-ion batteries. J Mater Chem A 2014;2:13801–4. [493] Li H-H, Li Z-Y, Wu X-L, Zhang L-L, Fan C-Y, Wang H-F, et al. Shale-like Co3O4 for high performance lithium/sodium ion batteries. J Mater Chem A 2016;4:8242–8. [494] Qiu K, Yan H, Zhang D, Lu Y, Cheng J, Zhao W, et al. Hierarchical 3D mesoporous conch-like Co3O4 nanostructure arrays for high-performance supercapacitors. Electrochim Acta 2014;141:248–54. [495] Zhao W, Zhou X, Kim IJ, Kim S. Self-assembled Co3O4 hexagonal plates by solvent engineering and their dramatically enhanced electrochemical performance. Nanoscale 2017;9:940–6. [496] Chou SL, Wang JZ, Liu HK, Dou SX. Electrochemical deposition of porous Co3O4 nanostructured thin film for lithium-ion battery. J Power Sources 2008;182:359–64. [497] Zhu S, Li J, Deng X, He C, Liu E, He F, et al. Ultrathin-nanosheet-induced synthesis of 3D transition metal oxides networks for lithium ion battery anodes. Adv Funct Mater 2017;27:1605017. [498] Needham SA, Wang GX, Konstantinov K, Tournayre Y, Lao Z, Liu HK. Electrochemical performance of Co3O4–C composite anode materials. Electrochem SolidState Lett 2006;9:A315–9. [499] Uhlig LM, Stewart D, Sievers G, Brüser V, Dyck A, Wittstock G. Electrochemical characterization of temperature dependence of plasma-treated cobalt-oxide catalyst for oxygen reduction reaction in alkaline media. Int J Hydrogen Energy 2016;41:22554–9. [500] Pan GX, Xia XH, Cao F, Chen J, Zhang YJ. Construction of Co/Co3O4-C ternary core-branch arrays as enhanced anode materials for lithium ion batteries. J Power Sources 2015;293:585–91. [501] Zhu G, Jiao B-J, Li J-T. Synthesis of hierarchical Co3O4@C composite and its lithium storage property as anode material for rechargeable lithium ion batteries. Mater Res Innov 2014;18:528–34. [502] Zhou K, Lai L, Zhen Y, Hong Z, Guo J, Huang Z. Rational design of Co3O4/Co/carbon nanocages composites from metal organic frameworks as an advanced lithium-ion battery anode. Chem Eng J 2017;316:137–45. [503] Liu W, Yang H, Zhao L, Liu S, Wang H, Chen S. Mesoporous flower-like Co3O4/C nanosheet composites and their performance evaluation as anodes for lithium ion batteries. Electrochim Acta 2016;207:293–300. [504] Marzuki NS, Taib NU, Hassan MF, Idris NH. Enhanced lithium storage in Co3O4/carbon anode for Li-ion batteries. Electrochim Acta 2015;182:452–7. [505] Zhang P, Guo ZP, Huang Y, Jia D, Liu HK. Synthesis of Co3O4/carbon composite nanowires and their electrochemical properties. J Power Sources 2011;196:6987–91. [506] Liu J, Jiang L, Tang Q, Zhang B, Su DS, Wang S, et al. Coupling effect between cobalt oxides and carbon for oxygen reduction reaction. ChemSusChem 2012;5:2315–8. [507] Park SH, Lee WJ. Hierarchically mesoporous flower-like cobalt oxide/carbon nanofiber composites with shell-core structure as anodes for lithium ion batteries. Carbon 2015;89:197–207. [508] Zhou S, Wang G, Xie Y, Wang H, Bai J. Synthesis of carbon-coated Co3O4 composite with dendrite-like morphology and its electrochemical performance for lithium-ion batteries. J Nanopart Res 2013;15:1740. [509] Liu X, Or SW, Jin C, Lv Y, Li W, Feng C, et al. Co3O4/C nanocapsules with onion-like carbon shells as anode material for lithium ion batteries. Electrochim Acta 2013;100:140–6. [510] Yu BC, Lee JO, Song JH, Park CM, Lee CK, Sohn HJ. Nanostructured cobalt oxide-based composites for rechargeable Li-ion batteries. J Solid State Electrochem 2012;16:2631–8. [511] Chaikittisilp W, Torad NL, Li C, Imura M, Suzuki N, Ishihara S, et al. Synthesis of nanoporous carbon-cobalt-oxide hybrid electrocatalysts by thermal conversion
670
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
of metal-organic frameworks. Chem Eur J 2014;20:4217–21. [512] Xiong S, Chen JS, Lou XW, Zeng HC. Mesoporous Co3O4 and CoO@C topotactically transformed from chrysanthemum-like Co(CO3)0.5(OH) 0.11H2O and their lithium-storage properties. Adv Funct Mater 2012;22:861–71. [513] Yang Y, Sun Q, Li YS, Li H, Fu ZW. A CoOx/carbon double-layer thin film air electrode for nonaqueous Li-air batteries. J Power Sources 2013;223:312–8. [514] Han Y, Dong L, Feng J, Li D, Li X, Liu S. Cobalt oxide modified porous carbon anode enhancing electrochemical performance for Li-ion batteries. Electrochim Acta 2015;167:246–53. [515] Park CS, Kim KS, Park YJ. Carbon-sphere/Co3O4 nanocomposite catalysts for effective air electrode in Li/air batteries. J Power Sources 2013;244:72–9. [516] Silva R, Pereira GM, Voiry D, Chhowalla M, Asefa T. Co3O4 nanoparticles/cellulose nanowhiskers-derived amorphous carbon nanoneedles: sustainable materials for supercapacitors and oxygen reduction electrocatalysis. RSC Adv 2015;5:49385–91. [517] Kim DS, Kim SB. Electrochemical performance of surface modified CNF/Co3O4 composite for Li-air batteries. J Electroceramics 2014;33:246–51. [518] Fu B, Zhou X, Wang Y. Co3O4 carbon nanofiber mats as negative electrodes for sodium-ion batteries. Mater Lett 2016;170:21–4. [519] Kim DS, Park YJ. Buckypaper electrode containing carbon nanofiber/Co3O4 composite for enhanced lithium air batteries. Solid State Ionics 2014;268:216–21. [520] Hao F, Zhang Z, Yin L. Co3O4/carbon aerogel hybrids as anode materials for lithium-ion batteries with enhanced electrochemical properties. ACS Appl Mater Interfaces 2013;5:8337–44. [521] Wang H, Mao N, Shi J, Wang Q, Yu W, Wang X. Cobalt oxide-carbon nanosheet nanoarchitecture as an anode for high-performance lithium-ion battery. ACS Appl Mater Interfaces 2015;7:2882–90. [522] Song MJ, Kim IT, Kim YB, Shin MW. Self-standing, binder-free electrospun Co3O4/carbon nanofiber composites for non-aqueous Li-air batteries. Electrochim Acta 2015;182:289–96. [523] Zhi J, Deng S, Zhang YX, Wang YF, Hu AG. Embedding Co3O4 nanoparticles in SBA-15 supported carbon nanomembrane for advanced supercapacitor materials. J Mater Chem A 2013;1:3171–6. [524] Chen Z, He D, Xu X, Liu Z, Huang M, Wang X, et al. Coupled cobalt oxide/hollow carbon sphere as an efficient electrocatalyst for the oxygen reduction reaction. RSC Adv 2016;6:34159–64. [525] Zhang F, Yuan C, Zhu J, Wang J, Zhang X, Lou XW. Flexible films derived from electrospun carbon nanofibers incorporated with Co3O4 hollow nanoparticles as self-supported electrodes for electrochemical capacitors. Adv Funct Mater 2013;23:3909–15. [526] Abouali S, Garakani MA, Zhang B, Luo H, Xu Z. Co3O4/porous electrospun carbon nanofibers as anodes for high performance Li-ion batteries. J Mater Chem A 2014;2:16939–44. [527] Abouali S, Akbari Garakani M, Zhang B, Xu Z-L, Kamali Heidari E, Huang J, et al. Electrospun carbon nanofibers with in situ encapsulated Co3O4 nanoparticles as electrodes for high-performance supercapacitors. ACS Appl Mater Interfaces 2015;7:13503–11. [528] Leng X, Wei S, Jiang Z, Lian J, Wang G, Jiang Q. Carbon-encapsulated Co3O4 nanoparticles as anode materials with super lithium storage performance. Sci Rep 