Enhancing the electro catalytic activity of manganese ferrite through cerium substitution for oxygen evolution in KOH solutions

T. Pandiarajan, S. Ravichandran* and L. J. Berchmans
CSIR – Central Electrochemical Research Institute, Karaikudi-630006, India. E-mail: Srravi371@gmail.com

Received 5th September 2014 , Accepted 17th November 2014

First published on 18th November 2014


Abstract

Alkaline water electrolysis (AWE) is the simplest way of producing hydrogen, an attractive fuel for the future. In view of their cost-effectiveness and durability, non-noble metal oxides are promising catalysts for AWE. Here, we studied the effect of Ce substitution on the OER activity of manganese ferrite. Ce substituted MnFe2O4 (0 ≤ x ≤ 0.8) was synthesized by a combustion method. Characterization techniques such as SEM, XRD, EDAX and XPS were used to analyse the surface morphology and the chemical composition of CexMnFe2−xO4. Substitution of Ce3+ in the cubic lattice of MnFe2O4 increases the conductivity of CexMnFe2−xO4, which results in the negative shift in the OER onset potential. Among all the Ce3+ substituted manganese ferrites, Ce0.2MnFe1.8O4 was found to be more active for OER in terms of current and stability. Notably, Ce0.2MnFe1.8O4 affords a current density of 10 mA cm−2 at a small overpotential of ∼310 mV and a Tafel slope value of 31 mV per decade, these values are comparable to the well-investigated non-noble metal oxides.


1 Introduction

Our world has moved steadily towards a serious energy crisis, which has sparked dedicated research related to renewable energy, storage, and conversion technologies. In pursuit of the ultimate energy source, developing solar water-splitting cell, batteries, and fuel cells is a current research focus. As a result of the sluggish kinetics involved in the oxygen evolution reaction (OER) and hydrogen evolution reaction (HER) sustainable energy systems require a large input of energy.1–3 Metal oxides composed of iridium and ruthenium exhibit excellent catalytic activity for the oxidation of water into molecular hydrogen and oxygen; however, high cost and scarcity make these oxides undesirable for commercial applications.4

Although electrolytic water splitting was first identified in an acidic medium, because of cost-effective electrolyser fabrication and the simple control of corrosion, alkaline water electrolysis has dominated large-scale hydrogen production technology for several decades. Still, it has been challenge to find non-noble metal catalysts that exhibit a current density of j > 0.5 A cm−2 at an over potential of ηO2 < 0.3 V with long-term stability.5,6

The exhaustive search for non-noble catalysts has identified that the mixed metal oxides of transition elements, spinel ferrites, oxy hydroxides (such as MnO2, NiOx, MnFe2O4, NiFe2O4, CoFe2O4, Co3O4, etc.) and perovskites oxides, as catalysts for OER reaction.2,6–10 Importantly, inspired by the oxygen evolution centres (OEC) of photosystem II, several investigations of low-cost manganese oxides for electrochemical water splitting revealed the efficiency of MnO2 and Mn3O4 for OER, particularly the α-Mn2O3 phase. While MnO2 is a potential bifunctional catalyst in both acidic and basic mediums, changing the pH from alkaline to neutral results in a sharp increase in the onset potential.11–14

Surprisingly, spinel ferrites and mixed metal oxides NPs supported with graphene or carbon nanotube potentially support the ORR in alkaline solutions.2,9 However, their performance also drops significantly on conventional carbon supports. But, Spinel ferrites (MxFe3−xO4) NPs supported with conventional carbon exhibit better ORR activity compared to commercial Pt NPs in alkaline solution.15

However, the spinel ferrites possessing significant electro catalytic activity and stability in high pH, the low electrical conductivity, restricts its application as a catalyst. In general, the practical applications of ferrites are mainly depends on electronic conductivity, results from thermal activation of electrons or positive holes along chains of neighbouring cations in the ionic lattice. Due to the bimetallic nature of spinels, the conductivity of ferrites can be tuned by substituting with a foreign metal, using different preparation conditions or using heat treatment.16–18

