Hydrothermal synthesis and multiferroic properties of Y2NiMnO6

Chenyang Zhanga, Tingsong Zhanga, Lei Gea, Shan Wangb, Hongming Yuan*a and Shouhua Fenga
aState Key Laboratory of Inorganic Synthesis and Preparative Chemistry, Jilin University, Changchun, China. E-mail: hmyuan@jlu.edu.cn
bThe Department of Materials Science and Engineering, Jilin Institute of Chemical Technology, Jilin, China

Received 14th July 2014 , Accepted 3rd October 2014

First published on 3rd October 2014


Abstract

The double perovskite Y2NiMnO6 was prepared by a mild hydrothermal method. The room temperature powder X-ray diffraction analysis shows that it crystallizes in the monoclinic perovskite lattice with space group P21/n. Analysis by bond valence sum and X-ray photoelectron spectroscopy reveal its oxide states to be Ni2+/Mn4+. Magnetic behaviours were characterized by DC and AC magnetic susceptibility, as well as magnetization–hysteresis. After the as-synthesized sample was annealed at 1273 K, the order of the arrangement between Ni2+ and Mn4+ increased. Close to the magnetic ordering temperature, a dielectric anomaly was observed at 84 K. The ferroelectric polarization of about 35 μC m−2 at 77 K was determined by Positive-Up-Negative-Down method, which is induced by up-up-down-down magnetic arrangement as expected theoretically.


Introduction

Multiferroic materials that are both ferroelectric and magnetic have attracted much research interest due to their potential applications in multi-state data storage and electric-field controlled spintronics.1,2 Among all the well-studied multiferroic systems, a large number of them are transition metal oxides, especially with perovskite-related structures.3–7 According to the “d0-ness rule”, transition metal ions having partially filled d shells cannot induce the ferroelectricity directly. Their magnetic orders can sometimes break the spatial inversion symmetry and produce a spontaneous ferroelectric polarization (P) so that the problem of chemical incompatibility for achieving simultaneous ferroelectric and magnetic orderings could be overcome. For example, non-collinear spiral magnetic order in Tb(Dy)MnO3 is responsible for the generation of spontaneous polarization below the corresponding magnetic transition temperature (T).3,8 The reverse Dzyaloshinskii–Moriya interaction between non-collinear canted spins in Sm(Y)FeO3 also gives rise to ferroelectric polarization.6,9 The exchange striction interaction between Gd(Dy) and Fe ions in Gd(Dy)FeO3 results in the emergence of ferroelectricity.10,11 More recently theoretical calculations suggest that some double perovskites are multiferroics, in which collinear up-up-down-down magnetic arrangement along the alternating B–B′ chains can break the space inversion symmetry via an exchange striction mechanism and produce a net ferroelectric polarization along the chain direction.12 Similar up-up-down-down (↑↑↓↓) spin configuration driven ferroelectricity has also been observed in the Ising chain magnet Ca3MnCoO6 and double perovskite R2CoMnO6 (R = Y and Lu).13–15

Recently, multiferroicity has been predicted to exist in the low temperature magnetic phase of double perovskite Y2NiMnO6 with the up-up-down-down spin configuration.12 Booth et al. reported a synthesis by solid state reaction, but the dielectric constant was only measured at room temperature.16 Maiti et al. synthesized nanocrystal Y2NiMnO6 by sol–gel method.17 However, the ferroelectric transition was observed at 537 K, far above the Curie temperature. This is probably due to the possible existence of mixed valence Mn3+/Mn4+ or chemical inhomogeneity in the nanocrystal Y2NiMnO6. In contrast, hydrothermal synthesis has been successfully applied to prepare advanced solid materials, advantageous to form perfect crystals and introduce specified oxidation states into a solid.18,19 In this study we report hydrothermal synthesis and characterization of the double perovskite Y2NiMnO6 and confirm the existence of ferroelectric transition occurring at Tc = 84 K which agrees well with the magnetic transition temperature below which an up-up-down-down spin configuration presents.

Experimental section

Material synthesis

In a typical synthetic procedure, 10 mL Y(NO3)3 (0.4 mol L−1), 5 mL NiSO4 (0.4 mol L−1) and 0.1739 g MnO2 were initially mixed in a beaker by continuously stirring to form a solution. Then, KOH was added on stirring and cooling to room temperature. The final mixture was transferred into a 45 mL Teflon-lined stainless steel autoclave with a filling capacity of 60%. Crystallization was carried out under autogenous pressure at 533 K for 3 days. After the autoclave was cooled down and depressurized, the sample was washed thoroughly with distilled water and sonicated by the direct immersion of a titanium horn (Vibracess, 20 kHz, 200 W cm−2). The dark crystals were obtained at the bottom of the beaker.