2015;5:16629. [529] Song MJ, Kim IT, Kim YB, Kim J, Shin MW. Metal-organic frameworks-derived porous carbon/Co3O4 composites for rechargeable lithium-oxygen batteries. Electrochim Acta 2017;230:73–80. [530] Li D, Yang D, Zhu X, Jing D, Xia Y, Ji Q, et al. Simple pyrolysis of cobalt alginate fibres into Co3O4/C nano/microstructures for a high-performance lithium ion battery anode. J Mater Chem A 2014;2:18761–6. [531] Sun J, Li D, Xia Y, Zhu X, Zong L, Ji Q, et al. Co3O4 nanoparticle embedded carbonaceous fibres: a nanoconfinement effect on enhanced lithium-ion storage. Chem Commun 2015;51:16267–70. [532] Wang Y, Zhang HJ, Lu L, Stubbs LP, Wong CC, Lin J. Designed functional systems from peapod-like Co@carbon to Co3O4@carbon nanocomposites. ACS Nano 2010;4:4753–61. [533] Gu D, Li W, Wang F, Bongard H, Spliethoff B, Schmidt W, et al. Controllable synthesis of mesoporous Peapod-like Co3O4/@carbon nanotube arrays for highperformance lithium-ion batteries. Angew Chem 2015;127:7166–70. [534] Pan GX, Xia XH, Cao F, Chen J, Tang PS, Zhang YJ, et al. High-performance asymmetric supercapacitors based on core/shell cobalt oxide/carbon nanowire arrays with enhanced electrochemical energy storage. Electrochim Acta 2014;133:522–8. [535] Chen J, Xia X, Tu J, Xiong Q, Yu Y-X, Wang X, et al. Co3O4–C core–shell nanowire array as an advanced anode material for lithium ion batteries. J Mater Chem 2012;22:15056–61. [536] Tan YL, Gao QM, Yang CX, Yang K, Tian WQ, Zhu LH. One-dimensional porous nanofibers of Co3O4 on the carbon matrix from human hair with superior lithium ion storage performance. Sci Rep 2015;5:12382. [537] Yan C, Chen G, Zhou X, Sun J, Lv C. Template-based engineering of carbon-doped Co3O4 hollow nanofibers as anode materials for lithium-ion batteries. Adv Funct Mater 2016;26:1428–36. [538] Sun J, Liu H, Chen X, Evans DG, Yang W. An oil droplet template method for the synthesis of hierarchical structured Co3O4/C anodes for Li-ion batteries. Nanoscale 2013;5:7564–71. [539] Wang J, Gao B, Zhang L, Li R, Shen J, Qiao Z, et al. Controlled synthesis of porous Co3O4-C hybrid nanosheet arrays and their application in lithium ion batteries. RSC Adv 2014;4:30573–8. [540] Yu YX, Liu XY, Xia XH, Xiong QQ, Wang XL, Gu CD, et al. Net-structured Co3O4/C nanosheet array with enhanced electrochemical performance toward lithium storage. Mater Res Bull 2014;51:112–8. [541] Zhong Y, Wang X, Jiang K, Zheng JY, Guo Y, Ma Y, et al. A facile synthesis and lithium storage properties of Co3O4–C hybrid core-shell and hollow spheres. J Mater Chem 2011;21:17998–8002. [542] Jayaprakash N, Jones WD, Moganty SS, Archer LA. Composite lithium battery anodes based on carbon@Co3O4 nanostructures: synthesis and characterization. J Power Sources 2012;200:53–8. [543] Liu J, Wan Y, Liu C, Liu W, Ji S, Zhou Y, et al. Solvothermal synthesis of uniform Co3O4/C hollow quasi-nanospheres for enhanced lithium ion intercalation applications. Eur J Inorg Chem 2012;2012:3825–9. [544] Shen L, Wang C. Hierarchical Co3O4 nanoparticles embedded in a carbon matrix for lithium-ion battery anode materials. Electrochim Acta 2014;133:16–22. [545] Lan D, Chen Y, Chen P, Chen X, Wu X, Pu X, et al. Mesoporous CoO nanocubes @ continuous 3D porous carbon skeleton of rose-based electrode for highperformance supercapacitor. ACS Appl Mater Interfaces 2014;6:11839–45. [546] Wang L, Zheng Y, Wang X, Chen S, Xu F. Nitrogen-doped porous carbon/Co3O4 nanocomposites as anode materials for lithium ion batteries. ACS Appl Mater Interfaces 2014;6:7117–25. [547] Hu W, Wang Q, Wu S, Huang Y. Facile one-pot synthesis of nitrogen-doped mesoporous carbon architecture with cobalt oxides encapsulated in graphitic layers as a robust bicatalyst for oxygen reduction and evolution reactions. J Mater Chem A 2016;4:16920–7. [548] Wang K, Wang R, Li H, Wang H, Mao X, Linkov V, et al. N-doped carbon encapsulated Co3O4 nanoparticles as a synergistic catalyst for oxygen reduction reaction in acidic media. Int J Hydrogen Energy 2015;40:3875–82. [549] Liu K, Zhou Z, Wang H, Huang X, Xu J, Tang Y, et al. N-doped carbon supported Co3O4 nanoparticles as an advanced electrocatalyst for the oxygen reduction reaction in Al-air batteries. RSC Adv 2016;6:55552–9. [550] Liu S, Li L, Ahnb HS, Manthiram A. Delineating the roles of Co3O4 and N-doped carbon nanoweb (CNW) in bifunctional Co3O4/CNW catalysts for oxygen reduction and oxygen evolution reactions. J Mater Chem A 2015;3:11615–23. [551] Hou Y, Li J, Wen Z, Cui S, Yuan C, Chen J. Co3O4 nanoparticles embedded in nitrogen-doped porous carbon dodecahedrons with enhanced electrochemical properties for lithium storage and water splitting. Nano Energy 2015;12:1–8. [552] Salunkhe RR, Tang J, Kamachi Y, Nakato T, Kim JH, Yamauchi Y. Asymmetric supercapacitors using 3D nanoporous carbon and cobalt oxide electrodes synthesized from a single metal-organic framework. ACS Nano 2015;9:6288–96. [553] Lee KJ, Kim T-H, Kim TK, Lee JH, Song H-K, Moon HR. Preparation of Co3O4 electrode materials with different microstructures via pseudomorphic conversion of Co-based metal–organic frameworks. J Mater Chem A 2014;2:14393–400.
671
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
[554] Han Y, Zhao M, Dong L, Feng J, Wang Y, Li D, et al. MOF-derived porous hollow Co3O4 parallelepipeds for building high-performance Li-ion batteries. J Mater Chem A 2015;3:22542–6. [555] Li C, Chen T, Xu W, Lou X, Pan L, Chen Q, et al. Mesoporous nanostructured Co3O4 derived from MOF template: a high-performance anode material for lithiumion batteries. J Mater Chem A 2015;3:5585–91. [556] Chaikittisilp W, Ariga K, Yamauchi Y. A new family of carbon materials: synthesis of MOF-derived nanoporous carbons and their promising applications. J Mater Chem A 2013;1:14–9. [557] Shen K, Chen X, Chen J, Li Y. Development of MOF-derived carbon-based nanomaterials for efficient catalysis. ACS Catal 2016;6:5887–903. [558] Zheng Y, Li Z, Xu J, Wang T, Liu X, Duan X, et al. Multi-channeled hierarchical porous carbon incorporated Co3O4 nanopillar arrays as 3D binder-free electrode for high performance supercapacitors. Nano Energy 2016;20:94–107. [559] Xu J, Yu Q, Wu C, Guan L. Oxygen reduction electrocatalysts based on spatially confined cobalt monoxide nanocrystals on holey N-doped carbon nanowires: the enlarged interfacial area for performance improvement. J Mater Chem A 2015;3:21647–54. [560] Li L, Liu S, Manthiram A. Co3O4 nanocrystals coupled with O- and N-doped carbon nanoweb as a synergistic catalyst for hybrid Li-air batteries. Nano Energy 2015;12:852–60. [561] Kim H, Kim Y, Noh Y, Lee S, Sung J, Kim WB. Thermally converted CoO nanoparticles embedded into N-doped carbon layers as highly efficient bifunctional electrocatalysts for oxygen reduction and oxygen evolution reactions. ChemCatChem 2017;9:1503–10. [562] Tan Y, Xu C, Chen G, Fang X, Zheng N, Xie Q. Facile synthesis of