Electric, magnetic and optical limiting (short pulse and ultrafast) studies in phase pure (1 − x)BiFeO3xNaNbO3 multiferroic nanocomposite synthesized by the pechini method

Rehana P. Ummera, P. Sreekanthd, B. Raneeshb, Reji Philipd, Didier Rouxele, Sabu Thomasbc and Nandakumar Kalarikkal*ab
aSchool of Pure and Applied Physics, Mahatma Gandhi University, Kottayam, Kerala-686560, India. E-mail: nkkalarikkal@mgu.ac.in; Fax: +91-481-2731669; Tel: +91-944-7671962
bInternational and Inter University Centre for Nanoscience and Nanotechnology, Mahatma Gandhi University, Kottayam, Kerala-686560, India
cSchool of Chemical Sciences, Mahatma Gandhi University, Kottayam, Kerala-686560, India
dUltrafast and Nonlinear Optics Lab, Light and Matter Physics Group, Raman Research Institute, Bangalore, 560080, India
eInstitut Jean Lamour-UMR CNRS 7198, Faculté des Sciences et Techniques, Campus Victor Grignard-BP 70239, 54506, Vandoeuvre-les-Nancy Cedex, France

Received 2nd June 2015 , Accepted 30th July 2015

First published on 30th July 2015


Abstract

The perovskite (1 − x)BiFeO3xNaNbO3 nanocomposite was successfully synthesized by Pechini method and crystallographic information was obtained from XRD and TEM analysis. Structural analysis using XRD and TEM shows that phase pure samples were obtained with reduced particle size. The observations reveal that particle size plays a crucial role in deciding the electric and magnetic properties. The multiferroic character of nanoparticles is confirmed through magneto electric (ME) coupling studies. The good coexistence of ferroelectric and ferromagnetic behaviors in the composite provides the possibility to achieve a measurable ME effect. The highest value of the magneto electric coefficient (α) is observed for x = 0.1 (α = 0.13 V cm−1 Oe−1) and it reduces for higher x values. Magnetization measurements for x = 0.1 show a small hysteresis at 6 K which also confirms the presence of the magnetic phase at low temperature while the usual antiferromagnetic behavior of BFO is found at 300 K. The reduced size and absence of impurity could be the reason for the enhancement in electrical properties. Low loss tangent values observed in the samples display a remarkable improvement. Open aperture Z-scan measurements reveal a nonlinear absorption behavior, which results in good optical limiting when excited with short pulse (nanosecond) as well as ultrafast (femtosecond) laser pulses.


1. Introduction

In recent years, the study of multiferroic materials is a time honored research subject due to their potential applications for future technologies in information storage, sensors, optical limiters1–5 and for exploring physical phenomena of the coupling mechanism between electric and magnetic order parameters.6–9 The combination of more than one ferroic property in a single material is also interesting and fascinating. There are few materials that can exhibit both magnetization and electric polarization together. Among them BiFeO3 (BFO) is the most extensively studied compound.10 It is one of the well known single phase multiferroic material with distorted perovskite (ABO3) structure. The perovskite BiFeO3 possesses a ferroelectric transition Tc ∼ 830 °C and antiferromagnetic transition at TN ∼ 370 °C. BFO is unique amongst many multiferroics, as its ferroelectric and magnetic transition temperatures are well above the room temperature raising the possibility of potential devices based on magneto electric (M-E) coupling operating at room temperature.7 The BFO also has other exceptional applications such as photo catalytic, ultrafast, optoelectronic and infrared detector devices due to its small band gap.11,12 In spite of the above qualities, this material has some inherent problems that limit its applications. For example, BFO has high leakage current that allows current to pass through it when a high voltage is applied. It has high dielectric loss and low polarization. In order to solve these inherent problems, several attempts have been made in the recent past, such as doping it with rare earth elements like samarium, lanthanum etc. at the appropriate atomic sites and/or fabricating its composites. NaNbO3 is one of the representative lead-free piezoelectric perovskite oxides which have large practical importance and composite with BFO is of considerable interest.13 Multiphase multiferroics are successful approach to overcome the limitations of single phase materials such as the small magneto electric coupling.2,4,13–17 The composites of BFO made with NaNbO3 showed reduction in the leakage current and enhancement of electric properties.18 A lot of attempts have been concentrated on the research and development of high performance lead-free piezoelectric ceramics such as alkali niobates. NaNbO3 is one among them which give rise to a set of phase transitions.19,20 NaNbO3 is highly soluble in any solid solution of perovskite materials.21

