N-Doped carbon supported Co3O4 nanoparticles as an advanced electrocatalyst for the oxygen reduction reaction in Al–air batteries

Kun Liua, Zhi Zhoub, Haiyan Wang*a, Xiaobing Huangc, Jingyan Xua, Yougen Tang*a, Jingsha Lia, Hailiang Chud and Jiajie Chena
aCollege of Chemistry and Chemical Engineering, Central South University, Changsha, 410083, P. R. China. E-mail: wanghy419@126.com; ygtang@csu.edu.cn; Fax: +86 0731 8879616; Tel: +86 0731 8830886
bCollege of Science, Hunan Agricultural University, Changsha, 410128, P. R. China
cCollege of Chemistry and Chemical Engineering, Hunan University of Arts and Science, Changde, 415000, P. R China
dGuangxi Key Laboratory of Information Materials, Guilin University of Electronic Technology, Guilin, 541004, P. R. China

Received 22nd April 2016 , Accepted 1st June 2016

First published on 2nd June 2016


Abstract

Low-cost and high-performance catalysts are highly desirable for the oxygen reduction reaction (ORR) in metal–air batteries. Herein, a Co3O4/N-doped Ketjenblack (Co3O4/N-KB) composite is proposed as a high performance catalyst for Al–air batteries. The synergistic effect between Co3O4 and N-KB enables the Co3O4/N-KB composite to have a much higher cathodic current, a much more positive half-wave potential and more electron transfer in comparison with Co3O4 or N-KB alone. The Co3O4/N-KB composite favors a direct four electron pathway in the ORR process, and its ORR current density even outperforms the commercial Pt/C (20 wt%). Al–air batteries using the as-prepared catalyst in the air electrode were constructed, which displayed a high discharge voltage plateau of ∼1.52 V, comparable to that of the commercial Pt/C. The developed Co3O4/N-KB composite here could meet the requirements of large-scale application of metal–air batteries due to the much cheaper cobalt source and more economically commercialized carbon source compared to the commercial Pt/C.


Introduction

To meet the demands of modern society, energy conversion and storage systems with low-cost, high performance and environmental benignity are greatly needed.1–3 The performance of these devices depends on the properties of their electrode materials.4 The current bottleneck of fuel cells and metal–air batteries lies in the sluggish ORR on the cathode side.5,6 Despite tremendous effort, developing catalytic materials for the ORR with high activity remains a great challenge.5,7 Although the precious metal Pt has been considered as the most promising catalyst, the scarce resources and high cost limit its long-term practical applications.8,9 Accordingly, it is very important to develop non-precious metal catalysts with high electrocatalytic activities for ORR.10,11

Surface chemistry and structure of the support materials can greatly influence the activity of the resultant catalysts.12,13 Recently, catalyst supports such as Ketjenblack (KB), BP2000, and others carbon materials have been widely used in ORR because of their good electrical conductivity, surface properties, high surface area and stability.14 Note that the catalytic activities of these kinds of carbon supports are far insufficient for ORR. Nitrogen-doped carbon materials could affect the electron transportation in electrodes and increase active sites for ORR, leading to improved catalytic activity.15 These properties have been attributed to electronic or structural changes caused by the nitrogen incorporation into carbon framework.16 Although the compositions and structures of the active sites remain unclear, it has been well known that the doped nitrogen atoms (such as pyridine-like, pyrrole-like, graphite-like, and quaternary nitrogen atoms, etc.) play a decisive role for ORR.17 Nam et al.18 reported metal-free Ketjenblack incorporated nitrogen-doped carbon sheets for efficient catalysts of ORR in alkaline solution. They considered that the N dopant sites could provide positive catalytic effect for ORR. The nitrogen-doped gelatin with Ketjenblack carbon composites (GK) could adsorb more oxygen molecules owing to more exposed edge sites of GK sheets, where the oxygen adsorption was more favorable than that in basal plane. Lee et al.19 synthesized the carbonized melamine foam (Fe/Fe3C–melamine/N-KB) with higher positive onset potential than pure Ketjenblack. It is reasonable to infer that C–N group showed higher performance for ORR.

Recently, low cost transition metal oxides (such as manganese oxides, iron oxides, cobalt oxides, and nickel oxides, etc.) with good catalytic abilities towards ORR have been considered as effective components for composite electrocatalysts.20 Cobalt oxides have been investigated as non-precious ORR catalysts and applied in ORR in terms of their low cost, low electrical resistance and environmental friendship. The active sites of Co3+ ions in Co3O4 play a determinant role in the performance for ORR.21,22 Li et al.23 reported a series of M-doped polypyrrole-modified BP2000 via the hydrothermal method, the synergy among Co, N, and C effectively enhances the performance of ORR in alkaline medium, which depends on the nature of the transition element and the corresponding anion. Cobalt oxide has been well known as a promising electrocatalyst for metal–air batteries such as Li–air and Al–air batteries.9,24,25 Note that the natural abundance of aluminium in the earth's crust is much higher and is non-toxic and environmental friendly.26 In addition, aluminum has highly negative standard electrode potential [E0 = −1.7 V vs. standard hydrogen electrode (SHE)].27 Therefore, Al–air battery has been considered as a promising energy storage and conversion device because of its high practical energy density (600–800 W h kg−1).28–30 Phinergy and Alcoa developed an electric car using this kind of battery, which could run more than 1600 km by only adding water several times.31 The success of this technology has drawn intensive attention.32,33 In short, aluminum–air batteries turn out to be the more efficient energy systems for electric vehicles. One of the critical challenges is to create a stable and highly efficient air electrode under certain discharge current density.

