Structural, Mössbauer spectroscopy, magnetic properties, and thermal measurements of Y3−xDyxFe5O12
Lataifeh Mahdi1, Mohaidat Qassem I1, †, H Mahmood Sami2, Bsoul Ibrahim3, Awawdeh Mufeed1, Abu-Aljarayesh Ibrahim1, Altheeba Mu’ath1
Physics Department, Yarmouk University, Irbid 21163, Jordan
Physics Department, The University of Jordan, Amman 11942, Jordan
Physics Department, Al al-Bayt University, Mafraq 13040, Jordan

 

† Corresponding author. E-mail: q.muhaidat@yu.edu.jo

Project supported by the Deanship of Research and Graduate Studies of Yarmouk University (Garnet No. 33/2015).

Abstract

Yttrium iron garnet powder samples (Y3−xDyxFe5O12, where part of yttrium ions are substituted by dysprosium ions with different concentrations are prepared by the solid state reaction method. The properties of the prepared samples are examined by different methods such as x-ray diffraction (XRD), Mössbauer spectroscopy, macroscopic magnetization measurements, and thermal measurements. The XRD measurements show that all the samples reveal the presence of a single garnet phase with a BCC structure. Room temperature Mössbauer spectra indicate that iron ions occupy three magnetic sites, i.e., two octahedral sites and one tetrahedral site. The saturation magnetization and the initial magnetic susceptibility decrease with the increase of Dy3+ substitution. The Curie temperature obtained from the thermal measurements seems to be independent of Dy3+ substitution.

1. Introduction

Ferrimagnetic garnets are isomorphic with the naturally occurring garnet Ca3Fe2(SiO4)3. Specifically, Yttrium iron garnet (Y3Fe5O12, YIG) was discovered in 1956, and demonstrated suitability for high frequency microwave applications.[13] The discovery of YIG led to a revolution in the technological advancement of passive microwave devices including tunable filters, circulators, and gyrators, as well as magnetic-bubble-domain-type digital memory devices and other applications.[46]

The unit cell which has a cubic symmetry contains eight formula units of garnet, where the metal ions occupy three different interstitial sites. The Fe3+ ions occupy the 24d tetrahedrally coordinated sites and the 16a octahedrally coordinated sites, whereas the RE ions occupy the 24c dodecahedral sites. These latter sites are not all crystallographically equivalent, where six different c sites were reported.[7]

Since Y3+ is a nonmagnetic ion, the magnetization of YIG arises from the superexchange interaction between the moments of Fe3+ ions at a and d sublattices, resulting in a net magnetization of per molecule at 0 K. When nonmagnetic Y3+ ions are substituted by heavy rare-earth (RE) ions from the series Gd3+ through Yb3+, however, the magnetization as a function of temperature exhibits a compensation temperature (Tcomp) at which the magnetization vanishes.[1,6] This is because these heavy RE elements are magnetic due to their unfilled 4f orbitals, introducing a third magnetic sublattice (c) which couples antiferromagnetically with the net magnetization of the Fe3+ sublattices due to the fact that the ad superexchange interaction of Fe3+ ions is much stronger than the ac or dc interactions between Fe3+ and RE ions. Accordingly, such a substitution would facilitate the modifications of the magnetic properties of the garnet compounds, which renders these compounds attractive from the scientific and technological point of view. Further, the strong Fe–Fe interaction is responsible for the magnetic ordering of rare earth iron garnets (RIG), and the weak RE–Fe interactions make the Curie temperature of all rare-earth iron garnets nearly the same (∼550 K).[810]

The effects of RE substitution for Y on the magnetic properties and hyperfine interactions in YIG garnets were investigated by several researchers. The saturation magnetization of Y3−xDyxFe5O12 nanoparticles prepared by the sol-gel method was found to decrease linearly with increasing Dy3+ concentration (x) in a range between , and increase with the increase of particle size.[11,12] Also, room temperature 57Fe Mössbauer spectroscopy study of rare earth iron garnets (RxY1−x)3Fe5O12, where R is either Ho3+ or Gd3+, has shown that the hyperfine field at the iron nucleus in the octahedral and tetrahedral sites increase with rare earth substitution.[13] On the other hand, room temperature Mössbauer spectroscopy indicated that Ho substitution for Y in YIG did not induce noticeable changes in the hyperfine fields of the octahedral site nor tetrahedral sites.[14]

