Lu Zhang,
Wanhong He,
Xu Xiang*,
Ying Li and
Feng Li
State Key Laboratory of Chemical Resource Engineering, Beijing University of Chemical Technology, Beijing, 100029, China. E-mail: xiangxu@mail.buct.edu.cn; Fax: +86-10-64425385; Tel: +86-10-64412109
First published on 14th August 2014
High-efficiency and Earth-abundant electrocatalysts for the oxidation of water are required in the production of clean energy from the electrolysis or photolysis of water. Spinel Co3O4 microcrystals with a windmill shape were grown on a flexible metal substrate. The microcrystals were then roughened by a surface impregnation treatment. A secondary nanostructure grew out of the blades of the windmills to form a micro/nano hierarchical structure. The as-grown micro/nano Co3O4 had an excellent electrochemical performance in the oxidation of water. The onset overpotential of the micro/nano Co3O4 electrocatalyst for the oxidation of water was about 0.29 V in alkaline solution and the overpotential of the optimum Co3O4 electrocatalyst was 0.41 V at a current density of 10 mA cm−2. These results suggest that the electrochemical performance is associated with the roughness and active surface of the Co3O4 electrodes. The turnover frequency of the optimized Co3O4 reached 0.39 s−1 at an overpotential of 0.6 V, about 1.4 times higher than that for the pristine Co3O4 microcrystals. The turnover frequency of micro/nano Co3O4 is higher than, or comparable to, that previously reported for high-efficiency nanosized Co3O4 in alkaline solution. Stability tests indicated that these micro/nano Co3O4 electrocatalysts were highly durable towards the oxidation of water, with no structural change and no decrease in noticeable activity after operating for 12 h in oxygen-evolving reactions. This work verifies the contribution of surface roughness and an active surface to the electrochemical oxidation of water by the design of an optimum micro/nano Co3O4 electrocatalyst. Our current understanding of the catalytic role of a specific micro/nano structure is also strengthened.
Many attempts have been made to find high-efficiency and Earth-abundant alternatives to the noble metal catalysts, such as Co, Mn, Ni and Fe-based metal oxides or molecular complexes.10–15 Although molecular or heterogeneous catalysts have shown a high electrocatalytic performance for the oxidation of water in various electrolytes, they usually occur as molecular units or as dispersive particles separated from the electrode materials.16–18 Tedious post-treatment is required to bind or cast the powders onto the surface of the electrode and peel-off is a common problem. As a result, it is challenging to integrate the catalysts directly onto the electrodes for a practical and engineered product with no significant loss of catalytic activity and durability. Some workers have explored constructions based on a concept of structured catalysts.19–21
As an efficient, Earth-abundant and easily made electrocatalyst, Co3O4 has received much attention over the last few decades, although it is slightly less active than the noble metal oxides for the oxidation of water. A variety of spinel Co3O4 or doped cobalt oxide structures have been synthesized.18,19,22–25 Koza et al.23 synthesized crystalline and amorphous Co3O4 films using an electrodeposition method, which showed high OER activity in alkaline solution. Han et al.18 reported the deposition of nanostructured cobalt oxide catalysts from molecular cobaloximes for the oxidation of water in an alkaline buffer solution. Zou et al.25 prepared porous Ni-doped Co3O4 nanomaterials via the thermal treatment of graphitic C3N4 embedded with metal ions.25 The porous structure and Ni doping led to a superior electrocatalytic activity for the OER. Nanostructuring or roughening the pristine surface of catalysts are the most common ways of increasing the number of surface active sites in catalytic reactions. For example, Li et al.24 reported NixCo3−xO4 nanowire arrays for electrocatalytic oxygen evolution, in which the surface was roughened by constructing an Ni-doped mesoporous structure. A greater roughness resulted in a higher catalytic activity. Inspired by these studies, we thought that roughening the surface of pristine Co3O4 by a facile means could be a promising way of enhancing its catalytic activity for the oxidation of water.