We have prepared solid solutions of BFO and NaNbO3 in different composition with a general formula (Bi1−xNax) (Fe1−xNbx)O3 where x = 0, 0.1, 0.5 and 0.7 using sol–gel reaction. In this paper, we detail the electric, magnetic and optical study of the as prepared samples. As per our knowledge there is no previous work reported on the physical properties of nano structured BiFeO3–NaNbO3 composite.25 Nonlinear study is previously observed on bismuth ferrite by Monte Carlo simulation and density functional theory.27 Also bismuth based glasses exhibit large third order optical nonlinearity in oxide glasses, indicating they are promising materials for nonlinear optical devices.23

2. Materials and methods

The solid solution ceramics of Bi1−xNaxFe1−xNbxO3 with x = 0, 0.1, 0.5, and 0.7 were prepared by modified sol–gel method called Pechini method.24,26 AR grade of ammonium niobate oxalate (Sigma-Aldrich > 99% pure), Bi(NO3)3 (Sigma-Aldrich > 99% pure), Fe(NO3)3 (Sigma-Aldrich > 99% pure), NaNbO3 (Sigma-Aldrich > 99% pure) were used as raw materials. Firstly, the raw materials were carefully weighted in stoichiometric ratio and dissolved in citric acid aqueous solution (in 1[thin space (1/6-em)]:[thin space (1/6-em)]1 molar ratio with respect to the total metal cation) and pH value was adjusted to 5 using NH3OH. The clear solution thus obtained was dried at 100 °C to form gel and the gel was burnt at 500 °C to get the ceramic powders. After that the powder was ground and pressed into pellets. Finally the pellets were sintered at 850 °C for 1 hour to get the final sample. The crystal structures of the samples were examined by Phillips X'Pert Pro XRD with Cu-Kα radiation (1.54056 Å). Step scanned powder XRD data was collected in the 2θ range 10°–80° at room temperature on the finely ground sample. Detailed structural analysis was performed using Scanning Electron Microscope (JEOL JSM 6390) and Transmission Electron Microscope (JEOL JEM 2100). The relative dielectric constant ε, and dissipation factor tan[thin space (1/6-em)]δ at room temperature were measured using an LCR meter (Agilent E4980A). The magnetoelectric coupling was determined using a lock in amplifier technique and magnetization measurements were performed using SQUID magnetometer. The absorption spectrum was taken using a UV-Vis spectrophotometer (SHIMADZU, UV-2450). The nonlinear optical measurements were carried out using the conventional open aperture Z-scan technique. The powder samples suspended in DMF by sonication were taken in a 1 mm cuvette. For excitation, 5 ns laser pulses were obtained from a frequency doubled Nd:YAG laser (Minilite Continuum, 532 nm) and 100 fs pulses were obtained from a regeneratively amplified mode-locked Ti-Sapphire laser (Spectra physics, 800 nm). In the Z-scan set up, the laser beam is focused using a lens and the sample is translated along the beam axis (z axis) through the focus (z = 0). At each position ‘z’ the sample sees a different laser fluence, and the transmission is measured using a pyroelectric laser energy probe which is placed after the sample. Inorder to monitor the pulse-to-pulse laser energy variation we used a reference beam picked off from the main beam. All measurements were done in the single-shot mode and there is an interval of about 10 seconds between successive laser pulses. Using the data obtained from Z-scan measurements, nonlinear optical parameters could be calculated by numerically fitting the data points using the appropriate nonlinear transmission equations.