In this study, Y3−xDyxFe5O12 powders are prepared by the solid state reaction method with part of Y ions being substituted by Dy ions in the full range between 0.0 and 3.0 (x=0.0, 0.5, 1.0, 1.5, 2.0, 2.5, and 3.0). The structural and magnetic properties of the prepared compounds are investigated by x-ray diffraction (XRD), Mössbauer spectroscopy, and vibrating sample magnetometry (VSM). Also, differential scanning calorimetry is employed to investigate the Curie temperature of the compound. The interpretation of the magnetic results is based on the structural data and on the thermal measurements. This work is a continuation of our project involving the investigation of rare-earth iron garnets by several techniques such as nuclear magnetic resonance (NMR) of holmium in different RIG hosts,[10] Mössbauer spectroscopy,[13,14] and magnetization measurements.[8,9]

2. Experimental procedures

Y3−xDyxFe5O12 powders with x =0.0, 0.5, 1.0, 1.5, 2.0, 2.5, and 3.0 were prepared by the solid state reaction method. Metallic oxides Y2O3, Dy2O3, and Fe2O3 were used to fabricate Y3−xDyxFe5O12 powder samples. Stoichiometric ratios of the metallic oxides were loaded into a hardened stainless steel cup with a ball-to-powder ratio of 8:1. The milling process was carried out at 250 rpm for 16 h. The resulting precursor was annealed at 1300 °C for 2 h. The XRD patterns were collected using a Philips X’pert PRO X-ray diffractometer (PW3040/60) operating at (45 kV, 40 mA), with Cu– radiation (λ = 1.5405 Å). The samples were scanned over an angular range of with 0.02° scanning step and speed of 1°/min. The XRD patterns were analyzed using X’pert HighScore software to identify the phases present in each sample. Rietveld refinement of the XRD patterns for all fabricated samples was carried out using FullProf suite 2000 software to obtain the values of refined structural parameters.

Samples for Mössbauer spectroscopy studies were prepared by gently pressing a thin layer of the powdered sample between two circular Teflon disks with diameter 2 cm, which is small compared with the distance between the -ray source and the sample, in order to avoid the angular broadening of the spectrum. Room temperature (RT) Mössbauer spectra were collected using a standard constant acceleration Mössbauer spectrometer over 1024 channels. The -ray source was a 50-mCi 57Co in a rhodium matrix. The spectra were then analyzed using fitting routines based on least-squares analysis.

The magnetic measurements were carried out using a vibrating sample magnetometer (VSM MicroMag 3900, Princeton Measurements Cooperation), providing a maximum applied magnetic field of 10 kOe (1 Oe =79.5775 A/m). The measurements were performed at room temperature.

Differential scanning calorimetry (DSC) was used to obtain quantitative and qualitative information about phase transitions involving endothermic or exothermic processes, or changes in the heat capacity of the sample under investigation. This technique was therefore employed to investigate the phase transitions and related critical temperatures of the prepared garnets using a NETZSCH–Gerätebau GmbH 204 F1 Phoenix calorimeter. The DSC measurements were performed in a temperature range from 30 °C to 350 °C at a heating rate of 5 °C per minute.

3. Results and discussion

The structural characteristics of all Y3−xDyxFe5O12 samples are investigated using XRD measurements. Rietveld analysis of the diffraction patterns (shown in Fig. 1) is carried out to obtain the refined structural parameters of crystalline phases in the samples, and the refined structural results are summarized in Table 1. Figure 1 indicates that all samples reveal the presence of a single garnet phase, crystallizing in a body-centered cubic (BCC) structure with space group. The lattice parameter a of the two end compounds (Y3Fe5O12 and Dy3Fe5O12) is in very good agreement with previously reported values[6] and the low values of χ2 obtained from the Rietveld analysis indicate good agreement between the theoretical pattern (black continuous line) and the experimental data (red dots), where the line in blue representing the residual difference between the observed and calculated patterns, was a straight horizontal line with small ripples for all samples. Moreover, the enlarged views of the patterns of the two end compounds (Fig. 2) reveal that the two samples are crystallographically identical, indicating a complete substitution of Dy3+ ions at the Y3+ sites.