Windmill-shaped spinel Co3O4 microcrystals were therefore grown directly on a metal substrate and subsequently roughened by a facile post-treatment of the surface. Film electrodes consisting of a micro/nano structure were formed and these electrodes showed a roughness-dependent electrochemical OER performance. These findings suggest that the OER activity may be enhanced by roughening pristine Co3O4 microcrystals into an optimized micro/nano hierarchical structure. The enhancement in activity was attributed to the increased roughness and active surface of the electrodes, which resulted in more active sites being accessible to the reactants during the OER.
Electrochemical impedance spectroscopy (EIS) measurements were carried out at an applied potential of 0.56 V (vs. SCE) in 0.1 M KOH solution on a CHI660C electrochemical workstation (CH Instrument Co.) with an AC amplitude of 10 mV. The impedance spectra were recorded in the frequency range 100 kHz to 0.01 Hz at room temperature. Before the measurement, KOH solution was bubbled with high-purity N2 at a flow-rate of 25 mL min−1 for 30 min. Before recording each impedance spectrum, the electrode was first equilibrated at a constant potential for 120 s. The experimental impedance data were fitted to a specific equivalent circuit using ZSimPWin software to derive the resistance of the electrodes.
The amount of oxygen evolved was measured using a NeoFOX oxygen-sensing system (Ocean Optics Inc.) equipped with a FOXY probe inserted into the head-space of a gas-tight cell. The cell was purged with nitrogen for 20 min before measurement. The oxygen fluorescence signal was allowed to stabilize once the nitrogen had stopped. After the residual oxygen signal became constant, the desired potential was applied. Oxygen evolution was monitored at 1.64 V (vs. RHE) for 1 h. The fluorescence data measured was transformed into the amount of O2 in μmol.
![]() | ||
Fig. 1 XRD patterns of Co3O4 grown on a metal foil by a surface treatment method. (A) Co3O4, (B) Co3O4-1, (C) Co3O4-3 and (D) Co3O4-5 (closed triangle = reflections from Ni). |
Raman spectroscopy is a useful tool for revealing the delicate structure associated with the vibration modes of bonds. Fig. 2 shows the Raman spectra of the Co3O4 film electrodes. Three intensive vibration bands at about 194, 476 and 681 cm−1 are observed and also two weak bands at 520 and 615 cm−1. These Raman shifts are characteristic of the spinel structure Co3O4, consistent with the assignments reported previously.26 Specifically, the band at 681 cm−1 is assigned to the A1g mode, the band at 476 cm−1 is the Eg mode and the bands at 194, 520 and 615 cm−1 are assigned to the F2g mode.27 No vibration band related to either the Ni–O nor the Cr–O species was observed. The spectrum of commercially available Co3O4 powder is also shown. The A1g mode appears at 689 cm−1 and the Eg mode at 481 cm−1, with a higher frequency shift of 5–8 cm−1 compared with the micro/nano Co3O4 (Co3O4-1, Co3O4-3, Co3O4-5). This finding suggests that the micro/nano hierarchical structure may cause a slight structural distortion in spinel Co3O4, leading to a medium shift in the vibration modes towards a lower frequency.26
The morphology of the Co3O4 grown on the foil was observed by SEM. The windmill-like microcrystals were found to grow vertically or were tilted towards the substrate with respect to the pristine Co3O4 (Fig. 3A). Detailed observations show a six-bladed windmill structure with a lateral size of about 10–12 μm. This structure has rarely been reported previously for the Co3O4 or doped Co3O4; nanowires or rods are more usually observed.20,24 After treatment with an impregnation–calcination procedure, the structure of the microcrystals evolves (Fig. 3B–D). The needle-like branches grow up from the blades of the windmill structure. The density of the needles increases and secondary growth on the primary needle-like structure occurs when the amount of cobalt precursor impregnated increases from 1 to 3% (Fig. 3C). Such a growth mode could roughen the surface of the microcrystals, resulting in more catalytically active sites. However, the needles fuse together and become thicker when the amount of cobalt precursor impregnated is increased to 5% (Fig. 3D), which decreases the roughness and the number of redox-active species. When a pure Ni foil was used as the substrate, only densely grown Co3O4 nanowire bundles were observed (Fig. 4), significantly different from the windmill-like structure. Nevertheless, the digital photos of these samples show no visible difference; they all have a black appearance (Fig. S2†). Elemental analyses using EDS measurements suggest that the substrate for Co3O4 growth consists of Ni of 73%, Cr of 21% and a small amount of Fe and Mn (Fig. S3 and Table S1†). Therefore the presence of Cr (or Mn, Fe) in the foil could be a key factor in the formation of this structure. It has been reported that Co3O4 shows distinct morphologies when grown on different substrates.28 We suggest that the growth mode of Co3O4 is affected by the presence of Cr (or Fe, Mn) during hydrothermal growth and subsequent calcination. The preferential adsorption of Cr, Fe or Mn species on the specific facets of Co3O4 at the initial stage possibly results in favourable development along the directions of the six branches in the structure. The exact growth mechanism needs to be studied further.
