Introduction

Graphene-based nanocomposite is one of the most currently investigated materials in the fields of chemistry, physics, materials science and nanotechnology because of its intriguing physicochemical properties and promising functionalities1,2,3,4. This family of materials boasts excellent functionalities for many energy-related applications such as secondary batteries, supercapacitors, photocatalysts, photovoltaics and fuel cells4,5,6,7,8,9,10,11,12. One of the most promising applications of the graphene-based nanocomposites is an electrode for secondary batteries. An increasing demand for the large-scale application of secondary batteries evokes intense research efforts for the exploration of novel graphene-based electrode materials showing excellent rate characteristics and high electrochemical stability11. The hybridization of electrode materials with highly conductive graphene nanosheets leads to a significant improvement of electrode performance at high current density via the increase of electrical conductivity12. Additionally the porous stacking structure of the graphene-based nanocomposite can relieve the drastic volume change and electrical disconnection of electrode materials upon electrochemical cycling, leading to the improvement of cyclability12. Thus, there is a great deal of research activity for the nanocomposite electrode materials composed of reduced graphene oxide (rG-O) nanosheets and electrochemically active metal oxides like Co3O4, Mn3O4 and SnO2 and elemental metals/semimetals like Si and Ge13,14,15,16,17,18. However, the rG-O nanosheet suffers from a strong tendency to form tightly packed structure due to a strong π–π interaction between sp2 carbon arrays. This prevents the intimate nanoscale mixing between electrode crystals and rG-O nanosheets and the formation of the open stacking structure of nanocomposite, which diminish the beneficial effect of the hybridization with rG-O nanosheets.

Currently intense research interest on graphene nanosheet is extended to 2D nanosheets of inorganic solids such as layered metal oxide, layered metal chalcogenide and layered metal hydroxide19,20,21. Like the rG-O nanosheets, the subnanometer-thick nanosheets of layered inorganic compounds can be synthesized by the chemical exfoliation of their pristine materials22,23. The obtained metal oxide nanosheets can form homogeneous colloidal mixture with rG-O nanosheets24. Taking into account the stiffness and the absence of π electron clouds of these inorganic nanosheets, the incorporation of metal oxide nanosheets is supposed to be effective in enhancing the porosity and homogeneity of metal oxide–rG-O nanocomposite via the weakening of π–π interactions between rG-O nanosheets. Among diverse metal oxide nanosheets, redoxable transition metal oxide nanosheets such as CoO2, [Mn1/3Co1/3Ni1/3]O2 and MnO2 show sufficiently high electrical conductivity and high electrochemical activity25,26,27. Such redoxable metal oxide nanosheets can be suitable additives for optimizing the composite structure, pore structure and performance of graphene-based nanocomposite electrode materials. Yet at the time of publication of this study, we are aware of no report about the use of the mixed colloidal suspension of layered metal oxide and graphene nanosheets as a precursor for the optimization of the electrode performance of graphene-based nanocomposites for secondary batteries.

Here we report an effective and universal way to improve the electrode functionality of graphene-based nanocomposites using the colloidal mixture of inorganic and graphene nanosheets. The effects of the intervention of layered CoO2 nanosheets on the composite structure, pore structure and the electrode activity of Co3O4–graphene nanocomposite are systematically investigated. The present strategy is also extended by the incorporation of layered MnO2 nanosheets into the Mn3O4–graphene nanocomposite.