Previous studies of MnFe2O4 employed several synthesis methods, mostly to examine structural, electrical conductivity, magnetization and OER properties.19,20 Recently, Shouheng Sun and his co-works found that similar ferrite system exhibits the ORR activity in alkaline solution, comparable to that of Pt.15 The focus of our current study involves determining the catalytic behaviour of cerium substituted manganese ferrite, for OER in alkaline solution. We prepared manganese ferrite and cerium-substituted manganese ferrites by a combustion method.18 Electro catalytic analysis showed that Ce0.2MnFe1.8O4 is a promising catalyst for OER in alkaline solutions.

2 Experimental section

2.1 Synthesis of catalysts

Manganese ferrite and cerium substituted manganese ferrites were synthesized by a combustion method. Aliquot amounts of analytical grade metal nitrates, Mn(NO3)2·6H2O, Fe(NO3)2·9H2O and Ce(NO3)3·6H2O were mixed with hexamine and dissolved in deionized water to obtain a precursor solution. The solution was preconcentrated in a quartz crucible until the free water evaporated, after which spontaneous combustion of the dried powder occurred. This resulted in the formation of black porous ash in the container. This ash was then sintered at 1100 °C for 24 hours in an electrical furnace. The catalyst was milled for 8 hours in a Retsch PM 100 ball mill at room temperature at 600 rpm. For milling experiments, the ball-to-powder weight ratio was taken as 40[thin space (1/6-em)]:[thin space (1/6-em)]1.

2.2 Physiochemical characterization

Catalyst crystallinity and structure examined by a powder X-ray diffractometer (Model D8 Advance, Bucker) with a Cu-Kα1 X-ray radiation (λ = 0.15406 nm) source, operating at 40 kV and 30 mA. Catalyst morphology and surface composition investigated using a VEGA3 TESCAN Scanning Electron Microscope, with Energy Dispersive X-ray Spectroscopy (SEM/EDS) with X-Flash Detector 410M with Bruker ESPRIT QUANTAX EDS analysing software. X-ray photoelectron spectroscopy (XPS) was carried out on MULTILAB 2000 Base system with X-ray, Auger and ISS attachments (Thermo scientific) with Twin Anode Mg/Al (300/400 W) X-ray Source.

2.3 Working electrode preparation

The suspension of catalyst was prepared by sonicating 3 mg of the catalyst, in 60 μL isopropyl alcohol for 15 minutes.30 μL of the suspension was pipetted out and dried on a metal substrate (1 cm × 1 cm) by heating, in a pre-heated oven at 80 °C. Subsequently the substrate was sintered, at 1200 °C in the presence of air for 2 hours.

2.4 Electrochemical measurements

Electrochemical experiments were conducted at room temperature using a VMP3 multi-channel potentiostat (Bio-Logic Science Instrument Company). In the three-electrode system, a Pt wire was used as counter electrode, and all potentials were measured with respect to 1 M KOH Hg/HgO reference electrode that was housed in a custom glass Luggin capillary. The potential of the 1 M KOH Hg/HgO reference electrode was 0.929 V. All potentials during electrochemical characterization were IR-compensated and reported versus RHE. The scan rate was 1 mV s−1 for all electrochemical measurements except during the cyclic voltammetry studies. All cyclic voltammograms were measured in a 1 M KOH electrolyte with a sweep rate of 10 mV s−1. Three KOH solutions, 0.01 M, 0.1 M, and 1.0 M, were used as electrolytes. Electrochemical impedance spectroscopy (EIS) measurements were conducted in the frequency range 1 MHz to 10 Hz, 1.6 V versus RHE, with an AC voltage amplitude of 10 mV.