In view of its low resistance, high mesoporous area and porosity, surface clean, KB has been employed as a promising catalyst support in alkaline fuel cell.25,34 In our recent work, we reported a MnOx–CeO2/KB hybrid, which exhibited a comparable activity and better stability towards ORR in comparison with the commercial Pt/C.35 Here, we reported a composite of Co3O4 nanoparticles grown on N-KB as a high-performance catalyst for ORR. Although KB alone has little ORR activity, N-KB exhibits surprisingly higher performance in ORR in alkaline solutions. Synergetic effect between Co3O4 and N-KB enables the as-prepared composite superior catalytic activity towards ORR. The high performance Co3O4/N-KB with low cost shows a good potential for large-scale commercial applications in metal air batteries.

Experimental section

Synthesis of Co3O4/N-KB

2.0 g of KB (Ketjenblack carbon, EC-300J) was treated in HNO3 at 80 °C for 8 h to remove other impurities and introduce functional groups on the carbon surface. After that, the mixture was centrifuged at 8000 rpm for 5 min to remove the high concentration of HNO3 at least repeated 10 times until the solution became neutral, and then dried in air at 110 °C. For nitrogen doped KB (N-KB), 1 g of acid-treated KB and 3.0 g of urea were added into agate mortar and firstly ground for 10–15 min then heated at 700 °C for 2 h under an argon atmosphere at a rate of 5 °C min−1. Co3O4/N-KB composite was prepared by a simple hydrothermal method. Briefly, a certain quantity of Co(NO3)2·6H2O was first dissolved into 80 mL distilled water, then 0.3 g of N-KB was added into the Co(NO3)2 solution. This suspension was stirred for at least 10 min and then 0.0668 g of NH4HCO3 was added. Finally, the mixed suspension was rapidly transferred into a 100 mL Teflon-lined autoclave and then kept at 140 °C for 12 h. The as-prepared Co3O4/N-KB was filtered and dried in air at 110 °C. After grinding in agate mortar, the samples were heated at 300 °C for 2 h in a muffle furnace with a slow heating rate of 5 °C min−1, finally Co3O4/N-KB was obtained. For comparison, Co3O4 was synthesized by the similar method as Co3O4/N-KB composite but without adding N-KB. Co3O4/KB was prepared through the similar procedure as Co3O4/N-KB using KB instead of N-KB. Control experiments with different contents of Co3O4 (5 wt%, 10 wt%, 15 wt%, 20 wt%) in Co3O4/N-KB composite were also performed to optimize the catalytic activity of the composite.

Material characterizations

X-ray diffraction (XRD) patterns of as-prepared samples were recorded by X-ray diffractometer (Rigaku D/Max 2500) utilizing a CuKα source with a step of 0.02°. X-ray photoelectron spectroscopys (XPS, K-Alpha1063) were applied to measure the chemical compositions of the samples. Scanning electron microscope (SEM) images of as-prepared Co3O4/N-KB were conducted using a Nova NanoSEM 230 SEM. Transmission electron microscope (TEM), high resolution TEM (HRTEM) images, scanning TEM (STEM) and elemental mapping of as-prepared Co3O4/N-KB were obtained using a FEI Tecnai G2 F20 S-TWIX TEM. The Raman spectra were performed using a Jasco Laser Raman Spectrophotometer NRS-3000 Series, with an excitation laser wavelength of 532 nm, at a power density of 2.9 mW cm−2. Differential scanning calorimetry and thermogravimetric analysis (DSC/TGA) was performed on a STA 449C with a heating rate of 10 °C min−1 from 30 to 800 °C in air.

Electrode preparation and electrochemical test

For the rotating disk electrode (RDE) measurements, 4 mg of as-prepared catalyst was dissolved in 1900 μL ethanol and sonicated 20 min, then 100 μL of Nafion (5 wt%) was added into the above solution and sonicated 20 min again to form a homogeneous ink. Then 10 μL of the catalyst ink was loaded onto on a glassy carbon rotating disk electrode of 5 mm in diameter. Cyclic voltammetry (CV) and linear sweep voltammetry (LSV) were conducted using saturated calomel electrode (SCE) as the reference electrode, a platinum wire as the counter electrode and the sample modified glassy carbon electrode as the working electrode. Electrolyte was 0.1 M KOH solution, oxygen (99.999%) was accessed for 20 min to make the electrolyte saturated with oxygen. All potentials initially in this study measured by the SCE were calibrated with respect to reversible hydrogen electrode (RHE). LSV were measured at a scan rate of 10 mV s−1 with varying rotating rates of 400, 625, 900, 1225, and 1600 rpm. CV were measured at a scan rate of 50 mV s−1. The obtained current data were normalized to the area of the catalyst used (i.e. given in mA cm−2) for direct comparison between samples. Koutecky–Levich plots (j−1 vs. ω−1/2) were analyzed at various electrode potentials. The slopes of their linear fit lines were used to calculate the electron transfer number (n) on the basis of the Koutecky–Levich equation. The number of electrons transferred per oxygen molecule was calculated by the Koutecky–Levich equation given below:36–38
image file: c6ra10486j-t1.tif

B = 0.62nFC0D02/3ν−1/6

For the Tafel plot, the kinetic current was calculated as follows:

image file: c6ra10486j-t2.tif
where j the measured current density of the experiment, jL is the diffusion-limiting current density, jK is the kinetic current density, ω is the electrode rotating speed in rpm, F is the Faraday constant (F = 96[thin space (1/6-em)]485 C mol−1), C0 the bulk concentration of O2 (1.2 × 10−3 mol L−1), D0 the diffusion coefficient of O2 in 0.1 M KOH (1.9 × 10−5 cm2 S−1), and ν is the kinematic viscosity of the electrolyte (0.01 cm2 S−1). B is the slope of K–L plots, n represents the number of electrons transferred per oxygen molecule.

