Guohui Qin*ab,
Xuan Wuab and
Hongjuan Zhangab
aSchool of Chemical Engineering & Technology, Tianjin University, Tianjin 300072, China. E-mail: guohuiq163@sina.com; Fax: +86 22 27890481; Tel: +86 22 27890481
bSynergetic Innovation Center of Chemical Science and Engineering, Tianjin 300072, China
First published on 30th September 2014
Highly ordered mesoporous crystalline C–TiO2–V2O5 core–shell microspheres encapsulated by porous carbon further embedded in N-doped graphene network (GN–TV–C) nanostructures were fabricated by a simple combination of hydrothermal-calculation method. Such material exhibits a highly efficient photocatalytic activity for water splitting, as well as a high specific capacity and exceptional cycle ability in LIBs. The extrusive features of such material include the assembly of components in a manner that enables an effective integration between the constituents and the ability to modify the electronic properties of GN–TV–C. The positive synergistic incorporation between TiO2 and V2O5, the high electrical conductivity, and the three-dimensional hierarchically mesoporous nanostructure of these composites result into a highly active photocatalytic ability, showing patterns of increased light harvesting ability and promoted exciton dissociation, and excellent electrochemical performance in terms of a high rate capability and stable cycling. Profiting from the dual insurance of the flexible carbon layer derived from glucose and elastic GN walls with superior specific surface areas, a significant enhancement in the electron transfer and electronic diffusion channels, and a highly enhanced structural stability of the TiO2-based electrode material were simultaneously obtained. In addition, the synergistic function among TiO2, V2O5 and GN involving the optimized energy gap, the compromised particle assembly and surface defects, as well as their distinctive core–shell nanostructures were extensively studied. The carbon shell serves as a blocking layer that retards the interfacial recombination during photocatalysis, thereby protecting the active materials from pulverization during the superior cycles in the energy store. The current study may provide us an alternative approach for improving the performances of TiO2 nanocrystals used in energy storage and photocatalysis applications.
However, numerous unfavorable factors, such as the short lifetime of the charge carriers9 and poor ionic and electronic conductivity, restrict its capacity and rate capability.10–12 To facilitate the separation and transfer of photogenerated charge carriers so as to improve the kinetics of water-splitting processes and enhance the conductivity of the electrode material in order to build blocks for preferable channels for electron and ion separation and utilization.13–16 Quite recently, several studies have shown that TiO2-based materials exposed to highly-reactive facets exhibit superior photocatalytic activities, as well as enhanced lithium storage capabilities.17–21 N-doped graphene (GN) introduced into the GN-semiconductor photocatalysts has been certified to extend the visible light harvest ability. It also prolongs the lifetime of the exciton and facilitates the transfer of the photogenerated charge carriers; thus, improving the adsorption capacity of the reactant, for which GN is supposed to behave as an electron reservoir in order to capture/shuttle the electrons photogenerated from the semiconductor.13–17 On the other hand, the GN are also observed to show a stable cycle ability and excellent rate capability as anodic materials in lithium ion batteries. Therefore, the heteroatom doping further generates a narrow band gap of photocatalysts and extrinsic defects in the interwall space for Li storage.
However, the practical application of TiO2–V2O5 is restricted by the instability of the interaction between the TiO2–V2O5 nanoparticles and the GN surface both for the photocatalyst and anodic material. Zhang recently reported that the rendering of glucose can act as the facet-controlling agent and as a connection linker owing to its large number of exposed hydroxyl groups.18 The addition of glucose during the preparation of TiO2-based materials increases the surface active sites, which serves as an electron trap by capturing electrons and promotes charge separation for the photocatalytic reaction. This subsequently promotes the surface redox reactions and inhibits the photocorrosion problem of V2O5 in the photocatalysis case19,20 and prohibits the direct contact of V2O5 within electrolytes. This in turn prevents its dissolution in electrolytes that can cause unfavorable side reactions; thus, leading to rapid capacity fading during cycling in the energy store problem, which is a bottleneck problem for V2O5 as an anodic material.