3 Results and Discussions

The X-ray diffraction patterns of the as prepared MnFe2O4 with different concentrations of Ce3+ ions from 0.0 M to 0.8 M are as shown in Fig. 3a. The reflection patters are distinct, clearly delineating the cubic phase of the spinels. Moreover, no typical diffraction peak for MnO2 (110) or Ce2O3 (111), (220) is seen in the XRD pattern for MnFe2O4, which reveals that the samples are not composites. The sharp, well-defined peaks display the crystalline nature of the spinel ferrites. As shown in Fig. 3a no changes occurred in the position of the reflections due to the inclusion of Ce3+ for Fe3+, only a variation in the lattice constant21 values were observed because of a great deviation in the ionic radius of metal ions (Ce3+ = 1.3 Å; Fe3+ = 0.64 Å). The variation of lattice constant values derived from XRD data are given in Fig. 3b. A similar behaviour of lattice parameter with La3+ substitution is observed by L. J. Berchmans et al. in a similar ferrite system.18
image file: c4ra09806d-f1.tif
Fig. 1 SEM images of as prepared CexMnFe2−xO4(X=0.0,0.2) samples. SEM images of sintered samples at 1200 °C.

image file: c4ra09806d-f2.tif
Fig. 2 EDAX spectrum of catalysts after sintering at 1200 °C.

image file: c4ra09806d-f3.tif
Fig. 3 (a) XRD patterns of (i) MnFe2O4 (ii) Ce0.2MnFe1.8O4 (iii) Ce0.4MnFe1.6O4 (iv) Ce0.6MnFe1.4O4 (v) Ce0.8MnFe1.2O4. (b) The variation of lattice constant (a) with x of CexMnFe2−xO4.

Fig. 2 shows the chemical composition of Pt/MnFe2O4 electrodes (sintering at 1200 °C, in the presence of air for 2 hours) estimated by energy-dispersive X-ray (EDX) analysis, which revealed that the catalysts contained only Mn, Fe and O atoms in a ratio of 1[thin space (1/6-em)]:[thin space (1/6-em)]2[thin space (1/6-em)]:[thin space (1/6-em)]4. The absence of Pt in the EDX profile shows that no diffusion of platinum occurs during the high temperature treatment (Fig. 5). Moreover, the EDS measurements confirm the presence of Ce in the MnFe2O4 lattice in the calculated ratio (Fig. 2). The XPS spectrum, shown in Fig. 4, was also used to verify the presence of Ce3+. The characterized peaks at 845 eV and 905 eV in the XPS spectra occurs due to the presence of Ce3+.22 Based on the above results, we conclude that Ce3+ is present in our samples.


image file: c4ra09806d-f4.tif
Fig. 4 XPS spectra of the catalysts. The insert (a) is survey spectra of as prepared Ce0.2MnFe1.8O4, (b) XPS spectra of Fe before sintering (c) XPS spectra of Ce before sintering (d) XPS spectra of Fe (after sintered at 1200 °C).

image file: c4ra09806d-f5.tif
Fig. 5 Cyclic voltammograms of Pt supported MnFe2O4 and CexMnFe2−xO4 electrodes at a potential scan of 10 mV s−1 in 1 M KOH at 25 °C.

The morphology of the as-prepared MnFe2O4 and substituted MnFe2O4 are confirmed by the SEM images Fig. 1a. All the particles have discrete crystals with a cube shape. Each sample is then dispersed in isopropyl alcohol (IPA) and coated on Pt (1 cm × 1 cm) electrode. As shown in Fig. 1b, sintering the electrode to 1200 °C leads to the melting of discrete MnFe2O4 particles, resulting in the agglomeration of ferrite particles together to form a single, thin, film-like layer on the Pt surface.10 The sintering process favours the formation of a thin flim of spinel ferrites on Pt surface; however, no characterise peak for Pt is observable in XRD analysis, which accounts for absence of Pt diffusion into the spinel ferrites. Thin flim layer insulates the Pt surface completely. Also, no redox peak appeared for Pt on cyclic voltammetry studies of samples in KOH solution.