manganese-oxide-containing mesoporous nitrogen-doped carbon for efficient oxygen reduction. Adv Funct Mater 2012;22:4584–91. [563] Huang Q, Zhou P, Yang H, Zhu L, Wu H. CoO nanosheets in situ grown on nitrogen-doped activated carbon as an effective cathodic electrocatalyst for oxygen reduction reaction in microbial fuel cells. Electrochim Acta 2017;232:339–47. [564] Zhang C, Antonietti M, Fellinger TP. Blood ties: Co3O4 decorated blood derived carbon as a superior bifunctional electrocatalyst. Adv Funct Mater 2014;24:7655–65. [565] Liu K, Huang X, Wang H, Li F, Tang Y, Li J, et al. Co3O4-CeO2/C as a highly active electrocatalyst for oxygen reduction reaction in Al–air batteries. ACS Appl Mater Interfaces 2016;8:34422–30. [566] Sultana I, Rahman MM, Mateti S, Ahmadabadi VG, Glushenkov AM, Chen Y. K-ion and Na-ion storage performances of Co3O4–Fe2O3 nanoparticle-decorated super P carbon black prepared by a ball milling process. Nanoscale 2017;9:3646–54. [567] Wang Y, Ma X, Lu L, He Y, Qi X, Deng Y. Carbon supported MnOx-Co3O4 as cathode catalyst for oxygen reduction reaction in alkaline media. Int J Hydrogen Energy 2013;38:13611–6. [568] Wang Y, Lu X, Liu Y, Deng Y. Silver supported on Co3O4 modified carbon as electrocatalyst for oxygen reduction reaction in alkaline media. Electrochem Commun 2013;31:108–11. [569] Ahmed J, Yuan Y, Zhou L, Kim S. Carbon supported cobalt oxide nanoparticles-iron phthalocyanine as alternative cathode catalyst for oxygen reduction in microbial fuel cells. J Power Sources 2012;208:170–5. [570] Park CS, Kim JH, Park YJ. Catalytic activity of carbon-sphere/Co3O4/RuO2 nanocomposite for Li-air batteries. J Electroceramics 2013;31:224–30. [571] Kong D, Cheng C, Wang Y, Wong JI, Yang Y, Yang HY, et al. Three-dimensional Co3O4@C@Ni3S2 sandwich-structured nanoneedle arrays: towards highperformance flexible all-solid-state asymmetric supercapacitors. J Mater Chem A 2015;3:16150–61. [572] Zhao S, Li C, Liu J, Liu N, Qiao S, Han Y, et al. Carbon quantum dots/SnO2-Co3O4 composite for highly efficient electrochemical water oxidation. Carbon 2015;92:64–73. [573] Wang X, Zhang J, Kong X, Huang X, Shi B. Increasing rigidness of carbon coating for improvement of electrochemical performances of Co3O4 in Li-ion batteries. Carbon 2016;104:1–9. [574] De Volder MFL, Tawfick SH, Baughman RH, Hart AJ. Carbon nanotubes: present and future commercial applications. Science 2013;339:535–9. [575] Datsyuk V, Kalyva M, Papagelis K, Parthenios J, Tasis D, Siokou A, et al. Chemical oxidation of multiwalled carbon nanotubes. Carbon 2008;46:833–40. [576] Landi BJ, Ganter MJ, Cress CD, DiLeo Ra, Raffaelle RP. Carbon nanotubes for lithium ion batteries. Energy Environ Sci 2009;2:638–54. [577] Li L, Yang H, Zhou D, Zhou Y. Progress in application of CNTs in lithium-ion batteries. J Nanomater 2014;2014. [578] Balasubramanian K, Burghard M. Chemically functionalized carbon nanotubes. Small 2005;1:180–92. [579] Tasis D, Tagmatarchis N, Bianco A, Prato M, Tasis D, Tagmatarchis N, et al. Chemistry of carbon nanotubes. Chem Rev 2006;106:1105–36. [580] Xu Y, Dong Y, Wang X, Wang Y, Jiao L, Yuan H, et al. Electrochemical performances of cobalt oxide-carbon nanotubes electrodes via different methods as negative material for alkaline rechargeable batteries. RSC Adv 2015;5:73410–5. [581] Fang Z, Xu W, Huang T, Li M, Wang W, Liu Y, et al. Facile scalable synthesis of Co3O4/carbon nanotube hybrids as superior anode materials for lithium-ion batteries. Mater Res Bull 2013;48:4419–23. [582] Nguyen VH, Kang C, Roh C, Shim JJ. Supercritical CO2-mediated synthesis of CNT@Co3O4 nanocomposite and its application for energy storage. Ind Eng Chem Res 2016;55:7338–43. [583] Deng Q, Wang L, Li J. Electrochemical characterization of Co3O4/MCNTs composite anode materials for sodium-ion batteries. J Mater Sci 2015;50:4142–8. [584] Li J, Tang S, Lu L, Zeng HC. Preparation of nanocomposites of metals, metal oxides, and carbon nanotubes via self-assembly. J Am Chem Soc 2007;129:9401–9. [585] Jiang K, Wang J, Li Q, Liu L, Liu C, Fan S. Superaligned carbon nanotube arrays, films, and yarns: a road to applications. Adv Mater 2011;23:1154–61. [586] Liu K, Jiang K, Wei Y, Ge S, Liu P, Fan S. Controlled termination of the growth of vertically aligned carbon nanotube arrays. Adv Mater 2007;19:975–8. [587] Jiang K, Li Q, Fan S. Spinning continuous carbon nanotube yarns. Nature 2002;419:801. [588] Zhang X, Jiang K, Feng C, Liu P, Zhang L, Kong J, et al. Spinning and processing continuous yarns from 4-inch wafer scale super-aligned carbon nanotube arrays. Adv Mater 2006;18:1505–10. [589] Feng C, Liu K, Wu JS, Liu L, Cheng JS, Zhang Y, et al. Flexible, stretchable, transparent conducting films made from superaligned carbon nanotubes. Adv Funct Mater 2010;20:885–91. [590] He XF, Wu Y, Zhao F, Wang JP, Jiang KL, Fan SS. Enhanced rate capabilities of Co3O4/carbon nanotube anodes for lithium ion battery applications. J Mater Chem A 2013;1:11121–5. [591] Zhuo L, Wu Y, Ming J, Wang L, Yu Y, Zhang X, et al. Facile synthesis of a Co3O4–carbon nanotube composite and its superior performance as an anode material for Li-ion batteries. J Mater Chem A 2013;1:1141–7. [592] Ming J, Wu Y, Yu Y, Zhao F. Steaming multiwalled carbon nanotubes via acid vapour for controllable nanoengineering and the fabrication of carbon nanoflutes. Chem Commun 2011;47:5223–5. [593] Fang H, Zhang S, Liu W, Du Z, Wu X, Xing Y. Hierarchical Co3O4@multiwalled carbon nanotube nanocable films with superior cyclability and high lithium storage capacity. Electrochim Acta 2013;108:651–9. [594] Abbas SM, Hussain ST, Ali S, Ahmad N, Ali N, Munawar KS. Synthesis of carbon nanotubes anchored with mesoporous Co3O4 nanoparticles as anode material for lithium-ion batteries. Electrochim Acta 2013;105:481–8. [595] Abdolmaleki A, Kazerooni H, Gholivand MB, Heydari H, Pendashteh A. Facile electrostatic coprecipitation of f-SWCNT/Co3O4 nanocomposite as supercapacitor material. Ionics 2014;21:515–23. [596] Fu L, Liu Z, Liu Y, Han B, Hu P, Cao L, et al. Beaded cobalt oxide nanoparticles along carbon nanotubes: towards more highly integrated electronic devices. Adv Mater 2005;17:217–21. [597] Xu M, Wang F, Zhang Y, Yang S, Zhao M, Song X. Co3O4-carbon nanotube heterostructures with bead-on-string architecture for enhanced lithium storage performance. Nanoscale 2013;5:8067–72. [598] Thangasamy PKS, Marappan SSMSK, Rangasamy T. Anchoring of ultrafine Co3O4 nanoparticles on MWCNTs using supercritical fluid processing and its performance evaluation towards electrocatalytic oxygen reduction reaction. Catal Sci Technol 2017;7:1227–34. [599] Al-Hakemy AZ, Nassr ABAA, Naggar AH, Elnouby MS, Soliman HMAEF, Taher MA. Electrodeposited cobalt oxide nanoparticles modified carbon nanotubes as a non-precious catalyst electrode for oxygen reduction reaction. J Appl Electrochem 2017;47:183–95.