Material characterization

Sample compositions were determined by inductively coupled plasma spectroscopy (ICP) on a Perkin-Elmer Optima 3300DV spectrometer. The X-ray diffraction (XRD) patterns were obtained from 20° to 120° in a step of 0.02° with a counting time of 2 s per step on a Rigaku D/Max 2500 V/PC X-ray diffractometer with Cu Kα radiation (λ = 1.5418 Å) at 40 kV and 200 mA. The morphology of the sample was checked with a JEOL JSM-6700F scanning electron microscope (SEM). The valence states of the sample were determined using X-ray photoelectron spectroscopy (XPS, ESCALAB250, America). DC and AC magnetic measurements were performed in a superconducting quantum interference device vibrating sample magnetometer (Quantum Design). For electric measurements, the sample was ground, pelletized and then annealed at 1273 K for 10 hours. The silver paste was attached on the two faces of the pellet as electrodes. Capacitance measurement was carried out with LCR meter Agilent E4980A using homemade setup. The polarization–electric field (PE) hysteresis loop was measured using Positive-Up-Negative-Down (PUND) method by a Radiant Precision II.

Results and discussion

Synthesis

In hydrothermal process, precursor, mineral reagent, alkaline concentration, appropriate reaction temperature and time are key factors for the growth of transition metal oxides.19,20 In one respect, the amount of alkaline is a key factor in producing the double perovskite Y2NiMnO6. In our specific case, when the alkalinity was less than 3 M, we obtained a mixture containing target crystals and impurities including Y(OH)3, NiO, and KxMnO2. In the other respect, once alkalinity is above 12 M, the high alkalinity will not be able to remove small amount of NiO in the products. When alkaline concentration was only within the range between 3 and 10 M, pure dark crystals could be attained. KOH not only maintains the alkalinity in the solution but also acts as mineralizing agent. We also tried to use NaOH or the combination of KOH and K2CO3 as mineralizer. But impurities mentioned above cannot be removed. Therefore, pure KOH is necessary to obtain the desired products. In addition, the reaction temperature and time are responsible for the formation of the title compound. When the temperature is below 533 K, and time over 5 days, Y(OH)3 and NiO impurities were also present in the products. These facts indicate there is competition between thermodynamically and kinetically stabilized phases. Zhang et al. synthesized R2NiMnO6 (R = Pr, Sm and Ho) using MnCl2 as manganese source which is not favorable for the formation of Y2NiMnO6 with smaller rare earth radius.21 We also tried to substitute MnCl2 & KMnO4 at a molar ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]4 for MnO2 as manganese source, which results only in nonstoichiometric sample with the Y[thin space (1/6-em)]:[thin space (1/6-em)]Ni[thin space (1/6-em)]:[thin space (1/6-em)]Mn ratio 1[thin space (1/6-em)]:[thin space (1/6-em)]0.36[thin space (1/6-em)]:[thin space (1/6-em)]0.46.

Structure and thermal stability analysis

Fig. 1a shows the powder X-ray diffraction (XRD) pattern of the as-grown sample. The XRD peaks can be indexed as a single phase monoclinic structure with space group P21/n. The Rietveld refinement was done with GSAS/EXPGUI system.22,23 The lattice parameters obtained from the analyses (Table 1) are in good agreement with those reported previously.16
image file: c4ra07099b-f1.tif
Fig. 1 Rietveld refinement of the XRD powder patterns of (a) Y2NiMnO6, (b) Y2NiMnO6 annealed at 1273 K. Orange circles, black lines, blue lines and red lines show the observed, calculated, background and the difference patterns respectively. The sticks correspond to Bragg positions. (c) The magnetic (left) and crystal (right) structure of Y2NiMnO6.
Table 1 Lattice parameters, structure refinement parameters and bond valence sum (BVS) of Y2NiMnO6 with P21/n space group
  As grown After annealing
Lattice parameters
a 5.2247(2) 5.2192(1)
b 5.5739(2) 5.5485(1)
c 7.4863(3) 7.4781(1)
â 89.788(3) 89.666(1)
V 218.00(1) 216.60(1)
Rwp 0.0788 0.0569
Rp 0.0578 0.0441
÷2 4.804 2.584
[thin space (1/6-em)]    
Bond valence sum
BVS Ni 1.89 2.12
BVS Mn 4.16 4.05
BVS Y 3.1 3.05