Results

The precursors of exfoliated CoO2 and G-O nanosheets can form stable mixture colloidal suspensions with variable ratios of CoO2/G-O, since they possess very similar surface charge and hydrophilicity each other (Supplementary Information, Fig. S1 and Table S1). The effect of NH4OH addition on the colloidal stability of G-O/CoO2 mixture as well as on the pure colloidal suspensions of layered CoO2 and G-O nanosheets is examined. Upon the addition of NH4OH, all the present colloidal suspensions remain unchanged without the formation of aggregated precipitates, clearly demonstrating the excellent colloidal stability of these suspensions (Supplementary Information, Fig. S1A). The size distribution of exfoliated CoO2 nanosheet is determined by a standard dynamic light scattering (DLS) analysis (Supplementary Information, Fig. S1B). Most of the exfoliated CoO2 nanosheets possess the lateral size of several hundreds of nanometers, which is comparable with the reported lateral dimension of G-O nanosheet28. As illustrated in Fig. 1A, the hydrothermal treatment of Co2+ ions and NH4OH dissolved in the mixture colloidal suspensions of the layered CoO2 and G-O nanosheets makes possible the incorporation of layered CoO2 nanosheets into the Co3O4–N-doped rG-O nanocomposite. Since both the exfoliated CoO2 and rG-O nanosheets are negatively-charged, the precursor Co2+ ions can be easily adsorbed on the surface of both the anionic nanosheets, which is followed by the crystal growth of Co3O4 phase. The resulting Co3O4–layered CoO2–N-doped rG-O nanocomposites with different CoO2/G-O ratios (0, 0.5, 1 and 2wt%) are denoted as CCG0, CCG5, CCG10 and CCG20, respectively. The reduction of precursor G-O to N-doped rG-O during the synthesis is confirmed by C 1s and N 1s X-ray photoelectron spectroscopic (XPS) analysis (Supplementary Information, Fig. S2 and Table S2). As presented in the powder X-ray diffraction (XRD) patterns of Fig. 1B, all of the present nanocomposite materials show typical Bragg reflections of spinel-structured Co3O4 phase, indicating the formation of mixed valent cobalt (II,III) oxide phase during the hydrothermal reaction. On the basis of Scherrer equation, the size of Co3O4 particle in the present materials is calculated to be 7.7, 8.2, 8.5 and 9.0 nm for CCG0, CCG5, CCG10 and CCG20, respectively, highlighting a slight increase of particle size upon the incorporation of layered CoO2 nanosheets. The observed minute variation of the particle size of Co3O4 upon the incorporation of CoO2 nanosheets underscores the limited influence of layered CoO2 nanosheets on the crystal growth of Co3O4 nanoparticles.

Figure 1
figure 1

(A) Schematic diagram for the synthesis of the CCG nanocomposites. (B) Powder XRD patterns of (a) CCG0, (b) CCG5, (c) CCG10 and (d) CCG20.

As illustrated in the field emission-scanning electron microscopy (FE-SEM) images of Fig. 2A, all the present CCG nanocomposites commonly exhibit porous morphology formed by the house-of-cards-type stacking of nanosheet crystallites, indicating the formation of many mesopores. Such a mesoporous stacking structure is commonly observed for the self-assembled nanocomposite materials synthesized by the restacking of 2D nanosheets with 0D nanoparticles29. The nanoscale hybridization of cobalt oxide and graphene nanosheets is confirmed by energy dispersive X-ray spectroscopy (EDS)–elemental mapping analysis (Supplementary Information, Fig. S3), showing the uniform distribution of cobalt, oxygen and carbon in entire parts of the nanocomposite materials.

Figure 2
figure 2

(A) FE-SEM images of (a) CCG0, (b) CCG5, (c) CCG10 and (d) CCG20. (B) TEM/HR-TEM/FFT data of (a,b,c) CCG0 and (d,e,f) CCG10 nanocomposites.

The composite formation of the rG-O and CoO2 nanosheets with Co3O4 nanocrystals is obviously evidenced by high resolution-transmission electron microscopy (HR-TEM), see Fig. 2B. The fast Fourier transform (FFT) image of the CoO2-free CCG0 material in Fig. 2B–(b) clearly demonstrates the diffraction spots of graphene component. From the enlarged HR-TEM image of the CoO2-free CCG0 material in Fig. 2B–(c), the distance between two consecutive fringes is determined to be ~0.29 nm, which is in good agreement with the interplanar distance of {220} planes of Co3O4 phase, indicating the immobilization of the spherical Co3O4 particles on the surface of the nanosheets. For the case of CoO2–incorporated CCG10 nanocomposite, the FFT images of Fig. 2B–(e) provide clear evidence for the co-existence of layered CoO2 nanosheets with Co3O4 and graphene. This is further confirmed by the observation of clear lattice fringes corresponding to the {110} plane of graphene, the {104} plane of layered CoO2 and the {440} plane of Co3O4, as illustrated in Fig. 2B–(f). The present HR-TEM/FFT results obviously demonstrate the homogeneous hybridization of layered CoO2, graphene and Co3O4 in the present CCG10 nanocomposite. As a reference, the rG-O-free Co3O4–layered CoO2 nanocomposite is also synthesized by the same synthetic method as that for the CCG nanocomposites except for G-O and ammonia. Like the CCG nanocomposites, the obtained Co3O4–layered CoO2 nanocomposite displays the formation of spinel-structured Co3O4 nanoparticles anchored on the surface of layered CoO2 nanosheets (Supplementary Information, Fig. S4). This result confirms that, like the N-doped rG-O nanosheet, the layered CoO2 nanosheets can play a role of support for the anchoring of Co3O4 nanoparticles. During the anchored growth of Co3O4 phase, the layered CoO2 nanosheets remains intact without any notable damage. This result provides clear evidence for the high stability of layered CoO2 nanosheets against the hydrothermal synthesis.