3. Results and discussion

3.1 Structural analysis

Fig. 1 shows the XRD pattern of (1 − x)BiFeO3xNaNbO3 composite system for x = 0, 0.1, 0.5 and 0.7 at room temperature. The peaks that correspond to BiFeO3 and NaNbO3 are indicated by B and N respectively. XRD data reveals that both the phases of BiFeO3 and NaNbO3 are present in the composite system. Non-perovskite phases such as Bi2Fe4O9 and Bi2O3/Fe2O3 were not detected in the XRD spectrum which is a common occurrence when we adopt other techniques for the preparation. In the XRD spectra it could be observed that as x increases the peaks corresponding to orthorhombic NaNbO3 appear. Also, as the content of NaNbO3 is increased from x = 0.1 to x = 0.7, the diffraction peaks corresponding to NaNbO3 phase increases remarkably and the diffraction peaks corresponding to BiFeO3 phase decreases gradually. The obvious peak splitting in the XRD spectrum shows the rhombohedral structure of nanoparticles, consistent with the structure of BFO ceramics.23,27,28 The crystal structure changes from rhombohedral to orthorhombic by increasing the amount of NaNbO3 phase in the composite. The crystallite size was calculated from the full width at half maximum (FWHM) of the diffraction peaks using the Scherrer's equation. The calculated values of particle size are 6.2 nm, 10.9 nm, 41.7 nm and 79.2 nm for x = 0, x = 0.1, x = 0.5 and x = 0.7 respectively.
image file: c5ra10422j-f1.tif
Fig. 1 XRD patterns of (1 − x)BiFeO3xNaNbO3 samples for x = 0, 0.1, 0.5 and 0.7.

A typical TEM image of the BiFeO3–NaNbO3 samples prepared by the present sol–gel method is shown in Fig. 2. HRTEM and SAED patterns of x = 0.1, 0.5 and 0.7 are shown in Fig. 3. Both HRTEM and SAED pattern confirm the polycrystalline nature of the samples. The average particle size estimated from HRTEM images came out to be 6 nm, 10 nm, 43 nm and 80 nm for x = 0, 0.1, 0.5 and 0.7 samples respectively and is consistent with the XRD analysis. The lattice spacing (d) calculated from the HRTEM image (Fig. 3(a)) matches with JCPDS values corresponding to BiFeO3 and NaNbO3. The d value 0.23 and 0.34 corresponds to (2,2,1) and (1,2,1) planes of orthorhombic NaNbO3 (JCPDS 89-8957). Similarly d = 0.39 corresponds to (1,0,0) plane of rhombohedral BiFeO3 (JCPDS 74-2016). The SAED patterns show the presence of sharp diffraction spots, which is a clear indication of well developed, crystalline nanoparticles. The diffuse diffraction spots indicate the nanosize of the synthesized material.


image file: c5ra10422j-f2.tif
Fig. 2 TEM images of (1 − x)BiFeO3xNaNbO3 samples (a) for x = 0, (b) for x = 0.1, (c) for x = 0.5 and (d) for x = 0.7.

image file: c5ra10422j-f3.tif
Fig. 3 HRTEM images of (1 − x)BiFeO3xNaNbO3 samples for (a) x = 0.1, (b) x = 0.5, (c) x = 0. 7. Inset shows corresponding SAED patterns.

3.2 Electrical studies

The frequency dependence of dielectric constant for (1 − x)BiFeO3xNaNbO3 (x = 0, 0.1, 0.5, 0.7) at room temperature in the frequency range 100 Hz to 2 MHz is shown in Fig. 4. The inset shows the dielectric loss (tan[thin space (1/6-em)]δ) with frequency. The variation of dielectric constant with frequency is very much consistent with that of other compounds/composites. It is evident that the dielectric constant and loss tangent decreases with increasing frequency. In the starting low frequency range both ε and tan[thin space (1/6-em)]δ have higher values.
image file: c5ra10422j-f4.tif
Fig. 4 The dielectric spectra of (1 − x)BiFeO3xNaNbO3 samples (a) for x = 0, (b) for x = 0.1, (c) for x = 0.5 and (d) for x = 0.7. (Inset shows the dielectric loss as a function of frequency).