Fig. 5 shows the corresponding cyclic voltammograms of Pt/CexMnFe2−xO4 with x = 0.0, 0.2, 0.4, 0.6, 0.8 at a potential scan rate of 10 mV s−l in 1 M KOH solution at 25 °C. The electrode surface of the Pt/CexMnFe2−xO4 did not show any redox peaks either for Pt or for catalyst on cyclic voltammetry studies.19 It implies that ferrite does not undergo further oxidation. R. N. Singh and co-workers reported the similar behaviour for Mn-substituted ferrites coated on Pt and Ti substrates.19

However, with each subsequent CV cycle, OER current increases and OER onset potential shifted towards more negative region due to incorporation of Ce3+. It is often considered that the effect of Ce3+ incorporation on MnFe2O4 enhances the electrical conductivity of materials in turn influencing the OER activity.

The polarisation curves of Pt/CexMnFe2−xO4 electrodes were performed in 1 M KOH solution Fig. 7. The Pt/MnFe2O4 electrode prepared by current method shows an over potential of 360 mV for OER which is lower than the reported value of 580 mV for MnFe2O4 electrodes by Iwakura et al.23 Related to pure MnFe2O4 a negative shift in the OER onset potential is noticed with all Ce substituted electrodes. Among all, the Ce0.2MnFe1.8O4 exhibits a maximum negative shift of 60 mV for OER lower than MnFe2O4. Comparison of over potentials of the various catalysts required to achieve a current density of 10 mA cm−2 for CexMnFe2−xO4 are shown in Fig. 7. In a similar condition, Iwakura et al.23 observed an over potential of 440 mV and 580 mV for CoFe2O4 and MnFe2O4 in 1 M KOH respectively. Similarly, Orehotsky et al. obtained the same current density at the over potential of 340 mV for NiFe2O4 in 30 wt% KOH. In addition, our results are comparable to well investigated Cobalt-based electrocatalysts (Table 2).


image file: c4ra09806d-f6.tif
Fig. 6 Electrochemical impedance spectra of catalysts as a function of Ce substitution, measured at 1.6 V during oxygen evolution.

image file: c4ra09806d-f7.tif
Fig. 7 Polarisation curves of CexMnFe2−xO4 sintered on Pt (with catalyst loading of 1 mg cm2) measured in 1 M KOH.
Table 1 Tafel slope values of CexMnFe2−xO4
S. no. Catalyst Tafel slope (b) mV per decade
1 MnFe2−O4 38
2 Ce0.2MnFe1.8O4 31
3 Ce0.4MnFe1.6O4 33
4 Ce0.6MnFe1.4O4 35
5 Ce0.8MnFe1.2O4 37


Table 2 Catalytic activity of various electrodes in 1 M KOH
S. no. Electrodes KOH concentration & temperature OER potential (mV) at 10 mA cm−2 Tafel slope value (b) mV per decade Ref.
1 Pt/Ce0.2MnFe1.8O4 1 M, 25 °C 310 31 This work
2 Ni/MnFe2O4 1 M, 25 °C 300 36–42 19
3 Pt/MnFe2O4 1 M, 25 °C 580 110–115 23
4 Pt/CoFe2O4 1 M, 25 °C 440 110–115 23
5 NiFe204 30 wt%, 25 °C 340 19
6 Ni/Co3O4/N-rmGO 1 M, 25 °C 310 67 2
7 NiLaOx 1 M, 25 °C 400 3