The chemical bonding nature of rG-O and cobalt oxide components in the present nanocomposites is examined with micro-Raman spectroscopy. As illustrated in Fig. 3A, all the present nanocomposites show two intense Raman features D and G in high wavenumber region of >1000 cm−1, characteristic of graphene species, confirming the incorporation of graphene nanosheets in these materials30. In contrast to N-undoped G-O and rG-O nanosheets, all the CCG nanocomposites as well as N-doped graphene demonstrate a distinct shoulder peak D', indicating N-doping for graphene component30. This peak D' originates from a significant perturbation of the carbon sp2 network of graphene upon the incorporation of nitrogen element. The peak 2D reflecting the degree of the structural disorder and stacking of graphene is observed at ~2700 cm−1 for all the present materials. As the content of layered CoO2 nanosheets increases, this 2D peak shows a slight red-shift with the increase of spectral weight, clearly demonstrating the decreased numbers of stacked graphene layers upon the intervention of layered CoO2 nanosheets in-between the graphene nanosheets31 (Fig. 3B). This spectral variation provides strong evidence for the weakening of π–π interaction between the graphene nanosheets upon the incorporation of layered CoO2 nanosheets, leading to the prevention of the irreversible restacking or agglomeration of graphene nanosheets during electrochemical Li+ insertion/extraction. Such a depression of the restacking or aggregation of graphene nanosheets would be beneficial in enhancing the pore structure of the present CCG nanocomposites. In the low wavenumber region, typical Raman features of spinel Co3O4 phase are discernible commonly for all the CCG nanocomposites, confirming the formation of Co3O4 crystals in these materials. The incorporation of layered CoO2 nanosheets as well as the formation of Co3O4 particles in the present CCG nanocomposites is cross-confirmed by Co K-edge X-ray absorption near-edge structure (XANES) spectroscopy, see Fig. 3C. As can be seen clearly from the expanded views of edge jump region, a slight but distinct blue shift of edge position is clearly observed after the incorporation of layered CoO2 nanosheets, highlighting an increase of average Co oxidation state caused by the increase of CoO2 content. This result confirms the HR-TEM results showing the presence of CoO2 nanosheets in the present CCG nanocomposites. The successful incorporation of tetravalent CoO2 nanosheets in the present materials is further evidenced by the Co 2p XPS result (Supplementary Information, Fig. S5), in which the tetravalent Co4+ ions are identified and the concentration of Co4+ ion increases with increasing the concentration of CoO2 nanosheet incorporated. Both the XANES and XPS results provide strong support for the HR-TEM results showing the presence of CoO2 nanosheets in the CCG nanocomposites (Fig. 2B).

Figure 3
figure 3

(A) Micro-Raman spectra, (B) their expanded views of high wavenumber region and (C) Co K-edge XANES spectra and their expanded views near edge jump for (a) Co3O4, (b) G-O, (c) rG-O, (d) N-doped rG-O, (e) CCG0, (f) CCG5, (g) CCG10 and (h) CCG20.