Rotating ring-disk electrode (RRDE) technique was used to further estimate the electron transfer number (n). The peroxide percentage and the transferred electron number (n) were calculated based on the following equations:16

image file: c6ra10486j-t3.tif

image file: c6ra10486j-t4.tif
where Id is disk current, Ir is ring current, and N is current collection efficiency of the Pt ring (0.37).

Al–air battery fabrication and testing

Al–air full batteries in home-made model were built to investigate the practical electrochemical activities of the as-prepared catalysts. It consists of a polished aluminum electrode and an air electrode as the anode and cathode, respectively. 6 mol L−1 KOH adding 0.01 mol L−1 Na2SnO3, 0.0005 mol L−1 In(OH)3, 0.0075 mol L−1 ZnO as corrosion inhibitors was used as the electrolyte. The air electrodes with three layered structure, consisting of catalytic layer, current collector (nickel foam), and gas diffusion layer was fabricated using the hot pressure method. The catalytic layers were fabricated as follows: catalysts, Ketjenblack carbon and acetylene black, polytetrafluoroethylene (PTFE) were mixed well in a weight ratio of 3[thin space (1/6-em)]:[thin space (1/6-em)]3[thin space (1/6-em)]:[thin space (1/6-em)]1[thin space (1/6-em)]:[thin space (1/6-em)]3, then agitated for 30 min to form a homogeneous slurry. When the above slurry turned into paste, it was rolled to the layer until the thickness was ∼0.2 mm.39 The thickness of this air electrode was pressed to a range of 0.4–0.6 mm (3 cm × 5 cm) at 15 MPa. Finally, the air electrode was dried in vacuum at 60 °C for 12 h. The home-built electrochemical cells was used for Al–air battery measurements on a Neware battery testing system (Shenzhen, China). The details have been given in our previous work.40

Results and discussion

Fig. 1a shows typical XRD patterns of the as-prepared N-KB and Co3O4/N-KB composite. A broad peak (2θ = 24°) can be observed in the XRD pattern of N-KB, which is related to the characteristic peak of graphitic carbon. The tiny broad peak (2θ = 45°) is indexed to an amorphous structure;19,41 whereas, for what concerns Co3O4, the XRD pattern indicates a more ordered structure compared to N-KB and all other sharp crystalline peaks can be well indexed to the crystalline spinel structure of Co3O4 [JCPDS: 43-1003]. However, the intensities of the peaks related to Co3O4 species are weak owing to low Co3O4 content. According to the Scherrer's equation, the crystal size of Co3O4 particle on N-KB is determined to be 15.6 nm. The lattice parameters are calculated to be a = b = c = 8.08 Å. XRD analyses indicate that Co3O4 and N-KB co-exist in the as-prepared Co3O4/N-KB composite.
image file: c6ra10486j-f1.tif
Fig. 1 (a) XRD patterns of N-KB and Co3O4/N-KB composite, XPS spectra of (b) survey spectrum, (c) Co 2p, (d) O 1s, (e) N 1s and (f) C 1s for Co3O4/N-KB composite.

XPS measurement was performed to obtain more detailed elemental composition and the chemical element valence of Co3O4/N-KB composite. The full survey of Co3O4/N-KB composite in Fig. 1b indicates the existence of Co 2p, O 1s, N 1s and C 1s. By using a Gaussian fitting method, the Co 2p emission spectrum (Fig. 1c) was well fitted considering two spin–orbit doublets characteristic of Co2+ and Co3+. The fitting peaks at binding energies of 795.1 eV and 780.1 eV are ascribed to Co2+, while the peaks at 796.8 eV and 781.7 eV are due to Co3+.42–44 The high resolution spectrum for O 1s shows three oxygen species (Fig. 1d), marked as OI, OII and OIII. According to previous reports, the fitting peak of OI at 533.2 eV can be ascribed to a multiplicity of physically and chemically bonded water on and within the surface,45 that of OII at 531.5 eV corresponds to a high number of defect sites with a low oxygen coordination in the material with small particle size.46 And that of OIII at 530.2 eV is a typical metal–oxygen bond.45,47,48 The N 1s emission spectrum can be divided into three main peaks (Fig. 1e), 398.5, 400.3 and 401.3 eV, corresponding to pyridinic nitrogen, pyrrolic nitrogen and graphitic nitrogen.49–51 It should be noted that the edge plane nitrogen groups (pyrrolic or pyridinic) in Co3O4/N-KB composite may provide additional active sites for catalytic process due to the lone pair electrons.52 The high resolution C 1s (Fig. 1f) peak is centered at 284.8 eV, related to the sp2 graphitic carbon.53 Apparently, the electron couples of Co3+/Co2+ in spinel Co3O4/N-KB, which may provide good electrocatalytic activity towards ORR.

