Yiming
Xu†
a,
Kaicai
Fan†
a,
Yu
Zou
a,
Huaiqin
Fu
a,
Mengyang
Dong
a,
Yuhai
Dou
a,
Yun
Wang
a,
Shan
Chen
a,
Huajie
Yin
ab,
Mohammad
Al-Mamun
*a,
Porun
Liu
*a and
Huijun
Zhao
*a
aCentre for Catalysis and Clean Energy, School of Environment and Science, Griffith University, Gold Coast Campus, Queensland, 4222, Australia. E-mail: m.al-mamun@griffith.edu.au; p.liu@griffith.edu.au; h.zhao@griffith.edu.au
bKey Laboratory of Materials Physics, Centre for Environmental and Energy Nanomaterials, Anhui Key Laboratory of Nanomaterials and Nanotechnology, CAS Center for Excellence in Nanoscience Institute of Solid State Physics, HFIPS, Chinese Academy of Sciences, Hefei 230031, P. R. China
First published on 23rd November 2021
Electrocatalytic energy conversion between electricity and chemical bonding energy is realized through redox reactions with multiple charge transfer steps at the electrode–electrolyte interface. The surface atomic structure of the electrode materials, if appropriately designed, will provide an energetically affordable pathway with individual reaction intermediates that not only reduce the thermodynamic energy barrier but also allow an acceptably fast kinetic rate of the overall redox reaction. As one of the most abundant and stable forms, oxides of transitional metals demonstrated promising electrocatalytic activities towards multiple important chemical reactions. In this topical review, we attempt to discuss the possible avenues to construct the electrocatalytic active surface for this important class of materials for two essential chemical reactions for water splitting. A general introduction of the electrochemical water splitting process on the electrocatalyst surface with applied potential will be provided, followed by a discussion on the fundamental charge transfers and the mechanism. As the generally perceived active sites are chemical reaction dependent, we offer a general overview of the possible approaches to construct or create electrocatalytically active sites in the context of surface atomic structure engineering. The review concludes with perspectives that summarize challenges and opportunities in electrocatalysis and how these can be addressed to unlock the electrocatalytic potentials of the metal oxide materials.
Transition metal oxides are a group of materials abundantly available in Earth's crust. Several features such as mechanical and chemical stability, relatively high charge transfer property even in the amorphous state, and particularly tuneable electronic structure, make metal oxides one of the most promising candidatures for water splitting.5 Moreover, even though metal phosphides, sulfides, carbides, nitrides are more active materials towards HER, when considering overall water splitting, metal oxides are generally more stable, especially under OER conditions.6 Also, the fabrication approaches toward metal oxides can be facile and free from harmful chemicals or harsh reaction conditions. Based on these unique properties, metal oxides have been identified as crucial candidates in practical applications, such as thin-film transistors, solar cells, catalysis, energy storage. It is expected that the versatility and functionality of the metal oxides could provide unique opportunities for water splitting. The unexplored potentials of metal oxide as catalysts have inspired and driven many recently published research articles on these fundamentally essential groups of materials as electrocatalysts. Part of the key findings has been summarized in a number of insightful reviews on HER,7–9 OER10–14 and overall water splitting.15–17 A comprehensive review summarizing the active sites, in situ characterization techniques, and material engineering strategies for earth-abundant metal oxides for the application of water splitting (HER and OER) has not been published to date. The information provided in the proposed review will allow a timely overview of the useful approaches and also inspirations towards the design of high-performance electrocatalysts for water splitting applications.
In this topical review, we provide an overview of the construction of electrocatalytic active sites (primarily for HER and OER) on metal oxide surfaces. In section 1, we introduce the nature of electrocatalytic active sites in general and their expected features for heterogeneous electrocatalytic water splitting. This is followed by a summary of the characterization techniques for the metal oxides in section 2. The aim of section 3 is to summarize current approaches that could achieve the targeted features of metal oxides. The review is concluded in section 4 with perspectives and outlooks on the future development of high-performance metal oxide electrocatalysts. The review aims to provide colleagues in nanoscience and nanotechnology with a well-established baseline by summarising recent knowledge advance on structure–activity relationship and developments in material engineering strategies for high-performance metal oxide electrocatalysts. We believe the included advancements from theories, characterizations, and strategies surveyed in the review will inspire ideas and innovations in real-world electrolysis. The information included herein has broader implications for the fundamental researches on nanoscale functional material, surface chemistry, electrochemistry that extendable to the aforementioned widespread applications of metal oxides.
Water splitting occur in two different pathways, in which the reactions are expressed as:
In acidic solution:
Cathode: 2H+ + 2e− ↔ H2 | (1) |
Anode: H2O ↔ 2H+ + 1/2O2 + 2e− | (2) |
In neutral and alkaline solutions:
Cathode: 2H2O + 2e− ↔ H2 + 2OH− | (3) |
Anode: 2OH− ↔ H2O + 1/2O2 + 2e− | (4) |
The thermodynamic voltage of water splitting is 1.23 V at 25 °C, 1 atm regardless the water media yet it decreases with increase of temperature. In practise, nevertheless, to achieve the electrochemical water splitting, the applied voltages are higher than the thermodynamic potential value. The overpotential (η) is primarily applied to overcome the intrinsic activation barriers on both anode (ηa), cathode (ηc) and some other resistances, e.g., ohmic drop (ηother). The practical operational voltage (EOP) for water splitting can be described as:2
EOP = 1.23 V + ηa + ηc + ηother | (5) |
(1) Electrochemical adsorption (Volmer or discharge reaction)
H+ + M* + e− ↔ M–H* (acidic solution) | (6) |
H2O + M* + e− ↔ M–H* + OH− (alkaline solution) | (7) |
(2) Electrochemical desorption (Heyrovsky or ion + atom reaction)
M–H* + H+ + e− ↔ M* + H2 (acidic solution) | (8) |
M–H* + H2O + e− ↔ M* + OH− + H2 (alkaline solution) | (9) |
(3) Chemical desorption (Tafel or combination reaction)
2M–H* ↔ 2M* + H2 (acid and alkaline solutions) | (10) |
Fig. 1 (a) The schematic of electrocatalytic turn-over of hydrogen evolution on the surface of an electrode.21 (b) The adsorbate evolution mechanism (AEM) and lattice oxygen mediated mechanism (LOM) of oxygen evolution turnover on the surface of catalyst.23 |
Although the two mechanisms are applicable in both acid and alkaline solutions, it should be noticed that reactants and products are different in various pH values. For example, in acid, only the reduction of a proton into an H* is involved in the Volmer step. While in alkaline solution, besides the creation of the H*, water dissociation is also a critical step for the entire HER process, which requires a breaking of the strong H–O–H covalent bond.
In acidic solution:
H2O + M* ↔ M–OH* + H+ + e− | (11) |
M–OH* ↔ M–O* + H+ + e− | (12) |
M–O* + H2O ↔ M–OOH* + H+ + e− | (13) |
M–OOH* ↔ M + O2 + H+ + e− | (14) |
In alkaline solution:
OH− + M* ↔ M–OH* + e− | (15) |
M–OH* + OH− ↔ M–O* + H2O + e− | (16) |
M–O* + OH− ↔ M–OOH* + e− | (17) |
M–OOH* + OH− ↔ M* + O2 + H2O + e− | (18) |
The single-site adsorption mechanism of OER is schematically shown in Fig. 1b.23 It is generally acknowledged that the OER activity of an electrocatalyst depends strongly on the bonding strength of the absorbents on the M active site (M–OH*, M–O* and M–OOH*), of which the adsorption energies are linearly correlated.
Moreover, another OER mechanism (lattice oxygen mediated mechanism, LOM, Fig. 1b) that involves the formation of O–O bonds on adjacent metal ions have also been proven.23,24 At first, two HO* on the two neighbouring metal ions deprotonate and form two M–O*. Then, the two adjacent O* couple directly with each other, resulting in the formation of O–O bond. O2 is eventually released, and two vacant metal sites are occupied by OH−. In contrast to AEM, intermediate HOO* is not involved in LOM.
Ea1 = α1 ΔE + β1 | (19) |
Ea2 = α2 ΔE + β2 | (20) |
As the perfectors α1 is positive and α2 is negative, one can see the plots of the curves for overall reaction rate (by converting the Ea and reaction rate r with Arrhenius equation) exhibits a Sabatier volcano shape, i.e., with a maximal rate curve at the top of the full solution to the microkinetic model.