Fig. 1. (color online) Refined XRD patterns for Y3−xDyxFe5O12 samples.
Fig. 2. (color online) Refined XRD patterns for samples Y3−xDyxFe5O12 with x = 0 and x =3, with residual difference between the experimental and calculated data given in the lower part of the plots (blue line).
Table 1.

Lattice constant a, unit cell volume V, χ 2, and x-ray density of Y3−xDyxFe5O12 garnets.

.

The data in Table 1 reveal that the lattice parameter a and the cell volume increase with the Dy3+ substitution increasing, which can be attributed to the fact that the ionic radius of Dy3+ (1.027 Å) is greater than that of Y3+ (1.019 Å) at dodecahedral sites.[15] The monotonic increase of the x-ray density with increasing x, however, is opposite to what is expected based on the increase of the cell volume. This behavior can be associated with the increase of the molecular weight of the compound with Dy3+ substitution level increasing. The x-ray density is defined as

where Z(= 8) is the number of formula units in a unit cell, M is the molecular weight, NA=6.023 × 1023 is Avogadro’s number, and V is the cell volume. Further, the molecular weight can be calculated from the atomic weights of the constituents of the molecule, which turns out to be

Combining Eqs. (1) and (2) one can expect that a plot of versus x should yield a perfect straight line. Such a plot is shown in Fig. 3, where the linear fit to the experimental data presents an intercept of about 738.5 and a slope of about 73.5 (see the inset table in Fig. 3), which are in excellent agreement with the theoretical values in Eq. (2). This is an indication that the behaviors of the cell volume and x-ray density can be well explained on the basis of the statistical distribution of Dy3+ ions at the 24c sites.

Fig. 3. (color online) Linear increase of the product of the x-ray density and cell volume with x value increasing.

Figure 4 shows Mössbauer absorption spectra of all Y3−xDyxFe5O12 samples, and the hyperfine parameters are listed in Table 2. These parameters are obtained by fitting Mössbauer spectra with three magnetic sextets, two of which correspond to Fe3+ ions at two non-equivalent octahedral (16a) sites, and one corresponds to Fe3+ ions at the tetrahedral (24d) site. The intensity ratio of the components associated with the octahedral sites to that with the tetrahedral site does not deviate from the value of ∼37:63 for each of all compounds, which is in good agreement with the theoretical ratio of 2:3. The hyperfine fields of the Dy3+ substituted compounds do not deviate appreciably from the value of Bhf = 497 kOe nor 487 kOe nor Bhf= 401 kOe for the pure YIG. This is an indication that the ad superexchange interactions between Fe3+ ions at the octahedral and tetrahedral site are much stronger than ca or cd interactions between Fe3+ and rare-earth ions.[1] The difference in hyperfine field between the two sublattices is due to the difference in the number of O2− ions coordinated with Fe+3 ions in the octahedral sites and tetrahedral site of the iron garnet crystal lattice. In addition, the values of the center shifts (isomer shifts) are nearly the same for all samples. In view of the fact that the center shift is dependent on the s-electronic density at the nucleus,[16] the constancy of the center shifts for the octahedral sublattice (0.38 ±0.02 mm/s) and tetrahedral sublattice (0.15 ±0.02 mm/s) is an indication that the substitution of Y3+ ions by Dy3+ ions does not perturb the wave function of the Fe3+ ions, which is a further indication of the weak ca and cd coupling. This conclusion is also supported by the fact that the observed average hyperfine field of the octahedral sites (∼492 kOe) and that of the tetrahedral site (∼401 kOe) are practically equal to those of 492 kOe–495 kOe and 397 kOe–399 kOe, respectively, for Ho-substituted YIG.[14]

Fig. 4. (color online) Mössbauer spectra for Y3−xDyxFe5O12 samples (solid circles) together with theoretical spectra obtained from fitting software (red solid curves). Components corresponding to octahedral sites are represented by blue curves, while the component corresponding to tetrahedral sites is represented by green curves.
Table 2.

Hyperfine parameters and percentage relative intensity (I) values obtained at room temperature for the Y3−xDyxFe5O12 samples.

.