A negative shift in the OER onset is associated with incorporation Ce3+ in the MnFe2O4 spinel lattice. As shown in Fig. 6 the electrochemical impedance plots of CexMnFe2−xO4 exhibit low charge transfer resistance compared to pure MnFe2O4 electrodes. Because of the reduced charge transfer resistance the electronic conductivity increases and favours in the enhancement of electro catalytic activity Ce3+ substituted MnFe2O4. Typically, manganese ferrite has inverse spinel structure in which 80% of Mn2+ occupies site A and 20% occupies site B. Site B can also be occupied by Fe2+ and Fe3+ ions.24,25 The electrical conductivity of ferrites are explained by Verwey-de Boer mechanism and the polaron effect. Although the electrical conduction in ferrites originates between Fe2+ and Fe3+ ions at site B, conduction mainly depends on the Fe2+ concentrations. In the sintering process, some lattice oxygen escapes from the oxides, causing an oxygen deficiency in the crystal lattice. Therefore, to balance the electrical charge created in the lattice unit, Fe3+ is reduced to Fe2+.24,26 The X-ray photoelectron spectroscopy (XPS) spectrum, shown in Fig. 4, confirms the existence of Fe2+ ions in the spinel lattice, and the presence of two main peaks with binding energies of 710.4 and 724.8 eV, which clearly proves that Fe2+ coexists with Fe3+.27,28 Moreover, it was found that the created Fe2+ ions prefer to occupy the octahedral B-sites. This reduction facilitates the formation of excess Fe2+ ions in the system (as shown in Fig. 4(b)) such that the hopping rate of electrons in ferrites increases which is the cause of their higher electronic conductivity.24,29 If it is the only cause then, we could expect a gradual increase in catalytic activity of substituted MnFe2O4 with higher concentrations of Ce3+. In contrast, the OER onset potential increased at concentrations higher than 0.2 M. Because, the replacement of rare earth ions forms grain boundaries in ferrites;30 if high molar concentration is used, then it leads to the formation of secondary phases (i.e. ABO3) on the materials.16,17 Most importantly, the secondary phase exists at concentration >0.4. Although the density of these grain boundaries is reduced upon treating the powders at high temperature, secondary phase are still present in the catalyst layer at high concentrations of cerium; which may hinder the mobility of charge carriers and increases the electrical resistivity.31 Similar behaviour was found in Al doped manganese ferrites.19 This coincides with the negative shift of OER onset potential.

As a result, of the cooperative effect32 between the electron-rich Mn2+ and Ce3+ centres in the cubic spinel lattice also enhances the OER activity of Ce0.2MnFe1.8O4.17 Nevertheless, both effects work simultaneously and increase the electro catalytic activity of the substituted manganese ferrite. It is clear from the above discussion that the electronic and structural properties of spinels play an important role in the OER.

The iR corrected Tafel slope values of the CexMnFe2−xO4 are calculated from the steady state polarization plots shown in Fig. 8 using the following equation:

 
η = a + b[thin space (1/6-em)]log[thin space (1/6-em)]j (1)


image file: c4ra09806d-f8.tif
Fig. 8 Tafel plots of MnFe2O4 and CexMnFe2−xO4 coated on Pt recorded at 1 mV s−1 in 1 M KOH.

We also obtained a range of Tafel slope values varying from 31–45 mV per decade for all oxides, which is similar to the previous reports (Table 1).19,23

The reaction order (p) of Ce0.2MnFe1.8O4 and pure MnFe2O4 with respect to [OH] are calculated by the method described in ref. 18. The calculated p and b values are in good agreement with recently reported spinel ferrites such as MnFe2O4 (b = 36 mV per decade, p = 1.9), CrxMnFe2−xO4 (b = 35 mV per decade, p = 2) and CrxCoFe2−xO4 (b1 = 40–51, b2 = 60–89 mV per decade, p = 1.2–1.4). Besides, similar electrode kinetic parameters also exhibited by NiCo2O4 film (b = 42 mV per decade, p = 1.7) and CuCo2O4 (b = 40 mV per decade) sprayed on Ni.19,23

Based on the Tafel slope value and oxygen evolution potential, the sequential order of the electro catalytic activity of the electrodes is given as Ce0.2MnFe1.8O4 > Ce0.4MnFe1.6O4 > Ce0.6MnFe1.4O4 > Ce0.8MnFe1.2O4 > MnFe2O4.