Based on the Sabatier principle, for an efficient HER catalyst, its surface ought to allow an interaction strong enough to absorb H* for facilitating the proton–electron-transfer process, while also weak enough to promote the bond breakage and gaseous H2 release.26 Nevertheless, it is difficult to establish the quantitative relationship between the energetics of the H* intermediate and the reaction rate. From the perspectives of physical chemistry, the Gibbs free energy change for H* adsorption on a catalyst surface (ΔGH*) is a widely accepted descriptor for the catalysts. According to the Sabatier principle (a full microkinetic solution for a simple generalized heterogeneous catalytic reaction25), if ΔGH* is zero, the overall reaction has the maximum rate. Furthermore, the relationship between the experimentally measured exchange current density (j0) and the quantum chemistry-derived ΔGH* form a volcano curve for a wide range of electrode surfaces.27 Following this trend, the relationship between the intrinsic electrochemical nature and HER kinetics could be built. Pt-based metals are at the summit of the volcano, as shown in Fig. 2a,27 which means they have the best activities, and the hydrogen adsorption energy is close to zero. Furthermore, the metals to the left of Pt have strong adsorption to the hydrogen atoms, which blocks the active site and hinders hydrogen generation. While metals to the right of Pt adsorb the hydrogen too weakly, which will not effectively stabilize the intermediate state, thus hindering the occurrence of following hydrogen generation.25,27
Fig. 2 (a) A volcano plot of experimentally measured exchange current density as a function of the calculated ΔGH.27 (b) The free energy diagram for OER steps (from left to right) at different applied potentials (U). Due to the scaling relation between the free energy of the intermediates (OH*, O*, OOH*), they are correlated and thus moved in a concerted manner. At U = 2.55 V, all the OER steps are exothermic (downhill from left to right).34 (c) Activity trends of the perovskite OER electrocatalysts. The negative theoretical overpotential was plotted against the difference of standard free energy of the step.22 |
Even though the trends in activity for the HER in alkaline media have never been established independently from those results for the HER in acidic media, differences between the descriptors for alkaline and acid HER have been identified. In acid solutions, the reaction is mainly controlled by the hydrogen recombination (the Tafel step), indicating the strong relationship between the activity and the ΔGH*. However, for HER that occurs in alkaline solutions, the kinetics are determined by the balance between the water dissociation (Volmer step) and the accompanying interaction of the water dissociation product (OH*) on the catalysts.28–30 In this regard, the optimal ΔGH* as the sole descriptor is insufficient to describe the HER activity (exchange current in alkaline media is 2–3 orders of magnitude smaller). Thus, other parameters should be further involved and optimized to explore the highly active catalysts. Gong et al. suggested that the water adsorption energy and the hydroxide anion adsorption energy can also influence the activity.31 Low water adsorption decreases the number of reactants, and high hydroxide adsorption energy may result in active site poisoning. Alonso-Vante and co-workers suggested that the kinetics of HER in alkaline is determined by OH adsorption.32 Thereby, modification of Pt by inducing oxophilic metal centres or sp2 carbon sites favours the adsorption of OH at low potentials, removes Hads intermediate produced in Volmer step, and thus improves the HER kinetics. Markovic et al., established the relationship overall catalytic activities for HER in alkaline media as a function of OH–M2+δ bond strength using 3d-M hydro(oxy)oxides and determined that the activity for the established systems follows the order Ni > Co > Fe > Mn.29 Moreover, Markovic's group identified three parameters based on recent researches, (i) the proton donor, (ii) the energy of formation of the activated complex from proton donor, (iii) the availability of active sites.33 They believed these parameters could be applied in any aqueous system. These results provided a foundation for the rational design of catalysts for practical HER in alkaline electrolytes.
In contrast, for the AEM of OER, free energy (ΔG) diagrams have been used to investigate the rate-determining steps based on Density-functional theory (DFT) calculations (Fig. 2b).34 The ΔG values of the intermediates at different steps have a scaling relation and thus change correlatedly.34 In other words, one free parameter can determine both the free energy diagram and the activity. Later, the scaling relation between the adsorption energies of OOH* and OH* was found to be universal for most of the investigated metal oxide electrocatalysts (Fig. 2c).22 When the theoretical thermodynamic overpotentials were plotted against the difference of standard free energy of the O* and OH* , a volcano relation has been obtained. In line with the Sabatier principle, adsorbate binds to the electrocatalysts on the left branch of the plot too strongly and binds to the counterparts on the right branch too weakly. Also, the values of all the good OER catalysts are constantly about 3.2 eV, compared to the perfect value of 2.46 eV, indicating a minimum overpotential of 0.4–0.2 V. This explains the relatively large overpotential of OER compared to the HER. To tune the and values through varying the oxide surface or electrochemical interface may help to improve the OER activities.
On the other hand, in OER pathway following LOM, the intermediate HOO* is not involved (Fig. 1b).23 Instead, the direct coupling of the lattice oxygen occurs. It is generally believed that oxygen vacancy participation plays an important role in activating the LOM mechanism.35 Thereby, OER following LOM breaks the scaling relation of HOO* and HO*, and surpassing the maximal activities predicted in the volcano plot in AEM become possible. However, there are still some limitations in the LOM. The process of OER following LOM still involves oxygen intermediates, HO* and O*. Thereby the activity in the LOM is affected by the binding energy between oxygen intermediates and active sites,36 which means the optimization of adsorption energies of intermediates is important for LOM. For example, Shao-Horn's group performed DFT calculations and galvanostatic oxidation test and proposed that the deprotonation of the surface OH* group was the rate-limiting step.37 Therefore, a rational-designed OER catalyst following LOM with optimal activities requires both breaking the scaling relation between HOO* and HO* and optimizing the adsorption energies of intermediates.
A number of reaction descriptors have been proposed for the oxygen catalysis (OER and oxygen reduction reaction, ORR), such as the energy of the metal–OH bond (metals),40 enthalpy of transition from lower oxide to nominal oxide transition (metal oxides),41 the delocalization and filling (occupancy) of the antibonding σ* (eg) states (metal oxides),42–46 the metal–oxygen covalency (metal oxides),45,46 (DFT, metal oxides),34 the number of outer electrons (DFT, metals and oxides).47 For metal oxide catalyst in solid oxide fuel cell, the descriptor of O p-band centre relative to the Fermi level (DFT)48 and oxygen and hydroxyl adsorption energies22 have been proposed. A statistical machine learning and analysis of the descriptors has done by Shao-Horn et al., who found the number of d electrons, charge-transfer energy and optimality of eg are important in correlating the OER activity.49 It should be noted that the utilization of the above descriptors is limited by the structural type of catalysts, and it is vital to find the appropriate descriptor to predict and explain the OER activity. For example, the eg orbital filling and number of d electrons can be the effective descriptors for perovskite materials, while the eg occupancy of the cations at octahedral sites can be used for spinel materials.9,46,50
These as-mentioned descriptors will always provide the theoretical foundations to improve the electrocatalytic activity towards a specific reaction. Important parameters have been used to describe the performance of an electrocatalyst, such as overpotential (η), Tafel slope, areal activity, mass activity, faradaic efficiency (FE) etc.12 However, when comparing the electrochemical activity, the overpotentials under the same current density are always influenced by the mass loading. Compared with other parameters, turnover frequency (TOF) is the kinetic parameter indicating the intrinsic activity of each active site. TOF is defined as the number of reactant molecules that a catalyst can convert to the desired product per active site in a unit of time. It can be calculated by the equation: TOF = (j × NA)/(n × F × Γ). j stands for the current density, NA is the Avogadro number, n is the number of transferred electrons, F is the Faraday constant, and Γ represents the number of active sites. From this equation, it can be deduced that at a given TOF, the more active sites the catalyst own, the larger the current density becomes, representing faster kinetics of the specific reaction. Therefore, it is expected that efficient catalysts possess well-exposed abundant active sites.
These analyses have shown the possibility of fine-tuning of both intrinsic activity and the number of active sites on metal/metal oxides. Thus, in the coming section, we will summarize characterization techniques that have been used for such purposes to allow better constructing the active catalytic sites.
Even though these conventional methods have been proved to be adequate to gain details of the electrocatalyst, it is noteworthy that these details are strongly influenced by environmental conditions, such as atmosphere, temperature, and pressure. Further information is still needed to detect the actual active sites and reaction mechanism towards a specific electrochemical reaction those may guide towards the activity enhancement. Therefore, characterizations at specific voltages are required, and in situ experiments are designed for this purpose. Spectroscopic characterizations include Raman spectroscopy, X-ray absorption spectroscopy (XAS), X-ray diffraction (XRD), X-ray photoelectron spectroscopy (XPS), and UV/visible spectroscopy are useful to in situ capture the structural transformation of the electrode and oxidation state change of the catalytic species during a specific electrochemical reaction. While microscopic techniques such as atomic force microscope (AFM) and scanning tunnelling microscope (STM) provide direct morphological snap shoot of the evolving electrocatalysts.
The selection of the characterization technique from the tool box is also subject dependent. Currently, material scientists mainly focus on two aspects to improve the performance of electrocatalysts. One aspect is exposing or building more active sites. In this regard, strategies including nanoscale confinement, phase engineering, and facet engineering are utilized. Since some new active phases or facets may form during the electrochemical test, in situ Raman spectroscopy, XRD, AFM, and STM are effective tools to detect the newly formed active part and investigate the underlying mechanism. The other aspect is improving the intrinsic activity of each active site. Commonly used methods involve heteroatom doping, defect and strain engineering, etc. These methods often result in the changes in electronic structure, such as metal d band and oxygen p band. Therefore, in situ XAS, XPS, and UV/visible spectroscopy can be useful techniques to study the properties of metal oxides. The following part illustrates the specific applications of the aforementioned in situ techniques.