The ε and tan[thin space (1/6-em)]δ values decreases gradually as frequency increased from 100 Hz to 10 KHz and then decreases slowly and become almost constant up to 2 MHz for all compositions. The observations may be explained by the phenomenon of dipole relaxation.29 The dielectric constant (ε) of all the samples was found to decrease with increase of frequency in the low frequency region. This phenomenon can be attributed to the space charge relaxation effect. At low frequencies, the space charges are able to follow the frequency of the applied field, whereas these space charges do not find time to undergo relaxation at high frequency region.28 At high frequency electric dipoles are unable to switch with frequency of the applied field. This is a general feature of dielectric materials.22 It has been observed that at low frequency region, the dielectric constant is found to be dependent on different kinds of polarization (electronic, atomic, interfacial and ionic) whereas at high frequency region, only electronic polarization mainly contributes for the dielectric constant.30 So there is a sharp decrease in dielectric constant at high frequency.31 The values of the dielectric loss have been found to be very low at high frequencies and may find applications of these materials in high frequency microwave devices. The decrease of tan[thin space (1/6-em)]δ with increase in frequency can be explained on the basis of Koop's phenomenological theory.32,33 The low losses may be attributed to the nanosized grains.34,35 The low frequency dielectric dispersion increased with increase in NaNbO3 concentration and it was maximum for x = 0.7. It should be noted that the dielectric constant is found to increase with increasing NaNbO3 content. The dielectric constant is found to be increasing with decrease in particle size because of the presence of nanosized grains which act as a large insulating barrier for mobile charge carriers.36 The dielectric constants of sol gel synthesized samples are very much higher as compared to solid state synthesized samples. It is clearly observed that sol–gel synthesized samples have very high value of dielectric constant in low frequency region in comparison with the previous reports on other rare earth doped BiFeO3.36–38

AC conductivity studies were carried out for a better understanding of the frequency dependence of electrical property of the materials. Fig. 5 shows variation of AC conductivity as a function of frequency. The conductivity σac was calculated using the dielectric data and empirical relation.

 
σac = ωε0εrtan[thin space (1/6-em)]δ (1)
where ε0 is the permittivity of free space and ω is the angular frequency. The conductivity curve shows that σac increases with frequency which is a common feature of semiconductors.36 The increasing trend of σac with frequency in the low frequency region might be attributed to the disordering of cations between neighboring sites and presence of space charges.39


image file: c5ra10422j-f5.tif
Fig. 5 AC conductivity of (1 − x)BiFeO3xNaNbO3 samples (a) for x = 0, (b) for x = 0.1, (c) for x = 0.5 and (d) for x = 0.7.

