• Nenhum resultado encontrado

Mössbauer and magnetic study of nanocrystalline strontium hexaferrite prepared by an ionic coordination reaction technique

N/A
N/A
Protected

Academic year: 2021

Share "Mössbauer and magnetic study of nanocrystalline strontium hexaferrite prepared by an ionic coordination reaction technique"

Copied!
5
0
0

Texto

(1)

Mössbauer and magnetic study of nanocrystalline strontium

hexaferrite prepared by an ionic coordination reaction technique

J.H. de Araújo

a

, J.M. Soares

b,n

, M.F. Ginani

c

, F.L.A. Machado

d

, J.B.M. da Cunha

e a

Departamento de Física Teórica e Experimental, Universidade Federal do Rio Grande do Norte, 59072-970 Natal-RN, Brazil

bDepartamento de Física, Universidade do Estado do Rio Grande do Norte, 59610-010 Mossoró-RN, Brazil c

Departamento de Química, Universidade Federal do Rio Grande do Norte, 59072-970 Natal-RN, Brazil

d

Departamento de Física, Universidade Federal de Pernambuco, 50670-901 Recife-PE, Brazil

e

Instituto de Física, Universidade Federal do Rio Grande do Sul, 91501-970 Porto Alegre-RS, Brazil

a r t i c l e i n f o

Article history:

Received 19 November 2012 Received in revised form 21 April 2013

Available online 14 May 2013 Keywords: Sr-hexaferrite Chemical synthesis X-ray diffraction Mössbauer spectroscopy Magnetic property

a b s t r a c t

Hard-magnetic nanocrystalline strontium hexaferrites SrFe12O19 were synthesized using an ionic coordination reaction technique. In this sample preparation technique the biopolymer chitosan was used as a nanoreactor. The obtained precursor powders were calcined at temperatures in the range 600– 9001C. The samples were analyzed by X-ray diffraction, transmission electron microscopy, Mössbauer spectroscopy and vibrating sample magnetometry. A complementary study of X-ray Rietveld refinement and Mössbauer spectroscopy shown that the hexaferrite phase formation is accompanied by formation of maghemite and hematite as intermediate phases. It was found that hexaferrite is present in the samples calcined at and above 6001C but it is fully developed at 900 1C. For this sample the average particle size was found to be 41.6 nm. Magnetization measurements yielded squared hysteresis loops with a magnetization ratio (Mr=Ms) of 0.58 and a coercivefield of 6.48 kOe. The overall results indicated that the particles in these samples are in the single domain regime and that the magnetization reversal in these particles is mainly due to coherent rotation.

& 2013 Elsevier B.V. All rights reserved.

1. Introduction

The physical and chemical properties of hexaferrites make these materials very attractive for some technological applications. They present high magnetic anisotropy, moderate saturation magnetization and good chemical stability. A variety of processes has been used to synthesize fine powders of hexaferrites. The conventional solid state reaction route involves the mixing of oxide/carbonate and heat treatment at high temperature (12001C)

[1]. The solid state reaction method has some inherent disadvan-tages such as chemical inhomogeneity and coarser particle size distribution. Moreover, the hexaferrite must be milled to further reduce the particle size and to reach the single magnetic domain regime. The milling process introduces root mean square strain in the particles yielding low values for the saturation magnetization and for the intrinsic coercivity. In addition, the powder character-istics are usually poor and they have a wide particle size distribu-tion that include magnetic multi-domain particles. A number of chemistry-based processing routes has recently been devised for the preparation offine particles of hexaferrites. These chemistry-based novel synthesis routes include hydrothermal methods[2,3], sol–gel [4,5], combustion [6,7], chemical co-precipitation [8,9],

salt-melt[10], aerosol pyrolyzes[11,12], glass crystallization[13], cryochemical methods [14], colloidal synthesis [15] and micro-emulsion mediated processes [16]. However, these methods are complex and expensive to be used in large scale.

In the present work, hard-magnetic nanocrystalline strontium hexaferrites were synthesized. The formation process was investi-gated and their structures and morphologies were characterized by X-ray diffraction and transmission electron microscopy. Some of the magnetic parameters were determined making use of a vibrating sample magnetometer and by Mössbauer spectroscopy. The samples were produced by a synthesis process based on the ionic coordination reaction (ICR) technique[17]where the biopo-lymer chitosan was used as a nanoreactor. Chitosan is an amino-polysaccharide extracted by a deacetylation procedure from chitin, the most abundant biopolymer in nature after cellulose. The ICR technique is also attractive because of its cost, its low-temperature synthesis, the processing is somewhat simpler and it yields highly homogeneous samples. The particles exhibited a coercivity of 6.48 kOe and a saturation magnetization of 53.2 emu/g.