As plotted in Fig. 4A, N2 adsorption−desorption isotherm measurements clearly demonstrate the porous nature of the present CCG nanocomposites. All of the present materials display a significant N2 adsorption at low pressure region of ppo−1 < 0.4, reflecting the existence of micropores in these materials. A distinct hysteresis commonly occurs at high pressure region of ppo−1 > 0.45 for all the present CCG nanocomposites. The observed isotherm behavior corresponds to Brunauer–Deming–Deming–Teller (BDDT) type-IV shape and IUPAC H2-type hysteresis loop, suggesting the presence of open slit-shaped capillaries with very wide bodies and narrow short necks. The incorporation of layered CoO2 nanosheets enhances the adsorption of N2 molecule in the low pressure region and also the total amount of N2 molecules adsorbed, underscoring the remarkable increase of micropore volume and surface area. According to the calculation of surface area using the Brunauer–Emmett–Teller (BET) equation, the surface area of the present nanocomposite is estimated to be 32 m2g−1 for CCG0, 64 m2g−1 for CCG5, 97 m2g−1 for CCG10 and 82 m2g−1 for CCG20, respectively. This result demonstrates that the surface areas of the present CCG nanocomposites become greater with increasing the content of CoO2 nanosheets upto the composition of CCG10. However, the further addition of CoO2 nanosheets leads to the depression of surface area. The observed lowering of the surface area of the CCG20 nanocomposite is attributable to too high content of CoO2 nanosheet, which is much heavier than the graphene. That is, the increase of the sample mass caused by the addition of heavy CoO2 nanosheets outweighs the accompanying optimization of the pore structure of the nanocomposite. This result clearly demonstrates that, even at a small concentration of CoO2 nanosheets, the incorporation of layered CoO2 nanosheets is fairly useful in expanding the surface area of Co3O4–graphene nanocomposite. As evidenced by the micro-Raman spectroscopy (Fig. 3B), the incorporation of CoO2 nanosheets is effective in depressing the π–π interaction between rG-O nanosheets and also in preventing the formation of tightly packing structure of graphene. Taking into account the fact that the severe self-restacking of graphene nanosheets leads to the remarkable decrease of surface area, the observed increase of the surface area of CCG nanocomposites upon the incorporation of CoO2 nanosheets can be attributed to the depressed interaction between the graphene nanosheets. In fact, such a prominent increase of the surface area of restacked graphene nanosheets upon the incorporation of inorganic nanosheet is also observed for other cases like Pt–layered titanate–graphene nanocomposite, clarifying the effectiveness of the nanosheet addition in enhancing the porosity of graphene-based nanocomposite9. The calculation of pore size based on Barrett–Joyner–Halenda (BJH) method (Fig. 4B) clearly demonstrates that all the present materials possess uniform-sized mesopores with an average diameter of ~3.3–3.4 nm, which are formed by the house-of-cards-type stacking structure of nanosheet crystallites. The BJH analysis reveals the increase of pore volume upon the incorporation of layered CoO2 nanosheets, highlighting the positive effect of inorganic nanosheet in enhancing the porosity of graphene-based nanocomposites. The present results of N2 adsorption–desorption isotherm analysis clearly demonstrate that the incorporation of CoO2 nanosheets is quite powerful in increasing the surface area and pore volume of the restacked graphene nanosheets via the depression of π–π interaction between the graphene nanosheets.

Figure 4
figure 4

(A) N2 adsorption–desorption isotherms and (B) pore size distribution curves calculated on the basis of the BJH equation for the nanocomposites of (a) CCG0, (b) CCG5, (c) CCG10 and (d) CCG20.

The present CCG nanocomposites are applied as anode materials for lithium ion batteries. Fig. 5A shows the representative cyclic voltammogram (CV) curves of the CCG10 electrode, which are collected at a scan rate of 0.5 mV s−1 in the voltage range of 0.01–3 V vs. Li/Li+. In the first cycle, an irreversible reduction peak appears at ~0.7 V, which originates from the degradation of electrolyte caused by the formation of polymer/gel-like film around the electrode particles32. In the second cycle, there are two cathodic peaks at 1.41 and 0.9 V, which are ascribed to the reduction of Co3O4 to Co caused by the lithiation of Co3O4. The lithiation voltage of the second cycle is shifted to higher value than that of the first cycle, indicating the improved kinetics of the CCG10 nanocomposite12. Meanwhile, two anodic peaks at 1.46 and 2.15 V are attributable to the oxidation of Co element to Co3O4, which corresponds to the delithiation process. These redox peaks can be regarded as evidence for the electrochemical reaction of Co3O4 and Li12. Although very small quantity of layered CoO2 nanosheet makes it difficult to directly detect a redox peak corresponding to reaction between layered CoO2 and Li in the present CV data, the lithium insertion/extraction reactions in the present nanocomposites can be described by the following equations.

Figure 5
figure 5

(A) CV curve of CCG10, (B) discharge–charge capacity plots at current density of 200  mA g−1 and (C) rate-dependent capacity plots of (a) CCG0, (b) CCG5, (c) CCG10, (d) CCG20 and (e) layered CoO2 nanosheets. (D) discharge–charge capacity plots of CCG10 electrode at current density of 1000 mA g−1. (E) Potential profile at current density of 1000 mAg−1 for the nanocomposite of CCG10.

Of prime importance is that there is no significant difference in the CV data of the CCG10 nanocomposite for the 2nd and 3rd cycles, highlighting the good reversibility of this material.