In addition, the Tafel slope measurements attribute to the common mechanism for OER reaction as well:

 
S + OH ↔ SOH + e (2)
 
SOH + OH ↔ SO + H2O (3)
 
SO ↔ SO + e (4)
 
2SO ↔ 2S + O2 (5)
where, S indicates the active species on the catalyst surface.

R. N. Singh and co-workers19 observed a tafel slope (b) value of 36–47 mV per decade in 1 M KOH at 25 °C, suggesting that MnxFe2−xO4 follows second order kinetics with respect to [OH] and step (3) as the rate determining step (RDS) for OER. We also observed a similar trend in tafel slope values and reaction rate in case of MnFe2O4 (b = 38 mV per decade; p = 2.0) and Ce0.2MnFe1.8O4 (b = 31 mV per decade; p = 1.9). The long-term stability of Ce0.2MnFe1.8O4 under OER conditions was determined using chronoamperometric studies. The electrode is kept at a constant potential of 1.8 V for 10 h in 1 M KOH solution, while the current was measured as a function of time. Stability of the electrode found to be unchanged throughout the constant-potential studies with a constant current density of 130 mA cm−2 (Fig. 9).


image file: c4ra09806d-f9.tif
Fig. 9 Chronoamperometric curves of MnFe2O4 and Ce0.2MnFe1.8O4 loaded on Pt under potential of 1.8 V.

4 Conclusion

Ce-substituted MnFe2O4 powders with varying concentration of Ce were prepared by the combustion method. The X-ray diffraction patterns of the samples confirm the cubic structure of the spinels. Substitution of Ce3+ for Fe3+ decreases the OER onset potential for about 50 mV, which may be explained by the electronic conductivity and cooperative effect. Among all samples, Ce0.2MnFe1.8O4 showed the maximum electronic conductivity and OER activity. The chronoamperometric study of Ce0.2MnFe1.8O4 reveals the long-term stability of the catalyst.

Acknowledgements

The authors are thankful to DST (Technology Development and Transfer Division), New Delhi for its generous financial support under its Solar Hydrogen Programme (no: DST/TSG/SH/2011/106-G). We are greatly indebted to Dr Vijayamohanan K. Pillai, Director, CSIR-CECRI for his keen interest and encouragement. G. Sozhan, Head, Electroinorganic Chemicals Division for support. T. P thank DST and UGC for Providing JRF.