The morphology and microstructure of the as-synthesized Co3O4/N-KB were characterized by SEM, TEM and HRTEM. Fig. 2a shows the typical SEM images of Co3O4/N-KB composite. As observed, Co3O4 nanoparticles are well dispersed on the surface of N-KB. In order to deeply investigate the morphology and microstructure of the Co3O4/N-KB, low and high-magnification TEM images of the composites are presented in Fig. 2b–d. As seen from Fig. 2b, homogeneous distribution of nanosized Co3O4 with intimate contact on the N-KB is further confirmed. Note that there are some deep dark areas, which should be resulted from Co3O4 supported on N-KB, consistent well with the reference.5 HRTEM images of the Co3O4/N-KB composite show clearer lattice fringes of Co3O4 in Fig. 2c and d. The lattice spacing of 0.244 nm, 0.202 nm, and 0.143 nm correspond to Co3O4 (311), (400), and (440) planes, respectively (Fig. 2c). The lattice spacing of 0.286 nm in Fig. 2d corresponds to (220) planes of Co3O4 by lattice analysis. The Co3O4 particles promise their intimate contact on the surface of N-KB, leading to better synergistic effect. To determine the distribution of Co3O4 in N-KB composite, elemental identification of C, O, N and Co elements is performed by STEM-EDS. As seen from the EDS mapping, uniform elemental distribution of C, O, N and Co elements was achieved in the Co3O4/N-KB composite (Fig. 3).


image file: c6ra10486j-f2.tif
Fig. 2 (a) SEM images and (b) low magnification TEM images of Co3O4/N-KB composite, (c and d) show the magnified HRTEM images for clearer lattice fringes in corresponding square region marked in (b).

image file: c6ra10486j-f3.tif
Fig. 3 (a) STEM elemental mapping images of the Co3O4/N-KB composite taken from the left selected region marked in (a), (b–e) the corresponding elemental mapping images of (b) C–K, (c) O–K, (d) N–K, and (e) Co–K.

To further identify the structural changes of the carbon framework of KB, N-KB, and Co3O4/N-KB composite, Raman spectroscopy results are given in Fig. 4a. Clearly, the D and G bands of KB show a shift to 1325.8 and 1592.7 cm−1, respectively. N-KB shows the similar D and G bands at 1321.6 and 1591.4 cm−1, respectively. A general feature of the N-KB is that the D band and the G band shift to lower wavenumber than those of the KB.54 Co3O4/N-KB composite has two prominent peaks at 1329.2 and 1591.7 cm−1 which correspond to the well-documented D band and G band, respectively. After hydrothermal treatment, the D and G bands shift to higher wavenumber in the Co3O4/N-KB composite. From the Raman spectra, we know that the intensity ratio of D band to G band (i.e. ID/IG) for Co3O4/N-KB composite is 1.389, which is relatively higher in comparison with that of N-KB (ID/IG = 1.373) and KB (ID/IG = 1.347). The increasing of ID/IG ratio for N-KB is attributed to the structural defects and edge plane exposure caused by nitrogen atom incorporation into the carbon layers present on the surface of N-doped carbon materials.54–56 The ID/IG ratio is larger for Co3O4/N-KB composite than N-KB, indicates a decrease in the average size of the sp2 domains formed during a hydrothermal process.57 This can be allowed that the structure of the carbon layers has rebuilded in Co3O4/N-KB composite, which implies that the size of the in-plane-sp2 domains and the defects increased during Co3O4/N-KB composite.58,59 The increase of ID/IG in Co3O4/N-KB composite can also further increase defect sites of such systems.60 As a matter of fact, the sites exhibit high electrocatalytic activities for ORR.


image file: c6ra10486j-f4.tif
Fig. 4 (a) Raman spectra of undoped KB, N-KB, and Co3O4/N-KB composite, (b) DSC/TG curve of the as-prepared Co3O4/N-KB composite.

The mass ratio of N-KB in the composite was investigated by DSC/TG curve (Fig. 4b). A large weight loss of 88.6% is observed in the temperature range between 261 and 631 °C in TG curve, corresponding to the burning of N-doped Ketjenblack.61 There is a big exothermic peak at about 420 °C in DSC curve which should be related to the reaction between N-KB and O2.62 When the temperature is over 631 °C, there is no weight loss, indicating that the content of N-KB in the composite is about 88.6%.

To obtain further insight into the ORR performance of each catalyst, the as-prepared catalysts were loaded onto glassy carbon electrodes for the linear sweeping voltammograms (LSVs) at 1600 rpm in O2-saturated 0.1 M KOH in Fig. 5a. All samples were scanned cathodically at a rate of 10 mV s−1. It clearly indicates that the pure Co3O4 and KB exhibit very poor onset potential. Compared to pure Co3O4, Co3O4/KB shows higher positive onset potential and much higher limiting current density. The results suggest that cobalt oxides supported on carbon materials play a significant role in increasing the performance of ORR. N-KB also shows better activity towards ORR than pure KB. It has been well proved that N doping in KB could facilitate chemisorption of oxygen, leading to a relatively high catalytic activity towards ORR.63 Two plateaus (from 0.8 to 0.5 V and 0.5 to 0 V) are observed for Co3O4, KB and N-KB, indicating that Co3O4, KB and N-KB underwent a two-electron pathway (see Fig. 6a–c) and proceeded via a hydroperoxide anion (HO2) intermediate.64 The assessed H2O2 yield is found to be significantly high for Co3O4 (nearly 100%), KB (nearly 50%) and N-KB (nearly 30%) in the potential region of 0.3 to 0.85 V (vs. RHE). The derived electron transfer numbers from the RRDE estimation are found to be 1.9, 2.4 and 2.7 for Co3O4, KB and N-KB, respectively (see Fig. 5d). Interestingly, Co3O4/N-KB composite possesses a more positive half-wave potential (E1/2, 0.79 V), hence resulting in better ORR activity, which is higher than those of Co3O4 (0.28 V), KB (0.56 V), N-KB (0.64 V), Co3O4/KB (0.72 V) and only 40 mV negative shift compared with the commercial Pt/C catalyst. Moreover, the limiting current density of Co3O4/N-KB is up to about −5.6 mA cm−2, which is much higher than the other samples, and even outperforms the commercial Pt/C (−5.2 mA cm−2). In addition, the catalytic activity of Co3O4/N-KB composite depends crucially on the amount of Co3O4. Too many Co3O4 may cause the free N-KB, which will weaken the ORR performance. A control experiment was fulfilled to optimize the catalytic activity of the Co3O4/N-KB composite by methodically changing the content of Co3O4. The results in Fig. 5b demonstrate that the Co3O4/N-KB composite with 5 wt% Co3O4 displays the best ORR activity. It should be noted that N-KB and Co3O4 exhibit inferior ORR catalytic activity, while Co3O4/N-KB presents significantly improved electroactivity to ORR. After the incorporation of N-KB, the ORR kinetics of Co3O4 could be significantly enhanced. The strongly synergistic effect of Co3O4 and N-KB in the composite was the key towards improving electrochemical performance. It should be noted that the integration of Co3O4 and N-KB was feasible to improve the catalytic activity of cobalt oxides based cathode in fuel cells and metal–air batteries.65 The synergistic coupling between Co3O4 and N-doped carbon materials establish more active sites to obtain high ORR activity owing to interaction between carbon structure and other molecules.66 Meanwhile, it has revealed that the synergistic performance enhancement results from an improved carbon-catalyst binding and increased electrical conductivity, which leads to a large number of the active sites of the Co3O4/N-KB composite.23,67 We may conclude that the Co3O4 grown on N-KB prevented Co3O4 from restacking to lead to uniform nucleation during hydrothermal process, which made Co3O4 better dispersion on N-KB.68 Such a synergistic coupling led to not only enhanced current density but also plenty of active sites for ORR.