Fig. 5B shows the galvanostatic discharge–charge curves at a current density of 200 mA g−1 in the range of 0.01–3 V vs. Li/Li+. All the present nanocomposites exhibit promising electrode performance with the huge initial discharge capacity of 1505, 1963, 2262 and 1797 mAh g−1 for CCG0, CCG5, CCG10 and CCG20, respectively. Although notable capacity fading occurs at the second cycle due to the formation of solid–electrolyte–interphase (SEI) layer32, the discharge capacity of CCG nanocomposites becomes greater with proceeding the cycle. Such an increase of discharge capacity is frequently observed for porous nanostructured materials, which is related to the formation of stable diffusion paths of Li+ ions during the repeated electrochemical cycling12,33,34,35. Among the present nanocomposites, the CCG10 nanocomposite with the largest surface area exhibits the most prominent enhancement of discharge capacity during the cycle. After the 20th cycle, the discharge capacities of CCG nanocomposites are stabilized to ~1230 mAh g−1 for CCG0, ~1500 mAh g−1 for CCG5, ~1750 mAh g−1 for CCG10 and ~1530 mAh g−1 for CCG20, highlighting the promising electrode performance of the present nanocomposites with huge discharge capacity and good cyclability. To the best of our knowledge, the observed discharge capacity of the CCG10 nanocomposite is the largest reversible capacity of Co3O4-based materials ever-reported (Supplementary Information, Table S3). Since all of the components in the present nanocomposite including N-doped rG-O nanosheet are electrochemically active36, the observed huge capacity of the present material is a result of the synergistic combination of these electrochemically active materials. On the basis of the theoretical capacities of the component materials (i.e. 890 mAhg−1 for Co3O4, 1216 mAhg−1 for N-doped rG-O and 1178 mAhg−1 for CoO2)12,35, the theoretical capacity of the present CCG10 nanocomposite is estimated to be ~960 mAhg−1, which is much smaller that the observed reversible capacity of ~1750 mAhg−1. In fact, there are several reports about the larger capacity of graphene-based nanocomposite than the theoretical one33,34,37. On the basis of these researches, several factors are supposed to be responsible for the unusually large reversible capacity of the present CCG nanocomposite; (1) the expansion of surface area with the increase of pore volume upon the composite formation results in the additional storage of Li+ ions in the interfacial site of nanocomposite38. (2) The N-doping for the rG-O component makes another contribution to the large discharge capacity of the nanocomposite, since the N-doping can improve the diffusivity of Li ions in the electrode and can create defects in the graphene lattice providing more active sites for Li insertion and increasing Li adsorption energies at the vacancy sites36. (3) During the cycling process, the crystalline Co3O4 nanoparticles are changed to amorphous ones, leading to the formation of more accessible active sites for Li-ion insertion34. (4) The graphene nanosheets suffers from a strong tendency to form tightly packed structure due to the strong π–π interaction between sp2 carbon arrays. The incorporation of CoO2 nanosheets induces the formation of more open stacking porous structure providing more active sites for Li+ ions insertion9.

All the present CCG nanocomposites display high coulombic efficiency of >98%, reflecting the highly stable and reversible insertion/extraction of lithium ions. As shown in the potential profiles of the CCG nanocomposites (Supplementary Information, Fig. S6), all the present materials show nearly identical potential profiles, indicating the retention of the original electrochemical properties of Co3O4–graphene nanocomposite upon the incorporation of CoO2 nanosheets. Based on the present results of electrochemical measurements, it can be concluded that the incorporation of layered CoO2 nanosheets leads to the remarkable improvement of the electrode performance of Co3O4–rG-O (i.e. CCG0) nanocomposite.

As presented in Fig. 5C, the remarkable improvement of electrode performance upon the incorporation of layered CoO2 nanosheet is more distinct for higher current density condition. The CoO2-incorporated CCG10 nanocomposite exhibits larger reversible capacities for all the current densities applied than does the CoO2-free CCG0 material; the CCG10 nanocomposite shows the discharge capacities of 1684, 1656, 1562, 1288, 1002 and 817 mAh g−1 at the current density of 100, 200, 400, 800, 1600 and 3200 mA g−1, respectively. However, the CoO2-free CCG0 electrode delivers much smaller discharge capacities of 1232, 1207, 1044, 949, 642 and 428 mAh g−1 at the same current densities, respectively. While the discharge capacity of the CCG10 nanocomposite at 100 mA g−1 is larger by 136% than that of the CCG0 material, the CCG10 material shows even twice larger discharge capacity compared with the CCG0 one at a higher current density of 3200 mA g−1. This finding provides clear evidence for the improvement of rate performance upon the incorporation of CoO2 nanosheets, highlighting the improvement of charge transport property.