Fig. 3 (a) In situ Raman spectra collected in a large wavenumber region from NiFe LDH during HER process in 1 M KOH at various overpotentials vs. RHE. (b) Magnification of the corresponding orange wavelength region of (a). (c) Schematic illustration. (d–f) In situ Raman spectra of NiFe LDH with 532 nm excitation under OER condition. (d) In situ Raman spectra collected in a large wavenumber region from NiFe LDH during OER process in 1 M KOH at various overpotentials vs. RHE. (e) Magnification of the corresponding green wavelength region of (d). (f) Schematic picture.51 |
Yeo prepared different OER electrochemical catalysts by depositing Co3O4 on various metals and found that substrate metal with larger electronegativity owned better performance towards OER.53In situ Raman spectroscopy results indicated that at an OER active anodic potential, Co species were oxidized to the valence which was higher than 3 and acted as the active species for OER. By performing the in situ surface-enhanced Raman spectroscopy, Smith's group studied the electrocatalytic performance towards OER of two Ni-based catalysts (Ni(Fe)–Bi and Ni(Fe)OOH).54 The authors concluded from the region ca. 900–1150 cm−1 that at the potential before the actual OER occurred, some “active” oxygen sites in Ni(Fe)–Bi and Ni(Fe)OOH got negatively charged and acted as OER precursors. Bron's group also investigated Ni and Ni/Fe thin-film electrodes by in situ Raman spectroscopy.55 At OER potentials, the peaks at 475 and 557 cm−1 in the Raman spectroscopy results showed the formation of γ-NiOOH. With the increasing Fe contents, the peaks of γ-NiOOH deceased, and I475/I557 band ratio also decreased, which meant disorder was introduced into γ-NiOOH. The authors believed that certain amount of Fe can maintain the presence of γ-NiOOH and induce the occurrence of some other structural disorders, which were vital to the improved OER activity. 15% Fe content was determined to be optimum to obtain the best performance.
Fig. 4 (a and b) X-ray absorption near edge structure spectra of the three accessible oxidation states of cubane, performed in MeCN + 0.1 M n-Bu4NPF6.56 (a) Co K-edge X-ray absorption data for the Co(III)4 (green lines), Co(III)3(IV) (blue lines), and Co(III)2(IV)2 (red lines) states, obtained by applying no potential, 0.80, and 1.8 V vs. Fc+/0, respectively. (b) Co-facial Co(IV)2 sites in Co oxygen-evolving catalyst (Co-OEC) are modelled by a Co4O4 cubane. (c–e) In situ structural characterization of Co3O4 films.57 (c) In situ X-ray diffraction patterns. (d) In situ Fourier-transforms (FT) of quasi-in situ EXAFS spectra collected at the Co K-edge. (e) In situ XANES spectra. |
Fig. 5 (a and b) Operando XRD patterns collected on (a) Ni(OH)2/NF and (b) NiCeOxHy /NF electrodes at different potentials. (c) The OER proceeds more favourably on NiCeOxHy with larger interlamellar spacing and higher ECSA relative to Ni(OH)2.59 |
Fig. 6 (a) Normalized and background-subtracted O 1s XPS spectra of Ni–Fe electrocatalyst. (b) Fit results for spectra in (a): relative intensity of the M–O component in the MOx(OH)y catalyst, and measured H2O peak positions (black symbols). (c) Comparison of XPS measurements taken at identical applied potentials before (green) and after (red) the second electrochemical conditioning. Normalized XPS spectra of (d) Ni 2p and (e) Fe 2p transitions for a Ni–Fe electrocatalyst.61 |
As for determining the concentration of active sites towards HER, since it is difficult to differentiate current associated with the metal redox reactions (if any) superimposed in the cathodic faradaic current, the as-mentioned method is no longer suitable. For the determination of electrochemical active surface area, other electrochemical methods, such as carbon monoxide (CO) stripping, hydrogen underpotential deposition (H-UPD) and peroxide oxidation, have also been used.68,69 One can also assume that all the metal atoms are active towards HER or OER, and then calculate the active site number and TOF. However, it is noteworthy that this method produced numerous errors.
DFT is often coupled with characterizations to identify the real active sites. After characterizing the geometrical and electronic structure of electrocatalysts, one can build an appropriate model on the basis of the structural information acquired from the characterizations. The essential information such as Gibbs free energy diagram, density of states (DOS), band structure, and charge distribution, etc. are accessible with the help of DFT methods.70–72 These results will finally reflect in descriptors such as d band centre, M–O covalency, and O p-band centre, which regulate the reaction pathways by governing the adsorption energy of intermediates and potential-determining step (PDS). Therefore, the activity of active sites can be predicted with the calculated descriptors and compared with the experimental results.72 Notably, some electrocatalysts will undergo surface reconstruction and phase transformation, especially when working under OER conditions.23 Thus it is more accurate to call these catalysts used in the published work “pre-catalysts”. The calculation results based on the ex situ characterizations without surface reconstruction/activation could be misleading. Thereby, truthful structural information from in situ observations under operando conditions is more suitable for DFT calculation. Despite some challenges in combinatorial research, such as the lack of DFT tool to effectively simulate the reaction interface under operando conditions with affordable computational cost, or the structural complexity at the interface in atomic scale, computation modelling coupled with in situ characterizations have been the most effective approach to date and will accelerate exploring high-performance electrocatalysts.
To summarize, in situ characterization methods have been briefly summarized as the toolbox for electrocatalyst development that provide researchers with precise information under working conditions as well as guidance to deeper mechanism. The next section of the review is to illustrate that with the help of the characterization techniques, how specific material engineering strategies influence geometric and electronic structures of electrocatalysts and further construct the active catalytic sites on metal oxides.
The top-down creation of nanoparticles from the parent particles with larger sizes is a common strategy to create a large active area or abundant active sites. Wang et al. introduced lithium ions into transition metal oxides and fabricated nanoparticles with an ultrasmall diameter (2–5 nm).73 It was demonstrated that the ultrasmall NiFeOx and CoO nanoparticles exhibited superior catalytic activity towards HER and OER compared to nanoparticles with large diameters (∼20 nm). Moreover, the obtained catalyst achieved excellent stability of 200 h at 10 mA cm−2 without any decay. Different from the catalysts prepared by other traditional chemical synthetic methods,74 the nanoparticles in situ formed via this lithium-induced conversion reaction still maintain excellent electrical interconnection while introducing large surface areas and exposed active sites (Fig. 7a–e).73 From transmission electron microscopy (TEM) images, the interconnected ultrasmall nanoparticles were manifested to have abundant grain boundaries, which acted as the origin of increased surface area and active centres. The rich boundaries were further verified by calculating electrochemical double layer capacities from cyclic voltammetry results. The authors attributed the excellent stability to the in situ preparation method involving delithiation reaction process, which endowed excellent mechanical and electrical contacts between the catalyst and the substrate. On the other hand, the coordination environment of atoms in outer shells is significantly affected by the size. The electronic structure is also primarily altered. Zeng et al. managed to obtain optimal eg orbital occupation by controlling the particle size of perovskite LaCoO3.75 The authors found that when the size was down to 80 nm, LaCoO3 showed the best performance for OER with an overpotential of only 320 mV in 0.1 M KOH electrolyte. The authors explained that the cobalt ions have an optimal electron configuration of eg ∼1.2 according to Shao-Horn’ principle46 when the particle size was ∼80 nm which was beneficial to improve the OER activity. By conducting electron energy loss spectroscopy (EELS) spectra, the enhancement stemmed from the spin-state transition from the low-spin state to a high-spin state of cobalt ions at the surface.
Fig. 7 (a–f) Schematic of transition metal oxides (TMO) morphology evolution under galvanostatic cycles.73 (a–e) TMO particles gradually change from single crystalline to ultra-small interconnected crystalline NPs. Long-term battery cycling may result in the break-up of the particle. (f) The galvanostatic cycling profile of CoO/CNF galvanostatic cycling. (g–i) The XAS results and theoritical results of γ-CoOOH with different size.77 (g) Calculated density of states (DOS) for γ-CoOOH nanosheet, γ-CoOOH (001) thin film and bulk γ-CoOOH. (h) Electronic structure transformation of ultrathin γ-CoOOH nanosheet and model of water oxidation on the nanosheet surface. (i) Calculated adsorption energy for ultrathin γ-CoOOH nanosheet, γ-CoOOH (001) thin film and bulk γ-CoOOH. |
Two-dimensional materials, such as nanosheets and thin films, are also widely explored due to their large surface areas, good electrical conductivity, and abundant low-coordinated surface atoms which are essential to accelerate electrochemical kinetics. Sun et al. prepared atomically thin cobalt oxide porous sheets and carefully investigated the impact of the nanostructure on OER both theoretically and experimentally.76 The authors synthesized porous Co3O4 ultrathin sheets by heating CoO sheets. TEM and AFM results showed that the porous Co3O4 sheets possessed a thickness of only 0.43 nm, which was about half of a unit cell. This ultrathin thickness induced a larger surface area and the expose of all Co3+ of Co3O4, which was regarded as the active site for OER. More importantly, the coordination number of Co3+ atoms decreased from 6 and 5 to 4 or even 3 because of the ultrasmall thickness. According to DFT calculations, the adsorption energy of H2O molecules on lower coordinated Co3+ atoms was larger (0.45 eV), leading to a higher catalytic activity. Furthermore, the density of states at the edge of the valence band and conduction band increased due to the structure disorder on the surface of atomically thick sheets. Therefore, the electron transfer along the conducting channel was promoted, and thus the sluggish OER kinetics was significantly accelerated. Besides, the OER performance for the Co3O4 nanosheet was retained even after 10000 CV cycles, while the bulk counterpart experienced a dramatic drop in OER performance. Through the interfacial charge-transfer resistance test, the outstanding stability was attributed to the unique 2D configuration, which provided intimate contact between the glassy carbon (GC) electrode and the catalyst, and enabled a feasible release of evolved gas bubbles. In another work, ultrathin γ-CoOOH nanosheets were fabricated by atomic-scale phase transformation strategy and manifested superior performance for OER than the state-of-art catalyst.77 The ultrathin nanosheets were found to be half-metallic, and the electric conductivity was 52 times better than the bulk CoOOH. EXAFS and XPS results in Fig. 7g–i showed that the size confinement led to the existence of surface structural distortion of CoO6−x and rearrangement of Co 3d electrons population.77 This rearrangement resulted in forming an electron configuration of t2g5eg1.2, which ultimately improved the electron transfer between surface cations and adsorbed intermediates. The thickness effect of metal oxides on their electrical conductivity was also investigated. Viswanathan et al. investigated the relationship between the critical thickness of TiO2 and the location of valence band maximum relative to the limiting potential for OER to study the charge transport mechanism.78 With the help of atomic layer deposition (ALD), the authors managed to control the thickness of TiO2. Through theoretical analysis, it was predicted that the thermodynamic limiting potential for OER was in the band gap of TiO2, located close to the valence band, and a bias potential was needed to maintain the electrochemical current when the thickness of TiO2 surpassed a certain value. The experimental results also proved that the overpotential occurred when the thickness exceeded 4 nm.