To further investigate the structure and composition of the samples, elemental mapping was carried out using an EDS instrument attached to a transmission electron microscope. The results show that the micro/nano Co3O4 (Co3O4-3) consists of the elements cobalt and oxygen (Fig. 5). The amount of additional metal elements (Ni, Cr, Fe, Mn) is negligible, with an atomic ratio less than 0.2% for each element (Table S2†). The effect of the metal substrate on the EDS results cannot be completely avoided as the sample was mechanically scraped from the substrate for TEM characterization. As the micro/nano Co3O4 and untreated samples were grown on the same metal substrate, the effect of the substrate (i.e. the composition) on their performance could effectively be neglected. As a result, the difference in the electrochemical performance could be attributed to the structural effect alone, rather than being a compositional effect.
![]() | ||
Fig. 5 Elemental mapping of Co3O4-3 sample by an EDS instrument attached to the transmission electron microscope. |
The surface properties and elemental valences were further studied by XPS. The core level Co2p3/2 in the Co3O4-3 sample was deconvoluted to four fitted peaks (Fig. 6A). The peaks at binding energies (BEs) of 780.0 and 786.0 eV can be assigned to Co3+ and its satellite line and the other couple at a higher BE (781.6 and 789.5 eV) can be assigned to Co2+ and its satellite line.29 The BE value with a spin–orbit splitting of ca. 15.6 eV (795.6 eV in the Co2p1/2 region to 780.0 eV in the Co2p3/2 region) indicates that the Co3+ species are predominant on the surface.30 The O1s core level spectra show two fitted peaks OI and OII (Fig. 6B). The peak at 530.1 eV (OI) corresponds to the Co–O species and that at 531.5 eV (OII) originates from low-coordinated oxygen defects or vacancies.31,32 No Cr– (or Fe–, Mn–) related species were detected, which is indicative of the absence of such species on the surface of the windmill-like structures. The XPS spectra of Co3O4-NW grown on an Ni foil in a control measurement are also shown (Fig. 6C and D). The Co2p3/2 core level was fitted to four peaks. The BE values corresponding to Co3+, Co2+ and their satellite lines show little shift compared with those in the Co3O4-3 sample. These results indicate that heavy doping of the heteroatoms (Cr, Fe, Mn) to Co3O4-3 is not possible as the doping usually causes a BE shift in the core levels.
![]() | ||
Fig. 6 XPS core level spectra of Co3O4-3 sample. (A) Co2p, (B) O1s and Co3O4-NW, (C) Co2p and (D) O1s. |
The electrochemical performance of the Co3O4 film electrodes towards the oxidation of water was evaluated in a three-electrode cell configuration (0.1 M KOH, pH 13). LSV curves are shown in Fig. 7. The Co3O4 electrodes present a lower overpotential and higher current density than the substrate, indicating the catalytic role of Co3O4 itself. Compared with the untreated Co3O4, the surface treatment further lowers the overpotential for impregnation amounts of 1–3%. The Co3O4-3 has an onset overpotential of 0.29 V, lower than that of Co3O4 (0.33 V, Table 1). The overpotential of Co3O4-3 at a current density of 10 mA cm−2 is only 0.411 V, a remarkable decrease compared with Co3O4 (0.475 V, Table 1). This value is even lower than that of Co3O4-NW grown on an Ni foil (0.431 V). Nevertheless, the overpotential of Co3O4-5 increases to 0.45 V at the same current density (10 mA cm−2). The LSV curves of samples treated with 0.5 and 2% solute (cobalt salt) are shown in Fig. S4.† These results suggest that the sample treated with 3% solute is the optimum. A comparison of the mass activity of the Co3O4 samples shows that Co3O4-3 is the best catalyst with the highest mass activity (Table S3†).