Herein, we report a simple template-free self-assembly synthesis of uniform continuous carbon layer-encapsulated TiO2–V2O5 mesoporous nanostructures that closely anchor on the GN sheets (GN–TV–C). The close incorporation of TiO2–V2O5 within GN is via the bulk intercalative powder between the carbon layer from the glucose and from the GN sheets. This design possesses nanostructure engineering and combination project with which have been demonstrated two effective approaches for improving the TiO2 based materials both for photocatalysts and LIBs. The visible spectrum of the as-prepared material shows extended light absorption from 400 to 600 nm. It shows remarkable photocatalytic activity for the water splitting in aqueous solution under visible light irradiation. Moreover, the lithium storage performance of such material was significantly enhanced in terms of excellent rate capability and superior cycle endurance. A reversible lithium storage capacity of 288.52 mA h g−1 for up to 1000 cycles without any notable decrease was obtained. The excellent photocatalytic and electrochemical behaviors of such complicated material are attributed to the core–shell nature and well-connected nanoscale heterostructure (the strong positive synergistic effect between GN and high active TiO2–V2O5 nanoparticles) with the improvement of both photocatalysis and electrochemical performance. The configuration of the carbon layer retards the electron–hole recombination and prevents the structural deformation upon cycling. It also shortens the electron–ion transfer path, while the embedded GN serves as an electron trap and preserves the electrical integrity between the electrochemical active species and the current collector. This ensures a good contact among the carriers with water molecules, electrodes and electrolyte, respectively, during the progress of the above reaction.
Photoelectrochemical measurements were carried out in a three-electrode system with a potentiostat (HSV-110, Hokuto Denko) and an electrochemical cell at room temperature. All measurements were carried out under an inert helium atmosphere in 1 M K2SO4. Typically, 4 mg of the catalyst and 30 μL of the Nafion solution (Sigma-Aldrich, 5 wt%) were dispersed in 1 mL of the water–isopropanol solution with a volume ratio of 3:
1 by sonicating for 1 h to form a homogeneous ink. Then, 5 μL of the dispersion (containing 20 μg of catalyst) was loaded onto a FTO electrode with 3 mm diameter (loading 0.285 mg cm−2). Hg/Hg2SO4 in 1 M K2SO4 saturated solution was employed as the reference electrode, a Pt rod (AlfaAesar, 99.9995%) was designed as the counter electrode, and the GN–TV–C/FTO functioned as the working electrode. All the voltages were calculated relative to the reversible hydrogen electrode according to the Nernst equation:
V vs. RHE(volt) = V vs. Hg/Hg2SO4(volt) + 0.236(volt) | (1) |
The Nyquist plots were measured with frequencies ranging from 1000 kHz to 1 Hz at an overpotential of 200 mV. The impedance data was fitted to a simplified Randles circuit to extract the series and charge-transfer resistances. The irradiation wavelength was controlled by a 300 W Xe arc lamp (PLS-SXE300, Beijing Perfect light Co., Ltd.) equipped with a UV-cut filter to cut off light at certain wavelengths (λ > 420 nm). Light intensity, as measured by a visible-irradiance meter, was about 4.0 mW cm−2 at the position of the working electrode. The dark currents were orders of magnitude lower than photocurrents in all cases within the measured voltage ranges.
The water splitting experiment was conducted in H-type reactor with 50 mL of 0.1 M K2SO4 aqueous solution in each solution. A small overpotential as high as 200 mV was needed to support the stable H2 and O2 evolution rates. A bubbler purged the electrolyte in the reactor to push the evolved gases into a gas chromatograph for analysis. After that, the gas was fed back into the reactor via a recirculation pump. The gas chromatograph (SRI Instrument, Inc.) was equipped with a molecular sieve 13× packed column and a helium ionization detector. Before each reaction, the system was calibrated with H2 and O2 gases of accurate concentration.
It appeared that the V2O5 component in these samples was distributed mainly near the surface, but the degree of segregation was too small to be detected by XRD, owing to the relatively low concentration of the component (V/Ti = 0.01). Graphene and nitrogen were also not distinctively detected in the pattern of all the Ti samples because of their less content or amorphous feature. However, the Raman shift highlighted the presence of graphene in the GN–TV–C composite (Fig. 1b), whereas the XPS spectroscopy confirmed the existence of nitrogen. The two distinct peaks correspond to the D band at ca. 1350 cm−1 and the G band at ca. 1595 cm−1, respectively.22 The D band is commonly attributed to a series of structural defects, whereas the G band is observed for all graphitic structures. It is noted that the intensity ratio of the D and G bands, ID/IG, represents the relative concentration of the sp3 hybridized defects relative to the sp2 hybridized graphene domains. The average ID/IG values for GN–TV–C and GN–TV are 1.25 and 1.12, respectively, demonstrating that the presence of GN introduces some defects and the load of the carbon layer from glucose further increase the disorder of GN–TV–C to some extent.