3.3 Magnetic studies

The variations of ME voltage with AC and DC magnetic fields are plotted in Fig. 6(a) and (b) respectively. The ME output voltage measured up to 100 Oe (for Hac) shows linear increase with increasing magnetic field. When a magnetic field is applied to a magnetoelectric material, the material is strained. This strain induces a stress on the piezoelectric, which generates the electric field. This field could orient the ferroelectric domains, leading to an increase in polarization value. The magnetoelectric effect in multiferroics is fully described by the magnetoelectric coupling coefficient. From the slope of the graph, coupling coefficient (α) is determined and it is found to be decreasing with increasing the NaNbO3 content. The high value of ME coefficient, α = 0.13 V cm−1 Oe−1 (for x = 0.1) indicates the coexistence of electric and magnetic phases in this composition. To understand the magnetic behavior, the magnetization (M) versus applied field (H) is measured for x = 0.1 at two different temperatures (Fig. 6(c)). For bulk BFO the hysteresis loop is generally observed to be linear, indicating antiferromagnetic ordering of spins at the ground state. But here the hysteresis loop is similar to a typical weak ferromagnetic ‘S’ shaped curve. The small hysteresis loop observed at 6 K represents a soft magnetic phase with a low coercive field. It is found that the saturation is attained within the field of 40 kOe. The magnetic behavior at low temperature (6 K) can be due to reduced particle size and absence of impurity. Also, numerous magnetization studies of antiferromagnetic nanoparticles have shown that the magnetization in large applied fields is considerably larger than that in the corresponding bulk materials. It was suggested by Neel that this might be due to the finite number of magnetic moments in nanoparticles, which may lead to a difference in the number of spin in the two sub lattices because of random occupancy of lattice sites.40 This results in an uncompensated magnetic moment that leads to enhanced magnetic properties. At 300 K the hysteresis loop shows linear behavior which represents AFM ordering of spins characteristics of bulk BFO.10,36
image file: c5ra10422j-f6.tif
Fig. 6 (a) ME voltage as function of AC magnetic field, (b) ME voltage as a function of DC magnetic field, (c) magnetization (M) versus magnetic field (H) plot of x = 0.1 composition at 6 K and 300 K (inset shows a zoomed view of the central portion).

3.4 Linear optical studies

The UV-Visible absorption spectra of composites are shown in Fig. 7(a). The absorption cut-off wavelength of the as prepared composite samples lies between 500–600 nm which is close to the reported value for pure BFO (ie 560 nm),23 suggesting that the present material can absorb visible light in the wavelength range of 400–565 nm. The UV-Visible absorption spectroscopy is frequently used to determine the energy bandgap of the powder samples from their absorption spectra. The optical band gap was determined from the Tauc plot (Fig. 7(b)).42 The tangent line, which is extrapolated to ()2 = 0, gives the bandgap (Eg).42,43 The band gap values for each composite is shown in Table 1 which is in agreement with values from previous reports.41 Pure NaNbO3 have reported Eg value of 3.3–3.4 eV.36 Incorporation of NaNbO3 increases the band gap energy of the composite which is probably due to the interaction between BiFeO3 and NaNbO3.
image file: c5ra10422j-f7.tif
Fig. 7 (a) UV-Vis diffuse absorption spectra of the BiFeO3–NaNbO3 nanoparticles, where the dotted line is the division between UV and visible light. (b) Plot of ()2 versus photon energy (E).
Table 1 Bandgap energy of the samples
Sl. no Sample name Bandgap energy (eV)
1 X = 0 1.7
2 X = 0.1 1.9
3 X = 0.5 2.3
4 X = 0.7 2.8


3.5 Nonlinear optical studies

The open aperture Z-scan curves (inset) and the corresponding intensity dependent transmissions calculated from the Z-scan curves, measured for ultrafast laser pulse excitation (800 nm, 100 fs) in the BiFeO3–NaNbO3 composites (for x = 0, 0.1, 0.5, 0.7) are shown in Fig. 8, whereas those for short pulse excitation (532 nm, 5 ns) are shown in Fig. 9. The linear transmission is adjusted to be about 65% at both excitation wavelengths, and experiments were carried out at an average energy of 40 micro joules for nanosecond and 11 micro Joule for femtosecond excitations.
image file: c5ra10422j-f8.tif
Fig. 8 Open aperture Z-scan curves (inset) and the corresponding intensity dependent normalized transmission obtained in (1 − x)BiFeO3xNaNbO3 (for x = 0, 0.1, 0.5 and 0.7) for ultrafast (100 fs) pulse excitation. Open circles represents experimental data while the solid line represents numerical fit to eqn (3).

image file: c5ra10422j-f9.tif
Fig. 9 Open aperture Z-scan curves (inset) and corresponding intensity dependent normalized transmission obtained in (1 − x)BiFeO3xNaNbO3 composites (for x = 0, 0.1, 0.5 and 0.7) for 5 ns pulse excitation. Open circles represent experimental data while the solid line represents numerical fit to eqn (3).