672
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
[600] Zhao S, Rasimick B, Mustain W, Xu H. Highly durable and active Co3O4 nanocrystals supported on carbon nanotubes as bifunctional electrocatalysts in alkaline media. Appl Catal B Environ 2017;203:138–45. [601] So I-S, Kim N-I, Cho S-H, Kim Y, Yoo J, Seo Y, et al. Enhancement of bifunctional activity of the hybrid catalyst of hollow-net structure Co3O4 and carbon nanotubes. J Electrochem Soc 2016;163:F3041–50. [602] Park J, Moon WG, Kim GP, Nam I, Park S, Kim Y, et al. Three-dimensional aligned mesoporous carbon nanotubes filled with Co3O4 nanoparticles for Li-ion battery anode applications. Electrochim Acta 2013;105:110–4. [603] Ke Q, Tang C, Yang ZC, Zheng M, Mao L, Liu H, et al. 3D nanostructure of carbon nanotubes decorated Co3O4 nanowire arrays for high performance supercapacitor electrode. Electrochim Acta 2015;163:9–15. [604] Zhang Y-X, Guo X, Zhai X, Yan Y-M, Sun K-N. Diethylenetriamine (DETA)-assisted anchoring of Co3O4 nanorods on carbon nanotubes as efficient electrocatalysts for the oxygen evolution reaction. J Mater Chem A 2015;3:1761–8. [605] Wu FD, Wang Y. Self-assembled echinus-like nanostructures of mesoporous CoO nanorod@CNT for lithium-ion batteries. J Mater Chem 2011;21:6636–41. [606] Chen YM, Yu L, Lou XW. Hierarchical tubular structures composed of Co3O4 hollow nanoparticles and carbon nanotubes for lithium storage. Angew Chem Int Edit 2016;55:5990–3. [607] Tang Z, Li F, Zhang Y, Fu X, Xu Y. Composites of titanate nanotube and carbon nanotube as photocatalyst with high mineralization ratio for gas-phase degradation of volatile aromatic pollutant. J Phys Chem C 2011;115:7880–6. [608] Weng B, Liu S, Zhang N, Tang Z, Xu Y. A simple yet efficient visible-light-driven CdS nanowires-carbon nanotube 1D–1D nanocomposite photocatalyst. J Catal 2014;309:146–55. [609] Xiao F, Hung S, Tao HB, Miao J, Bin Yang H, Liu B. Spatially branched hierarchical ZnO nanorod-TiO2 nanotube array heterostructures for versatile photocatalytic and photoelectrocatalytic applications: towards intimate integration of 1D–1D hybrid nanostructures. Nanoscale 2014;6:14950–61. [610] Hu J, Jiang H, Chen Z, Sun Y, Bando Y, Sekiguchi T, et al. Uniform, thin and continuous graphitic carbon tubular coatings on CdS nanowires. J Mater Chem 2009;19:1093–7. [611] Lee Il T, Jegal JP, Park JH, Choi WJ, Lee JO, Kim KB, et al. Three-dimensional layer-by-layer anode structure based on Co3O4 nanoplates strongly tied by capillary-like multiwall carbon nanotubes for use in high-performance lithium-ion batteries. ACS Appl Mater Interfaces 2015;7:3861–5. [612] Venugopal N, Lee D-J, Lee YJ, Sun Y-K. Self-assembled hollow mesoporous Co3O4 hybrid architectures: a facile synthesis and application in Li-ion batteries. J Mater Chem A 2013;1:13164–70. [613] Jian Z, Liu P, Li F, Chen M, Zhou H. Monodispersed hierarchical Co3O4 spheres intertwined with carbon nanotubes for use as anode materials in sodium-ion batteries. J Mater Chem A 2014;2:13805–9. [614] Qiu S, Gu HB, Lu GX, Liu JR, Li XY, Fu Y, et al. Rechargeable Co3O4 porous nanoflake carbon nanotube nanocomposite lithium-ion battery anodes with enhanced energy performances. RSC Adv 2015;5:46509–16. [615] Rahman MM, Sultana I, Chen Z. Ex situ electrochemical sodiation/desodiation observation of Co3O4 anchored carbon nanotubes: a high performance sodiumion battery anode produced by pulsed plasma in a liquid. Nanoscale 2015;7:13088–95. [616] Wenelska K, Neef C, Schlestein L, Klingeler R, Kalenczuk RJ, Mijowska E. Carbon nanotubes decorated by mesoporous cobalt oxide as electrode material for lithium-ion batteries. Chem Phys Lett 2015;635:185–9. [617] Wei W, Wang Z, Liu Z, Liu Y, He L, Chen D, et al. Metal oxide hollow nanostructures: fabrication and Li storage performance. J Power Sources 2013;238:376–87. [618] Wang Z, Zhou L, Lou XW. Metal oxide hollow nanostructures for lithium-ion batteries. Adv Mater 2012;24:1903–11. [619] Huang G, Zhang F, Du X, Qin Y, Yin D, Wang L. Metal organic frameworks route to in situ insertion of multiwalled carbon nanotubes in Co3O4 polyhedra as anode materials for lithium-ion batteries. ACS Nano 2015;9:1592–9. [620] Zhou G, Li L, Zhang Q, Li N, Li F. Octahedral Co3O4 particles threaded by carbon nanotube arrays as integrated structure anodes for lithium ion batteries. Phys Chem Chem Phys 2013;15:5582–7. [621] Liu Y, Liu Y, Cheng SHS, Yu S, Nan B, Bian H, et al. Conformal coating of heterogeneous CoO/Co nanocomposites on carbon nanotubes as efficient bifunctional electrocatalyst for Li-air batteries. Electrochim Acta 2016;219:560–7. [622] Xiao J, Chen C, Xi J, Xu Y, Xiao F, Wang S, et al. Core–shell Co@Co3O4 nanoparticle-embedded bamboo-like nitrogen-doped carbon nanotubes (BNCNTs) as a highly active electrocatalyst for the oxygen reduction reaction. Nanoscale 2015;7:7056–64. [623] Seo B, Sa YJ, Woo J, Kwon K, Park J, Shin TJ, et al. Size-dependent activity trends combined with in situ X-ray absorption spectroscopy reveal insights into cobalt oxide/carbon nanotube-catalyzed bifunctional oxygen electrocatalysis. ACS Catal 2016;6:4347–55. [624] Li X, Fang Y, Lin X, Tian M, An X, Fu Y, et al. MOF derived Co3O4 nanoparticles embedded in N-doped mesoporous carbon layer/MWCNT hybrids: extraordinary bi-functional electrocatalysts for OER and ORR. J Mater Chem A 2015;3:17392–402. [625] Dou S, Li X, Tao L, Huo J, Wang S. Cobalt nanoparticle-embedded carbon nanotube/porous carbon hybrid derived from MOF-encapsulated Co3O4 for oxygen electrocatalysis. Chem Commun 2016;52:9727–30. [626] Yoon TH, Park YJ. Polydopamine-assisted carbon nanotubes/Co3O4 composites for rechargeable Li-air batteries. J Power Sources 2013;244:344–53. [627] Xie K, Masa J, Madej E, Yang F, Weide P, Dong W, et al. Co3O4-MnO2-CNT hybrids synthesized by HNO3 vapor oxidation of catalytically grown CNTs as OER electrocatalysts. ChemCatChem 2015;7:3027–35. [628] Yang S, Cui G, Pang S, Cao Q, Kolb U, Feng X, et al. Fabrication of cobalt and cobalt oxide/graphene composites: towards high-performance anode materials for lithium ion batteries. ChemSusChem 2010;3:236–9. [629] Qu Q, Gao T, Zheng H, Li X, Liu H, Shen M, et al. Graphene oxides-guided growth of ultrafine Co3O4 nanocrystallites from MOFs as high-performance anode of Li-ion batteries. Carbon 2015;92:119–25. [630] Sun F, Huang K, Qi X, Gao T, Liu Y, Zou X, et al. Enhanced 3D hierarchical double porous Co3O4/graphene architecture for superior rechargeable lithium ion battery. Ceram Int 2014;40:2523–8. [631] Lim HD, Gwon H, Kim H, Kim SW, Yoon T, Choi JW, et al. Mechanism of Co3O4/graphene catalytic activity in Li-O2 batteries using carbonate based electrolytes. Electrochim Acta 2013;90:63–70. [632] Rai AK, Gim J, Anh LT, Kim J. Partially reduced Co3O4/graphene nanocomposite as an anode material for secondary lithium ion battery. Electrochim Acta 2013;100:63–71. [633] Guo J, Li F, Sui J, Zhu H, Zhang X. Self-assembled 3D Co3O4-graphene frameworks with high lithium storage performance. Ionics 2014;20:1635–9. [634] Chandrasekaran S, Choi WM, Chung JS, Hur SH, Kim EJ. 3D crumpled RGO- Co3O4 photocatalysts for UV-induced hydrogen evolution reaction. Mater Lett 2014;136:118–21. [635] Wu X, Wang B, Li S, Liu J, Yu M. Electrophoretic deposition of hierarchical Co3O4@graphene hybrid films as binder-free anodes for high-performance lithiumion batteries. RSC Adv 2015;5:33438–44. [636] He G, Li J, Chen H, Shi J, Sun X, Chen S, et al. Hydrothermal preparation of Co3O4@graphene nanocomposite for supercapacitor with enhanced capacitive performance. Mater Lett 2012;82:61–3. [637] Xiang C, Li M, Zhi M, Manivannan A, Wu N. A reduced graphene oxide/Co3O4 composite for supercapacitor electrode. J Power Sources 2013;226:65–70. [638] Ujjain SK, Singh G, Sharma RK. Co3O4@reduced graphene oxide nanoribbon for high performance asymmetric supercapacitor. Electrochim Acta 2015;169:276–82. [639] Ghosh D, Lim J, Narayan R, Kim SO. High energy density all solid state asymmetric pseudocapacitors based on free standing reduced graphene oxide-Co3O4 composite aerogel electrodes. ACS Appl Mater Interfaces 2016;8:22253–60. [640] Wang B, Wang Y, Park J, Ahn H, Wang G. In situ synthesis of Co3O4/graphene nanocomposite material for lithium-ion batteries and supercapacitors with high capacity and supercapacitance. J Alloys Compd 2011;509:7778–83. [641] Mehmood Shahid M, Rameshkumar P, Jeffrey Basirun W, Joon Ching J, Ming Huang N. Cobalt oxide nanocubes interleaved reduced graphene oxide as an efficient electrocatalyst for oxygen reduction reaction in alkaline medium. Electrochim Acta 2017;237:61–8.