To examine the long-term stability of the CoO2-incorporated nanocomposite, the extended electrochemical cycling test is carried out for the CCG10 nanocomposite with high current density of 1000 mA g−1 and in the voltage window of 0.01–3 V vs. Li/Li+. As plotted in Figs. 5D, E, this material shows an outstanding cyclability and rate capability with the maintenance of large specific capacity of ~1150 mAh g−1 upto the 500th cycle. For the entire cycle, the coulombic efficiency of the CCG10 nanocomposite is well-maintained to ~99%, verifying the high electrochemical stability of this material. The observed beneficial effect of the incorporation of CoO2 nanosheets on the discharge capacity of the nanocomposite is attributable to the expansion of surface area and the change of pore structure, resulting in the additional storage of Li+ ions in interfacial site of the CoO2 nanosheet-scaffolded nanocomposite. The incorporation of layered CoO2 nanosheets also induces an enhanced nanoscale mixing between Co3O4 particles and rG-O nanosheets, which is responsible for the excellent cyclability and rate characteristics of the present CoO2-incorporated CCG nanocomposites.

To better understand the origin of the beneficial effect of CoO2 addition, the transport property of the present nanocomposites is investigated with electrochemical impedance spectroscopy (EIS). As plotted in Fig. 6, all the present nanocomposites demonstrate partially overlapping semicircles reflecting the charge transfer resistance (Rct) at high-to-medium frequencies and a line corresponding to Warburg impedance at low frequencies. The incorporation of layered CoO2 nanosheets gives rise to a significant reduction in the diameter of semicircle, indicating the decrease of Rct. Among the present nanocomposites, the CCG10 material displays the smallest diameter of the semicircle, indicating its most efficient transport property. A further increase of CoO2 content to the CCG20 nanocomposite degrades the electron transport property, which is attributable to the decrease of highly conductive graphene content. The relative order of Rct is in good agreement with the relative electrode performances of the present nanocomposites, underscoring the main role of the improvement of transport property in enhancing the electrode performance upon the incorporation of CoO2 nanosheets. Such a variation of transport properties is further confirmed by the change of the line slope in the low frequency region. The slope of this line for the present nanocomposites becomes steeper in the order of CCG0 < CCG20 < CCG5 < CCG10, reflecting the improvement of transport property caused by the promoted nanoscale mixing of Co3O4 nanoparticles and graphene nanosheets upon the incorporation of CoO2 nanosheets.

Figure 6
figure 6

EIS spectra of (a) CCG0, (b) CCG5, (c) CCG10 and (d) CCG20.

The effects of electrochemical cycling on the crystal structure and morphology of the present nanocomposites are examined with powder XRD and FE-SEM analyses. As demonstrated in Fig. 7A, the extended electrochemical cycling induces a decrease of the particle size of the CCG10 nanocomposite whereas the CoO2-free CCG0 nanocomposite shows a significant aggregation of electrode particles. Such an aggregation of electrode particles is negligible for the CoO2-incorporated CCG10 nanocomposite, confirming the beneficial role of CoO2 nanosheets in the maintenance of the open structure of nanocomposite. Since Co3O4 experiences severe volume change during lithiation–delithiation process, the depression of particle agglomeration is surely advantageous in enhancing the electrode performance of the nanocomposite. In addition, the electrochemical cycling induces an amorphization of both the CCG0 and CCG10 nanocomposites, see Figs. 7B,C. Such a formation of disordered structure during the electrochemical cycling causes the significant enhancement of Li+ diffusion via the provision of more diffusion paths39. The incorporation of layered CoO2 nanosheets does not induce any significant change in the XRD data of the cycled derivative, indicating negligible effect on the structural stability of the nanocomposite. The present finding strongly suggests that the beneficial role of CoO2 addition mainly originates from the improvement of the morphological stability of composite structure rather than the change of crystal structure. Even though the incorporated CoO2 nanosheets is transformed into cobalt oxide particles during the electrochemical cycling, an intimate mixing between cobalt oxide and rG-O in the CoO2-incorporated CCG10 nanocomposite provides improved diffusion paths for Li+ ions as well as a strong electronic coupling between cobalt oxide and graphene nanosheeets.

Figure 7
figure 7

(A) Discharge–charge FE-SEM images of (a) CCG0 and (b) CCG10. Powder XRD patterns of (i) as-prepared and (ii) electrochemically-cycled nanocomposites (after the 50th cycle) of (B) CCG0 and (C) CCG10.