The carbon and nitrogen contents in the GN–TV–C composites were determined by TG analysis (Fig. 1c). In the case of GN–TV, 5.25% of weight is achieved at a high temperature (800 °C) because of the complete combustion of GN under air atmosphere. The TGA of GN is nearly overlapped with that of GN–TV, indicating the complete reduction of GO to graphene. The oxidative decomposition of the carbon shell is believed to be attributed to the different weight loss of GN–TV–C and GN–TV. The content of the carbon shell in the GN–TV–C composites can be calculated to be 4.98%. It proves that the TV composites were effectively restricted in the pores of GN–TV–C, which is consistent with the results of BET measurements as shown in Fig. S2.†
The N2 adsorption/desorption isotherms were performed to study the porous feature of GN–TV–C microspheres. As shown in Fig. S2a,† a distinct hysteresis loop between 0.8 and 1.0 was observed, which is a characteristic of mesoporous materials. The pore-size distribution curve of GN–TV–C microspheres monitored by Barrett–Joyner–Halenda (t-plot) method have micropores relatively with a size distribution of 0.65–1.9 nm and a sharp peak at 30 nm (Fig. S2b†), demonstrating the mesoporous structure GN–TV–C nanoparticles. The Brunauer–Emmett–Teller (BET) surface area of GN–TV–C nanostructures is calculated to be 142 m2 g−1, a pore volume of 0.587 cm3 g−1, which is a significant improvement as compared with that of bulk TV nanoparticle (35 m2 g−1 and 0.139 cm3 g−1) and the GN–TV nanocrystals (102 m2 g−1 and 0.413 cm3 g−1), indicating the efficiency of the carbon shell and GN sheets in increasing the surface area of TiO2-based nanohybrids. The substantial increase of surface area and pore volume is mainly because of the contribution of GN nanosheets with a large outer surface and the porous carbon shells derived from the carbonization of glucose with inner specific area. It is believed that the larger specific surface area of mesoporous nanostructure promotes an increase in the photocatalytic and electrochemical active sites, which are beneficial for improving the photocatalytic and electrochemical performances of GN–TV–C.
Furthermore, the X-ray photoelectron spectroscopy (XPS) spectra with the total survey spectrum (Fig. 1d) also verified the existence of GN–TV–C. The C1s XPS (Fig. 2a) indicates the abundance of various oxygen-containing functional groups on the graphene surface. The strong binding energy peak at 284.7 eV is assigned to the C–C bonding, 287.8 eV is assigned to the CO bonding, and the weak binding energy peak at 289.3 eV is assigned to the O–C
O bonding, indicting the strong chemical reactivity between TiO2–V2O5 and graphene, in which the TiO2–V2O5 particles and GN in excellent interface binding were uniformly interconnected, which is further confirmed by SEM and TEM images of these materials (Fig. 3–5).