673
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
[642] Wang Y, Bai Y, Li X, Feng Y, Zhang H. A general strategy towards encapsulation of nanoparticles in sandwiched graphene sheets and the synergic effect on energy storage. Chem Eur J 2013;19:3340–7. [643] Xue P, Zhang L, Zhang L, Feng X, Zhang Y, Hao W, et al. In-plane vacancy-induced growth of ultra-high loading cobalt oxide-graphene composite for highperformance lithium-ion batteries. Electrochim Acta 2014;136:330–9. [644] Li B, Cao H, Shao J, Li G, Qu M, Yin G. Co3O4@graphene composites as anode materials for high-performance lithium ion batteries. Inorg Chem 2011;50:1628–32. [645] Hu T, Xin GQ, Sun HT, Sun X, Yu MP, Liu CS, et al. Electrospray deposition of a Co3O4 nanoparticles-graphene composite for a binder-free lithium ion battery electrode. RSC Adv 2014;4:1521–5. [646] Qiu D, Bu G, Zhao B, Lin Z, Pu L, Pan L, et al. In situ growth of mesoporous Co3O4 nanoparticles on graphene as a high-performance anode material for lithiumion batteries. Mater Lett 2014;119:12–5. [647] Jang K, Hwang DK, Auxilia FM, Jang J, Song H, Oh BY, et al. Sub-10-nm Co3O4 nanoparticles/graphene composites as high-performance anodes for lithium storage. Chem Eng J 2017;309:15–21. [648] Wu Z, Ren W, Wen L, Gao L, Zhao J, Chen Z, et al. Graphene anchored with Co3O4 nanoparticles as anode of lithium ion capacity and cyclic performance. ACS Nano 2010;4:3187–94. [649] Kim H, Seo DH, Kim SW, Kim J, Kang K. Highly reversible Co3O4/graphene hybrid anode for lithium rechargeable batteries. Carbon 2011;49:326–32. [650] Zhu X, Ning GQ, Ma XL, Fan ZJ, Xu CG, Gao JS, et al. High density Co3O4 nanoparticles confined in a porous graphene nanomesh network driven by an electrochemical process: ultra-high capacity and rate performance for lithium ion batteries. J Mater Chem A 2013;1:14023–30. [651] Choi BG, Chang S-J, Lee YB, Bae JS, Kim HJ, Huh YS, et al. 3D heterostructured architectures of Co3O4 nanoparticles deposited on porous graphene surfaces for high performance of lithium ion batteries. Nanoscale 2012;4:5924–30. [652] Bao L, Li T, Chen S, Peng C, Li L, Xu Q, et al. 3D graphene frameworks/Co3O4 composites electrode for high-performance supercapacitor and enzymeless glucose detection. Small 2017;13:1602077. [653] Xie L, Su F, Xie L, Li X, Liu Z, Kong Q, et al. Self-assembled 3D graphene-based aerogel with Co3O4 nanoparticles as high-performance asymmetric supercapacitor electrode. ChemSusChem 2015;8:2917–26. [654] Mao S, Wen Z, Kim H, Lu G, Hurley P, Chen J. A general approach to one-pot fabrication of crumpled graphene-based nanohybrids for energy applications. ACS Nano 2012;6:7505–13. [655] Peng C, Chen B, Qin Y, Yang S, Li C, Zuo Y, et al. Facile ultrasonic synthesis of CoO quantum dot/graphene nanosheet composites with high lithium storage capacity. ACS Nano 2012;6:1074–81. [656] Guan Q, Cheng J, Wang B, Ni W, Gu G, Li X, et al. Needle-like Co3O4 anchored on the graphene with enhanced electrochemical performance for aqueous supercapacitors. ACS Appl Mater Interfaces 2014;6:7626–32. [657] Garakani MA, Abouali S, Zhang B, Takagi CA, Xu Z, Huang J, et al. Cobalt carbonate/and cobalt oxide/graphene aerogel composite anodes for high performance Li-ion batteries. ACS Appl Mater Interfaces 2014;6:18971–80. [658] Wang Q, Zhang CY, Xia XB, Xing LL, Xue XY. Extremely high capacity and stability of Co3O4/graphene nanocomposites as the anode of lithium-ion battery. Mater Lett 2013;112:162–4. [659] Hu R, Zhang H, Bu Y, Zhang H, Zhao B, Yang C. Porous Co3O4 nanofibers surface-modified by reduced graphene oxide as a durable, high-rate anode for lithium ion battery. Electrochim Acta 2017;228:241–50. [660] Ryu W, Yoon T, Song SH, Jeon S, Park Y, Kim I. Bifunctional composite catalysts using Co3O4 nano fibers immobilized on nonoxidized graphene nanoflakes for high-capacity and long-cycle Li-O2 batteries. Nano Lett 2013;13:4190–7. [661] Cao L, Ma L, Xiao P, Zhang Y, Zhang S, Yang S. Vertically aligned cobalt oxide nanowires on graphene networks for high-performance lithium storage. Nanotechnology 2014;25. [662] Dong X, Xu H, Wang X, Huang Y, Chan-park MB, Zhang H. 3D graphene-cobalt oxide electrode for high-performance supercapacitor and enzymeless glucose detection. ACS Nano 2012;6:3206–13. [663] Song Z, Zhang Y, Liu W, Zhang S, Liu G, Chen H, et al. Hydrothermal synthesis and electrochemical performance of Co3O4/reduced graphene oxide nanosheet composites for supercapacitors. Electrochim Acta 2013;112:120–6. [664] Wu C, Shen Q, Mi R, Deng S, Shu Y, Wang H, et al. Three-dimensional Co3O4/flocculent graphene hybrid on Ni foam for supercapacitor applications. J Mater Chem A 2014;2:15987–94. [665] Sun H, Liu Y, Yu Y, Ahmad M, Nan D, Zhu J. Mesoporous Co3O4 nanosheets-3D graphene networks hybrid materials for high-performance lithium ion batteries. Electrochim Acta 2014;118:1–9. [666] Zhang J, Li P, Wang Z, Qiao J, Rooney D, Sun W, et al. Three-dimensional graphene–Co3O4 cathodes for rechargeable Li–O2 batteries. J Mater Chem A 2015;3:1504–10. [667] Zou Y, Kinloch IA, Dryfe RAW. Mesoporous vertical Co3O4 nanosheet arrays on nitrogen-doped graphene foam with enhanced charge-storage performance. ACS Appl Mater Interfaces 2015;7:22831–8. [668] Chen SQ, Wang Y. Microwave-assisted synthesis of a Co3O4–graphene sheet-on-sheet nanocomposite as a superior anode material for Li-ion batteries. J Mater Chem 2010;20:9735–9. [669] Suryanto BHR, Lu XY, Zhao C. Layer-by-layer assembly of transparent amorphous Co3O4 nanoparticles/graphene composite electrodes for sustained oxygen evolution reaction. J Mater Chem A 2013;1:12726–31. [670] Wang X, Liu S, Wang H, Tu F, Fang D, Li Y. Facile and green synthesis of Co3O4 nanoplates/graphene nanosheets composite for supercapacitor. J Solid State Electrochem 2012;16:3593–602. [671] Wang RH, Xu CH, Sun J, Liu YQ, Gao L, Lin CC. Free-standing and binder-free lithium-ion electrodes based on robust layered assembly of graphene and Co3O4 nanosheets. Nanoscale 2013;5:6960–7. [672] Dou Y, Xu J, Ruan B, Liu Q, Pan Y, Sun Z, et al. Atomic layer-by-layer Co3O4/graphene composite for high performance lithium-ion batteries. Adv Energy Mater 2016;6:1501835. [673] Chi X, Chang L, Xie D, Zhang J, Du G. Hydrothermal preparation of Co3O4/graphene composite as anode material for lithium-ion batteries. Mater Lett 2013;106:178–81. [674] Sun C, Li F, Ma C, Wang Y, Ren Y, Yang W, et al. Graphene–Co3O4 nanocomposite as an efficient bifunctional catalyst for lithium–air batteries. J Mater Chem A 2014;2:7188–96. [675] Yang S, Feng X, Ivanovici S, Müllen K. Fabrication of graphene-encapsulated oxide nanoparticles: towards high-performance anode materials for lithium storage. Angew Chem Int Edit 2010;49:8408–11. [676] Sun H, Sun X, Hu T, Yu M, Lu F, Lian J. Graphene-wrapped mesoporous cobalt-oxide hollow spheres anode for high-rate and long-life lithium ion batteries. J Phys Chem C 2014;118:2263–72. [677] Yang X, Fan K, Zhu Y, Shen J, Jiang X, Zhao P, et al. Tailored graphene-encapsulated mesoporous Co3O4 composite microspheres for high-performance lithium ion batteries. J Mater Chem 2012;22:17278–83. [678] Zhang M, Li R, Chang X, Xue C, Gou X. Hybrid of porous cobalt oxide nanospheres and nitrogen-doped graphene for applications in lithium-ion batteries and oxygen reduction reaction. J Power Sources 2015;290:25–34. [679] Geng H, Guo Y, Ding X, Wang H, Zhang Y, Wu X, et al. Porous cubes constructed by cobalt oxide nanocrystals with graphene sheet coatings for enhanced lithium storage properties. Nanoscale 2016;8:7688–94. [680] Kumar K, Canaff C, Rousseau J, Arrii-Clacens S, Napporn TW, Habrioux A, et al. Effect of the oxide-carbon heterointerface on the activity of Co3O4/NRGO nanocomposites toward ORR and OER. J Phys Chem C 2016;120:7949–58. [681] Jia X, Gao S, Liu T, Li D, Tang P, Feng Y. Controllable synthesis and bi-functional electrocatalytic performance towards oxygen electrode reactions of Co3O4/NRGO composites. Electrochim Acta 2017;226:104–12.