Such an advantageous effect of the incorporation of metal oxide nanosheet is further evidenced from MnO2 nanosheet-incorporated Mn3O4–N-doped rG-O nanocomposite. The obtained Mn3O4–layered MnO2–N-doped rG-O displays the typical XRD patterns of Mn3O4 phase and porous morphology formed by the house-of-cards-type stacking of sheet-like crystallites, as observed for the present CCG nanocomposites (Supplementary Information, Fig. S7). This material shows the discharge capacity of 1383 and 900 mAh g−1 for the 1st and 2nd cycles, respectively. The discharge capacity of this material becomes increasing with proceeding the cycling, leading to the huge discharge capacity of 1250 mAh g−1 at the 50th cycle. Even at high current density of 2000 mA g−1, the material can still deliver a high capacity of ~700 mAh g−1 (Supplementary Information, Fig. S7). The electrode performance of the present nanocomposite is the best one among the reported data of Mn3O4–graphene nanocomposites (Supplementary Information, Table S4), highlighting the excellent electrode activity of the present material. It is worthwhile to mention that, in comparison with the cobalt oxide-based CCG nanocomposites, the Mn3O4–layered MnO2–N-doped rG-O nanocomposite is much more promising electrode material because of the low price and low toxicity of Mn elements. This result underscores the usefulness of the present synthetic strategy for exploring novel efficient electrode materials highly suitable for practical use.

Discussion

In the present study, we are successful in developing a very efficient method to improve the electrode performance of graphene-based nanocomposite materials using the colloidal mixture of layered metal oxide and graphene nanosheet as a precursor. In fact, there are considerable numbers of reports about the cobalt oxide–graphene nanocomposite electrode materials for lithium secondary batteries12,35,40,41,42,43. In one instance, Wang et al. reports the large discharge capacity of 1200 mAh g−1 for the Co3O4–graphene film, which is greater than the data of other previous reports12,40. In comparison with these data, the present CCG nanocomposites show much larger discharge capacity of ~1550–1750 mAh g−1 with excellent cyclability and good rate characteristics, which outperforms all of the previously reported Co3O4-based electrode materials (Supplementary Information, Table S3). Similarly the electrode performance of the Mn3O4–layered MnO2–N-doped rG-O nanocomposite is superior to all the reported data of Mn3O4–graphene nanocomposite (Supplementary Information, Table S4). This result provides strong evidence for the unique merit of metal oxide nanosheets as an additive for graphene-based nanocomposite electrode materials. The observed dramatic enhancement of the electrode performance upon the incorporation of layered metal oxide nanosheet is attributable to the increase of surface area, the enhancement of the nanoscale mixing of components, the improvement of electrical transport properties and the enhancement of morphological stability upon the incorporation of layered metal oxide nanosheets. The present study obviously verifies a beneficial and universal role of exfoliated metal oxide nanosheets in optimizing the electrode performance of graphene-based nanocomposite. As mentioned in the introduction section, the graphene-based nanocomposites boast versatile applications such as electrodes for secondary batteries, supercapacitors, fuel cells and solar cells, photocatalysts, redox catalysts, nanobio materials, structural materials and so on4,5,6,7,8,11,15,40,41,42,43,44,45,46. For most applications of these graphene-containing materials, the homogeneous blending between graphene nanosheets and hybridized functional materials and the optimization of porous structure are commonly important in enhancing their functionalities. Taking into account the fact that the present synthetic strategy is readily applicable for many types of rG-O-based nanocomposites, the incorporation of exfoliated inorganic nanosheets can provide a powerful methodology to optimize the diverse functionalities of graphene-containing nanocomposites through the effective deterioration of the tight packing structure of graphene nanosheets. Our current research project is the use of the various colloidal mixtures of graphene and inorganic nanosheets for the exploration of novel graphene-based functional materials with various applicabilities for solar cells, photocatalyst, supercapacitors and so on.