![]() | ||
Fig. 2 XPS spectrum of GN–TV–C at 700 °C in NH3 atmosphere for 1 h: C1s narrow scan (a) N1s narrow scan (b) Ti2p narrow scan (c) O1s narrow scan (d). |
![]() | ||
Fig. 3 Images of pure TV samples at different magnifications (a) and (b) GN–TV–C composite at different magnifications (c) and (d). |
![]() | ||
Fig. 4 EDS map of the selected rectangular area of FESEM image of GN–TV–C nanocrystals and results the of EDS analysis. |
As shown in Fig. 2b, the low intensity of N1s signal indicate that N is more incorporated with graphene than with TiO2–V2O5. Two characteristic peaks of the Ti2p spectrum at 458.2 and 463.9 eV are attributed to the Ti2p3/2 and Ti2p1/2 (Fig. 2c), respectively. According to the relationship between the oxidation state of Ti and the separation of peak energy (Ti2p), such energy gap of 5.7 eV infers that the average oxidation state of elemental Ti in the GN–TV–C composite is around −3.9, showing the dominance of Ti4+ in the rutile TiO2. The two peaks located at ∼532.5 and ∼533.3 eV attributed to the O1s (Fig. 2d). In addition, for GN–TV–C core–shell–shell nanostructure, the TiO2 core has been fully coated with the V2O5 shell format and continuous carbon layer deriving from glucose, the binding energy of V mainly located at 517 eV and 525 eV are covered by the large amounts of O signal; thus, reducing the signal intensity of the inside element and the V element is obviously not distributed in the XPS spectrum. However, V is certified in the EDS file (Fig. 4). The heterojunction between the core–shell–shell TiO2–V2O5–C nanostructures embedded in continuous GN network can facilitate the quick transfer of the photogenerated electrons from TiO2–V2O5–C to GN. The shorten transfer time for long lived-exciton impedes the electron–hole recombination and their ability to participate in electron transfer is integral for these excitons to serve as light harvesting antennae. In addition, the flexible carbon layer and elastic GN sheets significantly improve the conductivity of the electrode material, accommodate the large volume change of GN–TV–C anode upon lithiation–delithiation, leading to an excellent cycling stability and high coulombic efficiency even at high current densities, protecting from V2O5 from dissolution in aqueous and electrolyte, prevent its photocorrosion under visible light irradiation and decay in long life cycle in LIBs, and thus, improve the structural stability. This investigation might facilitate the high-efficient and stable hybrid photocatalysts applications in water splitting using solar energy and energy store in the development of sustainable resources. In particular, the adjustable space in the core–shell accommodates well a dispersion of active nanocrystals and the retention of intact protective graphene layer finally contributes to the significantly improved cycle ability. The continuous carbon network can not only provide higher accessible surface area for the GN–TV–C nanostructures in the composite, but also allow the enhanced light absorption capability and the electrolyte ions to rapidly diffuse inside. Therefore, it is highly desirable to fabricate pore-rich materials for their potential applications in both catalysis and energy storage.
The particle morphology and size of GN–TV–C samples are confirmed by SEM and TEM, as shown in Fig. 3. As shown in Fig. 3a and b, the TV SEM images show microsized clusters with obvious agglomerated particle-like nanoparticles. The grain sizes are around 30 nm for pure TV after annealing at 700 °C. In the case of GN–TV–C (Fig. 3C), particle-like TiO2–V2O5–C nanoparticles with a diameter of 20 nm anchored on the GN surface with significantly decreased aggregation tendency. The uniform carbon encapsulation and the flexible GN sheets helped in preventing the agglomeration of the primary crystallites of TV during calcination.
To unravel the elemental composition and the amount of distribution of elements present in TV–C core–shell heterostructures, energy dispersive X-ray spectroscopy (EDS) elemental mapping was conducted. The elemental maps of this nanocomposite also suggest a uniform distribution of nitrogen, carbon, titanium, vanadium, oxygen, which is consistent with the XRD and XPS results.
The mapping results (Fig. 4) of the selected rectangular area of FESEM image of GN–TV–C clearly depicted the coexistence of T, V, O, N and C elements in the heterostructures. It is noted that the distribution of all the elements is homogeneous and uniform, but the distribution of Ti and V exhibit a considerably larger area than the other elements, indicating that TiO2 cores are densely and uniformly encased by V2O5 layer, thereby resulting into core–shell–shell heterostructures.
Fig. 5 shows the typical TEM images of the TiO2 samples, as shown in Fig. 5a, and pure TiO2 nanoparticles consist of densely packed irregular nanoparticles, which is consistent with the SEM results with a high degree of crystallinity. As shown in Fig. 5b, The observed HRTEM image clearly reveals the lattice spacing of 0.32 nm, which corresponds to the (110) crystal planes of the rutile TiO2 and the {101} orientation of the V2O5 nanoparticles, respectively. The TV nanocrystals exhibit a polycrystal of the TiO2 (Fig. 5c).
As for GN–TV–C, the aggregation phenomenon occurred in the case of the TV particles, which is in agreement with the SEM result. Moreover, an excessive amount of carbon led to a thick carbon-coating layer (10–20 nm). The TEM images in Fig. 5d show the primary crystallites of the TV material with sizes ranged from 5–10 nm. TV had smaller primary crystallites (5–10 nm) than pure TV, which agrees with the XRD result. It is clear that many obvious bright regions appear in the sphere, revealing an underlying hierarchical porous structure in Fig. 5d, which complies well with the result of BET. The image also shows the GN wrapping around the core–shell–shell TiO2–V2O5–C nanostructure.
