Arumugam Venkatesanab,
Nagamuthu Raja Krishna Chandard,
Arumugam Kandasamye,
Madhu Karl Chinnuf,
Kalusalingam Nagappan Marimuthug,
Rangasamy Mohan Kumar*a and
Ramasamy Jayavelc
aDepartment of Physics, Presidency College, Chennai 600 005, India. E-mail: mohan66@hotmail.com
bDepartment of Physics, Panimalar Engineering College, Chennai 600 123, India
cCentre for Nanoscience and Technology, Anna University, Chennai 600 025, India
dCrystal Growth and Crystallography Division, School of Advanced Sciences, VIT University, Vellore 632 014, India
ePG Department of Physics, L. N. Government College, Ponneri 601 204, India
fNano Optoelectronics Lab, Department of Applied Physics, Tunghai University, Taichung, 40704, Taiwan
gDepartment of Chemistry, Saveetha Engineering College, Chennai 602 105, India
First published on 9th February 2015
Vanadium pentoxide nanostructures have been obtained from an alkoxide sol–gel, prepared by a simple and inexpensive facile non-aqueous method. The progressive addition of rare earth (RE) ions (Gd3+, Nd3+) to pristine V2O5 and the structural, functional, morphological, optical and electrochemical properties were studied. XPS studies confirmed the presence of RE ions in the orthorhombic phase of pristine V2O5, which was supported by XRD. The doping of RE ions significantly altered the morphology of V2O5 into various nanostructures by the linkage of small V2O5 nanoparticles. A significant red shift from undoped V2O5 was observed from UV absorption and PL spectra. From the CV experiment, it was observed that the overall cell potential was increased for the doped samples. The specific capacity of the Gd3+ and Nd3+ doped V2O5 increased upto 10%, which is useful for secondary Li-ion rechargeable batteries.
A schematic illustration of the experimental procedure for the synthesis is depicted in Fig. 1. The chemical reaction formulae are given in eqn (1)–(3) for undoped, Gd and Nd doped V2O5 respectively.
2VOCl3 + 3C6H5CH2OH → V2O5 + 3C6H5CH2Cl + 3HCl↑ | (1) |
![]() | (2) |
![]() | (3) |
![]() | ||
Fig. 1 Schematic of non-aqueous sol–gel route for the preparation of pristine and rare earths (Gd, Nd) doped V2O5 nanostructures. |
Bondarenka et al. have done a useful statistical analysis of XPS data on the binding energies of V 2p3 lines for vanadium in different oxidation states.25 The reported binding energies for V5+, V4+, and V3+ species were found to match with three Gaussian distributions centered at 517.3, 516.5 and 515.6 eV, respectively, with the statistical deviation of ±0.25 eV. The broad and intense peak observed between 140 and 150 eV is attributed to the 4d state of gadolinium as shown in Fig. 2e. The spin–orbit splitting of 4.25 eV occurred due to the doublet Gd 4d3/2 and Gd 4d5/2 at 148 and 143.75 eV respectively.
In the case of Nd, the Nd 3d core level showed complex and asymmetrical features between 970 and 1020 eV, corresponding to the 3d5/2 and 3d3/2 doublet (shown in Fig. 2f). The peak asymmetry originated from the final state effects (with Nd 4f level) as well as from overlapping with O KLL Auger lines that are observed in the same region. The BE for Nd 3d5/2 are 983.3 ± 0.2 eV (4f3) and 975.3 ± 2.3 eV (4f4), which is in agreement with the available data for Nd3+.26 Because of the significant error for the calculation of atomic ratios, due to the overlapping between Nd 3d and O KLL lines, the Nd 4d core level is used for quantification. It presents a more simple shape at a BE of about 122 eV. Another less intense peak observed at 1006.7 eV is ascribed to the core level spectra of Nd 3d3/2.