Sample | ηonseta (V) | η10 mA cm−2 (V) | Rfb (103) | Active surfacec(cm2) |
---|---|---|---|---|
a Overpotentials (η) calculated using the formula η = ERHE − 1.23 V, where ERHE is the potential referred to the RHE. The onset potential was defined by a tangent method applied to the LSV curves.24b Roughness factor (Rf) calculated according to the formula Rf = Cdl/(60 × 10−6 F cm−2), where Cdl is the double-layer capacitor, i.e. the slope of the linear relationship between the peak current density and the scan rate; 60 × 10−6 F cm−2 is a reference value for the surface of oxides.24c Active surface (A) calculated based on the Randels–Sevcik equation: Ip = 2.687n3/2υ1/2D1/2AC, where n is the number of electrons transferred, υ1/2 is the square root of the scan rate, D is the diffusion coefficient (1.9 × 10−5 m2 s−1 in 0.1 M KOH), A is the active surface of the electrode and C is the concentration of the active substance.34 The linear relationship of Ip/υ1/2 is shown in Fig. S5.† | ||||
Co3O4 | 0.33 | 0.475 | 2.91 | 67 |
Co3O4-1 | 0.32 | 0.447 | 3.24 | 76 |
Co3O4-3 | 0.29 | 0.411 | 3.53 | 117 |
Co3O4-5 | 0.29 | 0.450 | 2.98 | 69 |
Co3O4-NW | 0.30 | 0.431 | 3.86 | 128 |
Substrate | 0.36 | 0.639 | — | — |
The difference in electrochemical performance may be associated with the structural effect or surface properties of the Co3O4. Consequently, the roughness factor (Rf) was calculated in terms of the double-layer capacitance in solution by assuming a standard value of 60 μF cm−2 for the oxide surface.24 The Rf value gradually increases with the amount of impregnation up to 3% (Table 1), although the Rf (103) of Co3O4-5 shows a decrease down to 2.98, close to the value for pristine Co3O4 (2.91). This finding indicates that Rf has an optimum value for Co3O4-3, consistent with the electrochemical performance towards the oxidation of water. This clearly shows that Rf plays a critical part in the OEC properties of the electrocatalysts. Enhancing the surface roughness is therefore an efficient way of improving the electrocatalytic performance of a specific material.33
The active surface (A) of each sample was calculated based on the Randels–Sevcik equation: Ip = 2.687n3/2υ1/2D1/2AC to compare the number of active sites.34 The calculated active surface presents a volcano-like curve, with the highest value of 117 cm2 corresponding to Co3O4-3, 1.75 times higher than that of untreated Co3O4. A higher active surface indicates that more active sites are accessible for catalytic reactions. The analyses of the roughness and active surface are in good agreement with the morphological observations by SEM, where Co3O4-3 showed a better developed micro/nano hierarchical structure. Co3O4-3 has a similar active surface to Co3O4-NW (117 vs. 128 cm2), suggesting that the treatment significantly increases the number of active sites on pristine Co3O4 microcrystals.