The particles become indiscernible and completely coated by a thick layer of carbon irrespective of the high crystalline of the TV core–shell particle (read by the XRD result) (Fig. 5d). There are many voids inner the carbon layer to alleviate the volume change during the lithium–delithium progress (inset Fig. 5d). This intermixing allows the TV and carbon matrix to provide especially effective mechanical reinforcement, accommodating the volume change of the active particles, while also trying to separate them and inhibit their agglomeration during extended cycling. Therefore, the reinforcing GN matrix and the elastic carbon layer guarantee the stability structure of GN–TV–C as intended. The brightness difference between the core and the shell in the magnified TEM image portrayed in the inset of Fig. 5d substantiates the existence of a core–shell configuration of the heterostructures.
By analyzing the interface of the core and shell in HRTEM taken from the selected rectangular region of GN–TV–C heterostructures (as marked in Fig. 5e), it is observed that the lattice fringes are continuous from core to the shell (Fig. 5e). Thus, TEM and HRTEM analysis, explicitly validate the core–shell morphology of TV-C heterostructures. As clearly seen from the TEM images (Fig. 5f), the polymeric morphologies were affected by thick carbon formation and the SAED in Fig. 5f showing a slightly clear six-fold pattern confirms that the crystal structure of the original graphite is retained in the exfoliated sheets and would improve by the hydrothermal reaction. This suggests that they are dispersed on the GN sheets with good contact, the intimate contact makes the interfacial charge transfer available at the interface of the composites. Thus, on the basis of the FESEM, FESEM, EDS elemental mapping, TEM, HRTEM and point-EDS analysis, it is safe to conclude the core–shell–shell morphology of TiO2–V2O5–C.
The GN–TV–C crystals were concluded to be formed during the dissolution and recrystallization of TiO2–V2O5 progress, which involved the Ostwald ripening process.23 As confirmed by the TEM images, the rutile TiO2 mesocrystals were formed through the oriented self-assembly process of rutile nanocrystal subunits elongated along the {110} direction under equilibrium conditions, while the V2O5 finally preserves its most highly active {101}. It is well-known that {110} and {101} are the most thermodynamically stable facets for rutile TiO2 and V2O5. A large amounts of oxygenated groups, hydroxyl, carbonyl and carboxyl groups can be induced on the glucose surface during the hydrothermal reation.24 Glucose can be expected to act as a role of surfactant,18 in order to verify the speculation, the surface tension of the aqueous solution was investigated. Consequently, the tension reduced from 68.96 mN m−1 without glucose to 98.73 mN m−1 after the addition of glucose. Therefore, the glucose indeed plays a predominant role as a surfactant during the configuration of the core–shell–shell of TiO2–V2O5–C, as a tension to impose on the pristine TV particles, and remains in a meta-stable phase in the aqueous form because of a higher surface energy of the inner core. The TV nanoparticles were dissolved from inside to outside intensely enclosed by the glucose macular. The rapid dissolution and slow expansion glucose–TV led to the formation of the porous carbon shell. Therefore, the TV particles were gradually encapsulated by the carbon microspheres because the surface tension of the sphere is the lowest, and the ripening process for the core–shell–shell TiO2–V2O5–C micronanostructures was completed. The glucose adsorbed encapsulating the active TV by interweaving the sides of GN at the same time could be competing with the face-growth inhibitors, slowing the growth of the involved face of TV and the production of mesoporous channels in the TV nanocrystals, which introduces favorable charge carrier and ion diffusion. The surface energy, which is responsible for the crystallization was involved during the whole progress of the crystals growth. The TV nanostructures appear in a uniformly distributed TiO2–V2O5–C morphology in the aqueous medium, whereas the nanocrystals incline to a homogeneously arranged porous core–shell–shell feature. Moreover, the nanocrystal subunits were partly recrystallized during the oriented self-assembly process. The well crystallized TiO2–V2O5–C densely grown on GN matrix with mesoporous nanostructure can facilitate the quick transfer of the photogenerated electrons from TiO2–V2O5–C to GN to cause the effective charge separation in the photocatalysis case, as well as provide the accommodation of volume change during the charge and discharge processes in LIBs application; thus, indicating its possible application as a visible-light driven photocatalyst and as an attractive anode material candidate for LIBs. Scheme 2 illustrates the entire progress of the reaction.