The typical XRD patterns for the pristine and RE (Gd & Nd) doped V2O5 nanomaterials are shown in Fig. 3. It is clearly observed that the synthesized samples exhibit a crystalline nature and the indexed planes are well matched with the standard JCPDS data (card no: 41-1426; space group: Pmmn). From the XRD patterns, no characteristic peaks of impurity phases were observed. It ensures that the structure of the host material is not modified by dopant elements. The most intense diffraction peaks of the doped samples (in Fig. 3b and c) show a slight shift towards a higher angle compared to that of the undoped V2O5 sample. This shifting in XRD lines of the doped V2O5 suggests that Gd3+ and Nd3+ have been successfully doped into the pure V2O5 host structure at the V5+ site and it is ascribed to the decrease in the values of ‘d’ spacing.27
The estimated lattice parameters are depicted in Table 1. When compared to the ‘d’ spacing of pristine V2O5, both the doped samples possess smaller ‘d’ spacing values for all the diffraction angles. This is purely due to the larger atomic radius of the dopant cations (Gd = 180 pm & Nd = 185 pm) than the host cation (V = 135 pm). Besides the larger atomic radius, the dopant cations are trivalent (Gd3+ and Nd3+). When the impurity ion has a larger radius than the radius of the host lattice, the lattice constant will decrease and the particle size will increase.28 But the host material is a pentavalent ion (V5+), so, it may be concluded that there is a reduction in the lattice fringes (‘d’ spacing) of the host material, caused by a deficiency of dopant ions to accumulate in the vacancy position of the oxygen ions in the pentavalent host lattice.29 The average crystallite sizes of the pristine and doped samples were estimated using Scherrer's formula and are listed in Table 1. These results substantiate that the decrement in lattice fringes caused the increment of particle sizes.
(hkl) planes | Pure V2O5 | Gd:V2O5 | Nd:V2O5 | |||
---|---|---|---|---|---|---|
2θ (deg) | ‘d’ spacing (Å) | 2θ (deg) | ‘d’ spacing (Å) | 2θ (deg) | ‘d’ spacing (Å) | |
200 | 15.46 | 5.729 | 15.53 | 5.701 | 15.56 | 5.692 |
001 | 20.32 | 4.367 | 20.40 | 4.349 | 20.42 | 4.346 |
110 | 26.21 | 3.397 | 26.28 | 3.389 | 26.31 | 3.385 |
400 | 31.08 | 2.875 | 31.15 | 2.869 | 31.18 | 2.866 |
310 | 34.38 | 2.606 | 34.44 | 2.602 | 34.48 | 2.599 |
a (Å) | 11.458 | 11.402 | 11.384 | |||
b (Å) | 3.566 | 3.569 | 3.567 | |||
c (Å) | 4.367 | 4.352 | 4.350 | |||
V (Å3) | 179.83 | 177.076 | 176.624 | |||
Avg. crystallite size ‘D’ (nm) | 16 | 19 | 27 |
V2O5 has orthorhombic symmetry and a layered structure: each vanadium atom is connected to five oxygen atoms to create pyramids that share their corners to build a double chain.30 The chains are connected along the edges to form layers that are then stacked to form the bulk structure. There are three structurally different oxygen atoms in each layer. One is coordinated to one vanadium atom, the second is found in a bridging position, and the third has a threefold-coordinated position. The binding of the atoms inside a layer is strong whereas the interactions that keep the layers stacked are weaker, resulting in V2O5 being easily cleaved (shown in Fig. 4a). When the dopant material with a higher ionic radii (Gd3+) intercalates into the interstitial layer of the V2O5 lattice it causes an increase in the distance between layers (shown in Fig. 4b).31,32
![]() | ||
Fig. 4 Crystal structure depicting (a) the orthorhombic phase of V2O5 (Pmmn) (b) insertion of Gd ions into the V2O5 layered structure. |
FTIR spectra of pristine and rare earth ion (Gd & Nd) doped V2O5 nanostructures are shown in Fig. 5. For all samples, a very broad and high intense peak observed at 3450–3350 cm−1 is attributed to the O–H (hydroxyl group) symmetric and bending vibrational modes. The position of the C–O stretch can be used to determine the type of alcohol. The peak observed around 1050 cm−1 confirms the presence of primary alcohols (M–CH2–OH), which is ascribed to the presence of the benzyl alcohol solvent, even after washing several times with ethanol. The terminal oxygen symmetric stretching mode of VO and the bridge oxygen asymmetric and symmetric stretching modes (νas and νs) of V–O–V were observed at 997, 759 and 553 cm−1, respectively.
The positions of the absorption bands in the RE doped V2O5 samples are similar to those observed in the undoped sample, which implies that the layer structure is not significantly altered by intercalation. The peak at 997 cm−1 in the high frequency region is due to vibration of VO atoms, which also gives information about the structural quality of the product.33 The vibration band, around 997 cm−1 for pristine V2O5 was shifted to 1014 and 1018 cm−1 in the Gd and Nd doped V2O5 spectra respectively.