2. Experimental

The precursor solution used to make samples was prepared by using 99.99% pure reagents. A mixture of 0.102 g of SrðNO3Þ2,

Contents lists available atSciVerse ScienceDirect

journal homepage:www.elsevier.com/locate/jmmm

Journal of Magnetism and Magnetic Materials

0304-8853/$ - see front matter& 2013 Elsevier B.V. All rights reserved. http://dx.doi.org/10.1016/j.jmmm.2013.04.077

nCorresponding author. Tel.:+55 218433152196.

(2)

2.350 g of FeðNO3Þ39H2O and 2.548 g of citric acid was added to a

volume of 50 ml of the 2% chitosan biopolymer-5% acetic acid aqueous solution for yielding about 1.250 g of SrFe12O19. The

procedure followed the ICR technique where the citric acid works as a chelating agent helping the link between the polymer chains

[18]. It was found that the amount in mass of the nitrates of Sr and of Fe and of the citric acid which optimizes the process follows the ratio 1:12:25, respectively. Due to the ionic coordination bonding between chitosan and metal cations, Sr and Fe ions can be homogeneously distributed in the precursor solution at an atomic level. The chitosan-metal ion complex solution is cross-linked with aldehyde by means of the amine groups, forming a gel. In linked chitosan, the polymeric chains are interconnected by cross-linkers leading to the formation of 3D permanent covalent net-work. The gel was then burned at ambient atmosphere for 4 h at a temperature of 3501C forming the precursor powder. Parts of the precursor powder were calcined for 1 h at temperatures varying in the range 600–900 1C. At the end of the calcination process it becomes a ultra-fine brown powder. The structural characteriza-tion of the samples was made by X-rays diffraccharacteriza-tion (XRD) using a Rigaku diffractometer and the CuKα radiation and by transmis-sion electron microscopy (TEM). Some room temperature para-meters of the samples were measured using a vibrating sample magnetometer and applied magnetic fields in the range of −13:5oH o13:5 kOe.57Fe Mössbauer effect measurements were

taken room temperature using a transmission geometry conven-tional constant acceleration spectrometer operating in triangular wave mode with a57Co source in a Rhodium matrix. The

spectro-meter was calibrated with a 25μm thick αFe foil.

3. Results and discussions

Fig. 1shows the XRD patterns of the Sr-hexaferrite precursor powder prepared using the chitosan solution. The Rietveld analy-sis of the XRD pattern of this precursor showed that the particles are composed of strontium carbonate SrCO3 with space group

Pmcn:bca and cell parameters a0¼5.0758Å, b0¼8.4462Å, and

c0¼6.0485Å, and of maghemite γ−Fe2O3 with space group

Fd-3 m and cell parameter a0¼8.3509Å.

Fig. 2 shows a TEM image for the sample calcined at 9001C while the inset shows the particle size distribution. From this distribution one gets an average particle size Dmfor the particles of

48.1 nm. This value is close to the one obtained from the X-ray

data reported inTable 1indicating that the crystallographic order extends throughout the nanoparticles.

Fig. 3shows the XRD patterns for the samples calcined at 600, 700, 800 and 9001C. The samples calcined at 600 1C and above show typical XRD patterns of M-type hexaferrites. The XRD pattern of the sample calcined at 6001C does also shows an unequivocal feature of a spinel structure characterized by an intense peak around 2θ ¼ 351. This spinel structured phase was identified with the aid of the Mössbauer spectrum to be maghe-mite. The results for the Rietveld refinement of the XRD patterns are summarized below inTable 1.

The refinements were made assuming the hexagonal symmetry and the P63/mmc space group for the Sr-hexaferrite. The obtained lattice constants were a0¼5.8805 Å and c0¼23.0530 Å for the

sample calcined at 8001C. It is also clearly seen in Table 1that the calcination of the precursor powder transforms maghemite in both hematite and Sr-hexaferrite phases. By increasing the calci-nation temperature the relative amount of the Sr-hexaferrite phase increases while the maghemite one diminishes. The XRD did not show the presence of the maghemite phase in the samples calcined at and above 7001C. However, as shown below, a better fit to the Mössbauer spectra is achieved with the introduction of a small amount of maghemite.