To investigate the thermal stability and whether the degree of order between Ni2+ and Mn4+ could be enhanced by annealing, the samples were annealed at selected temperatures within the range of 573–1673 K. As shown in Fig. 2a, there is considerable Y2O3 in the product when annealing temperature was above 1573 K, but only minor NiO impurity at 1473 and 1373 K. When the annealing temperature was between 773 K and 1273 K, no impurities can be observed but the pattern shifts to higher angles. For annealing temperature at 573 K, there is no obvious change of XRD pattern from that of as-grown sample, implying no evidence of structure change.


image file: c4ra07099b-f2.tif
Fig. 2 (a) Room temperature XRD patterns of the as-grown sample and the samples annealed up to 1673 K. Inset shows the MH hysteresis loops at 4 K of as-grown sample (black), sample annealed at 773 (red), 973 (blue), 1073 (green), and 1273 K (pink). (b) XRD patterns around 33.5° of the as-grown sample and the samples annealed up to 1673 K.

Fig. 2b shows the main (112) and (−112) peaks around 33.5° for different annealed samples. In the as-grown sample, the two Bragg peaks locate almost at the same angle, indicating a very small deviation in structure from the high symmetric Pnma space group. For the samples annealed between 773 K and 1273 K, such two peaks are clearly splitting (Fig. 2b), indicating the occurrence of octahedral distortion. The diffraction pattern of the sample annealed at 1273 K (Fig. 1b) could be indexed as a monoclinic structure with space group P21/n. As seen in Table 1, the volume of the unit cell and β angle are decreased due to annealing. The above differences in terms of lattice structure actually point to a difference in the degree of order in La2NiMnO6.24,25 The volume and β angle are 233.65 Å3 and 90.01° for partially ordered La2NiMnO6, whereas two parameters decrease to 233.173 Å3 and 89.899° for fully ordered sample respectively. Therefore, it seems that the proper annealing condition can enhance the degree of order in our Y2NiMnO6 samples.

The promotion of order arrangement by annealing is further evidenced by the magnetic field (H) dependent magnetization (M) measurement (inset of Fig. 2a). With increasing the annealing temperature up to 1273 K, the saturation magnetic moment increases monotonically. The sample annealed at 1273 K exhibits the highest magnetic moment, indicating higher degree of order compared with those annealed at lower temperatures. It has been reported that in ordered double perovskite Ni2+ and Mn4+ ions are alternatingly located in corner-shared octahedral environments and a ↑↑↓↓ arrangement along c axis forms as elucidated in Fig. 1c. In order to analyze in detail the degree of order, oxide states and magnetic behaviours for the as-grown and annealed samples, XPS and magnetic measurements were performed as following.

Oxide states analysis

The XPS spectra of Mn 2p and Ni 2p core level can determine the oxide states of Mn and Ni ions in Y2NiMnO6.

Both the as-grown and the 1273 K annealed sample are investigated to check whether the oxide states of Mn and Ni ions have been changed or not after 1273 K annealing procedure, as shown in Fig. 3.


image file: c4ra07099b-f3.tif
Fig. 3 XPS spectra of Mn 2p (a) and Ni 2p (b) for the as-grown (upper lines) and the 1273 K annealed sample (lower lines), respectively.

The XPS spectra of the two samples suggest the Mn 2p3/2 peaks at about 642.6 eV. Since the 2p3/2 core level appears at 641.8 eV and 642.4 eV for Mn3+ and Mn4+, respectively,26 our data reveal unambiguously that the oxide states of Mn ion is +4 for two samples. The examined Ni 2p3/2 peaks of the two samples are at 855.4 eV, in agreement with the value reported for Ni2+.27 Furthermore, bond valence sum (BVS) calculated by software SPuDs is consistent with the XPS results (Table 1).28

Joy et al. prepared LaMn0.5Ni0.5O3 through low-temperature glycine-nitrate method.29 However, charge disproportionation of Mn4+ + Ni2+ → Mn3+ + Ni3+ occurs when the as-grown sample was annealed at high temperature. XPS spectrum and BVS of our sample which was also synthesized at relative lower temperature clearly show that the oxide states still keeps Ni2+/Mn4+ after the sample was annealed at 1273 K. Therefore, hydrothermal method provides an effective route to synthesize double perovskite R2NiMnO6 with stable Ni2+/Mn4+ oxide states.