674
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
[682] Wang Q, Hu W, Huang Y. Nitrogen doped graphene anchored cobalt oxides efficiently bi-functionally catalyze both oxygen reduction reaction and oxygen revolution reaction. Int J Hydrogen Energy 2017;42:5899–907. [683] He B, Chen X, Lu J, Yao S, Wei J, Zhao Q, et al. One-pot synthesized Co/Co3O4-N-graphene composite as electrocatalyst for oxygen reduction reaction and oxygen evolution reaction. Electroanalysis 2016;28:2435–43. [684] He Q, Li Q, Khene S, Ren X, López-Suárez FE, Lozano-Castelló D, et al. High-loading cobalt oxide coupled with nitrogen-doped graphene for oxygen reduction in anion-exchange-membrane alkaline fuel cells. J Phys Chem C 2013;117:8697–707. [685] Bezerra CWB, Zhang L, Lee K, Liu H, Marques ALB, Marques EP, et al. A review of Fe-N/C and Co-N/C catalysts for the oxygen reduction reaction. Electrochim Acta 2008;53:4937–51. [686] Lai L, Zhu J, Li Z, Yu DYW, Jiang S, Cai X, et al. Co3O4/nitrogen modified graphene electrode as Li-ion battery anode with high reversible capacity and improved initial cycle performance. Nano Energy 2014;3:134–43. [687] Yao X, Guo G, Zhao Y, Zhang Y, Tan SY, Zeng Y, et al. Synergistic effect of mesoporous Co3O4 nanowires confined by N-doped graphene aerogel for enhanced lithium storage. Small 2016;12:3849–60. [688] Ma XX, He XQ. Cobalt oxide anchored on nitrogen and sulfur dual-doped graphene foam as an effective oxygen electrode catalyst in alkaline media. Appl Mater Today 2016;4:1–8. [689] Tong Y, Chen P, Zhou T, Xu K, Chu W, Wu C, et al. A bifunctional hybrid electrocatalyst for oxygen reduction and evolution: cobalt oxide nanoparticles strongly coupled to B,N-decorated graphene. Angew Chem Int Edit 2017;56:7121–5. [690] Zhao Y, Yang L, Chen S, Wang X, Ma Y, Wu Q, et al. Can boron and nitrogen co-doping improve oxygen reduction reaction activity of carbon nanotubes? J Am Chem Soc 2013;135:1201–4. [691] Liu L, Fang D, Jiang M, Chen J, Wang T, Wang Q, et al. Co3O4/C/graphene nanocomposites as novel anode materials for high capacity lithium ion batteries. RSC Adv 2015;5:73677–83. [692] Wang Q, Yuan H, Feng H, Li J, Zhao C, Liu J, et al. One-step solution-phase synthesis of Co3O4/RGO/acetylene black as a high-performance catalyst for oxygen reduction reaction. RSC Adv 2014;4:18286–93. [693] Zhang Y, Wang Q, Li J, Wang X, Liu K, Feng H, et al. Facile fabrication of sandwich-structured Co3O4/N-rGO/AB hybrid with enhanced ORR electrocatalytic performances for metal-air batteries. RSC Adv 2015;5:9057–63. [694] Park S, Kim S. Effect of carbon blacks filler addition on electrochemical behaviors of Co3O4/graphene nanosheets as a supercapacitor electrodes. Electrochim Acta 2013;89:516–22. [695] Lu XY, Chan HM, Sun CL, Tseng CM, Zhao C. Interconnected core-shell carbon nanotube-graphene nanoribbon scaffolds for anchoring cobalt oxides as bifunctional electrocatalysts for oxygen evolution and reduction. J Mater Chem A 2015;3:13371–6. [696] Yuan C, Yang L, Hou L, Li J, Sun Y, Zhang X, et al. Flexible hybrid paper made of monolayer Co3O4 microsphere arrays on rGO/CNTs and their application in electrochemical capacitors. Adv Funct Mater 2012;22:2560–6. [697] Lu H, Huang Y, Yan J, Fan W, Liu T. Nitrogen-doped graphene/carbon nanotube/Co3O4 hybrids: one-step synthesis and superior electrocatalytic activity for the oxygen reduction reaction. RSC Adv 2015;5:94615–22. [698] Ren G, Li Y, Guo Z, Xiao G, Zhu Y, Dai L, et al. A bio-inspired Co3O4-polypyrrole-graphene complex as an efficient oxygen reduction catalyst in one-step ball milling. Nano Res 2015;8:3461–71. [699] Huang W, Zhong H, Li D, Tang P, Feng Y. Reduced graphene oxide supported CoO/MnO2 electrocatalysts from layered double hydroxides for oxygen reduction reaction. Electrochim Acta 2015;173:575–80. [700] Dai L, Liu M, Song Y, Liu J, Wang F. Mn3O4-decorated Co3O4 nanoparticles supported on graphene oxide: dual electrocatalyst system for oxygen reduction reaction in alkaline medium. Nano Energy 2016;27:185–95. [701] Nguyen VH, Tran VC, Kharismadewi D, Shim JJ. Ultralong MnO2 nanowires intercalated graphene/Co3O4 composites for asymmetric supercapacitors. Mater Lett 2015;147:123–7. [702] Lee DH, Kim JE, Han TH, Hwang WJ, Jeon SW, Choi SY, et al. Versatile carbon hybrid films composed of vertical carbon nanotubes crown on mechanically compliant graphene films. Adv Mater 2010;22:1247–52. [703] Li S, Luo Y, Lv W, Yu W, Wu S, Hou P, et al. Vertically aligned carbon nanotubes grown on graphene paper as electrodes in lithium-ion batteries and dyesensitized solar cells. Adv Energy Mater 2011;1:486–90. [704] Hu Y, Li X, Wang J, Li R, Sun X. Free-standing graphene-carbon nanotube hybrid papers used as current collector and binder free anodes for lithium ion batteries. J Power Sources 2013;237:41–6. [705] Fan Z, Yan J, Zhi L, Zhang Q, Wei T, Feng J, et al. A three-dimensional carbon nanotube/graphene sandwich and its application as electrode in supercapacitors. Adv Mater 2010;22:3723–8. [706] Zhang Y-Z, Wang Y, Cheng T, Lai W-Y, Pang H, Huang W. Flexible supercapacitors based on paper substrates: a new paradigm for low-cost energy storage. Chem Soc Rev 2015;44:5181–99. [707] Wang X, Lu X, Liu B, Chen D, Tong Y, Shen G. Flexible energy-storage devices: design consideration and recent progress. Adv Mater 2014;26:4763–82. [708] Youn DH, Seol M, Kim JY, Jang JW, Choi Y, Yong K, et al. TiN nanoparticles on CNT-graphene hybrid support as noble-metal-free counter electrode for quantum-dot-sensitized solar cells. ChemSusChem 2013;6:261–7. [709] Lv R, Cui T, Jun MS, Zhang Q, Cao A, Su DS, et al. Open-ended, N-doped carbon nanotube-graphene hybrid nanostructures as high-performance catalyst support. Adv Funct Mater 2011;21:999–1006. [710] Seol M, Youn DH, Kim JY, Jang JW, Choi M, Lee JS, et al. Mo-compound/CNT-graphene composites as efficient catalytic electrodes for quantum-dot-sensitized solar cells. Adv Energy Mater 2014;4:1300775. [711] Zhang J, Wang X, Qin D, Xue Z, Lu X. Fabrication of iron-doped cobalt oxide nanocomposite films by electrodeposition and application as electrocatalyst for oxygen reduction reaction. Appl Surf Sci 2014;320:73–82. [712] Zhuang L, Ge L, Yang Y, Li M, Jia Y, Yao X, et al. Ultrathin iron-cobalt oxide nanosheets with abundant oxygen vacancies for the oxygen evolution reaction. Adv Mater 2017;29:1606793. [713] Yu Q, Yu Q, Sun W, Wu H, Wang Z, Rooney D, et al. Novel Ni@Co3O4 web-like nanofiber arrays as highly effective cathodes for rechargeable Li-O2 batteries. Electrochim Acta 