Since the hierarchical structure provides highly exposed active sites and facilitates the charge and electron transfer of active species,79,80 electrocatalysts with different hierarchical structures have been fabricated and presented high electrocatalytic performance. Hierarchical NiCo2O4 hollow microcuboids with 1D porous nanowires subunits were fabricated by Gao and co-workers.81 The electrode exhibited excellent activity for overall water splitting, with a small applied potential (1.65 V) to reach a current density of 10 mA cm−2 towards overall water splitting as well as excellent stability of 36 h at 20 mA cm−2. Zhu et al. prepared a class of novel nickel cobalt oxide hollow nanosponges, which showed excellent catalytic activity towards OER due to their porous and hollow nanostructures.82
Overall, the nanoscale confinement strategy, aiming to control the morphology and size of the metal oxide, benefit the electrocatalytic activity by the following four ways. Firstly, the engineering of morphology in nanoscales, such as nanoparticles and nanosheets, often improves the surface specific area and exposes more active sites. Secondly, downsizing into a scale less than a unit cell can dramatically influence the coordination environment of the surface atoms and further impact the intrinsic activity of metal sites. Thirdly, the modulation of size brings a substantial change in the energy band. Then, the modified metal oxides may transform from insulator or semi-conductor to conductor, and the electron transfer pathway is facilitated. Finally, confinement strategies could facilitate more contact between catalysts and substrates and optimize the release of gas bubble, which significantly improve the durability. Thus, nanoscale confinement can be an effective measure to improve catalytic performance.
Fig. 8 (a and b) The influence of phase engineering on the energy band of Ti2O3 polymorphs.85 (a) Schematic energy band diagram for the Mott insulator and charge-transfer insulator. (b) Evolution of the U and Δ in Ti2O3 polymorphs. (c–e) XANES spectra and schematic illustration of intermediate spin state transition.93 (c and d) Co L-edge (c) and O K-edge (d) XANES spectra for LCO (100), (110), and (111) films. (e) Schematic illustration of the transition of electrons from t2g to eg orbital and the evolution of intermediate spin state. |
To sum up briefly, surface atomic arrangements of a bulk crystal are readily controllable with crystal phase and facet engineering. This provides scientists with an essential “knob” to fine tune the electronic structure and electrocatalytic nature even with the similar elemental composition. Firstly, engineering of phase and facet endow changes in energy band of metal oxides, which might change insulating characteristics into half-metallic or metallic one and modulate electronic transportation. Secondly, the engineering in phase and facet often changes the coordination environment of metal site and the M–O electronic structure, and thereby modulates the intrinsic activity, especially for OER. Finally, even assuming that active sites are with equal activity, crystalline catalysts with specific phases and facets expose more active sites, which certainly improve the overall electrochemical performance.
It has been proven that HER in alkaline solution significantly depends on activating and cleaving O–H bonds of water molecules and facilitating the adsorption and desorption of the formed H atoms. In this regard, the novel concept of composite materials facilitating these two key aspects of HER process in alkaline solution has been reported. A typical example is a heterostructure composed of oxide and metal, in which the oxide acts as the active site for the dissociation of water, while the metal facilitates the adsorption and desorption of hydrogen atoms. Subbaraman et al. prepared nanometre-scale Ni(OH)2 clusters with high water dissociation properties on platinum electrode surfaces.30 With a dual effect, the edges of the Ni(OH)2 clusters facilitated the dissociation of water, and the produced hydrogen intermediates were then adsorbed on the nearby Pt surfaces and recombined into molecular hydrogen. Inspired by this encouraging work, various metal/metal oxide heterostructures have been fabricated. In 2014, Gong et al. prepared a nanoscale NiO/Ni heterostructure attached to a carbon nanotube (CNT) network, which exhibited excellent HER activity in a wide range of pH values as shown in Fig. 9a.94 The NiO/Ni was partially reduced from Ni(OH)2 through thermal decomposition. It was proposed that the exposed NiO/Ni nano-interfaces might be the synergistically active sites towards HER. Specifically, the OH− generated by H2O splitting could preferentially adsorb to a NiO site at the interface, while a nearby Ni site would facilitate H adsorption, boosting synergistic HER catalytic activity to NiO/Ni. In another work, Cr2O3 blended with NiO/Ni was later fabricated, in which the chemically stable Cr2O3 was essential to prevent oxidation of the Ni core and to preserve the catalytically active sites in the heterostructure, and the catalyst with heterostructure obtained great stability of more than 80 h at 200 mA cm−2.95 To date, a number of metal/metal oxide interface nanostructures have been successfully also explored, collectively showcasing that the strategy can be extended to boost the HER performance in alkaline media.96–99
Fig. 9 (a–c) Schematic structure of (a) NiO/Ni-CNT nano-hybrid, (b) NiO/CNT, (c) Ni/CNT structure. (d) High-resolution Ni XPS spectra of the three hybrid materials. (e) Ni L edge XANES spectra of the three hybrid materials. (f) Linear sweep voltametry of the three hybrid materials in 1 M KOH. (g–i) Linear sweep voltammetry of NiO/Ni-CNT and Pt/C in (g) 1 M KOH (h) NaHCO3-Na2CO3 buffer (pH = 10.0) and (i) potassium borate buffer (pH = 9.5).94 (j–l) OER theoretical analysis of IrOx/SrIrO3 heterostructure. (j) Theoretical overpotential volcano plot with O* and OH* binding energies as descriptors, using the scaling relationship between OH and OOH. (k) Visual representation of IrOx and SrIrO3 surfaces used for the DFT calculations. (l) OH* and OOH* binding energies (black circles) overlaid on the universal scaling relationship (gray line with shaded uncertainty) that has been found for many transition metals and transition metal oxides.109 |
Fabricating heterostructures based on metal oxides is also an efficient strategy to obtain remarkable OER electrocatalysts. Specifically, the metallic compounds with good conductivity and large surface area are intensively utilized as the support to improve the electron transfer process and expose active sites. For example, metallic CoSe2 nanobelts were integrated with other metal oxides as a highly active and stable electrocatalyst for OER.100,101 In addition, the OER activity of pristine metal oxide is often hampered by its poor OER kinetics and mass-transfer ability while fabricating heterostructure with other metal oxides/hydroxides is proved to be an efficient strategy to optimize the energy barriers of intermediates and thus boost the catalytic activities.102–104 Cerium(IV) oxide (CeO2) has reversible surface oxygen ion exchange, good electronic/ion conductivity and oxygen storage capacity, and these properties are expected to enhance the catalytic performance of the electrocatalysts by altering the electrochemical reaction kinetics. For instance, electrodeposited FeOOH/CeO2 hybrids were explored as efficient electrocatalysts for OER, exhibiting the synergistic effects between two compounds.105 According to DFT calculations, the free energy changes of intermediates and products were lower than those free energy for CeO2 and FeOOH. The adsorption energy of OH− on the heterostructure was lower than those on individual FeOOH and CeO2 surfaces, which indicated that the OH− anions tended to be adsorbed on FeOOH/CeO2 more efficiently, and this trait endowed the FeOOH/CeO2 with superior catalytic activity for OER. The authors also believed that the strong electronic interaction improved durability since this heterostructure owned great stability of 50 h under 80 mA cm−2. Additionally, CeO2/Ni-TMO hybrid and CeO2/CoSe2 hybrid have also been recently fabricated, exemplifying the efficacious heterostructure methodology for high-performance electrocatalyst development.106,107