The hysteresis loops (Fig. 5) indicate that all samples of Y3−xDyxFe5O12 are soft magnetic materials each with a very small coercivity. Also, the loops indicate that the saturation magnetization decreases monotonically with Dy3+ concentration increasing. As a consequence of the ad superexchange interaction between Fe3+ ions being much stronger than the superexchange interaction between the Fe3+ ion and the magnetic rare-earth ion,[1,6] the net magnetization of the iron sublattices couples antiferromagnetically with the c sublattice containing the rare earth ions, and the net magnetization of the compound is given by[6]

At 0 K, the saturation magnetization of Dy3Fe5O12(∼98 emu/g) is significantly higher than that of YIG (∼38 emu/g), indicating that the substitution of Dy3+ for Y3+ should increase the saturation magnetization at 0 K. However, due to the difference in temperature dependence between the net iron moment and the rare-earth moment, the magnetization of the compound drops sharply with temperature increasing, and becomes zero at the compensation point Tcomp(=220 K for Dy3Fe5O12). Above the compensation temperature, the magnetization increases slowly and reaches ∼5 emu/g at 300 K. Since this value is significantly lower than the saturation magnetization of ∼27 emu/g for YIG at 300 K, the following simple magnetic dilution model is proposed to explain the monotonic decrease of the RT saturation magnetization with x increasing.

Fig. 5. (color online) Hysteresis loops for Y3−xDyxFe5O12 samples.

Assuming phase separation in the Dy-substituted compound into two separated garnets according to the formula:

and using the molecular weights of 737.97 for YIG and 958.74 for Dy3Fe5O12, the saturation magnetization of the general compound is calculated by using the observed RT saturation magnetizations of 28.3 emu/g[17,18] and 7.65 emu/g for the two end compounds. After some mathematical manipulation this is given by

A plot (Fig. 6) of the left-hand term of Eq. (5) versus x gives a straight line with an intercept of ∼280 and a slope of −60.7. These values are in very good agreement with the calculated values from Eq. (5), indicating that the behavior of the saturation magnetization is well described by the proposed simple magnetic dilution model, resulting from the statistical distribution of the Dy3+ ions in the c sublattice, which was confirmed earlier by the behavior of the x-ray density and cell volume.

Fig. 6. (color online) Illustration of saturation magnetization behavior with Dy concentration (x) increasing.

The initial magnetization curves (Fig. 7) of the samples of Y3−xDyxFe5O12 show that both the initial magnetic susceptibility χin and saturation magnetization Ms decrease with rare earth substitution increasing. Figure 8 shows a plot of the observed Ms [in units (emu/g)] and χin [in units () as a function of dysprosium concentration x. The saturation magnetization exhibits nearly a linear decrease with x increasing, which is in agreement with the reported behavior of the saturation magnetization of Y3−xDyxFe5O12() prepared by the sol–gel method.[11] On the other hand, χin shows a sharp decrease with x increasing from x = 0 to x =1, and a slow decreasing tendency with fluctuations for x greater than 1.0. The weakening of the magnetic parameters with Dy substitution is a consequence of the weak antiferromagnetic coupling between the RE magnetic sublattice and the net magnetization of the Fe sublattices, which leads to a sharper drop of the RE-sublattice magnetization (in comparison with the decrease of the magnetization of the Fe sublattices) as the temperature increases toward room temperature.[19]

Fig. 7. (color online) Initial magnetization curves of Y3−xDyxFe5O12 garnets.
Fig. 8. (color online) Saturation magnetization Ms and initial magnetic susceptibility χin as a function of x for Y3−xDyxFe5O12 garnets.

Differential scanning calorimetry (DSC) measurements are performed on all samples and the results are illustrated in Fig. 9. The thermal curves of all samples are almost smooth with small kinks (marked by arrows) in a temperature range of . Since abrupt changes in the thermal curves are usually associated with structural phase transitions or changes of the specific heat of the material under investigation, the positions of the small kinks in the thermal curves can be associated with temperatures at which the samples must have passed through transitions between phases with slightly different specific heats. In view of the fact that the magnetic contribution to the specific heat is rather small as reported for other magnetic oxides,[20] the observed kink in the thermal curve is associated with the Curie temperature at which ferrimagnetic–paramagnetic phase transition takes place.

Fig. 9. (color online) DSC versus T curves of Y3−xDyxFe5O12 garnets.