2016;220:654–63. [714] Song W, Ren Z, Chen SY, Meng Y, Biswas S, Nandi P, et al. Ni- and Mn-promoted mesoporous Co3O4: a stable bifunctional catalyst with surface-structuredependent activity for oxygen reduction reaction and oxygen evolution reaction. ACS Appl Mater Interfaces 2016;8:20802–13. [715] Huang L, Mao Y, Wang G, Xia X, Xie J, Zhang S, et al. Ru-decorated knitted Co3O4 nanowires as a robust carbon/binder-free catalytic cathode for lithium–oxygen batteries. New J Chem 2016;40:6812–8. [716] Ren Y, Zhang S, Li H, Wei X, Xing Y. Mesoporous Pd/Co3O4 nanosheets nanoarrays as an efficient binder/carbon free cathode for rechargeable Li-O2 batteries. Appl Surf Sci 2017;420:222–32. [717] Ishihara T, Thapa AK, Hidaka Y, Ida S. Rechargeable lithium-air battery using mesoporous Co3O4 modified with Pd for air electrode. Electrochemistry 2012;80:731–3. [718] Cao J, Liu S, Xie J, Zhang S, Cao G, Zhao X. Tips-bundled Pt/Co3O4 nanowires with directed peripheral growth of Li2O2 as efficient binder/carbon-free catalytic cathode for lithium−oxygen battery. ACS Catal 2015;5:241–5. [719] Gao G, Bin Wu H, Ding S, Liu LM, Lou XW. Hierarchical NiCo2O4 nanosheets grown on Ni nanofoam as high-performance electrodes for supercapacitors. Small 2015;11:804–8. [720] Shang K, Li W, Liu Y, Zhang W, Yang H, Xie J, et al. A novel shuttle-like Fe3O4–Co3O4 self-assembling architecture with highly reversible lithium storage. RSC Adv 2015;5:70527–35. [721] Ye Y, Kuai L, Geng B. A template-free route to a Fe3O4–Co3O4 yolk–shell nanostructure as a noble-metal free electrocatalyst for ORR in alkaline media. J Mater Chem 2012;22:19132–8.
675
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
[722] Rakhi RB, Chen W, Hedhili MN, Cha D, Alshareef HN. Enhanced rate performance of mesoporous Co3O4 nanosheet supercapacitor electrodes by hydrous RuO2 nanoparticle decoration. ACS Appl Mater Interfaces 2014;6:4196–206. [723] Li Y, Huang K, Zeng D, Liu S, Yao Z. RuO2/Co3O4 thin films prepared by spray pyrolysis technique for supercapacitors. J Solid State Electrochem 2010;14:1205–11. [724] Jang HS, Yang YJ, Lee NS, Son B, Lee Y, Lee C, et al. Electrospun RuO2-Co3O4 hybrid nanotubes for enhanced electrocatalytic activity. Mater Lett 2015;139:405–8. [725] Cai D, Huang H, Wang D, Liu B, Wang L, Liu Y, et al. High-performance supercapacitor electrode based on the unique ZnO@Co3O4 core/shell heterostructures on nickel foam. ACS Appl Mater Interfaces 2014;6:15905–12. [726] Kim GP, Sun HH, Manthiram A. Design of a sectionalized MnO2-Co3O4 electrode via selective electrodeposition of metal ions in hydrogel for enhanced electrocatalytic activity in metal-air batteries. Nano Energy 2016;30:130–7. [727] Wang F, Wen Z, Shen C, Wu X, Liu J. Synthesis of α-MnO₂ nanowires modified by Co3O4 nanoparticles as a high-performance catalyst for rechargeable Li-O2 batteries. Phys Chem Chem Phys 2016;18:926–31. [728] Liu J, Jiang J, Cheng C, Li H, Zhang J, Gong H, et al. Co3O4 nanowire@MnO2 ultrathin nanosheet core/shell arrays: a new class of high-performance pseudocapacitive materials. Adv Mater 2011;23:2076–81. [729] Han X, Cheng F, Chen C, Li F, Chen J. A Co3O4@MnO2/Ni nanocomposite as a carbon- and binder-free cathode for rechargeable Li–O2 batteries. Inorg Chem Front 2016;3:866–71. [730] Huang M, Zhang Y, Li F, Zhang L, Wen Z, Liu Q. Facile synthesis of hierarchical Co3O4@MnO2 core-shell arrays on Ni foam for asymmetric supercapacitors. J Power Sources 2014;252:98–106. [731] Qiu K, Yan H, Zhang D, Lu Y, Cheng J, Lu M, et al. Hierarchical 3D Co3O4@MnO2 core/shell nanoconch arrays on Ni foam for enhanced electrochemical performance. J Solid State Electrochem 2015;19:391–401. [732] Pang H, Deng J, Du J, Li S, Li J, Ma Y, et al. Porous nanocubic Mn3O4–Co3O4 composites and their application as electrochemical supercapacitors. Dalt Trans 2012;41:10175–81. [733] Xia X, Tu J, Zhang Y, Wang X, Gu C, Zhao X, et al. High-quality metal oxide core/shell nanowire arrays on conductive substrates for electrochemical energy storage. ACS Nano 2012;6:5531–8. [734] Hu Q, Gu Z, Zheng X, Zhang X. Three-dimensional Co3O4@NiO hierarchical nanowire arrays for solid-state symmetric supercapacitor with enhanced electrochemical performances. Chem Eng J 2016;304:223–31. [735] He F, Liu K, Zhong J, Zhang S, Huang Q, Chen C. One dimensional nickel oxide-decorated cobalt oxide (Co3O4) composites for high-performance supercapacitors. J Electroanal Chem 2015;749:89–95. [736] Cui J, Zhang X, Tong L, Luo J, Wang Y, Zhang Y, et al. A facile synthesis of mesoporous Co3O4/CeO2 hybrid nanowire arrays for high performance supercapacitors. J Mater Chem A 2015;3:10425–31. [737] Sennu P, Park HS, Park KU, Aravindan V, Nahm KS, Lee Y-S. Formation of NiCo2O4 rods over Co3O4 nanosheets as efficient catalyst for Li-O2 batteries and water splitting. J Catal 2017;349:175–82. [738] Cai D, Wang D, Liu B, Wang L, Liu Y, Li H, et al. Three-dimensional Co3O4@NiMoO4 core/shell nanowire arrays on Ni foam for electrochemical energy storage. ACS Appl Mater Interfaces 2014;6:5050–5. [739] Gu Z, Wang R, Nan H, Geng B, Zhang X. Construction of unique Co3O4@CoMoO4 core/shell nanowire arrays on Ni foam by the action exchange method for high-performance supercapacitors. J Mater Chem A 2015;3:14578–84. [740] Lu L, Wang H, Wang J-G, Wang C, Jiang Q-C. Design and synthesis of ZnO–NiO–Co3O4 hybrid nanoflakes as high-performance anode materials for Li-ion batteries. J Mater Chem A 2017;5:2530–8. [741] Kuang M, Tao Li T, Chen H, Mao Zhang S, Li Zhang L, Xin Zhang Y. Hierarchical Cu2O/CuO/Co3O4 core-shell nanowires: synthesis and electrochemical properties. Nanotechnology 2015;26:304002. [742] Ma XJ, Bin Kong L, Bin Zhang W, Liu MC, Luo YC, Kang L. Design and synthesis of 3D Co3O4@MMoO4 (M=Ni, Co) nanocomposites as high-performance supercapacitor electrodes. Electrochim Acta 2014;130:660–9. [743] Hu XW, Liu S, Qu BT, You XZ. Starfish-shaped Co3O4/ZnFe2O4 hollow nanocomposite: synthesis, supercapacity, and magnetic properties. ACS Appl Mater Interfaces 2015;7:9972–81. [744] Byun J, Patel HA, Kim DJ, Jung CH, Park JY, Choi JW, et al. Nanoporous networks as caging supports for uniform, surfactant-free Co3O4 nanocrystals and their applications in energy storage and conversion. J Mater Chem A 2015;3:15489–97. [745] Liu L, Hou Y, Wang J, Chen J, Liu H-K, Wu Y, et al. Nanofibrous Co3O4/PPy hybrid with synergistic effect as bifunctional catalyst for lithium-oxygen batteries. Adv Mater Interfaces 2016;3:1600030. [746] Shao X, Zhang T, Wen Z. One step fabrication of Co3O4-PPy cathode for lithium-O2 