image file: c6ra10486j-f5.tif
Fig. 5 (a) ORR polarization curves for different catalysts at 1600 rpm, (b) the linear polarization curves of air electrodes with different contents of Co3O4 in Co3O4/N-KB composite, (c) CVs of Co3O4, KB, N-KB, Co3O4/KB, Co3O4/N-KB and Pt/C in Ar and O2 saturated 0.1 M KOH solution, (d) the electron transfer number (dotted line) and percentage of peroxide (solid line) of different catalysts at different potentials.

image file: c6ra10486j-f6.tif
Fig. 6 ORR polarization curves of (a) Co3O4, (b) KB, (c) N-KB, (d) Co3O4/KB, (e) Co3O4/N-KB and (f) Pt/C at different rotating speeds (400–1600 rpm) in O2-saturated 0.1 M KOH solution at a scan of 10 mV s−1. Insets: K–L plots at different potentials.

CV measurements were carried out in an O2-saturated or Ar-saturated 0.1 M KOH solution with a scan rate of 10 mV s−1. The tested voltage ranges from 1.2 to 0 V (vs. RHE). As shown in Fig. 5c, CV curves of all samples recorded in Ar-saturated electrolyte show no obvious peaks, however, when the electrolyte was saturated with O2, an obvious reduction peak clearly appears, confirming the electrocatalytic activity for ORR. The CV curve for Co3O4/N-KB electrode in O2 shows a reduction peak potential of ∼0.78 V (vs. RHE), which is much more positive than those for Co3O4 (0.49 V), KB (∼0.74 V), N-KB (∼0.75 V), Co3O4/KB (∼0.76 V), and only negative shift of about 20 mV compared to the commercial Pt/C. The positive shifting of the reduction peak of Co3O4/N-KB in comparison with Co3O4 and N-KB indicates the existence of synergistic effect between Co3O4 and N-KB. This was in agreement with the LSVs and CVs results. It was reported that Co3O4 was an efficient synergistic component for ORR catalysts owing to possessing the function of the intermediate disproportionation reaction in oxygen reduction.65 Liu et al.69 brought insights into the synergistic coupling effect between Co3O4 and MWCNT, and demonstrated that the enhanced stability of Co3O4/MWCNT composite material could be affected by Co3O4 in the composite assisting the prevention of carbon corrosion. Furthermore, the good synergistic interaction between N-doped carbon and Co NPs in the Co/N–C-800 composite has been demonstrated by Su et al.,70 it can be inferred that the smaller charge transfer resistance, higher specific surface area, and better synergistic interaction could be the causes for the better catalytic activity of Co/N–C-800 than other Co/N–C samples. Olson et al.5,71 have put forward a dual-site mechanism for cobalt–polypyrrole/C ORR catalyst, where oxygen is firstly reduced to HO2 at Co–N–C sites and further reduced to OH at CoxOy/Co sites in alkaline media. Herein, our current research system may be similar to the above mechanism, suggesting synergistic coupling between Co3O4 and N-KB is critical to ORR activity of the composite.

To obtain further information about ORR kinetics, the Koutecky–Levich plots (j−1 vs. ω−1/2) of each catalyst are obtained from LSVs at various potentials. All plots show good linearity (R2 > 0.99) at various rotation speeds (Fig. 6e). The n value for Co3O4/N-KB was calculated to be ∼4 at a large potential range, which suggests a four-electron pathway for ORR. The electron transfer numbers (n) of Co3O4/KB and the commercial Pt/C were also calculated to be ∼4 (Fig. 6d and f), indicating the existence of a four-electron pathway toward the formation of OH ions.8 To further prove H2O2 yield of Co3O4/KB, Co3O4/N-KB and Pt/C, we obtained reliable data from the RRDE. The amount is found to be in a low level of 10% in the potential region of 0.3 to 0.85 V (vs. RHE). The derived electron transfer numbers (Fig. 5d) are close to 4, indicating the direct 4e transfer reaction in ORR. It showed that cobalt oxides supported on carbon materials led to a clear enhancement of the electron-transfer kinetics of ORR, similar to the behavior of the commercial Pt/C. Tafel plots of Co3O4/KB and Co3O4/N-KB derived from LSVs data are shown in Fig. 7a. The Tafel slope of Co3O4/N-KB is 74.7 mV per decade, which is superior to the 83.1 mV per decade of the Co3O4/KB, indicating a good ORR kinetic process for Co3O4/N-KB. In addition, the stability of Co3O4/N-KB for the ORR was examined via the chronoamperometric method in O2-saturated at 1600 rpm. The potential was selected as 0.7 V. As seen in Fig. 7b, the Co3O4/N-KB catalyst is more stable than Co3O4/KB catalyst and the commercial Pt/C catalyst at the same testing condition. Apparently, the Co3O4/N-KB here exhibits high onset potential, high positive half-wave potential, high electron transfer number (∼4) and can be considered as a suitable catalyst for high power metal–air batteries.