Fig. 4shows Mössbauer spectra measured at room temperature of samples calcined at various temperatures. The spectra were fitted with sextets using a minimum least-square computer program. Five Lorentzian sextets corresponding to the 2a, 2b, 4f1, 4f2 and 12k sites in the hexaferrite phase and one Lorentzian sextet corresponding to hematite phase were used to fit the spectra for the samples calcined at 800 and at 9001C. For the sample calcined at 6001C, on the other hand, it was also necessary

Fig. 1. X-ray diffractogram for the precursor powder showing the maghemite γ−Fe2O3and SrCO3phases.

Fig. 2. TEM image for the sample calcined at 9001C. The inset shows the particle size distribution obtained from the TEM image.

Table 1

Rietveld Refinement Data – Calcination temperatures and the corresponding percent-weight phase-composition. SFO, Hem, Mag and SrCO3are the strontium

hexaferrite, the hematite, the maghemite and strontium carbonate phases, respec-tively. The last column gives the average particle diameter for the SFO phase.

T (1C) SFO (%) Hem (%) Mag (%) SrCO3(%) DSFO(nm)

600 27.0 9.2 60.8 3.0 17.2 700 88.0 8.7 0.0 3.3 32.8 800 92.5 5.3 0.0 2.2 43.1 900 94.6 3.9 0.0 1.5 41.6

(3)

to add two Lorentzian sextets corresponding to the maghemite phase for a good fit to be achieved. Moreover, for the sample calcined at 7001C, it was only necessary to add one maghemite Lorentzian sextet tofit the spectrum. The Mössbauer parameters for the calcined samples are shown in Table 2. In the M-type hexagonal structure of SrFe12O19, the Fe3+ ions occupy five

different crystallographic sites. Three are octahedral sites (2a, 4f2 and 12k), one site is tetrahedral (4f1) and the fifth one (2b) is trigonal bi-pyramidal. The point symmetry of these sites are consistent with an asymmetry parameter η ¼ 0, and an electric field gradient tensor with its principal axis parallel to the c-axis. The theoretical relative population of the sites are 0.0833 for 2a and 2b sites, 0.1666 for 4f1 and 4f2 sites and 0.5 for 12k site. In general, the relative spectral areas for SrFe12O19 sub-spectra

follows approximately the relative populations of the sites. For example in the sample calcined at 9001C, recalculating the absorptions relating only to SFO phase, we obtain for the site 12k an absorption of 0.491, for the sum of 2a and 2b a relative absorption of 0.191 and for 4f1 and 4f2, the sum is 0.317.

The magnetic moments of the iron ions are arranged parallel to the hexagonal c-axis, but with opposite spin directions of the sublattices, three spin up (12k, 2a and 2b) and two spin down (4f1 and 4f2)[19]. Note that for our samples there is an increasing in the absorption of the sites with spin up, at the expense of the absorptions of the sites with spin down, favoring a higher net magnetic moment of the ferrimagnetic structure. The relative magnitudes of the hyperfine fields Hhf, are in agreement with [20], evolving as follows: Hhfð2bÞoHhfð12kÞoHhfð4f 1Þo

Hhfð2aÞoHhfð4f 2Þ.

Isomer shifts are in the range 0.10–0.40 mm/s relative to metallicαFe indicating that the ionic states of iron ions on the five sites are ferric Fe3+. The sequence of relative isomer shifts

Fig. 3. X-ray diffractograms for samples calcined at various temperatures.

Fig. 4. Mössbauer spectra for samples calcined at various temperatures.

Table 2

Mössbauer parameters for SFO samples calcined at different temperatures. The numbers in parentheses are the estimated errors and theδ values are relative to metallic Fe.

T (1C) Phase Site Hhf(kOe) ΔEQ (mm/s) δ (mm/s) A (%)

900 SFO 12k 412(2) 0.38(1) 0.236(3) 48.0 4f1 493(2) 0.13(1) 0.102(3) 19.6 4f2 510(2) 0.09(1) 0.201(3) 11.4 2a 505(2) 0.32(1) 0.393(3) 12.1 2b 408(2) 2.22(1) 0.152(3) 6.5 Hem 517(2) −0.06(1) 0.163(3) 2.4 800 SFO 12k 412(2) 0.38(1) 0.236(3) 47.4 4f1 492(2) 0.12(1) 0.101(3) 18.5 4f2 508(2) 0.07(1) 0.205(3) 11.5 2a 504(2) 0.35(1) 0.400(3) 12.5 2b 407(2) 2.18(1) 0.172(3) 7.4 Hem 515(2) −0.01(1) 0.195(3) 2.7 700 SFO 12k 411(2) 0.40(1) 0.239(3) 41.5 4f1 491(2) 0.12(1) 0.102(3) 19.0 4f2 507(2) 0.02(1) 0.210(3) 12.7 2a 502(2) 0.33(1) 0.403(3) 12.0 2b 407(2) 2.14(1) 0.170(3) 7.0 Hem 515(2) −0.08(1) 0.209(3) 5.0 Mag 408(2) 0.14(1) 0.237(3) 2.7 600 SFO 12k 411(2) 0.39(1) 0.235(3) 10.0 4f1 488(2) 0.09(1) 0.122(3) 3.6 4f2 513(2) 0.19(1) 0.272(3) 4.0 2a 490(2) 0.30(1) 0.257(3) 3.5 2b 407(2) 2.18(1) 0.156(3) 1.3 Hem 498(2) −0.01(1) 0.224(3) 5.8 Mag1 408(2) 0.14(1) 0.231(3) 37.7 Mag2 476(2) 0.12(1) 0.203(3) 27.5