Methods

Materials preparation

The exfoliated layered CoO2 nanosheets was prepared by the proton exchange of LiCoO2 and intercalation of tetramethylammonium (TMA) ions into HCoO223. The colloidal suspension of graphene oxide (G-O) was synthesized from graphite by a modified Hummers' method, in which the concentration of KMnO4 was reduced to 1/6 of conventional concentration47. The as-prepared G-O was dispersed in anhydrous ethanol with the concentration of 0.32 mg mL−1 by ultrasonication for 0.5 h. An aqueous suspension of exfoliated G-O (48 mL) was reacted with 2.4 mL 0.2 M Co(Ac)2 and 1 mL 30% NH4OH aqueous solution, 1 mL H2O and layered CoO2 nanosheets (0–2 wt% to exfoliated G-O nanosheets). The mixture was stirred at 80 °C for 10 h. Then the mixture was transferred to an autoclave for hydrothermal reaction. The reaction condition was 3 h at 150 °C. In this step, G-O was reduced to N-doped rG-O. After the reaction, powdery precipitates were collected by centrifugation, washed with ethanol and distilled water and then freeze-dried. After the completion of the reaction, only transparent supernatant solution remained. No observation of Tyndall phenomenon for the supernatant solution clearly demonstrated the absence of any precursor colloidal particles in this solution. Additionally, no formation of precipitate upon the addition of hydroxide ions confirmed the complete incorporation of Co2+, CoO2 and graphene reactants into the precipitated nanocomposite materials. On the basis of the present findings, the weight ratios of the components in these materials could be estimated from the starting ratios of the reactants. The weight ratio of Co3O4:CoO2:rG-O components was estimated to 3.7:0:1 for CCG0, 3.7:0.007:1 for CCG5, 3.7:0.015:1 for CCG10 and 3.7:0.03:1 for CCG20, respectively. Since the weight of G-O component was much more convenient and precise to calculate than its molar concentration, the weight ratios of CoO2/G-O were applied for controlling the compositions of the present nanocomposites like many other studies about the graphene-based nanocomposites.

Materials characterization

The crystal structures of the as-prepared CCG nanocomposites and their electrochemically cycled derivatives were analyzed by powder XRD analysis (Rigaku D/Max-2000/PC, Cu Kα radiation, 298 K) analysis. The crystal morphology of the present samples was examined by FE-SEM (JEOL JSM-6700F) and HR-TEM/SAED (Jeol JEM-2100F, an accelerating voltage of 200 kV). The spatial elemental distribution of the present materials was probed with EDS–elemental mapping analysis. XANES spectroscopic experiment was carried out at Co K-edge at the beam line 10C at the Pohang Accelerator Laboratory (PAL) in Korea. The chemical bonding nature of nitrogen species was investigated with XPS analysis (Thermo VG, UK, Al Kα), in which a monochromated X-ray beams was used. All the XPS spectra were calibrated with a reference to the adventitious C 1s peak at 284.8 eV to rule out any possible spectral shift by the charging effect. To avoid the accumulation of charge during the measurement, all the samples were deposited on metallic copper foil. N2 adsorption–desorption isotherms were measured at 77 K using Micromeritics ASAP 2020 analyzer to determine the surface area. Before the measurements, the degassing of the samples was carried out at 150 °C for 3 h under vacuum. Micro-Raman spectra were obtained with a JY LabRam HR spectrometer using an excitation wavelength of 514.5 nm. The zeta potentials of the pure colloidal suspensions of G-O and layered CoO2 nanosheets and their colloidal mixtures were measured with Malvern Zetasizer Nano ZS (Malvern, UK).

Electrochemical measurement

The CV data were collected using an IVIUM analyzer with a scanning rate of 0.5 mV s−1 and a potential range of 0.01–3.0 V (vs. Li/Li+). The EIS data were collected in the frequency range of 0.01 Hz–100 KHz. Electrochemical measurements were carried out at room temperature using 2016 coin-type cell of 1 M LiPF6 in an equivolume mixture of ethylene carbonate/diethyl carbonate (EC/DEC = 50:50). The working electrodes were fabricated by mixing 80 wt% active material, 10 wt% Super P and 10 wt% polyvinylidene fluoride (PVDF) dissolved in N-methyl-2-pyrrolidinone (NMP). The composite electrodes were prepared by coating the anode slurry onto a copper foil as a current collector and drying under vacuum at 110 °C for 12 h. The test cells were assembled in an argon-filled glove box. All the galvanostatic charge–discharge tests were performed with Maccor (Series 4000) multichannel galvanostat/potentiostat in the voltage range of 0.01–3.0 V (vs. Li/Li+) at current density of 100–3200 mA.

Additional Information

How to cite this article: Jin, X. et al. An Effective Way to Optimize the Functionality of Graphene-Based Nanocomposite: Use of the Colloidal Mixture of Graphene and Inorganic Nanosheets. Sci. Rep. 5, 11057; doi: 10.1038/srep11057 (2015).