image file: c6ra10486j-f7.tif
Fig. 7 (a) Tafel plots of kinetic current for Co3O4/KB and Co3O4/N-KB, (b) current–time (it) chronoamperometric response of Co3O4/KB, Co3O4/N-KB and Pt/C at 0.7 V in O2-saturated 0.1 M KOH.

Al–air batteries with Co3O4/KB, Co3O4/N-KB and commercial Pt/C catalyst as the cathodes were fabricated for further comparison of their practical activity in 6 M KOH solution. We recorded the galvanostatic discharge curves (Fig. 8) at a constant current discharge of 20 mA cm−2 for 15 h. As can be seen, all cells show the increased discharge voltage in the beginning due to the activation process of Al anode, on which the passive film was dissolved gradually. Before ∼2.5 h, the working voltage plateau of the cells with Co3O4/KB as an electrocatalyst is close to that of Co3O4/N-KB, which is higher than that of Pt/C. After ∼2.5 h, the cell with the as-prepared Co3O4/KB catalyst always shows a lower discharge voltage than those of Co3O4/N-KB and commercial Pt/C catalyst. However, in the initial 5 h, it is interesting to note that the cell with the as-prepared Co3O4/N-KB catalyst even exhibits a little higher discharge voltage than that with Pt/C catalyst. As follows, their average working voltages (1.52 V) are almost the same. Generally, it could be concluded that the practical electrocatalytic activity of the Co3O4/N-KB composite here in Al–air battery is comparable to that of the commercial Pt/C, enabling it a very promising catalyst with high efficiency and low cost for Al–air batteries.


image file: c6ra10486j-f8.tif
Fig. 8 Galvanostatic discharge curves of the Al–air batteries with Co3O4/KB, Co3O4/N-KB and Pt/C as the cathode catalysts at a discharge current density of 20 mA cm−2.

Conclusions

In summary, Co3O4/N-KB composite was successfully prepared by a facile strategy and used as a high performance catalyst for ORR. In comparison with the alone Co3O4 and N-KB, the as-prepared Co3O4/N-KB showed much higher current density and more positive half-wave potential. Note that Co3O4/N-KB composite underwent a four-electron reaction mechanism, whose catalytic activity approached to the commercial Pt/C. The improved performance of Co3O4/N-KB should be attributed to the increased active sites resulted from synergetic chemical coupling effect between Co3O4 and N-KB. Al–air full battery using Co3O4/N-KB as catalyst was also evaluated, which exhibited comparable discharge voltage plateau to the commercial Pt/C counterpart at the same current density. Our work provides a simple but efficient approach to produce low-cost Co3O4/N-KB composite as efficient ORR electrocatalyst for Al–air battery.

Acknowledgements

This work was financially supported by the National Nature Science Foundation of China (No. 21571189, No. 21271187, No. 51304077 and No. 21301193), the Opening Project of Guangxi Key Laboratory of Information Materials (No. 131006-K), and the Open-End Fund for Valuable and Precision Instruments of Central South University (CSUZC201622).