(4)

values found isδð4f 1Þoδð2bÞoδð4f 2Þoδð12kÞoδð2aÞ. The quad-rupole splitting of the site 2b is very large because of the strongly distorted environment.

The room temperature magnetic hysteresis cycles are shown in

Fig. 5. The hysteresis loops are squared for samples annealed at and above 7001C. However, for the sample calcinated at 600 1C the hysteresis loop shows a behavior that is typical of a mixture of hard (Sr-hexaferrite) and soft (maghemite) magnetic materials. The magnetic parameters for all calcined samples are shown in

Table 3. In general, the magnetic properties improved increasing the calcination temperature. The overall results indicate that the particles in these samples are in the single magnetic domain regime. This interpretation is supported by a theoretical model used to describe the magnetic behavior of uniaxial single-domain particles. Within this model, the calculated critical diameter Dcrit

for a spherical single-domain particle of Sr-hexaferrite is about 940 nm[21]. The average particle size for our samples (17–42 nm), however, are much less than Dcrit. Moreover, the theoretical value

for the mean coercivefield at room temperature is 6.70 kOe, for a system of randomly distributed single domain SrFe12O19particles

under reverse magnetization [22,23]. This value is close to the measured in the samples calcinated at and above 7001C. The ratio Mr=Msfor the samples annealed at 7001C and above does also

agrees with the theoretical value (Mr=Ms¼ 0:5) for a system of

randomly distributed single domain particle. Here we used, as a reasonable approximation, the value of the magnetization mea-sured at 13.5 kOe as the saturation magnetization. We can see fromTables 2and3that the increasing of the coercivefield and the saturation magnetization is accompanied by an increasing in the population of 2a site. Furthermore, the results and the interpretation presented in here are in excellent agreement with

those seen by transverse susceptibility measurements[24]. Thus, based on the experimental results and on the Stoner-Wohlfarth model, one can suggest that the mechanism responsible for the magnetization reversal in these particles is the coherent rotation one.

4. Conclusions

In summary, the ICR technique was used to produce high quality samples of hard-magnetic Sr-hexaferrite. XRD and Möss-bauer effect measurements shown that the hexaferrite phase is already present in samples calcined at temperatures as low as 6001C. Also found that the relative amount of phases, the particle size, the coercivity and the saturation magnetizations can be controlled by the calcination temperature. An increasing in the absorption of the sites with spin up, at the expense of the absorptions of the sites with spin down, favored a higher net magnetic moment of the samples. These results indicate that the synthesis method used herein may be important for optimization of hard-magnetic hexaferrites. Moreover, was found that the Mr=Msratio for samples calcined at and above 7001C agrees well

with the theoretical values determined by the Stoner– Wohlfarth model.

Acknowledgments

Work partially supported by CNPq, FACEPE and FINEP (Brazilian Agencies).

References

[1] G.C. Bye, C.R. Howard, Journal of Applied Chemistry and Biotechnology 21 (1971) 319.

[2] D. Primc, M. Drofenik, D. Makovec, European Journal of Inorganic Chemistry 25 (2011) 3802.

[3] M. Jean, V. Nachbaur, J. Bran, J.-M. Le Breton, Journal of Alloys and Compounds 496 (2010) 306.

[4] X. Yang, Q. Li, J. Zhao, B. Li, Y. Wang, Journal of Alloys and Compounds 475 (2009) 312.

[5] H. Haiyan, Rare Metal Materials and Engineering 35 (2006) 1334. [6] X.F. Pan, G.H. Mu, N. Chen, K.K. Gan, K. Yang, M.Y. Gu, Materials Science and

Technology 23 (2007) 865.