References

  1. M. W. Kanan and D. G. Nocera, Science, 2008, 321, 1072–1075 CrossRef CAS PubMed.
  2. Y. Liang, Y. Li, H. Wang, J. Zhou, J. Wang, T. Regier and H. Dai, Nat. Mater., 2011, 10, 780–786 CrossRef CAS PubMed.
  3. C. C. McCrory, S. Jung, J. C. Peters and T. F. Jaramillo, J. Am. Chem. Soc., 2013, 135, 16977–16987 CrossRef CAS PubMed.
  4. T. Reier, M. Oezaslan and P. Strasser, ACS Catal., 2012, 2, 1765–1772 CrossRef CAS.
  5. E. Rasten, G. Hagen and R. Tunold, Electrochim. Acta, 2003, 48, 3945–3952 CrossRef CAS PubMed.
  6. Y. Leng, G. Chen, A. J. Mendoza, T. B. Tighe, M. A. Hickner and C. Y. Wang, J. Am. Chem. Soc., 2012, 134, 9054–9057 CrossRef CAS PubMed.
  7. W. Zhou and J. Sunarso, J. Phys. Chem. Lett., 2013, 4, 2982–2988 CrossRef CAS.
  8. R. D. Smith, M. S. Prevot, R. D. Fagan, Z. Zhang, P. A. Sedach, M. K. Siu, S. Trudel and C. P. Berlinguette, Science, 2013, 340, 60–63 CrossRef CAS PubMed.
  9. Y. Gorlin and T. F. Jaramillo, J. Am. Chem. Soc., 2010, 132, 13612–13614 CrossRef CAS PubMed.
  10. S. Ida, K. Yamada, T. Matsunaga, H. Hagiwara, Y. Matsumoto and T. Ishihara, J. Am. Chem. Soc., 2010, 132, 17343–17345 CrossRef CAS PubMed.
  11. T. Takashima, K. Hashimoto and R. Nakamura, J. Am. Chem. Soc., 2012, 134, 1519–1527 CrossRef CAS PubMed.
  12. T. Takashima, K. Hashimoto and R. Nakamura, J. Am. Chem. Soc., 2012, 134, 18153–18156 CrossRef CAS PubMed.
  13. D. M. Robinson, Y. B. Go, M. Mui, G. Gardner, Z. Zhang, D. Mastrogiovanni, E. Garfunkel, J. Li, M. Greenblatt and G. C. Dismukes, J. Am. Chem. Soc., 2013, 135, 3494–3501 CrossRef CAS PubMed.
  14. D. M. Robinson, Y. B. Go, M. Greenblatt and G. C. Dismukes, J. Am. Chem. Soc., 2010, 132, 11467–11469 CrossRef CAS PubMed.
  15. H. Zhu, S. Zhang, Y. X. Huang, L. Wu and S. Sun, Nano Lett., 2013, 13, 2947–2951 CrossRef CAS PubMed.
  16. J. Z. Msomi, T. Moyo and J. J. Dolo, Hyperfine Interact., 2010, 197, 59–64 CrossRef CAS PubMed.
  17. R. Singh, J. Singh, H. Nguyencong and P. Chartier, Int. J. Hydrogen Energy, 2006, 31, 1372–1378 CrossRef CAS PubMed.
  18. L. J. Berchmans, M. P. I. Devi and K. Amalajyothi, Int. J. Self-Propag. High-Temp. Synth., 2009, 18, 11–14 CrossRef CAS.
  19. N. K. Singh, S. K. Tiwari, K. L. Anitha and R. N. Singh, J. Chem. Soc., Faraday Trans., 1996, 92, 2397 RSC.
  20. K. Mujasam Batoo, Phys. B, 2011, 406, 382–387 CrossRef PubMed.
  21. J. Xiang, X. Shen and Y. Zhu, Rare Met., 2009, 28, 151–155 CrossRef CAS PubMed.
  22. J. Hou, W. Jiang, Y. Fang and F. Huang, J. Mater. Chem. C, 2013, 1, 5892 RSC.
  23. C. Iwakura, M. Nishioka and H. Tamura, Nippon Kagaku Kaishi, 1982, 7, 136 Search PubMed.
  24. Y. Fukuda, S. Nagata and K. Echizenya, J. Magn. Magn. Mater., 2004, 279, 325–330 CrossRef CAS PubMed.
  25. B. Liu, K. Zhou, Z. Li, D. Zhang and L. Zhang, Mater. Res. Bull., 2010, 45, 1668–1671 CrossRef CAS PubMed.
  26. A. Tawfiks, Y. Atassi, and I. S. Daewish, Latv. J. Phys. Tech. Sci., 2005, N3, 68–74 Search PubMed.
  27. Y. Fu, P. Xiong, H. Chen, X. Sun and X. Wang, Ind. Eng. Chem. Res., 2012, 51, 725–731 CrossRef.
  28. L. Zhang and Y. Wu, J. Nanomater., 2013, 2013, 1–6 Search PubMed.
  29. B. Parvatheeswara Rao and K. H. Rao, J. Mater. Sci., 1997, 32, 6049–6054 CrossRef.
  30. X. jun, S. Xiangqian and Z. Yongwei, Rare Met., 2009, 28, 151 CrossRef PubMed.
  31. A. A. Sattar and A. M. Samy, J. Mater. Sci., 2002, 37, 4499–4502 CrossRef CAS.
  32. D. W. stephan, Coord. Chem. Rev., 1989, 95, 41 CrossRef CAS.

This journal is © The Royal Society of Chemistry 2014