Notes and references

  1. Y. G. Guo, J. S. Hu and L. J. Wan, Adv. Mater., 2008, 20, 2878–2887 CrossRef CAS .
  2. G. H. Yu, L. B. Hu, M. Vosgueritchian, H. L. Wang, X. Xie, J. R. McDonough, X. Cui, Y. Cui and Z. N. Bao, Nano Lett., 2011, 11, 2905–2911 CrossRef CAS PubMed .
  3. J. Yan, Q. Wang, T. Wei and Z. J. Fan, Adv. Energy Mater., 2014, 4, 1400500–14005005 Search PubMed .
  4. A. S. Aricò, P. Bruce, B. Scrosati, J. M. Tarascon and W. V. Schalkwijk, Nat. Mater., 2005, 4, 366–377 CrossRef PubMed .
  5. Y. Y. Liang, Y. G. Li, H. L. Wang, J. G. Zhou, J. Wang, T. Regier and H. J. Dai, Nat. Mater., 2011, 10, 780–786 CrossRef CAS PubMed .
  6. J. W. Xiao, C. Chen, J. B. Xi, Y. Y. Xu, F. Xiao, S. Wang and S. H. Yang, Nanoscale, 2015, 7, 7056–7064 CAS .
  7. Y. Y. Liang, H. L. Wang, P. Diao, W. Chang, G. S. Hong, Y. G. Li, M. Gong, L. G. Xie, J. G. Zhou, J. Wang, T. Z. Regier, F. Wei and H. J. Dai, J. Am. Chem. Soc., 2012, 134, 15849–15857 CrossRef CAS PubMed .
  8. Z. S. Wu, S. B. Yang, Y. Sun, K. Parvez, X. L. Feng and K. Müllen, J. Am. Chem. Soc., 2012, 134, 9082–9085 CrossRef CAS PubMed .
  9. R. G. Cao, J. S. Lee, M. L. Liu and J. P. Cho, Adv. Energy Mater., 2012, 2, 816–829 CrossRef CAS .
  10. Z. W. Chen, D. Higgins, A. P. Yu, L. Zhang and J. J. Zhang, Energy Environ. Sci., 2011, 4, 3167–3192 CAS .
  11. Y. L. Yu, Y. P. Hu, X. W. Liu, W. Q. Deng and X. Wang, Electrochim. Acta, 2009, 54, 3092–3097 CrossRef CAS .
  12. Y. Y. Shao, J. H. Sui, G. P. Yin and Y. Z. Gao, Appl. Catal., B, 2008, 79, 89–99 CrossRef CAS .
  13. Y. Y. Shao, J. Liu, Y. Wang and Y. H. Lin, J. Mater. Chem., 2009, 19, 46–59 RSC .
  14. P. Kichambare, S. Rodrigues and J. Kumar, ACS Appl. Mater. Interfaces, 2012, 4, 49–52 CAS .
  15. S. B. Yang, X. L. Feng, X. C. Wang and K. Mullen, Angew. Chem., Int. Ed., 2011, 50, 5339–5343 CrossRef CAS PubMed .
  16. R. A. Sidik, A. B. Anderson, N. P. Subramanian, S. P. Kumaraguru and B. N. Popov, J. Phys. Chem. B, 2006, 110, 1787–1793 CrossRef CAS PubMed .
  17. D. Geng, Y. Chen, Y. Chen, Y. Li, R. Li, X. Sun, S. Ye and S. Knights, Energy Environ. Sci., 2011, 4, 760–764 CAS .
  18. G. Nam, J. Park, S. T. Kim, D. B. Shin, N. Park, Y. Kim, J. S. Lee and J. Cho, Nano Lett., 2014, 14, 1870–1876 CrossRef CAS PubMed .
  19. J. S. Lee, G. S. Park, S. T. Kim, M. Liu and J. Cho, Angew. Chem., 2013, 125, 1060–1064 CrossRef .
  20. H. Zhang and W. Yang, Chem. Commun., 2007, 7, 4215–4217 RSC .
  21. Y. Wang, X. Lu, Y. Liu and Y. Deng, Electrochem. Commun., 2013, 31, 108–111 CrossRef CAS .
  22. K. Song, E. B. Cho and Y. M. Kang, ACS Catal., 2015, 5, 5116–5122 CrossRef CAS .
  23. Z. P. Li, Z. X. Liu, K. N. Zhu, Z. Li and B. H. Liu, J. Power Sources, 2012, 219, 163–171 CrossRef CAS .
  24. W. H. Ryu, T. H. Yoon, S. H. Song, S. k. Jeon, Y. J. Park and I. D. Kim, Nano Lett., 2013, 13, 4190–4197 CrossRef CAS PubMed .
  25. G. Wang, G. Sun, Q. Wang, S. Wang, J. Guo, Y. Gao and Q. Xin, J. Power Sources, 2008, 180, 176–180 CrossRef CAS .
  26. D. Gelman, B. Shvartsev and Y. Ein-Eli, J. Mater. Chem. A, 2014, 2, 20237–20242 CAS .
  27. R. Mori, RSC Adv., 2014, 4, 30346–30351 RSC .
  28. Q. Li and N. J. Bjerrum, J. Power Sources, 2002, 110, 1–10 CrossRef CAS .
  29. L. Fan, H. Lu and J. Leng, Electrochim. Acta, 2015, 165, 22–28 CrossRef CAS .
  30. M. Wang, Y. Lai, J. Fang, J. Li, F. Qin, K. Zhang and H. Lu, Int. J. Hydrogen Energy, 2015, 40, 16230–16237 CrossRef CAS .
  31. L. Fan and H. Lu, J. Power Sources, 2015, 284, 409–415 CrossRef CAS .
  32. R. Revel, T. Audichon and S. Gonzalez, J. Power Sources, 2014, 272, 415–421 CrossRef CAS .
  33. D. Gelman, B. Shvartsev and Y. Ein-Eli, J. Mater. Chem. A, 2014, 2, 20237–20242 CAS .
  34. J. S. Lee, G. S. Park, H. L. Lee, S. T. Kim, R. Cao, M. Liu and J. Cho, Nano Lett., 2011, 11, 5362–5366 CrossRef CAS PubMed .
  35. J. J. Chen, N. Zhou, H. Y. Wang, Z. G. Peng, H. Y. Li, Y. G. Tang and K. Liu, Chem. Commun., 2015, 51, 10123–10126 RSC .
  36. Z. Lin, G. H. Waller, Y. Liu, M. Liu and C. P. Wong, Nano Energy, 2013, 2, 241–248 CrossRef CAS .