[7] J. Huang, H. Zhuang, W.-L. Li, Materials Research Bulletin 38 (2003) 149. [8] S. Tyagi, R.C. Agarwala, V. Agarwala, Transactions of the Indian Institute of

Metals 63 (2010) 15.

[9] A. Drmota, A. Znidarsic, A. Kosak, Journal of Physics: Conference Series 200 (2010) 082005.

[10] Z.-B. Guo, W.-P. Ding, W. Zhong, J.-R. Zhang, Y.-W. Du, Journal of Magnetism and Magnetic Materials 175 (1997) 333.

[11] T. González-Carreño, M.P. Morales, C.J. Serna, Materials Letters 43 (2000) 97. [12] Y. Senzaki, J. Caruso, M.J. Hampden-Smith, T.T. Kodas, L.-M. Wang, Journal of

the American Ceramic Society 78 (1995) 2973.

[13] P.E. Kazin, L.A. Trusov, D.D. Zaitsev, Y.D. Tret'yakov, Russian Journal of Inorganic Chemistry 54 (2009) 2081.

[14] T.G. Kuzmitcheva, L.P. OlKhovik, V.P. Shabatin, IEEE Transactions on Mag-netics 31 (1995) 800.

[15] S.E. Kushnir, A.I. Gavrilov, P.E. Kazin, A.V. Grigorieva, Y.D. Tretyakov, M. Jansen, Journal of Materials Chemistry 32 (2012) 18893,http://dx.doi. org/10.1039/C2JM33874B.

[16] D.H. Chen, Y.Y. Chen, Journal of Colloid and Interface Science 236 (2001) 41. Fig. 5. Hysteresis cycles for the Sr-hexaferrite samples calcined at different

temperatures.

Table 3

Coercivefield, magnetization for H¼13.5 kOe (MHmax) and the magnetization ratio

for the Sr-hexaferrite samples calcined at different temperatures.

T (1C) Hc(kOe) MHmax(emu/g) Mr=Ms

600 0.45 47.2 0.33

700 6.04 48.2 0.58

800 6.47 47.8 0.58

(5)

[17] J.M. Soares, F.L.A. Machado, J.H. de Araújo, M.F. Ginani, Journal of Magnetism and Magnetic Materials 310 (2007) 2529.

[18] V.N. Dhage, M.L. Mane, M.K. Babrekar, C.M. Kale, K.M. Jadhav, Journal of Alloys and Compounds 509 (2011) 4394.

[19] E.W. Gorter, Proceedings of the IEEE 104B (1957) 225.

[20] B.J. Evans, F. Grandjean, A.P. Lilot, R.H. Vogel, A. Grard, Journal of Magnetism and Magnetic Materials 67 (1987) 123.

[21] V. Pillai, P. Kumar, D.O. Shah, Journal of Magnetism and Magnetic Materials 116 (1992) L299.

[22] E.C. Stoner, E.P. Wohlfarth, IEEE Transactions on Magnetics 27 (1991) 3475. [23] K. Haneda, H. Kojima, Journal of Applied Physics 44 (1973) 3760.

[24] J.M. Soares, F.L.A. Machado, J.H. de Araújo, F.A.O. Cabral, H.A.B. Rodrigues, M. F. Ginani, Physica B 384 (2006) 85.

Referências

Documentos relacionados

All Mössbauer spectra (Figure 5a-d) obtained from samples in powder of natural and pillared clay with aluminum and aluminum - lanthanum, show the presence of Fe 3+ , and

Figure 4: X-ray diffraction patterns of samples synthesized using calcined kaolin at different temperatures... The samples presented less intense relections than the

As an example, for ACRON MC composite four different dilutions were prepared and the NIR infrared spectra were obtained using a transmission micro-liquid cell with KBr windows at a

To determine the effect of temperature on lipase activity, purified enzyme and substrate were incubated at various reaction temperatures before starting the experiment and the

The linewidths ( Γ ) of the Mössbauer spectra tended to increase across all samples with increasing Al- substitution (Figure 4a) in a manner similar to that observed with

Parameters obtained from room temperature and low temperature (lines with bold characters) Mössbauer spectra for roots (R) and shoots (S) of rice plants from cultivars BR-IRGA

The FTIR spectra showed in this work were obtained using samples prepared by keeping the Eth:TIP molar ratio at 16:1.. UV – vis absorption spectra were taken in the spectral range

According to Krogh-Moe [19], the structure of pure vitreous B 2 O 3 consists of a random network of boroxol rings and BO 3 triangles connected by B–O–B linkage (bridging oxygen).