The band gap energy of semiconductors can be determined on the basis of the following formula:
α(hν) = C(hν − Eg)n/2 | (2) |
To further investigate the characteristics of GN–TV–C arrays, photoelectrochemical measurements were performed in a three electrode. Fig. 6c presents the time courses for the photocurrent of TV based materials under visible light irradiation at +0.2 V (vs. Hg/Hg2SO4). Obviously, the current abruptly increases and decreases as modulating the light source on and off. The incorporation of GN in the TV supply enhanced the photocurrent under visible-light irradiation, the core–shell–shell configuration of TiO2–V2O5–C due to the addition of glucose was most effective for improving the visible photoresponse. It is to be noted that the co-doping GN and carbon shell substantially improves the photocurrent of the final GN–TV–C. Moreover, the photocurrent of GN–TV–C (3% V2O5) is higher than that of GN–TV–C (5% V2O5), thicker TiO2–V2O5 contacting surface may become a new carrier and holes recombination center. This indicates that the amount and ratio of Ti and V are associated with the enhancement in photoactivity under visible light.
The result agreed well with the following PL measurement (Fig. 6d): PL is a predominant parameter to weigh the charge separation property of the photocatalysts. The higher PL intensity means a high rate of electron–hole recombination,27 in other words, the weaker intense the PL emission peak is the more it suppress the photocatalytic activity. In comparison with pristine TV and GN–TV, obviously, GN–TV–C (3% V2O5) exhibits the weakest intense PL peak than others from Fig. 6d, demonstrating that a minimum recombination of electron–holes in the TiO2 based sample, leading to the improvement of the photocatalytic activity, further the addition of V2O5 up to 5% leads to increased electron–hole recombination. It can be shown that hybrid TiO2 with GN can cause a clear diminution of the PL emission because the high conductive GN can trap photogenerated electrons through the interface to facilitate the charge separation to achieve a higher separation efficiency. Moreover, the accumulated holes on the surface of carbon layer participate in the oxygen evolution reaction, which guarantee the high oxygen production efficiency. This result was because of the fact that the electrons were excited from the valence band of TV to the conduction band, and then transferred to high conductive GN, thereby preventing the direct recombination of electrons and holes. The mitigated exciton recombination leads to an increased photocurrent.
![]() | ||
Fig. 7 The evolution of H2 and O2 gases of TV samples (a) stability of photocatalytic water for GN–TV–C nanocomposite under visible light irradiation (b). |
The photocatalysis mechanism is shown in Fig. 8. The photo-excited electrons would transfer from TV to FTO, and then to Pt, and the separated electrons on the surface of Pt would combine with adsorbed H+ to produce H2. Simultaneously, the accumulated holes on the surface of TV would firmly retained within the surface of carbon layer because of its positive band layer,26 which would oxidize the H2O to release O2. Compared with bare TV, the GN–TV–C heterostructures had prolonged the life time of separated exciton pairs, resulting in the improvement of the H2 and O2 production.
Cyclic voltammograms (CV) of GN–TV–C from the 1st to 3rd cycle at a scan rate of 0.5 mV s−1 in the voltage range of 1.0–3.0 V is illustrated in Fig. 9b. It can be observed that the intensity of the redox peaks increased during the second and third scan, implying a possible activating process in the electrode material. The second and third cycles with higher cathodic peaks and lower anodic peaks compared to the initial cycle are nearly overlapped, supporting the excellent electrochemical reactivity of modified TiO2 composites. The plots replication certifies the good reversibility of the lithium extraction–insertion reactions in terms of small polarization, which well complies with the aforementioned discussions. Fig. 9c–e describes the galvanostatic charge–discharge voltage profiles of cells for all TV samples at progressively increased C rates from 0.1 C to 30 C between 1.00 and 3.00 V vs. Li+/Li. It should be noted that the voltage plateau is lengthened for the GN and C modified TV materials, which should be ascribed to the higher electrochemical reactivity of modified TV and excellent kinetics. This result is consistent with the CV profiles. The modified TV materials exhibit higher capacities than TV at a low rate C/10, and a maximum capacity of 345.59 mA h g−1 is observed in GN–TV–C, while that of the GN–TV and TV reach 300.50, 298.58 mA h g−1, respectively. In addition, it is found that a serious polarization tendency at high C rates, while the modified samples have a slight polarization tendency. Fig. S3† compared the charge–discharge files of 3% V2O5 and 5% V2O5, it is observed that that the capacity of later decreased obviously, too much V2O5 can introduce more structure instability due to the irreversible phase transitions outweigh its contribution to the capacity of TV during charge–discharge progress.29