The open aperture Z-scan curves obtained in both cases exhibit a smooth valley indicating typical reverse saturable absorption behavior. In order to find the nature and strength of nonlinear absorption we numerically fitted the measured data to different nonlinear transmission equations. The best fit was obtained for a model in which 2 PA (for fs excitation) or effective 2 PA (for ns excitation) occurred along with SA.44

In such a case the nonlinear absorption coefficient can be written as

 
α(I) = [α0/(1 + (I/Is))] + βI (2)
and the corresponding transmission equation is given by
 
dI/dz′ = −[(α0/1 + (I/Is)) + βI]I (3)
This equation can be numerically solved to obtain the best fit values of saturation intensity (Is) and two photon absorption coefficient (β).

The calculated nonlinear parameters for short-pulse and ultrafast laser pulse excitations are presented in Table 2. The optical limiting performance is quantified using the optical limiting threshold value, which is defined as the input fluence at which the sample transmission drops to 50% of its linear transmission.

Table 2 Calculated values of nonlinear optical parameters
Sl. no Sample name 5 ns excitation 100 fs excitation
β (×10−10 m w−1) Is (×1011 w m−2) Optical limiting threshold (J cm−2) β (×10−14 m w−1) Is (×1016 w m−2) Optical limiting threshold (J cm−2)
1 X = 0 1.8 7.99 1.1 1.1 2.0 1.7
2 X = 0.1 1.2 5.499 1.4 1.1 3.0 1.7
3 X = 0.5 1.0 2.499 1.9 1.0 3.4 1.8
4 X = 0.7 0.9 2.499 2.2 1.0 3.6 1.9


Z-scan experiments with relatively long (5 ns) laser pulses allow for multiple absorption from the same laser pulse, which greatly enhances excited state absorption (ESA). While the incorporation of NaNbO3 into BiFeO3 provides no significant modification to the ultrafast optical nonlinearity of BiFeO3, it obviously affects the nanosecond optical nonlinearity (Fig. 9). In this case nonlinear absorption is found to decrease with increase in the NaNbO3 concentration. This quenching of nonlinear optical absorption in the composite originates from factors associated with size reduction such as increase in grain boundaries and increase in the energy gap by increasing the NaNbO3 content. Pure BiFeO3 has an Eg of 1.72 eV, but the incorporation of NaNbO3 into BiFeO3 results in an increase in band gap energy to 2.53 eV when x = 0.7.

4. Conclusion

Phase pure BiFeO3–NaNbO3 nanocomposite samples were successfully synthesized using Pechini method. XRD and TEM analysis show that all the particles are in the nanometer size and no secondary phases were formed during the synthesis. The reduced size and absence of impurity could be the reason for high dielectric constant and good ME coupling. For the selected composition (x = 0.1) improved magnetization is found in the hysteresis loop and we obtained reduced Eg values for the samples. The reduced band gap and reduction in particle size is found as the reason for enhanced nonlinear optical properties of the present set of samples compared to their bulk form. From exciting the samples using nanosecond and femtosecond laser pulses we conclude that 2 PA and effective 2 PA are the dominant absorptive nonlinearities in the present set of samples. These samples exhibit efficient optical limiting and can have potential applications in photonic devices.

Acknowledgements

The authors would like to acknowledge the financial support from DST – Govt. of India through the Nano Mission, PURSE, FIST Programs, and UGC–Govt. of India for the SAP program.