batteries. Chin J Chem 2017;35:35–40. [747] Xia X, Chao D, Qi X, Xiong Q, Zhang Y, Tu J, et al. Controllable growth of conducting polymers shell for constructing high-quality organic/inorganic core/shell nanostructures and their optical-electrochemical properties. Nano Lett 2013;13:4562–8. [748] Hai Z, Gao L, Zhang Q, Xu H, Cui D, Zhang Z, et al. Facile synthesis of core-shell structured PANI-Co3O4 nanocomposites with superior electrochemical performance in supercapacitors. Appl Surf Sci 2016;361:57–62. [749] Yang X, Xu K, Zou R, Hu J. A hybrid electrode of Co3O4@PPy core/shell nanosheet arrays for high-performance supercapacitors. Nano-Micro Lett 2016;8:143–50. [750] Raj RP, Ragupathy P, Mohan S. Remarkable capacitive behavior of a Co3O4–polyindole composite as electrode material for supercapacitor applications. J Mater Chem A 2015;3:24338–48. [751] Anh LT, Rai AK, Thi TV, Gim J, Kim S, Mathew V, et al. Enhanced electrochemical performance of novel K-doped Co3O4 as the anode material for secondary lithium-ion batteries. J Mater Chem A 2014;2:6966–75. [752] Xu L, Wang Z, Wang J, Xiao Z, Huang X, Liu Z, et al. N-doped nanoporous Co3O4 nanosheets with oxygen vacancies as oxygen evolving electrocatalysts. Nanotechnology 2017;28:165402. [753] Yu H, Li Y, Li X, Fan L, Yang S. Electrochemical preparation of N-doped cobalt oxide nanoparticles with high electrocatalytic activity for the oxygen-reduction reaction. Chem Eur J 2014;20:3457–62. [754] Wang X, Zheng Y, Yuan J, Shen J, Wang A, Niu L, et al. Uniform deposition of Co3O4 nanosheets on exfoliated MoS2 nanosheets as advanced catalysts for water splitting. Electrochim Acta 2016;212:890–7. [755] Dong S, Wang S, Guan J, Li S, Lan Z, Chen C, et al. Insight into enhanced cycling performance of Li-O2 batteries based on binary CoSe2/CoO nanocomposite electrodes. J Phys Chem Lett 2014;5:615–21. [756] Wang Z, Li B, Ge X, Goh FWT, Zhang X, Du G, et al. Co@Co3O4@PPD core@bishell nanoparticle-based composite as an efficient electrocatalyst for oxygen reduction reaction. Small 2016;12:2580–7. [757] Zhong J-H, Wang A-L, Li G-R, Wang J-W, Ou Y-N, Tong Y-X, et al. Co3O4/Ni(OH)2 composite mesoporous nanosheet networks as a promising electrode for supercapacitor applications. J Mater Chem 2012;22:5656–65. [758] Tang C, Yin X, Gong H. Superior performance asymmetric supercapacitors based on a directly grown commercial mass 3D Co3O4@Ni(OH)2 core-shell electrode. ACS Appl Mater Interfaces 2013;5:10574–82. [759] Qorbani M, Naseri N, Moshfegh AZ. Hierarchical Co3O4/Co(OH)2 nanoflakes as a supercapacitor electrode: experimental and semi-empirical model. ACS Appl Mater Interfaces 2015;7:11172–9. [760] Xu J, Wang Q, Wang X, Xiang Q, Liang B, Chen D, et al. Flexible asymmetric supercapacitors based upon Co9S8 nanorod//Co3O4@RuO2 nanosheet arrays on carbon cloth. ACS Nano 2013;7:5453–62. [761] Xia H, Zhu D, Luo Z, Yu Y, Shi X, Yuan G, et al. Hierarchically structured Co3O4@Pt@MnO2 nanowire arrays for high-performance supercapacitors. Sci Rep 2013;3:2978.
676
Progress in Materials Science 103 (2019) 596–677
J. Mei, et al.
[762] Lu Z, Yang Q, Zhu W, Chang Z, Liu J, Sun X, et al. Hierarchical Co3O4@Ni-Co-O supercapacitor electrodes with ultrahigh specific capacitance per area. Nano Res 2012;5:369–78. [763] Singh AK, Sarkar D, Gopal Khan G, Mandal K. Designing one dimensional Co-Ni/Co3O4-NiO core/shell nano-heterostructure electrodes for high-performance pseudocapacitor. Appl Phys Lett 2014;104. [764] Chen H, Zhang Q, Wang J, Xu D, Li X, Yang Y, et al. Improved lithium ion battery performance by mesoporous Co3O4 nanosheets grown on self-standing NiSix nanowires on nickel foam. J Mater Chem A 2014;2:8483–90. [765] Zhang D, Sun W, Chen Z, Zhang Y, Luo W, Jiang Y, et al. Two-dimensional cobalt-/nickel-based oxide nanosheets for high-performance sodium and lithium storage. Chem Eur J 2016;22:18060–5. [766] Wei C, Cheng C, Zhou B, Yuan X, Cui T, Wang S, et al. Hierarchically porous NaCoPO4-Co3O4 hollow microspheres for flexible asymmetric solid-state supercapacitors. Part Part Syst Charact 2015;32:831–9. [767] Zhu J, Jiang J, Sun Z, Luo J, Fan Z, Huang X, et al. 3D carbon/cobalt-nickel mixed-oxide hybrid nanostructured arrays for asymmetric supercapacitors. Small 2014;10:2937–45. [768] Wu L-K, Hu J-M. A silica co-electrodeposition route to nanoporous Co3O4 film electrode for oxygen evolution reaction. Electrochim Acta 2014;116:158–63. [769] Zhou J, Dou Y, Zhou A, Guo RM, Zhao MJ, Li JR. MOF template-directed fabrication of hierarchically structured electrocatalysts for efficient oxygen evolution reaction. Adv Energy Mater 2017;7:1602643. Jun Mei is a Ph.D. candidate at the School of Chemistry, Physics and Mechanical Engineering (CPME) of Queensland University of Technology (QUT) under the supervision of Prof. Ziqi Sun. His major research interests include the design and synthesis of novel nanomaterials for energy conversion and storage devices.
Dr. Ziqi Sun is an Associate Professor at the School of Chemistry, Physics and Mechanical Engineering, Queensland University of Technology, Australia. He received his PhD in Materials Science and Engineering in 2009 at Institute of Metal Research, Chinese Academy of Sciences. After finishing his NIMS postdoctoral fellowship (Japan) that he held for one year on solid oxide fuel cells, he joined the University of Wollongong (UOW) and then moved to Queensland University of Technology (QUT) in 2015. He was awarded with some prestigious awards and fellowships, including ARC Future Fellowship (2018), JSPS Invitational Fellow by Japan Society for the Promotion of Science (2019), TMS Young Leaders Award from the Minerals, Metals and Materials Society, USA (2015), Discovery Early Career Research Award from Australian Research Council (2015), Australian Postdoctoral Fellowship from Australian Research Council (2010), Alexander von Humboldt Fellowship from AvH Foundation Germany (2009), the Vice-Chancellor’s Research Fellowship from University of Wollongong (2013), etc. He also received some honorary positions, such as Chair of TMS Energy Committee, Editor of Sustainable Materials & Technologies (Elsevier), Principal Editor of Journal of Materials Research (SCI, MRS), Associate Editor of Surface Innovations (SCI, ICE), Editorial Board Member of Journal of Materials Science and Technology (SCI, Elsevier) and Scientific Reports (SCI, Nature Group). He held the roles as symposium organizers and session chairs in the prestigious conferences such as AM&ST18, TMS conference, AcerS annual conferences, etc.
677