Fig. 9f shows the cycling performances at various current density ranging from 100 mA g−1 to 30 A g−1. It can be clearly shown that, as compared to the GN–TV and pure TV samples, the specific capacity of the GN–TV–C sample is substantially higher at all the investigated charge–discharge rates. For example, the GN–TV–C electrode displays a considerably superior rate performance of 303.69 and 293.81 mA h g−1,which is 1.3 times that of the GN–TV electrode (243.79 and 238.95 mA h g−1) and is almost 2 times of that for pure TV (201.41, 179.26 mA h g−1) at current densities of 20 and 30 A g −1, respectively. Recovering the current density back to 50 mA g−1 and 100 mA g−1, reversible capacities of up to their initial capacities, verifying the good reversibility and structural stability of the GN–TV–C anode composite and the capacity keeps ascending during subsequent cycling, approaching its initial capacity. The excellent rate capability arises from the 3D interconnected porous core–shell–shell carbon structure and the GN interweaved favorable channels, which not only offer multidimensional channels for the electrolyte access and sufficient void space to buffer the volume change of TV during cycling, but also effectively shorten the electronic–ionic transport path. In particular, the excellent mechanical stability of the carbon layer greatly confines the active TV particles in the core–shell nanostructure impeding the corrosion from electrolytes, and moreover, the good conductivity of the 3D interconnected GN matrix results in the long cycle ability of GN–TV–C. The reversible formation of flexible carbon buffering layer and elastic graphene sheets restrain the volume change of TV during the charge and discharge processes.
The heterostructure GN–TV–C exhibits an excellent rate capability compared to the GN–TV and TV samples; however, the difference in conductivity between them is be taken into account. Electrochemical impedance spectroscopy (EIS) measurements were performed to investigate the conductivity of the cathodes composed of TV based electrodes. As shown in Fig. 10a, the GN–TV–C electrode possesses a considerably lower resistance than that of the GN–C and TV electrodes (ca. 73.2 vs. 145.9 and 330.4 Ω, respectively), indicating that the encapsulation of core–shell–shell nanostructure and GN matrix can significantly enhance the electrochemical kinetics. The synergistic effect of the successful integration of the TV structure and interweaved GN matrix is beneficial to the structure stability during the photocatalysis and charge and discharge progresses. Ideal electrode materials are both good electronic and ionic conductors. The electrons and Li ions must reach or leave the reaction point simultaneously. However, for TV, it not only suffers from extremely low intrinsic electronic conductivity because of the lacking of effective conductive carbon layers on its surface. Thus, the larger particle size for pure TV may decrease the electronic conductivity in the bulks and result in the higher charge-transfer impedance Rct.30
![]() | ||
Fig. 10 Electrochemical impedance spectroscopy (EIS) result of the TV samples (a) and long-term stability for GN–TV–C material 30 Ag−1 (b). |
In particular, the lower electron transfer resistance can effectively accelerate the transfer mechanism of photogenerated charges in the hetero-structured samples, positively suppressing the combination of the photogenerated electron–hole pairs and further enhances the reactive activities of these samples, which is beneficial for the enhancement of photocatalytic performance and extending the sunlight harvest ability. Therefore, the synergistic morphology and composite engineering together with the particular mesoporous nanostructure introduce a promising candidate for both photocatalysis, as well as energy storage.
The robust nanostructure of GN–TV–C was further investigated (Fig. 10b), the morphology of GN–TV–C were examined after the 2000 times cycling test. Fig. S4† show that the TEM image of the core–shell structure of TiO2–V2O5–C is remained escaping from pulverization after 1000 times cycling, which is attractively pursued when designing anode materials. With the dual protections from GN and carbon layer, the severe volume changes of TV can be effectively accommodated. The carbon layer can also protect the TV electrode from electrolyte corrosion and maintain the structural stability of the TV; thus, resulting into a considerably improved rate and low temperature capability and cycle stability of the GN–TV–C, in good agreement with the aforementioned results.
Footnote |
† Electronic supplementary information (ESI) available. See DOI: 10.1039/c4ra08931f |
This journal is © The Royal Society of Chemistry 2014 |