The heterostructure with abundant interfaces is also illustrated to improve charge transfer and to facilitate electrochemical turnover. For instance, ultrafine NiO nanosheets stabilized by TiO2 with abundant interfaces were fabricated by annealing a monolayer layered NiTi LDH precursor.108 It has been previously demonstrated that NiO nanosheets with a high proportion of exposed (110) facets presented efficient catalytic performance. The overall enhanced OER activity is originated from three contributing aspects. Firstly, for 3d transition metal-based electrocatalysts, the one with a surface cation eg orbital occupancy approaching unity, shows the best performance for OER. Different from Ni2+ atoms in the hexagonal ultrathin bulk NiO with t2g6eg2 electron configuration, the Ni2+ atoms on the ultrafine and ultrathin NiO nanosheet were partially oxidized to Ni3+ with t2g6eg1 electronic configuration. The Ni3+ atoms highly exposed on (110) facets were stabilized by TiO2 and acted as active sites with higher OER activity. Secondly, according to the DOS and the partial DOS (PDOS) results, the Ni/TiO2 heterostructure is featured with a continuous DOS around the Fermi level, strongly indicating a high carrier concentration and good electrical conductivity. Thirdly, the adsorption energy of H2O for this kind of heterostructure experienced a dramatic increase compared with other materials leading to the significantly enhanced OER activity. It is noteworthy that the heterostructure might form during the electrochemical OER. Thus extra attention is required when investigating the reaction mechanism and identifying the possible active sites. Seitz et al. thoroughly studied the formation of IrOx/SrIrO3 heterostructure in the OER test, during which Sr ions leached from the surface and minor surface rearrangement occurred.109 The authors utilized DFT calculations to examine the activity and stability of catalyst with possible heterostructures, such as IrO2-anatase/SrIrO3, IrO4/SrIrO3, IrO3/SrIrO3 with different layer thicknesses (Fig. 9k).109 Theoretical results indicated that IrO2-anatase/SrIrO3 and IrO3/SrIrO3 with larger IrO3 layer thickness owned both good activity and stability. The OER performance of IrOx/SrIrO3 heterostructure is outstanding with a small overpotential of 270 to 290 mV (at 10 mA cm−2), which was very near to the peak of the OER volcano. Further AFM and XPS characterizations also proved that Sr leached during OER test and thus IrOx film in situ formed as active sites. Besides the metal compound hybrids, integrating the metal oxide catalysts with carbon-based materials can also enhance electrochemical reaction kinetics and dramatically boost activity through tailoring the adsorption energy of intermediates on catalysts. Li et al. fabricated ultrathin edge-rich FeOOH@carbon nanotubes for efficient OER.110 The carbon nanotubes not only improved the electron transfer, but also promoted the electrochemical oxidation of oxygen species. Specifically, due to the p, π-conjugation effect of carbon atoms in graphene, oxygen atoms in FeOOH decreased the density of electron clouds around the Fe atoms. Herein, hybrid to carbon materials will enhance the adsorption of reaction intermediate on central Fe atoms, and their good electronic conductivity can also contribute to improved OER performance. However, since oxidation of carbon may occur under the OER condition, it is important not to mistake the oxidation of carbon as the OER activity.
To sum up, the synergetic effects of heterostructure can be categorized into three aspects. Firstly, the electronic structures of each composite can be efficiently modulated upon heterostructure formation. Thereby, the disassociation of water, the adsorption and desorption of the intermediates and the overall intrinsic electrochemical reaction kinetic process on the active sites are tuned and facilitated. Secondly, the heterostructures combine the unique advantages from each component and provide dual active sites at the interface, which has contributed to the excellent electrochemical performance. The support materials, especially the ones with high conductivity, can also facilitate charge transfer. Finally, some supports may mediate the growth of loaded catalytically active materials and hinder agglomerations of the nanostructured catalysts, which will contribute to higher durability. Because of the aforementioned multiple merits, heterostructure engineering has been one of the primary strategies toward the development of catalysts with high activity and excellent stability.
Fig. 10 (a) Different oxygen sites on W18O49 (010) and Mo–W18O49 (010). (b) The calculated Gibbs free energy of HER on different oxygen sites of W18O49 (010) and Mo–W18O49 (010).113 (c) Schematic illustrations of different RuO6 along with the splitting of Ru 4d orbitals. (d) Schematic evolution of band alignments through Sr2+ substitution for Y3+ in Y2Ru2O7.127 (e and f) Relationship between oxygen vacancy concentration and Co–O bond.115 |
For OER, although the role of cation dopant is still unclear in improving the electrochemical reaction, cation atoms doping has been indicated to efficiently lower the kinetic energy barrier of the water dissociation step and desorption of the formed OH− from the surface.114–116 Recently, researches about cation doped metal oxides, such as ZnxCo3−xO4 nanoarrays, Fe doped BaCo0.9Sn0.1O3−δ and Na1−xNiyFe1−yO2, demonstrated the importance of foreign cations in enhancing OER performance of metal oxides.117–119 Fominykh et al. compared the OER activity of the NiO samples with different Fe dopant concentrations.120 Fe0.1Ni0.9O was found to demonstrate the highest electrocatalytic performance towards OER with an overpotential of 297 mV at a current density of 10 mA cm−2. The Fe3+ doped nickel oxide for significantly enhanced OER has been well studied, and theoretical calculation demonstrated that Fe3+ sites provided optimal binding energy for the OH* and OOH* intermediates and worked as the active site other than Ni3+ sites. Meanwhile, through a in situ XAS experiment, Mukerjee et al. confirmed Fe substitutes stabilized Ni in +2 oxidation state and acted as the active sites for OER.121 However, Li et al. utilized the Lewis acid–base theory to study the role of Fe dopant and they observed that Fe3+ can also function as a Lewis acid and promoted the formation of Ni4+ resulting to an improved catalytic performance.122 The Fe dopant influences the Ni valency, and the formal oxidation state of Ni(IV)-oxo had a character of Ni(III)–O* resonance with an increased covalency. The oxyl radical character played a key role as oxygen radicals in O–O bond formation according to the proton-coupled electron transfer (PCET) mechanism. The research highlighted the role of Fe dopant in boosting OER activity in metal oxide catalyst. Huang and co-worker used Fe dopant to improve the OER performance of NiCo2O4 spinel materials.123 XAS studies showed that Fe ions mainly occupied tetrahedral site of spinel lattice and induced more Ni3+ and Co2+. The increased ratio of Ni3+/Ni2+ and decreased ratio of Co3+/Co2+ made the eg occupancies of Ni and Co closer to unity, respectively, which was the optimal eg filling for OER. Cation doping is also effective in promoting OER performance of perovskite materials. Shao-Horn's group found that the optimum eg filling should be ∼1.2 to achieve the top of the volcano plot.46 Cation doping can finely tune the eg filling of perovskites to achieve better performance towards OER. Based on this guidance, SrNb0.1Co0.7Fe0.2O3, CaCu3Fe4O12 perovskites were explored as catalysts with good activity and stability towards OER, which were comparable or even better than state-of-art OER catalysts.124,125 Xu's group also investigated the influence of Fe dopant on the OER performance of LaCoO3.126 With 10% Fe substitution, the spin state of Co3+ was transformed from a low spin state (LS: t2g6eg0) to a higher spin state, altering the catalyst characteristics from insulator to half-metal. Moreover, the change also increased overlap between Co 3d and O 2p state. Thus, M–O covalency was enhanced, which finally facilitated the OER activity. Zhang et al. improved the OER performance of pyrochlore ruthenate Y2Ru2O7 in acidic medium by substituting Y3+ with Sr2+.127 The mass activity was 1018 A gRu−1 at an overpotential of 300 mV in 0.5 M H2SO4 electrolyte. Compared to standard RuO6 octahedra, the RuO6 octahedra in Y2Ru2O7 are severely distorted, and the 4d electrons of Ru in Y2Ru2O7 are distributed into a lower Hubbard band (LHB), resulting in the upper Hubbard band (UHB) empty. This Mott–Hubbard splitting phenomenon impedes the charge transfer between active sites and intermediates, resulting in an inferior OER performance. For the Y1.7Sr0.3Ru2O7 catalyst, the exotic Sr2+ cations on one hand can regulate coordination geometry and charge redistribution. From the results of XRD Rietveld refinement and Raman spectra, the bond angle of Ru–O–Ru increased from 128.2° to 129.8° after Sr substitution, which alleviated the distortion of RuO6 moieties and improved the overlap of Ru 4d and O 2p orbitals (Fig. 10c).127 On the other hand, the Sr substitution also increased the Ru valence and facilitated the charge transfer from oxygen ligand to Ru centre, which also resulted in enhanced orbital hybridization and bond covalency (Fig. 10d).127
It should be noticed that substitution is the effective method to produce OER electrocatalysts beyond the top of the volcano in AEM. For example, substitution by metals with higher electronegativity can act as proton acceptors. Halck et al. utilized the incorporation of Ni and Co into RuO2 surface to improve the OER performance.128 According to the reaction pathway and scaling relationship in AEM, the smallest overpotential for OER would be 0.4 V. However, after substitution of Ru by Ni and Co, the oxygen atoms on the bridge positions were activated and acted as a proton acceptor. Unlike the traditional scaling relation of HOO* and HO*, the adsorption free energy difference between HOO* and HO* was changed. As a result, the overpotentials for Co- and Ni-substituted RuO2 were only ca. 0.1 V and 0.25 V, respectively. In addition, cation substitutions can also promote OER with the LOM mechanism. Stevenson et al. first proposed the LOM mechanism by using Sr substitution into La1−xSrxCoO3−δ.115 The covalency of the Co–O bond and the concentration of oxygen vacancies can be regulated by controlling Sr substitution (Fig. 10e and f).115 It was demonstrated that the transition from AEM to LOM occurred when x ≥ 0.4. With increasing Sr substitution, the oxidation state of Co was also increased, which led to a more significant overlap between Co 3d and O 2p orbitals and the formation of π* and σ* bands. When the overlap was great enough, ligand holes (oxygen vacancies) were formed, and the Co 3d π* band cannot be treated as an energy level isolated from the oxygen 2p π* band. Therefore, the Fermi energy can be modulated by controlling Co–O covalency, and thus Sr substitution provided the opportunity for lattice oxygen redox activity. The formation of superoxide-like –OO via the coupling of lattice oxygen and adsorption oxygen finally resulted in fast OER kinetics.