The Tafel slope could reflect information about the reaction mechanism. The Tafel plots of the Co3O4 electrodes are shown in Fig. S6,† where the Tafel slopes were derived by fitting the data in the linear part of the plots. The slopes are 130, 121, 115 and 130 mV dec−1 for Co3O4, Co3O4-1, Co3O4-3 and Co3O4-5, respectively. Tafel slopes ranging from about 40 to 120 mV dec−1 have been reported for thermally prepared cobaltites35–37 and this kinetic parameter is highly dependent on the preparation conditions. The slope of about 120 mV dec−1 obtained in this work for the Co3O4 samples suggests that it could follow a mechanism involving a rate-limiting one-electron transfer with no kinetically relevant electrochemical pre-equilibrium.38 The reactions involved are as follows:38–40
Co3O4 + H2O + OH− ⇄ 3CoOOH + e− | (1) |
CoOOH + OH− ⇄ CoO2 + H2O + e− | (2) |
Co(IV)s + OH− ⇄ Co(IV)(OH−)ad | (3) |
Co(IV)(OH−)ad + OH− → Co(IV)s + (1/2)O2 + H2O + e− | (4) |
Co(IV)s and Co(IV)(OH−)ad represent the Co(IV) species on the oxide surface and the adsorbed species, respectively. Oxygen evolution occurs at potentials more positive than the oxidation of Co(III) to Co(IV) (about 1.5 V vs. RHE; Fig. 7), suggesting that Co(IV) is the active catalyst for the OER. The Co3O4-3 sample shows a lower Tafel slope, indicative of its improved kinetics for the oxidation of water.
To further evaluate the catalytic activity of the micro/nano Co3O4, the TOFs were calculated on the basis of the assumption that all Co ions within the structure are catalytically active.41 It was found that the oxidation peak current of the Co redox species has a linear dependence on the scan rate (Fig. S7†). The line slope is proportional to the surface concentration of the Co species (Γ0) in terms of the equation: slope = n2F2AΓ0/4RT. As a result, the TOF values can be calculated from the equation TOF = JA/4Fm, where J is the current density at a constant overpotential, A is the area of the electrode, 4 represents the number of moles of electrons consumed to evolve one mole of oxygen, F is Faraday's constant and m is the number of moles of active sites. TOF values as a function of overpotential are shown in Fig. 8. The TOF values of each sample show a linear increase for overpotentials from 0.35 to 0.65 V, in which the TOFs of Co3O4-3 remain higher than the other samples beyond an overpotential of 0.4 V (Fig. 8A). For example, the TOF of Co3O4-3 reaches 0.39 s−1 at an overpotential of 0.6 V, about 1.4 times higher than pristine Co3O4 (0.28 s−1). However, the TOFs of Co3O4-5 are lower than those of pristine Co3O4 at overpotentials beyond 0.55 V. This tendency agrees with the results of the electrochemical LSV measurements. It is an indication that Co3O4-3 is the optimum micro/nano structure with the highest electrocatalytic activity towards the oxidation of water. The release of O2 bubbles from the surface of the Co3O4-3 electrode was visible with the naked eye at a current density of 10 mA cm−2 (video 1 in the ESI†). Measurements of the amount of O2 were made using an O2 fluorescent probe. The oxygen evolution was monitored in a gas-tight cell at 1.64 V (vs. RHE) for 1 h. Around 73.2 μmol of O2 were evolved after 1 h of electrolysis under these conditions (Fig. S8†). These results are evidence of the evolution of O2 from the electrodes.
![]() | ||
Fig. 8 (A) TOFs with respect to Co atoms in Co3O4 electrocatalysts as a function of the overpotential tested in 0.1 M KOH. (B) Comparison of TOFs of Co3O4-3 and Co3O4-NW tested in 1 M KOH. |
The TOFs of Co3O4-3 and Co3O4-NW tested in 1 M KOH are compared (Fig. 8B). The former presents higher TOFs than the latter throughout the overpotential range 0.3–0.65 V. At an overpotential of 0.6 V, the TOF of Co3O4-3 is 1.26 times higher than that of Co3O4-NW (0.54 vs. 0.427 s−1). This suggests that an optimum micro/nano hierarchical structure could be even more active than the nanowire.