Magnetic analysis

The temperature dependence of zero field cooled (ZFC) and field-cooled (FC) under 100 Oe applied magnetic field were measured in temperature range of 4 to 300 K (Fig. 4).
image file: c4ra07099b-f4.tif
Fig. 4 Temperature dependence of ZFC and FC susceptibility of (a) as grown Y2NiMnO6, and (b) Y2NiMnO6 annealed at 1273 K. Inset of (a) and (b) show χ−1 versus temperature with Curie–Weiss fitting.

A ↑↑↓↓ antiferromagnetic arrangement has been theoretically predicted in Y2NiMnO6 and confirmed by neutron diffraction experiment in R2CoMnO6 (R = Y and Lu). FC curves of the as-grown and annealed samples show ferromagnetic-like behaviour. This implies ferromagnetic instability accessed by magnetic-field cooling, as proposed by S. Yáñez-Vilar et al.14 The Curie temperature was extracted from the ZFC curves, i.e., the inflection point where dM/dT is a minimum. For the both samples, the Curie temperatures are 84 K and a cusp appears at 76 K. The ZFC and FC magnetizations have the same values around and above Tc, below which the irreversibility is observed. This feature could be attributed to either the low measuring field or spin glass behaviour. The Curie–Weiss fitting (inset of Fig. 4a and b) yields C = 3.77 emu mol−1 Oe−1 K−1, Tcw = 90 K for the as-grown sample and C = 3.82 emu mol−1 Oe−1 K−1, Tcw = 100 K for the annealed sample. The positive Curie–Weiss temperature indicates long range canted-spin correlation.15 The extrapolated effective moments increases from 5.49 μB to 5.58 μB after the sample was annealed, which are essentially consistent with the theoretical magnitude 5.9 μB,25,30 corresponding to one Ni2+ (S = 1) and one Mn4+ (S = 3/2) per formula.

Fig. 5 shows the MH curves at different temperatures, in the range of −30 kOe to +30 kOe. The magnetizations of both samples tend to reach saturated at 4 K and a 30 kOe applied field. At temperatures higher than Tc, the MH behaviours are Brillouin-function-like, corresponding to a paramagnetic state. The highest magnetic moment at 4 K and a 30 kOe magnetic field is 3.37 μB for the as-grown sample, indicating the presence of about 16% antisite disorder. It is noteworthy that the magnetic moment significantly increases to 5.11 μB after the sample was annealed at 1273 K, which is close to the summation of full ordered Ni2+ (S = 1) and Mn4+ (S = 3/2) magnetic moments in a formula unit.


image file: c4ra07099b-f5.tif
Fig. 5 The MH hysteresis of (a) Y2NiMnO6 and (b) Y2NiMnO6 annealed at 1273 K.

In order to shed more light on the magnetic behaviours of the two samples, we performed temperature dependence of susceptibility measurements in AC mode. Fig. 6 presents the frequency dependence of ac susceptibility, including the real component χ′(T) and imaginary components χ′′(T). Although it is not evident in the χ′(T), there exists a peak at 84 K corresponding to Tc in the χ′′(T), which is visible at higher frequency. For the as-grown sample, the susceptibilities show frequency dependent peak below the Curie temperature. In addition, a shoulder appears in the as-grown sample at about 68 K, which is absent in the annealed sample. As discussed above, the as-grown sample is partially ordered and short interaction exists in it.30 Thus we believe an antiferromagnetic matrix coexists with some ferromagnetic clusters in the as-grown sample. However, for the annealed sample, the peak shows no shift with frequency (Fig. 6d), indicating the elimination of antisite disorder and a greater degree of long range antiferromagnetic order. Besides, small non-interacting ferromagnetic clusters, which may still exist in the annealed sample, give rise to the ferromagnetic-like susceptibility (Fig. 4b). Similar behaviour may be also resulted from the antiferromagnetically ordered spins that are slightly canted in low field.


image file: c4ra07099b-f6.tif
Fig. 6 Real χ′ and imaginary χ′′ parts of AC susceptibility of (a) Y2NiMnO6, and (b) Y2NiMnO6 annealed at 1273 K. A zoomed view of AC susceptibility from 70 to 90 K of (c) Y2NiMnO6, and (d) Y2NiMnO6 annealed at 1273 K.