  37. Y. Ma, R. Wang, H. Wang, J. Key and S. Ji, J. Power Sources, 2015, 280, 526–532 CrossRef CAS .
  38. R. Liu, D. Wu, X. Feng and K. Mullen, Angew. Chem., Int. Ed., 2010, 49, 2565–2569 CrossRef CAS PubMed .
  39. Y. Tang, H. Qiao, H. Wang and P. Tao, J. Mater. Chem. A, 2013, 1, 12512–12518 CAS .
  40. H. Zhang, H. Qiao, H. Wang, N. Zhou, J. Chen, Y. Tang and C. Huang, Nanoscale, 2014, 6, 10235–10242 RSC .
  41. X. Yang, Y. Xu, H. Zhang, Y. A. Huang, Q. Jiang and C. Zhao, Electrochim. Acta, 2013, 114, 259–264 CrossRef CAS .
  42. J. Zhu and Q. Gao, Microporous Mesoporous Mater., 2009, 124, 144–152 CrossRef CAS .
  43. J. Li, S. Xiong, Y. Liu, Z. Ju and Y. Qian, Nano Energy, 2013, 2, 1249–1260 CrossRef CAS .
  44. L. E. Gómez, A. V. Boix and E. E. Miró, Catal. Today, 2013, 216, 246–253 CrossRef .
  45. X. F. Lu, D. J. Wu, R. Z. Li, Q. Li, S. H. Ye, Y. X. Tong and G. R. Li, J. Mater. Chem. A, 2014, 2, 4706–4713 CAS .
  46. R. Ding, L. Qi, M. Jia and H. Wang, J. Appl. Electrochem., 2013, 43, 903–910 CrossRef CAS .
  47. J. F. Marco, J. R. Gancedo, M. Gracia, J. L. Gautier, E. Ríos and F. J. Berry, J. Solid State Chem., 2000, 153, 74–81 CrossRef CAS .
  48. T. Choudhury, S. O. Saied, J. L. Sullivan and A. M. Abbot, J. Phys. D: Appl. Phys., 1989, 22, 1185–1195 CrossRef CAS .
  49. H. Wang, T. Maiyalagan and X. Wang, ACS Catal., 2012, 2, 781–794 CrossRef CAS .
  50. N. P. Subramanian, X. Li, V. Nallathambi, S. P. Kumaraguru, H. C. Mercado, G. Wu, J. W. Lee and B. N. Popov, J. Power Sources, 2009, 188, 38–44 CrossRef CAS .
  51. Y. Mao, H. Duan, B. Xu, L. Zhang, Y. Hu, C. Zhao, Z. Wang, L. Chen and Y. Yang, Energy Environ. Sci., 2012, 5, 7950–7955 CAS .
  52. Z. Chen, D. Higgins and Z. Chen, Carbon, 2010, 48, 3057–3065 CrossRef CAS .
  53. Z. Zhang, L. Su, M. Yang, M. Hu, J. Bao, J. Wei and Z. Zhou, Chem. Commun., 2014, 50, 776–778 RSC .
  54. A. Zahoor, M. Christy, Y. J. Hwang, Y. R. Lim, P. Kim and K. S. Nahm, Appl. Catal., B, 2014, 147, 633–641 CrossRef CAS .
  55. L. S. Panchakarla, A. Govindaraj and C. N. R. Rao, ACS Nano, 2007, 1, 494–500 CrossRef CAS PubMed .
  56. Z. H. Sheng, L. Shao, J. J. Chen, W. J. Bao, F. B. Wang and X. H. Xia, ACS Nano, 2011, 5, 4350–4358 CrossRef CAS PubMed .
  57. M. S. Arif Sher Shah, A. R. Park, K. Zhang, J. H. Park and P. J. Yoo, ACS Appl. Mater. Interfaces, 2012, 4, 3893–3901 Search PubMed .
  58. H. W. Wang, Z. A. Hu, Y. Q. Chang, Y. L. Chen, H. Y. Wu, Z. Y. Zhang and Y. Y. Yang, J. Mater. Chem., 2011, 21, 10504–10511 RSC .
  59. D. W. Chang, E. K. Lee, E. Y. Park, H. Yu, H. J. Choi, L. Y. Jeon, G. J. Sohn, D. Shin, N. Park, J. H. Oh, L. Dai and J. B. Baek, J. Am. Chem. Soc., 2013, 135, 8981–8988 CrossRef CAS PubMed .
  60. D. Sen, R. Thapa and K. K. Chattopadhyay, ChemPhysChem, 2014, 15, 2542–2549 CrossRef CAS PubMed .
  61. Y. Su, A. Pan, Y. Wang, J. Huang, Z. Nie, X. An and S. Liang, J. Power Sources, 2015, 295, 254–258 CrossRef CAS .
  62. J. W. Min, A. K. Kalathil, C. J. Yim and W. B. Im, Mater. Charact., 2014, 92, 118–126 CrossRef CAS .
  63. S. Wang, E. Iyyamperumal, A. Roy, Y. Xue, D. Yu and L. Dai, Angew. Chem., Int. Ed., 2011, 50, 11756–11760 CrossRef CAS PubMed .
  64. I. Hwang, E. Ahn and Y. Tak, Int. J. Electrochem. Sci., 2014, 9, 5454–5466 Search PubMed .
  65. J. Ahmed, Y. Yuan, L. Zhou and S. Kim, J. Power Sources, 2012, 208, 170–175 CrossRef CAS .
  66. M. Prabu, P. Ramakrishnan and S. Shanmugam, Electrochem. Commun., 2014, 41, 59–63 CrossRef CAS .
  67. G. Vijayaraghavan and K. J. Stevenson, Langmuir, 2007, 23, 5279–5282 CrossRef CAS PubMed .
  68. S. Zhang, H. Zhang, Q. Liu and S. Chen, J. Mater. Chem. A, 2013, 1, 3302–3308 CAS .
  69. Y. Liu, D. Higgins, J. Wu, M. Fowler and Z. Chen, Electrochem. Commun., 2013, 34, 125–129 CrossRef CAS .
  70. Y. Su, Y. Zhu, H. Jiang, J. Shen, X. Yang, W. Zou, J. Chen and C. Li, Nanoscale, 2014, 6, 15080–15089 RSC .
  71. T. S. Olson, S. Pylypenko, P. Atanassov, K. Asazawa, K. Yamada and H. Tanaka, J. Phys. Chem. C, 2010, 114, 5049–5059 CAS .

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.