It is known that the efficiency of catalysts for the oxidation of water depends on the pH of the solution. Co3O4 or cobalt oxide catalysts are commonly used in alkaline solution. Under such alkaline conditions, hydroxyl ions (OH−) are the dominant anions in solution and the O2-evolving reaction is as follows:
4OH−(aq.) → O2(g) + 2H2O(aq.) + 4e− |
Consequently, a higher pH (i.e. a larger OH− concentration) promotes the evolution of oxygen. It can be seen that the current density and TOF values for Co3O4-3 at the same overpotential (300 mV) remarkably increase when the pH is increased from 13 to 14 (Table 2).
The TOFs of Co3O4-3 were compared with previously reported Co3O4 electrocatalysts (Table 2). The optimized catalyst Co3O4-3 has a TOF value of 0.0452 s−1 at a 300 mV overpotential (pH 14), 2.4 times higher than the previously reported value for a Co3O4 nanocatalyst (0.0187 s−1 at 328 mV, pH 14).42 The TOF value shows an increase of 13% compared with that of 0.04 s−1 reported for a high-efficiency Co3O4 electrocatalyst under the same conditions.43 The TOF results verify that the structured catalysts are able to retain the high efficiency of their powder counterpart.
The electrochemical impedance spectra can be used to analyze the electron transport and recombination properties.44 The electrochemical impedance spectra of micro/nano Co3O4 were measured at an applied potential of 0.56 V (vs. SCE) in a 0.1 M KOH solution. Fig. 9 shows the Nyquist plots and the corresponding equivalent circuit. The semicircle geometry corresponds to the Faraday process and shows a Faraday impedance. The equivalent circuit indicates that the EIS profiles are composed of three components, i.e. a solution resistance (Rs), a charge transfer resistance across the electrode/solution interface (Rct) and a constant phase (CP). The fitted value of Rct is 5.5 Ω for the Co3O4-3 sample, which is the smallest resistance of all the Co3O4 electrodes. This indicates that Co3O4-3 has better charge transport properties during the oxygen-evolving reactions. The Rct values reasonably reflect the catalytic activity, i.e. the smaller charge transfer resistance results in a higher OER activity.
![]() | ||
Fig. 9 Electrochemical impedance spectra of the Co3O4 electrodes measured at a potential of 0.56 V (vs. SCE) in 0.1 M KOH solution. |
To study the stability of the micro/nano Co3O4 electrodes, the I–t curves were measured at a constant applied bias (1.58 V vs. RHE) in alkaline solution (0.1 M KOH). The optimum Co3O4-3 electrocatalyst reached an initial current density of 3.4 mA cm−2, which remained almost constant over a 12 h OER test with no noticeable decrease (Fig. 10A). In comparison, the untreated Co3O4 had a current density of 1.4 mA cm−2 at the beginning. After a 1 h OER test the value slightly decreased to about 1.3 mA cm−2 (a decrease of 7%), showing a comparable stability to Co3O4-3, although with a lower current density. It is understandable that the Co3O4 displays a high OER stability, as spinel-type Co3O4 has been reported to be a stable electrocatalyst towards the oxidation of water under alkaline conditions.23,43 SEM observations revealed that the micro/nano hierarchical structure of Co3O4-3 showed no change, even after a 12 h OER test (Fig. 10B). The SEM images of other Co3O4 samples after a 12 h OER test are shown in Fig. S9† and show little morphological change. XRD characterization also confirmed that the Co3O4-3 kept its spinel structure after the OER test, consistent with that before the reactions (Fig. S10†). In addition, the LSV curve for Co3O4-3 after a 12 h OER operation (0.1 M KOH, 1.58 V vs. RHE) overlaps that of unused Co3O4-3 (Fig. S11†), suggesting excellent recyclability for the oxygen-evolving reactions. Consequently, the micro/nano Co3O4 grown directly on the metal substrate could be used as highly efficient, durable and engineered electrocatalysts and are promising for use in the production of clean fuels from the electrolysis of water.
Footnote |
† Electronic supplementary information (ESI) available: XRD pattern, EDX spectra, elemental analysis, additional electrochemical measurements, and stability studies. See DOI: 10.1039/c4ra07082h |
This journal is © The Royal Society of Chemistry 2014 |