Dielectricity and ferroelectricity

The dielectricity and PE measurement were performed on sample annealed at 1273 K. For ferroelectricity, the transition from ferroelectricity to paraelectricity is always accompanied by the changes in the dielectric constant ε with temperature. The dielectric constant ε as a function of T is shown in Fig. 7.
image file: c4ra07099b-f7.tif
Fig. 7 Temperature dependence dielectric constant at 1 MHz.

The ε(T) exhibits an anomaly at 84 K, in agreement well with the Curie temperature. To check whether the sample is ferroelectricity, PUND measurement5,31 was performed at 77 K. This method could subtract non-ferroelectric components, such as the accumulated charge at the grain boundaries and leakage current.5 The PE loop is presented in Fig. 8, showing typical electric hysteresis and giving a saturated polarization of 35 μC m−2, which is comparable with that of magnetic related multiferroics.


image file: c4ra07099b-f8.tif
Fig. 8 PE hysteresis loop measured at 77 K.

To understand the mechanism of the ferroelectricity driven by magnetical order, we propose a model similar to double perovskite Lu2CoMnO6 and Y2CoMnO6 with identical structure with Y2NiMnO6, as discussed subsequently.14,15 As shown in Table 2, we compare the Ni–O–Mn bond angles of Y2NiMnO6 annealed at 1273 K with that of full ordered La2NiMnO6. The bond angles of Y2NiMnO6 are smaller than that of La2NiMnO6. With decreasing the bond angles, the superexchange interaction between Ni2+ and Mn4+ weakens. When the next-nearest neighbour Ni–O–Ni antiferromagnetic interaction is comparable with the nearest neighbour Ni–O–Mn ferromagnetic interaction, a frustrated ↑↑↓↓ magnetic arrangement is stabilized. Such ↑↑↓↓ arrangement of Ni2+ and Mn4+ breaks inversion symmetry and generates ferroelectricity, as theoretically predicted by Kumar et al.12

Table 2 Bond angles obtained from the room-temperature refinementa
  Bond angle (deg)
a The parameters of La2NiMnO6 were taken from ref. 26.
Y2NiMnO6 Ni–O1–Mn 144.43(26)
Ni–O2–Mn 148.31(32)
Ni–O3–Mn 144.49(32)
La2NiMnO6 Ni–O1–Mn 161.7(3)
Ni–O2–Mn 162.3(3)
Ni–O3–Mn 158.5(2)


Conclusion

In summary, double perovskite Y2NiMnO6 was synthesized by mild hydrothermal method. Bond valence sum and XPS show the transition metals are Ni2+ and Mn4+. After the sample was annealed at 1273 K, the degree of order is greatly enhanced and approach to fully ordered state, as proved by the magnetic measurements. For the annealed sample, a dielectric constant anomaly at magnetic transition temperature was observed, and a spontaneous polarization about 35 μC m−2 was determined by PUND method.

Acknowledgements

The authors are grateful to Yisheng Chai for the capacitance measurement.