In 2017, Xiao et al. developed a facile strategy to fill the oxygen vacancies in the Co3O4 with phosphorus by thermal coupled plasma modification approach.129 The electrocatalytic performance of P-Co3O4 for HER and OER was much better than these of pristine Co3O4 and VO-Co3O4 (VO = oxygen vacancies). The overpotential of overall water splitting was only 420 mV at 100 mA cm−2 in 5 M KOH electrolyte. The electronic properties of the P-Co3O4 have been investigated through both theoretical calculation and experiments. By comparing XAS results of P-Co3O4 and VO-Co3O4, P dopants were confirmed to fill into the O-vacancies. When VO was formed, electrons transferred into Co 3d orbitals, preferably forming Co3+Oh, instead of Co2+Td. While with P dopants in the O-vacancy site, electrons transferred out of Co 3d states, and thereby the number of Co2+Td sites (with better HER and OER catalytic activity) increased. More importantly, P-filling can also significantly improve the intermediate binding, hence remarkably boost the HER and OER activity. According to the calculated free energy diagram for HER, P-Co3O4 has a favourable ΔGH* of −0.08 eV. The free energy profiles along the OER process also demonstrated enhanced binding strength of O* and OOH* intermediates on P-Co3O4. Co3O4 with P-doping showed superior OER performance to pristine Co3O4. Nitrogen doped molybdenum trioxide (N-doped MoO3) was fabricated by heating commercial MoO3 at NH3 atmosphere and exhibited 6 times higher activity towards HER than intact MoO3.130 Positron annihilation spectrometry (PAS) confirmed that the N-doping induced increased concentration of VOVO divacancies, which might be the active sites for HER. Further DFT calculations verified that the existence of VOVO divacancies improved the state density near the valence band edge and thereby boosted the catalyst's electronic conductivity.
Anion doping can be facile to improve electrocatalytic performance of metal oxides in terms of activity and durability of OER. Through solid-state reaction, P doped SrCoO3−δ (SCP) perovskite has been fabricated and showed high OER performance benefiting from its high electrical conductivity and large amount of OO2−/O− species.131 The doping of high-valence-state P5+ in SCP resulted in charge compensation and boosted the electrical conductivities. Importantly, compared to SrCoO3, P doped SrCoO3−δ showed better durability which can be attributed to the stable tetragonal structure after P-doping. Controllable nitrogen doped cobalt oxides (N-Co3O4 and N-CoO) were synthesized from cobalt-alanine complexes by calcination at different atmosphere.132 The authors highlighted the OER performance of N-Co3O4, which required overpotential of only 190 mV at 10 mA cm−2 and showed the smallest Tafel slope of 29.8 mV dec−1 in 1.0 M KOH. The authors, through DFT calculations, claimed that N-doping can not only improve the OH-adsorption capability, but also facilitate OH-cleavage. Moreover, N-doping also reduced the bandgap from 1.80 eV to 1.42 eV, which made the excitation of charge carriers to the conduction band easier and improve the electronic conductivity. Sulphur doped cobalt oxide (CoO0.87S0.13/GN) was also investigated to explore the benefit of anion doping to the electrochemical reaction performance.133
It is acknowledged that the heterogeneous electrocatalytic process only occurred on the surface of the electrocatalysts, and the activities of the electrocatalysts are determined by their size and the accessible active sites. Therefore, preparing electrocatalyst in atomic scale will contribute to the enlarged exposure of the active sites and thus enhance electrochemical activity. Recently, Dou et al. prepared atomic-scale CoOx species through on-site transformation of the atomically distributed Co2+ in ZIF-67 by O2 plasma treatment.134 Benefiting from the large surface area and abundant exposed active species, the electrocatalyst exhibited remarkable OER performance.
In addition, metal oxide is usually utilized as the support to promote the distribution of the electrocatalytic active species. Especially used for selective activation of methane or CO oxidation, various single atom noble metals supported on metal oxide are intensively explored with remarkably enhanced catalytic performance.135,136 More recently, Nong et al. prepared Ru doped TiO2, in which the penta-valence Ru atom was homogeneously distributed in the lattice.137 The experiment and PDOS calculation confirmed that the Ru5+ existed in a row of –RuO3 in the surface of TiO2, which was defined as TiO2:Rusurf. According to the DFT calculation, the adsorption of hydrogen is preferred on the TiO2:Rusurf structure with increase of hydrogen coverage and the calculated ΔGH on these sites was −0.28 eV, suitable for HER.
To sum up, elemental doping can be a facile and effective method in boosting the electrochemical performance of metal oxides. Dopants can regulate the charge transfer properties and intermediate adsorption energy by modifying the bulk electronic structure, spin state of metal site, and M–O covalency. Besides, some dopants, especially the metals with higher electronegativity, can serve as proton acceptor. These dopants can introduce lattice oxygen, and thus produce efficient OER catalyst following LOM instead of AEM, providing opportunities to break the scaling relation. Finally, some dopants are the efficient active sites by themselves, especially for cases of cation doping and single atom doping.
Fig. 11 (a and b) Plausible reaction mechanism and DFT calculation results of electrocatalytic H2 evolution on WO2.9.143 (c) The projected density of states (PDOS) on pristine CoO and (d) CoO with O-vacancies.144 (d–f) Computational predictions for the strain effect on the HER activity of CoO.146 (d) Schematic illustration of H2O adsorption and dissociation on the CoO (111) surface with O-vacancies (e) hydrogen adsorption free energy, ΔGH*, vs. tensile strain for the CoO (111)-Ov surface. (f) Schematic illustration of the effect of strain on the electronic structure of (111)-Ov surface of CoO. |
In general, defect engineering in metal oxides mainly focus on the control over oxygen vacancies and metal site vacancies. By tuning vacancies, the charge transfer ability and adsorption energy of reactants and intermediates are optimized, thus, the intrinsic catalytic activity of active site is improved. In addition, when the defects accumulate to an exact extent, amorphization may be introduced, and more active sites can be exposed. Strain engineering can also control the vacancies and achieve similar results of anion defect engineering. Also, expansion and compression of catalyst lattice allows modulation in electronic structure towards optimal catalytic activity.
Indra et al. compared the electrochemical catalytic activity of amorphous and crystalline cobalt iron mixed metal oxides for OER and ORR.154 It was found that the amorphous CoFe2O4 can achieve a current density of 10 mA cm−2 with a small overpotential of 490 mV for OER, which was 70 mV lower than that of crystalline one. The authors considered that the better activity was from amorphous surface which exposed more Co3+ active sites and owned higher specific surface area. In 2014, Geng's group reported an aerosol-spray-assisted fabrication of amorphous metal oxides with precise control over compositions and structures.155 Among all the Fe–Ni–Ox prepared, Fe6Ni10Ox showed the best performance for electrochemical water oxidation, with a low overpotential of 0.286 V at 10 mA cm−2 under alkaline conditions. The effect of crystalline phase on electrochemical water oxidation by calcining the samples in different temperatures was further investigated. The results showed that the superior electrochemical catalytic activity of amorphous phase derived from its larger capacitance, which indicated the amorphous phase possessed a larger ECSA. Other authors also drew the similar conclusion that the large surface specific area of amorphous phase was the origin of improvement in electrochemical catalytic performance for OER.156,157
In general, compared with the crystalline phase, the amorphous phase obtains a larger surface specific area and thereby exposes more active sites.156 Besides, the amorphous surface possesses a high concentration of active sites, which play an important role in enhancing the adsorption of reactants and optimizing the coverage of intermediates.158
Hu's group reported Ni2P/NiOx catalysts with the core–shell structure derived from Ni2P particles towards overall water-splitting.163 The Ni2P/NiOx delivered a current density of 10 mA cm−2 with an overpotential of only 290 mV in alkaline electrolyte. The partial oxidation of Ni2P generated NiOx shell, which served as the active part towards OER. The Ni2P core acted as conducting support and provided favourable electron transfer pathway to the NiOx. The author declared there might be synergistic effects between these two parts, which needed further study. In another work, NiFeOx nanosheet was in situ transformed from NiFeSe2 nanoplate by implying a galvanostatic scan of 5 mA cm−2.164 The as-prepared catalyst exhibited a small overpotential of 195 mV at 10 mA cm−2 and a good stability of 24 h under 10 mA cm−2. From TEM images and EDS mapping results, the precursor nanoplate NiFeSe2 was entirely converted into NiFeOx, and the obtained oxide maintained the nanosheet morphology with only 1–2 nm thickness. Benefiting from the ultrathin nanosheet morphology and Ni–Fe synergic effect, the NiFeOx outperformed NiFe LDH and even other benchmark catalysts. In addition, Cui et al. utilized in situ oxidation to develop binary, ternary, and quaternary transition metal oxide from sulfides.165 Impressively, CoNiFeOx grown on carbon fibre exhibited a small overpotential of 232 mV at reach 10 mA cm−2 and extraordinary durability of more than 100 h, outperforming most of other non-noble OER catalysts.