The frequency shift to a higher wavenumber indicates the presence of impurity ions in the lattice, such as rare earth oxides (REO) in the vicinity of oxygen, and distortion of the VO bond.34 The absorption peaks around 1000 cm−1 are related to the vibration of isolated V
O vanadyl groups in [VO5] trigonal bipyramids.35 The νs(V–O–V) and νas(V–O–V) modes are shifted to a higher wavenumber. These chemical shifts could be related to the reduction of the oxidation state of vanadium from V5+ to V4+ and an increase of electronic conductivity.36 The shifting of triply coordinated oxygen (V3–O) from 553 cm−1 to a lower wavenumber in the vanadium oxide lattice may be caused by a distortion in the coordination geometry.37 In addition, some minor spectroscopic differences and displacements were observed, showing the physical interaction between trivalent dopant cations (Gd3+ & Nd3+) and V5+ ions.
The morphological evolution by doping was examined by high resolution scanning electron microscopy (HRSEM). HRSEM images of undoped V2O5, Gd:V2O5 and Nd:V2O5 are shown in Fig. 6a–c respectively. It can be seen that the pure sample consists of uniformly dispersed spherical particles within the nanoscale regime in the order of 10–15 nm (Fig. 6a). The Gd:V2O5 sample has a morphology with randomly oriented nanoflakes (Fig. 6b). The nanoparticle diameter was found to be about 20 ± 5 nm. The HRSEM image of the Nd:V2O5 sample showed a sugar cube-like, interlinked and porous nanostructure with particle size in the range of 20–25 nm. These suggest that the change in particle size is attributed to the doping with rare earth elements, which is consistent with the XPS and XRD results.
The SAED pattern of pristine V2O5 nanoparticles (Fig. 6j), illustrates the diffraction rings, which can be indexed to an orthorhombic structure. The SAED analysis of Gd and Nd doped vanadia samples indicates highly oriented nanocrystalline grains due to the ordered diffraction spots as compared to the X-ray diffraction results. The morphology and qualitative degree of dispersibility and agglomeration of the doped V2O5 nanoparticles were studied by TEM. Fig. 6g–i show the TEM overview images of pristine and Gd and Nd doped V2O5 nanostructures respectively. Improved dispersibility was achieved by using pure benzyl alcohol as the solvent. A slightly lower degree of agglomeration was observed in the case of Nd:V2O5. The particle sizes of all the samples are in accordance with the HRSEM results. Fig. 6d–f show the EDX spectra of pure and Gd and Nd doped V2O5 nanostructured materials. From Fig. 6d, it can be seen that the undoped V2O5 nanospheres contain only V and O. Apart from the vanadium and oxygen, an amount of gadolinium and neodymium was also present in the Gd and Nd doped samples respectively (Fig. 6e and f). HRTEM images reveal that the products are highly crystallized nanomaterials (Fig. 6m–o), and the lattice fringes corresponding to d spacings of 0.47 nm, 0.37 nm and 0.3 nm are in agreement with the d001, d110 and d110 spacing in the XRD pattern of undoped and Gd and Nd doped V2O5 samples. These results demonstrate that, the inorganic material synthesized through organic medium is a highly preferential way to achieve the products in the nanoscale regime for potential use in a variety of applications.38
UV-vis absorption spectra of pristine and RE doped V2O5 nanoparticles are shown in Fig. 7. The UV-vis absorption spectrum of pristine V2O5 shows a strong absorption in the UV region, which extends to the visible region, which is mainly associated to the charge transfer transition from the O 2p valence band to the empty V 3d orbitals. Also the UV-visible spectra of doped V2O5 nanoparticles exhibit a similar pattern to that of the V-oxide system and it is dominated by a broad envelope around 400 nm, ascribed to O2− → V5+ charge-transfer transitions and V4+ → V5+ intervalence transitions in the near infrared region.39 It was observed that the absorption edges of Gd and Nd doped V2O5 shift slightly towards the higher wavelength region compared to that of pristine V2O5, for which the samples were active under visible light of the solar spectrum.40 This red shift attributed to the charge transfer transition between the f electron of rare earth ions and the V2O5 conduction or valence band.41 The optical band gap was obtained by plotting the relation of (αhν)2 versus hν by using the relation,
αhν = A(hν − Eg)m |
![]() | ||