References

  1. J. F. Scott, Nat. Mater., 2007, 6, 256–257 CrossRef CAS PubMed.
  2. N. A. Spaldin and M. Fiebig, Science, 2005, 309, 391–392 CrossRef CAS PubMed.
  3. T. G. T. Kimura, H. Shintani, K. Ishizaka, T. Arima and Y. Tokura, nature, 2003, 426, 55–58 CrossRef CAS PubMed.
  4. S. Ishiwata, Y. Tokunaga, Y. Taguchi and Y. Tokura, J. Am. Chem. Soc., 2011, 133, 13818–13820 CrossRef CAS PubMed.
  5. Y. S. Chai, Y. S. Oh, L. J. Wang, N. Manivannan, S. M. Feng, Y. S. Yang, L. Q. Yan, C. Q. Jin and K. H. Kim, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 85, 184406 CrossRef.
  6. J.-H. Lee, Y. K. Jeong, J. H. Park, M.-A. Oak, H. M. Jang, J. Y. Son and J. F. Scott, Phys. Rev. Lett., 2011, 107, 117201 CrossRef.
  7. M. C. Weber, J. Kreisel, P. A. Thomas, M. Newton, K. Sardar and R. I. Walton, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 85, 054303 CrossRef.
  8. S. S. Tokura Yoshinori, Adv. Mater., 2010, 22, 1554–1565 CrossRef PubMed.
  9. M. Shang, C. Zhang, T. Zhang, L. Yuan, L. Ge, H. Yuan and S. Feng, Appl. Phys. Lett., 2013, 102, 062903 CrossRef PubMed.
  10. Y. Tokunaga, N. Furukawa, H. Sakai, Y. Taguchi, T.-h. Arima and Y. Tokura, Nat. Mater., 2009, 8, 558–562 CrossRef CAS PubMed.
  11. Y. Tokunaga, S. Iguchi, T. Arima and Y. Tokura, Phys. Rev. Lett., 2008, 101, 097205 CrossRef CAS.
  12. S. Kumar, G. Giovannetti, J. van den Brink and S. Picozzi, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 82, 134429 CrossRef.
  13. Y. J. Choi, H. T. Yi, S. Lee, Q. Huang, V. Kiryukhin and S. W. Cheong, Phys. Rev. Lett., 2008, 100, 047601 CrossRef CAS.
  14. S. Yáñez-Vilar, E. D. Mun, V. S. Zapf, B. G. Ueland, J. S. Gardner, J. D. Thompson, J. Singleton, M. Sánchez-Andújar, J. Mira, N. Biskup, M. A. Señarís-Rodríguez and C. D. Batista, Phys. Rev. B: Condens. Matter Mater. Phys., 2011, 84, 134427 CrossRef.
  15. G. Sharma, J. Saha, S. D. Kaushik, V. Siruguri and S. Patnaik, Appl. Phys. Lett., 2013, 103, 012903 CrossRef PubMed.
  16. R. J. Booth, R. Fillman, H. Whitaker, A. Nag, R. M. Tiwari, K. V. Ramanujachary, J. Gopalakrishnan and S. E. Lofland, Mater. Res. Bull., 2009, 44, 1559–1564 CrossRef CAS PubMed.
  17. R. P. Maiti, S. Dutta, M. Mukherjee, M. K. Mitra and D. Chakravorty, J. Appl. Phys., 2012, 112, 044311 CrossRef PubMed.
  18. S. Feng and R. Xu, Acc. Chem. Res., 2001, 34, 239–247 CrossRef CAS PubMed.
  19. Y. Chen, H. Yuan, G. Li, G. Tian and S. Feng, J. Cryst. Growth, 2007, 305, 242–248 CrossRef CAS PubMed.
  20. J. J. Urban, L. Ouyang, M.-H. Jo, D. S. Wang and H. Park, Nano Lett., 2004, 4, 1547–1550 CrossRef CAS.
  21. G. Zhang, G. Li, F. Liao, Y. Fu, M. Xiong and J. Lin, J. Cryst. Growth, 2011, 327, 262–266 CrossRef CAS PubMed.
  22. A. Larsen and R. Von Dreele, Report LAUR 86-748, Los Alamos National Laboratory, New Mexico, USA, 1994.
  23. B. H. Toby, J. Appl. Crystallogr., 2001, 34, 210–213 CrossRef CAS.
  24. D. Choudhury, P. Mandal, R. Mathieu, A. Hazarika, S. Rajan, A. Sundaresan, U. V. Waghmare, R. Knut, O. Karis, P. Nordblad and D. D. Sarma, Phys. Rev. Lett., 2012, 108, 127201 CrossRef CAS.
  25. N. S. Rogado, J. Li, A. W. Sleight and M. A. Subramanian, Adv. Mater., 2005, 17, 2225–2227 CrossRef CAS.
  26. J. C. Carver, G. K. Schweitzer and T. A. Carlson, J. Chem. Phys., 1972, 57, 973–982 CrossRef CAS PubMed.
  27. K. K. Lian, D. W. Kirk and S. J. Thorpe, J. Electrochem. Soc., 1995, 142, 3704–3712 CrossRef CAS PubMed.
  28. M. W. Lufaso and P. M. Woodward, Acta Crystallogr., Sect. B: Struct. Sci., 2001, 57, 725–738 CrossRef CAS PubMed.
  29. V. L. J. Joly, P. A. Joy, S. K. Date and C. S. Gopinath, Phys. Rev. B: Condens. Matter Mater. Phys., 2002, 65, 184416 CrossRef.
  30. Y. Mao, J. Parsons and J. S. McCloy, Nanoscale, 2013, 5, 4720–4728 RSC.
  31. M. Fukunaga and Y. Noda, J. Phys. Soc. Jpn., 2008, 77, 064706 CrossRef.

Footnote

Electronic supplementary information (ESI) available: The structural parameters of the as-grown and annealed samples. See DOI: 10.1039/c4ra07099b

This journal is © The Royal Society of Chemistry 2014
Click here to see how this site uses Cookies. View our privacy policy here.