References

  1. R. Yao, C. Bao Cao, C. Zheng and Q. Lei, RSC Adv., 2013, 3, 24231–24236 RSC.
  2. B. Raneesh, A. Saha, D. Das, R. Sreekanth, R. Philip and N. Kalarikkal, RSC Adv., 2015, 5, 12480 RSC.
  3. J. Hemberger, P. Lunkenheimer, R. Fichtl, H. A. Nidda, V. Tsurkan and A. Loidl, Nature, 2005, 434, 364 CrossRef CAS PubMed.
  4. J. Job Thomas, S. Krishnan, K. Sridharan, R. Philip and N. Kalarikkal, Mater. Res. Bull., 2012, 47, 1855–1860 CrossRef PubMed.
  5. J. Job Thomas, A. B. Shinde, P. S. R. Krishna and N. Kalarikkal, J. Alloys Compd., 2013, 546, 77–83 CrossRef PubMed.
  6. J. Hemberger, P. Lunkenheimer, R. Fichtl, H. A. Nidda, V. Tsurkan and A. Loidl, Nature, 2005, 434, 364 CrossRef CAS PubMed.
  7. N. A. Hill, J. Phys. Chem. B, 2000, 104, 6694 CrossRef CAS.
  8. W. Eerenstein, N. D. Mathur and J. F. Scott, Nature, 2006, 442, 759 CrossRef CAS PubMed.
  9. J. H. Xu, H. Ke, D. C. Jia, W. Wang and Y. Zhou, J. Alloys Compd., 2009, 472, 473–477 CrossRef CAS PubMed.
  10. V. F. Freitas, O. A. Protzek, L. A. Montoro, A. M. Gonçalves, D. Garcia, J. A. Eiras, R. Guo, A. S. Bhalla, L. F. Coticaad and I. A. Santos, J. Mater. Chem. C, 2014, 2, 364 RSC.
  11. S. R. Basu, L. W. Martin, Y. H. Chu, M. Gajek, R. Ramesh, R. C. Rai, X. Xu and J. L. Musfeld, Appl. Phys. Lett., 2008, 92, 091905 CrossRef PubMed.
  12. A. J. Hauser, J. Zhang, L. Mier, R. A. Ricciardo, P. M. Woodward, T. L. Gustafson, L. J. Brillson and F. Y. Yang, Appl. Phys. Lett., 2008, 92, 222901 CrossRef PubMed.
  13. G. Li, Z. Yi, Y. Bai, W. Zhang and H. Zhang, Dalton Trans., 2012, 41, 10194–10198 RSC.
  14. A. Perejon, P. E. Sanchez-Jimenez and L. A. Perez-Maqueda, J. Mater. Chem. C, 2014, 2, 8398 RSC.
  15. B. Raneesh, H. Soumya, J. Philip, S. Thomas and K. Nandakumar, J. Alloys Compd., 2014, 611, 381–385 CrossRef CAS PubMed.
  16. B. Y. Wang, H. T. Wang, S. B. Singh and Y. C. Shao, RSC Adv., 2013, 3, 7884–7893 RSC.
  17. C. A. F. Vaz, J. Hoffman, C. H. Ahn and R. Ramesh, Adv. Mater., 2010, 22, 2900–2918 CrossRef CAS PubMed.
  18. Y. Ma and X. M. Chen, J. Appl. Phys., 2009, 105, 054107 CrossRef PubMed.
  19. E. Hollenstein, M. Davis, D. Damjanovic and N. Setter, Appl. Phys. Lett., 2005, 87, 182905 CrossRef PubMed.
  20. G. A. Smolenski, V. A. Bokov, V. A. Isupov, N. N. Krainink, R. E. Pasynkov and A. I. Sokolov, Ferroelectric and related materials, Gordon and Breach Sci. Publishers, New York, 1984 Search PubMed.
  21. C. Ching Wua and C. Fu Yang, CrystEngComm, 2013, 15, 9097–9103 RSC.
  22. H. O. Rodrigues, G. F. M. Pires Jr, J. S. Almeida, E. O. Sancho, A. C. Ferreira, M. A. S. Silva and A. S. B. Sombra, J. Phys. Chem. Solids, 2010, 71, 1329 CrossRef CAS PubMed.