Overall, in situ transformation is an effective method to obtain metal–oxide-based catalysts with good performance. Firstly, the metal oxides normally possess analogous morphology to their pre-catalyst counterparts, and the specific morphology would substantially boost the electrocatalytic activity. Secondly, partial in situ oxidation may produce specific core–shell heterostructures, and the synergic effect between these two different parts may be beneficial to the enhanced overall electrocatalytic performance. In addition, the anion elements of the pre-catalysts inevitably undergo leaching into the electrolytes. Thus, the resultant true electrocatalysts are endowed with higher specific areas and more exposed active sites which benefit the electrocatalytic turnovers.
Taken together, this part of the review has surveyed strategies towards modulations of the morphology, composition, and structure of metal oxide electrocatalysts. A summary of the high-performance metal oxide electrocatalysts towards water splitting developed by these strategies is given in Table 1. Generally, by tuning the electronic structures and surface properties, these strategies can boost the intrinsic activity of each active site, improve the number of active sites, and optimize the charge transfer process.
Strategy | Electrocatalyst/substratea | Mass loading (mg cm−2) | Reaction | Electrolyte | η 10 (mV) | Tafel slope (mV dec−1) | Stability | Ref. |
---|---|---|---|---|---|---|---|---|
a CFP = carbon fibre paper; GC = glassy carbon; FTO = fluoride doped tin oxide; CNT = carbon nanotube; NC = nitrogen-doped carbon. | ||||||||
Nanoscale confinement | NiFeOx nanoparticles/CFP | 3 | HER | 1 M KOH | 88 | 150 | 100 h (@10 mA cm−2) | 73 |
OER | 230 | 31.6 | ||||||
Water splitting | 280 | N.A. | ||||||
LiCo0.33Ni0.33Fe0.33O2 nanoparticles/CFP | ∼0.1 | OER | 0.1 M KOH | 340 | 73.6 | 1000 cycles | 74 | |
80 nm-LaCoO3/GC | 0.25 | OER | 0.1 M KOH | 490 | 69 | 3 h (@10 A g−1) | 75 | |
Co3O4 nanosheets/GC | 0.34 | OER | 1.0 M KOH | 790 (@341.7 mA cm−2) | 25 | 10000 cycles | 76 | |
γ-CoOOH nanosheets/GC | ∼0.15 | OER | 1.0 M KOH | 300 | 38 | 13 h (@10 mA cm−2) | 77 | |
Co3O4 microtubes/Ni foam | N.A. | HER | 1.0 M KOH | 170 | 98 | 12 h (@20 mA cm−2) | 79 | |
OER | 260 (@150 mA cm−2) | 84 | ||||||
Water splitting | 400 | N.A. | ||||||
Co3O4 nanotubes/GC | 0.208 | OER | 0.1 M KOH | 390 | 76 | 2000 cycles/40 h (@20 mA cm−2) | 80 | |
NiCo2O4 hollow microcuboids/Ni foam | ∼1 | HER | 1.0 M NaOH | −110 | 49.7 | 36 h (@20 mA cm−2) | 81 | |
OER | 290 | 53 | ||||||
Water splitting | 420 | N.A. | ||||||
Ni–Co2–O hollow nanosponges/GC | 0.2 | OER | 0.1 M KOH | 362 | 64.4 | 500 cycles | 82 | |
Crystal phase and facet engineering | γ-Ti2O3 film/(001)SrTiO3 | N.A. | HER | 0.5 M H2SO4 | 271 | 199 | N.A. | 85 |
MnOx film/FTO or Au | N.A. | OER | 1.0 M NaOH | N.A. | 63 | N.A. | 86 | |
Co3O4@CoO nanocubes/GC | ∼0.025 | OER | 0.5 M KOH | 430 | 89 | 1000 h (@8 mA cm−2)/6000 cycles | 88 | |
0.5 M Na2SO4 | 851 | 375 | ||||||
BaNi0.83O2.5/GC | 0.295 | OER | 0.1 M KOH | ∼380 | N.A. | 1000 cycles | 89 | |
(110) NiCo2O4 nanosheets/Ni foam | 1.12 | HER | 1.0 M KOH | 157 (@5 mA cm−2) | 71.2 | 50 h (@5 mA cm−2) | 90 | |
OER | 330 (@5 mA cm−2) | 59.2 | ||||||
Water splitting | 360 | N.A. | ||||||
(112) Co3O4 nanoparticles/GC | 0.15 | OER | 0.1 M KOH | 380 | 62 | ∼7 h(@1 mA cm−2) | 91 | |
(012)-O α-Fe2O3/CFP | 0.014 | OER | 1.0 M NaOH | 317 | 58.5 | 30 h (@10 mA cm−2) | 92 | |
LaCoO3 (100) film/GC | N.A. | OER | 1.0 M KOH | 470 | 180 | 20000 s | 93 | |
Heterostructure | Ni(OH)2-Pt/GC | N.A. | HER | 0.1 M KOH | 100 | 100 | N.A. | 30 |
NiO-Ni/CNT | 0.28 | HER | 1.0 M KOH | <100 | 82 | 20 h (@−20 mA cm−2) | 94 | |
Ni@Cr2O3-NiO/Ni foam | 8 | HER | 1.0 M KOH | 115 (@100 mA cm−2) | N.A. | 80 h (@−200 mA cm−2) | 95 | |
CoOx@CN/GC | 1 | HER | 1.0 M KOH | 232 | 115 | 9000 s (@10 mA cm−2) | 96 | |
OER | 260 | N.A. | ||||||
Water splitting | 320(@20 mA cm−2) | N.A. | ||||||
Ni(OH)2-Pt/GC | N.A. | HER | 0.1 M KOH | 100 | 100 | N.A. | 30 | |
NiO-Ni/CNT | 0.28 | HER | 1.0 M KOH | <100 | 82 | 20 h (@−20 mA cm−2) | 94 | |
Ni@Cr2O3-NiO/Ni foam | 8 | HER | 1.0 M KOH | 115 (@100 mA cm−2) | N.A. | 80 h (@−200 mA cm−2) | 95 | |
CoOx@CN/GC | 1 | HER | 1.0 M KOH | 232 | 115 | 9000 s (@10 mA cm−2) | 96 | |
OER | 260 | N.A. | ||||||
Water splitting | 320(@20 mA cm−2) | N.A. | ||||||
Ni/NiO-NSAs/Ni foam | 0.59 | HER | 0.1 M KOH | ∼230 | 114 | 10 h | 97 | |
Ni-CeO2/CNT | 0.14 | HER | 1.0 M KOH | <100 | N.A. | 10 h(@−0.153 V) | 98 | |
Co-Co3O4/Ni foam | 0.85 | HER | 1.0 M KOH | 90 | 44 | 6000 s (@−0.120 V) | 99 | |
Mn3O4-CoSe2/GC | ∼0.2 | OER | 0.1 M KOH | ∼350 | 49 | 12000 cycles | 100 | |
CoO-CoSe2/Ti | ∼2 | HER | 0.5 M PBS | 337 | 131 | 10 h | 101 | |
OER | 510 | 137 | ||||||
Water splitting | 950 | |||||||
Mn3O4@MnxCo3−xO4/Ni foam | 0.3 | OER | 1.0 M KOH | 246 | 46 | 40 h | 102 | |
Co3O4-NiCo2O4/Ni foam | 1 | OER | 1.0 M KOH | 340 | 88 | 12000 cycles | 103 | |
NiO-Co3O4/NC | 0.2 | OER | 1.0 M KOH | 240 | 73 | 48 h | 104 | |
FeOOH-CeO2/Ni foam | N.A. | OER | 1.0 M NaOH | ∼240 | N.A. | 50 h (@80 mA cm−2) | 105 | |
CeO2-CoSe/GC | 0.2 | OER | 0.1 M KOH | 288 | 44 | 10 h (@10 mA cm−2) | 106 | |
CeO2-TMO/Ni | N.A. | HER | 1.0 M KOH | 93 | 69 | 30 h (@10 mA cm−2) | 107 | |
OER | 220 | 38 | ||||||
Water splitting | 350 | N.A. | ||||||
NiO/TiO2 | 0.34 | OER | 1.0 M KOH | 320 | 52 | 10 h | 108 | |
IrOx/SrIrO3 film | N.A. | OER | 0.5 M H2SO4 | 270 | N.A. | 30 h (@10 mA cm−2) | 109 | |
FeOOH/CNTs | 0.05 | OER | 1.0 M KOH | 206 | 31 | 10 h (@10 mA cm−2) | 110 | |
Heteroatom doping | Pr0.5(Ba0.5Sr0.5)0.5Co0.8, Fe0.2O3−δ/GC | 0.232 | HER | 1.0 M KOH | 237 | 45 | 25 h (@−50 mA cm−2) | 111 |
Ni,Zn-CoO/CFP | N.A. | HER | 1.0 M KOH | 53 | 47 | 24 h (@−10 mA cm−2, 6 M KOH) | 112 | |
Mo–W18O49/GC | 0.16 | HER | 0.5 M H2SO4 | 45 | 54 | 10 h (@−0.14 V) | 113 | |
Sr0.90Na0.10RuO3/GC | 0.1 | OER | 0.1 M HClO4 | 170 | 40 | N.A. | 114 | |
La1−xSrxCoO3−δ/GC | 0.051 | OER | 0.1 M KOH | ∼370 | 31 | 10 h (@10A g−1) | 115 | |
LiCo0.8Fe0.2O2/GC | 0.232 | OER | 0.1 M KOH | 340 | 50 | 6 h (@10 mA cm−2) | 116 | |
Na1−xNiyFe1−yO2/GC | 0.13 | OER | 1.0 M KOH | 260 | 44 | 30 h (@10 mA cm−2) | 117 | |
ZnxCo3−xO4/Ti | N.A. | OER | 1.0 M KOH | 320 | 51 | ∼2 h (@1.63 V) | 118 | |
BaCo0.9−xFexSn0.1O3−δ/GC | 0.232 | OER | 0.1 M KOH | ∼390 | 69 | 2 h (@5 mA cm−2) | 119 | |
Fe0.1Ni0.9O/Au | N.A. | OER | 0.5 M KOH | 297 | 37 | 10 h(@10 mA cm−2) | 120 | |
Ni0.8Fe0.1Co0.1O/GC | 0.25 | OER | 0.1 M KOH | ∼410 | N.A. | N.A. | 121 | |
Fe-NiCo2O4/FTO | N.A. | OER | 1.0 M NaOH | 201 | 39 | N.A. | 122 | |
SrNb0.1Co0.7Fe0.2O3/GC | 0.232 | OER | 0.1 M KOH | 420 | 76 | 1000 cycles | 124 | |
CaCu3Fe4O12/GC | 0.25 | OER | 0.1 M KOH | 310 | 51 | 100 cycles | 125 | |