The Tc values for Y3Fe5O12 and Dy3Fe5O12 (276.0 °C and 278.3 °C, respectively) are in good agreement with previous experimental results on rare earth iron garnets.[6] Further, Tc does not seem to change appreciably with Dy concentration, and did not differ from the values found previously for Gd3−xErxFe5O12 either,[21] indicating that Tc is independent of rare-earth substitution and is determined only by the strong superexchange interaction between the two iron sublattices. If, however, we believe the small increase of Tc (from 276.0 °C to 278.3 °C) with x increasing, this increase could then be attributed to the slight strengthening of the superexchange interactions as a result of the additional weak interaction between the RE and Fe sublattices.

4. Conclusions

Dy-substituted YIG garnets are successfully synthesized by using the conventional solid state reaction method, where XRD patterns reveal the presence of a single garnet phase in each sample with no secondary phases. The lattice constant a and cell volume V increase with increasing Dy3+ substitution for Y3+ due to the larger ionic radius of Dy3+, whereas the increase of the molecular mass is responsible for the observed increase of the x-ray density of the substituted compound. RT Mössbauer spectra indicate that the substitution of Dy3+ for Y3+ at the dodecahedral c site does not perturb the Fe3+ wave function appreciably, keeping the hyperfine parameters practically unchanged. The saturation magnetization decreases almost linearly with Dy3+ substitution increasing, which is explained in terms of a simple magnetic dilution model. The Curie temperature determined from the DSC measurements is found to be nearly the same for all Dy substituted compounds.

Acknowledgment

The authors Mahdi Lataifeh and Ibrahim Abu-Aljarayesh would like to thank the Deanship of Research and Graduate Studies of Yarmouk University under Garnet No. 33/2015 used to execute part of the measurements (XRD and the thermal differential scanning calorimetry measurements) at Jordan University of Science and Technology.

Reference
[1] Özgür Ü Alivov Y Morkoç H 2009 J. Mater. Sci. Mater. Electron. 20 789
[2] Özgür Ü Alivov Y Morkoç H 2009 J. Mater. Sci. Mater. Electron. 20 911
[3] Mahmood S H Abu-Aljarayesh I 2016 Hexaferrite Permanent Magnetic Materials Millersville PA Materials Research Forum LLC p. 153 https://doi.org/10.21741/9781945291074
[4] Paoletti A 1978 Physics of Magnetic Garnets: Proceedings of the International School of Physics ‘Enrico Fermi’, Course LXX, Varenna on Lake Como, Villa Monastero, 27th June–9th July 1977 Elsevier Science & Technology https://doi.org/10.1002/crat.19790141215
[5] Winkler G D 1981 Magnetic garnets/Gerhard Winkler Braunschweig Vieweg
[6] Gilleo M A 1980 Ferromagnetic insulators: Garnets Ferromagnetic materials Wohlfarth P E Amstrdam North-Holland p. 1 https://doi.org/10.1016/S1574-9304(05)80102-6
[7] Parida S Rakshit S Singh Z 2008 J. Solid State Chem. 181 101
[8] Lataifeh M S 2008 Appl. Phys. 92 681
[9] Lataifeh M S Al-sharif A 1995 Appl. Phys. 61 415
[10] Lataifeh M McCausland M 1994 Mu’tah Lil-Buhooth Wa Al-Dirasat 9 23
[11] Cheng Z Yang H Cui Y Yu L Zhao X Feng S 2007 J. Magn. Magn. Mater. 308 5
[12] Xu H Yang H 2007 Phys. Status Solidi 204 1203
[13] Lataifeh M S Lehlooh A F D 1996 Solid State Commun. 97 805
[14] Lataifeh M S Mahmood S Thomas M F 2002 Physica B: Condens Matter 321 143
[15] Shannon R T 1976 Acta Cryst. 32 751
[16] Omari I Saleh A S Mahmood S H 1989 J. Magn. Magn. Mater. 78 183
[17] Xu H Yang H Xu W Yu L 2008 Curr. Appl. Phys. 8 1 http://dx.doi.org/10.1016/j.cap.2007.04.002http://dx.doi.org/10.1016/j.cap.2007.04.002
[18] Bouziane K Yousif A Widatallah H M Amighian J 2008 J. Magn. Magn. Mater. 320 2330
[19] Smit J Wijn H P J 1959 Ferrites New York Wiley
[20] Turki D Remenyi G Mahmood S Hlil E Ellouze M Halouani F 2016 Mater Res. Bull. 84 245
[21] Mohaidat Q I Lataifeh M Mahmood S H Bsoul I Awawdeh M 2017 J. Supercond. Novelmagn. 30 2135