  23. S. T. Zhang, M. H. Lu, D. Wu, Y. F. Chen and N. B. Ming, Appl. Phys. Lett., 2005, 87(26), 2907 Search PubMed.
  24. S. Krishnan and N. Kalarikkal, J. Sol-Gel Sci. Technol., 2013, 66, 6–14 CrossRef CAS.
  25. I. P. Raevski, S. P. Kubrin, J. L. Dellis, S. I. Raevskaya, D. A. Sarychev and V. G. Smotrakov, Ferroelectrics, 2008, 371, 113–118 CrossRef CAS PubMed.
  26. S. X. Huo, S. L. Yuan, Y. Qiu, Z. Z. Ma and C. H. Wang, J. Mater. Lett., 2012, 68, 8–10 CrossRef CAS PubMed.
  27. S. Ju, T. Yi Cai and G. Yu Guo, J. Chem. Phys., 2009, 130, 214708 CrossRef PubMed.
  28. M. Kumar and K. L. Yadav, J. Phys.: Condens. Matter, 2007, 19, 242202 CrossRef PubMed.
  29. G. Li, T. Kako, D. Wang, Z. Zou and J. Ye, J. Phys. Chem. Solids, 2008, 69, 2487 CrossRef CAS PubMed.
  30. J. C. Anderson, Dielectrics, Chapman & Hall, 1964, London, vol. 1 Search PubMed.
  31. D. R. Patil, S. A. Lokare, R. S. Devan, S. S. Chougule, C. M. Kanamadi, Y. D. Kokekar and B. K. Chougule, Mater. Chem. Phys., 2007, 104, 254 CrossRef CAS PubMed.
  32. B. Kumari, P. R. Mandal and T. K. Nath, Adv. Mater. Lett., 2014, 5, 2 Search PubMed.
  33. S. R. Kulkarni, J. Phys. Chem. Solids, 2006, 67, 1607 CrossRef CAS PubMed.
  34. C. G. Koops, Phys. Rev., 1951, 83, 121 CrossRef CAS.
  35. Qian, F. Z. Jiang, J. S. Guo, S. Z. Jiang and D. M. Zhang, J. Appl. Phys., 2009, 106, 084312 CrossRef PubMed.
  36. Y. Du, Z. X. Cheng, M. Shahbazi, E. W. Collings, S. X. Dou and X. L. Wang, J. Alloys Compd., 2010, 490, 637 CrossRef CAS PubMed.
  37. G. Dhir, P. Uniyal and N. K. Verma, J. Supercond., 2014, 27, 1569–1577 CrossRef CAS.
  38. K. Sen, K. Singh, A. Gautam and M. Singh, Ceram. Int., 2012, 38, 243 CrossRef CAS PubMed.
  39. A. K. Jonscher, Dielectric Relaxation in Solids, Chelsca Dielectric Press Limited, London, 1986 Search PubMed.
  40. L. Neel, Magnetic nanoparticle assemblies, CBC press, Taylor and Francis group, 1961, vol. 1, ch. 2, p. 253 Search PubMed.
  41. B. Bhushan, A. Basumallick, S. K. Bandopadhyay, N. Y. Vasanthacharya and D. Das, J. Phys. D: Appl. Phys., 2009, 42, 065004 CrossRef.
  42. J. Tauc, Amorphous and Liquid Semiconductors, Plenum Press, New York, 1974, vol. 1, ch. 4, p. 239 Search PubMed.
  43. Y. Kim, S. J. Atherton, E. S. Brigham and T. E. Mallouk, J. Phys. Chem., 1993, 97, 11802 CrossRef CAS.
  44. B. Karthikeyan, M. Anija, P. Venkatesan, C. S. Suchand Sandeep and R. Philip, Opt. Commun., 2007, 280, 482 CrossRef CAS PubMed.

This journal is © The Royal Society of Chemistry 2015