Y2−xSrxRu2O7/GC | 0.071 | OER | 0.5 M H2SO4 | 264 | 44.8 | 28 h (@−10 mA cm−2) | 127 | |
P-Co3O4/Ti mesh | N.A. | HER | 1.0 M KOH | 120 | 52 | ∼2 h (@1.64 V) | 129 | |
OER | 280 | 51.6 | ||||||
N-MoO3/GC | 0.694 | HER | 0.5 M H2SO4 | 210 | 101 | 1000 cycles | 130 | |
SrCo0.95P0.05O3−δ/GC | 0.232 | OER | 0.1 M KOH | 480 | 84 | 1000 cycles | 131 | |
N-Co3O4/Ni foam | 0.25 | OER | 1.0 M KOH | 190 | 30 | 1.11 h | 132 | |
CoO0.87S0.13/graphene nanomesh | 0.36 | OER | 0.1 M KOH | 360 | N.A. | 3000 cycles | 133 | |
CoOx-ZIF/GC | ∼0.2 | OER | 1.0 M KOH | 318 | 70.3 | 2000 cycles | 134 | |
TiO2:Ru/GC | ∼0.2 | HER | 0.1 M KOH | 150 | 97 | 1000 cycles | 137 | |
Defect and strain engineering | Co3−xO4/GC | N.A. | OER | 1.0 M KOH | 268 | 38.2 | 2000 cycles | 141 |
Sn1−xCo0.9Fe0.1(OH)6/GC | ∼0.1 | OER | 1.0 M KOH | 270 | 42.3 | 2000 cycles | 142 | |
WO2.9/GC | 0.285 | HER | 0.5 M H2SO4 | 70 | 50 | ∼4 h (@−1.64 V) | 143 | |
CoO/CFP | ∼0.19 | OER | 1.0 M KOH | 330 | 44 | N.A. | 144 | |
3.0% S-CoO/CFP | ∼0.46 | HER | 1.0 M KOH | 73 | 83 | 28 h (@−0.073 V) | 146 | |
4.2% S-SrCoO3−δ/GC | N.A. | OER | 0.1 M KOH | ∼330 | 40 | ∼3 h (@5 μA) | 147 | |
S-LaNiO3/GC | N.A. | OER | 0.1 M KOH | ∼350 | N.A. | N.A. | 150 | |
Amorphization | IrOx/FTO | 0.0001 | OER | 1.0 M H2SO4 | 190 | 34 | 24 h (@1 mA cm−2) | 151 |
CoFeOx/GC | 0.051 | OER | 0.1 M KOH | 490 | N.A. | N.A. | 154 | |
Fe6Ni10Ox/GC | 0.1 | OER | 1.0 M KOH | 286 | 48 | N.A. | 155 | |
VFeCoOx/GC | 0.28 | OER | 1.0 M KOH | 307 | 36 | 12 h (@−1.55 V) | 156 | |
In situ transformation | Ni2P/NiOx/GC | 0.14 | OER | 1.0 M H2SO4 | 290 | 47 | 10 h (@10 mA cm−2) | 163 |
NiFeOx nanosheet/Ni foam | N.A. | OER | 0.1 M KOH | 195 | 28 | 24 h (@10 mA cm−2) | 164 | |
CoNiFeOx/carbon fibre cloth | N.A. | OER | 1.0 M KOH | 232 | 38 | 100 h (@20 mA cm−2) | 165 |
Firstly, although a variety of high-performance electrocatalysts have been developed, a mechanistic understanding of the surface charge transfer and catalytic sites during the electrochemical reaction is still highly demanding. Recently, some descriptors different from conventional ΔGH* and (ΔGO* − ΔGHO*) for HER and OER have been reported to evaluate the catalytic performance as well as rationally design materials. For example, the descriptor Φ considering the topological, bonding, and electronic structures of catalytic centre, the eg-filling descriptor in perovskite oxides, and the one involved d band position are applied to predict the electrocatalytic activities towards HER and OER.46,166,167 Additionally, advance in mechanistically understanding fundamental reaction steps has been realized by theoretical calculation, along with well-recognized principles providing useful avenues to study various electrochemistry towards other applications. However, mechanistic studies lack the understanding of reaction kinetics, which may need delicate experimental design. Therefore, combining with theoretical elucidations and experimental exploration, robust, high-performance electrocatalysts will be developed through effective catalyst activation strategies, such as single-atom or dual-atom sites creation, crystal phase engineering, heteroatoms doping, and interface manipulation.168–171
Secondly, the poor stability of the metal oxides electrocatalysts severely blocks their industrial application, compared with noble metal counterparts. Under long-term electrochemical test, the surface of metal oxides could undergo surface reconstruction, such as elemental enrichment or depletion.172 Meanwhile, the dissolution or deposition of electrode materials/impurities at the electrocatalyst/substrate interface may hamper the physical contact or restrict the reaction. For instance, Ru-based oxides are regarded as the most active OER materials, while they suffer from severe performance degradation due to Ru dissolution experimentally.66 Therefore, improvement in structural robustness is essential to drive HER or OER for long-term operation. Moreover, in practical application, the current density required is usually higher than 1000 mA cm−2 for proton-exchange membrane (PEM) electrolyser, requiring catalysts with robust stability up to 1000 h.173 To improve the long-term durability, incorporating molecular catalyst with carbon materials has been reported, significantly accelerating water oxidation half-reaction.174 Although endowing with many advantages, the biggest challenge of PEM is that only noble metal-based materials can be used as working electrodes so far, wherein Pt-group metals (PGM) serve as cathode and Ir or Ru-based oxides as the anode. Recently, the development of alkaline anion exchange membrane (AEM) electrolysers provides possibility for the use of PGM-free catalysts, while the limited durability remains a challenge to be solved.175 Over the last decade, alkaline membrane fuel cells (AMFCs) have been achieved with improved current densities and stability.176 An AMFC operated continuously for over 1000 h at 600 mA cm−2 has been reported by developing gas diffusion electrodes (GDL) with hydrophobic PTFE in both the GDL and catalyst layers.177 Thus, the structural integrity of the metal oxides should be assessed with a detailed investigation on the structural degradation or mechanical breakdown.178
Thirdly, accompanying with the development of theoretical methods, complementary operando characterization techniques are also highly desirable to precisely detect structures vibration and identify active sites during electrocatalysis process. Therefore, a rational design of well-defined electrochemically active components needs to comprehensively be tested with various parameters combining with in situ probing techniques for the metal oxides (as shown in Part 2). These technologies are also of paramount importance for understanding the reaction mechanism more profoundly. For example, accurate in situ detection of the atomic or electron structure could be realized by employing well-defined atomically thin electrode materials or adopting surface-sensitive probing techniques.65,179 Additionally, reactant isotope labelling experiments that utilize the kinetic isotope effects are also useful to probe the forming mechanisms of intermediates as well as rate-limiting step.180
In summary, the future industrial application of robust metal oxides for energy conversion and storage demands comprehensive advancement from techniques, theories, characterizations, and devices. We believe with a great development of these key elements, impactful breakthroughs in real-world large-scale water electrolysis are becoming indispensable in the expectable future.
Footnote |
† Equal contributions. |
This journal is © The Royal Society of Chemistry 2021 |