Fig. 7 UV-vis absorbance spectra of pristine (a) and Gd3+ (b) and Nd3+ (c) doped V2O5 nanoparticles. Tauc's plot for direct allowed band gap transition (inset figure). |
Fig. 8 shows the photoluminescence spectra of pristine and RE doped V2O5 nanoparticles. In the PL spectra, an excitation wavelength of 475 nm was used for all the samples.42–44 The emission peaks were observed at 502, 528 and 532 nm for pristine and Gd and Nd doped samples respectively.45 Peaks around 520 nm in VM oxides are assigned to the V
O double bonds.46 These results confirmed the shift in absorption edges towards the higher wavelength region (red shift), which is in good agreement with the UV-vis absorbance spectra. The total PL intensity decreased for doped samples, because of the increase in non-radiative traps produced by structural disorder. Furthermore, an absence of secondary emission peaks (defect peaks) in the PL spectra suggests that the dopant cations are successfully intercalated into the host lattice. The obtained results suggest that these materials are applicable to blue InGaN phosphor conversion.47
![]() | ||
Fig. 8 Room temperature photoluminescence spectra of pristine and Gd3+ and Nd3+ doped V2O5 nanoparticles. |
Initially, the green color of the as-prepared products suggests that the V5+ is partly reduced to V4+ during preparation in an organic environment. Later, it becomes yellow, due to the calcination at 400 °C. This is attributed to the oxidation during the calcination to V5+. The CV measurements were performed on the calcinated samples, using three electrode systems with a potential range from −2.0 to +2.0 V versus Li/Li+, in acetonitrile solution containing 0.1 M LiClO4 (Fig. 9). In this solution media, the electrochemical process can be attributed to the VV/VIV redox pair (yellow5+ ↔ green5+/4+ ↔ blue4+), with concomitant insertion and deinsertion of alkaline ions in order to maintain electroneutrality.48,49
xe− + xLi+ + V2O5 → LixV2O5 | (4) |
![]() | ||
Fig. 9 The typical cyclic voltammogram curves of (a) undoped V2O5 (b) Gd:V2O5 and (c) Nd:V2O5 with a scanning rate of 10 mV−1 between −2.0 and +2.0 V. |
Fig. 9 compares the cyclic voltammograms (20th cycle) of pure and Gd and Nd doped V2O5 nanostructured materials. All the prepared samples exhibited three well-defined reduction peaks at ∼−0.40, 0.20 and 0.63 V corresponding to the formation of ε, δ and γ phases, respectively.50 However, two peaks appeared at ∼0.62 and 1.40 V in the anodic region. This indicates that the complex structural changes occurred in the cathode material during the oxidation process. The slight shift in the cathodic and anodic peaks depends upon the dopant content. The redox current is more for the doped samples than the pristine V2O5. Moreover, the cathodic/anodic peak shifts to a more positive voltage range, which leads to an increase of the overall average cell potential compared with pristine V2O5.
Fig. 10 illustrates the 25 complete galvanostatic discharge cycles of the battery. When the batteries were cycled between −2.0 to +2.0 V at a constant current density of 6 mA g−1, the maximum discharge capacity of undoped V2O5 was 284 mA h g−1 and 188 mA h g−1 after 25 cycles, and the initial discharge capacities of Gd and Nd doped V2O5 were 302 mA h g−1 and 305 mA h g−1 respectively; and 247 mA h g−1 and 251 mA h g−1 after 25 cycles respectively. It was observed that the cyclic performance of the doped samples is obviously more stable after the 7th cycle and the specific capacity is higher than that of the pristine sample, which is in agreement with the cyclic voltammetry results. These results are due to the increased active area and the flexible structure of doped samples, which can provide more Li+ ion intercalation sites and accommodate a large volume variation, supported by the XRD results.51 The specific capacity of the doped samples increased upto 10% and meanwhile, the capacity loss decreased upto 16% from the pristine V2O5 nanoparticles. The successful incorporation of rare earth ions within the host lattices also enhanced the electronic conductivity and reduced the resistance within the nanostructures, which facilitates the faster mobility of Li+ ion in the cathode material.
![]() | ||
Fig. 10 Discharge curves of (a) undoped V2O5 (b) Gd:V2O5 and (c) Nd:V2O5 at a constant current density 6 mA g−1 when sweeps between −2.0 and +2.0 V. |
This journal is © The Royal Society of Chemistry 2015 |