Unequivocal evidence of enhanced magnetodielectric coupling in Gd3+ substituted multiferroic Bi2Fe4O9

S. R. Mohapatraa, A. Swainb, C. S. Yadavc, S. D. Kaushikd and A. K. Singh*a
aDepartment of Physics and Astronomy, National Institute of Technology, Rourkela-769008, Odisha, India. E-mail: singhanil@nitrkl.ac.in
bSchool of Physics, University of Hyderabad, Hyderabad-500046, India
cSchool of Basic Sciences, Indian Institute of Technology Mandi, Mandi-175001, H.P., India
dUGC-DAE Consortium for Scientific Research Mumbai Centre, R-5 Shed, BARC, Mumbai-400085, India

Received 2nd October 2016 , Accepted 22nd November 2016

First published on 22nd November 2016


Abstract

We report an enhanced magnetodielectric (MD) coupling in antiferromagnetic Bi2Fe4O9 (BFO) lightly substituted by gadolinium (Gd3+). Rietveld refined X-ray diffraction and Raman spectra reveal the formation of orthorhombic phase for all the synthesized samples. Gd3+ substitution results in a notable increase of antiferromagnetic transition (TN) from 250 K (x = 0, BFO) to 256 K (x = 0.02, BGFO2). At the same time, deviation from linear MH behaviour in the substituted samples at T ≤ 30 K implies the onset of weak ferromagnetic ordering, along with a sudden rise in magnetisation around ∼30 K. This result indicates the magnetically active nature of Gd3+ ions possessing a large magnetic moment of ∼8 μB and this is explained by considering the two sub-lattice mode. The temperature dependence of the dielectric study shows unusual oppression in ε′ and tan loss plots at TTN which hints a plausible MD coupling in the above studied samples. A contrasting behaviour in the temperature dependant MD% is seen for BFO and the substituted samples have an enhancement of ∼30% for BGFO2 at 300 K. Further, confirmation to this coupling is drawn by studying frequency and magnetic field dependant MD% at 300 K, where it is found that both obey a power law of the form ∝exp(ω)n and ∝(H)m respectively. Interestingly, the values of ‘n’ and ‘m’ obtained from fitting are found to lie in the range 0.45–0.5, thus signifying a similar dependence of MD on frequency as well as magnetic field. Lastly, the effect of Gd3+ substitution on PE loop of BFO and its substituted samples at room temperature have been discussed.


1. Introduction

In recent decades, magnetoelectrically coupled multiferroic materials have been under the intensive focus of contemporary academic interest owing to their diversified physical aspects along with remarkable multifunctional device applications.1–5 These materials manifest the interaction of magnetic and electrical order parameters, often known as ‘magnetoelectric (ME) effect’.2 The coupling of these two ferroic order parameters lends an additional degree of freedom in device designing where magnetic and electric polarization can be sensitively tuned with electric and magnetic field, respectively.6 Moreover, the simultaneity of these ferroic orders at room temperature is of both fundamental and technological interest. However, the natural occurrence of room temperature ME materials possessing novel technological applications is very uncommon due to their mutually exclusive features.2 Although, several single-phase systems such as BiFeO3, BiFeO3-based compositions, hexagonal RMnO3 (R = rare earth) and LuFe2O4 have been explored, they are found to exhibit weak coupling between the order parameters due to the occurrence of magnetic and electric ordering at different temperatures.7–11 Usually, most of the known ME multiferroic systems till date are functional much below room temperature. So, the quest for developing single phase multiferroic materials with optimum ME effect at ambient temperature still continues and is a matter of practical challenge.

The compound of our interest is multiferroic Bi2Fe4O9 (BFO), a well-known antiferromagnetically ordered system and a material prototype of Cairo spin lattice.12–14 BFO has been widely studied over the past few decades and shows numerous functional applications.15–17 The most peculiar feature exhibited by BFO is the unique pentagon spin frustration which arises due to the interactions among different kind of Fe3+ ions present in the BFO unit cell, thus leading to a non-collinear magnetic structure.18 Fig. 1(a) shows the unit cell of BFO along with the direction of spins (arrows) of different kind of Fe3+ ions. In fact, BFO comprises of two stacked octahedral coordinated Fe atoms (denoted as Fe(O)). Here, we observe four tetrahedral coordinated Fe3+ ions (Fe(T)) in the interior of the unit cell whereas on the sides of the cell there are four octahedral coordinated Fe3+ ions. It can also be seen that the spins of tetrahedral Fe3+ ions interact antiferromagnetically among themselves denoting the superexchange interaction while spins of octahedral Fe3+ ions interact ferromagnetically (FM) within a pair of octahedral spins. These competing exchange interactions of different kinds of Fe atoms generate pentagon spin frustration in the BFO system. Moreover, Fe(T) being antiferromagnetically interacting among themselves are considered to display the strongest coupling.15 However, despite being well-known for its multiferroicity, the urge for further improving its properties and achieve room temperature magnetodielectric effect has motivated researchers to chemically substitute Bi and Fe-site of BFO.19–22


image file: c6ra24525k-f1.tif
Fig. 1 (a) A unit cell of Bi2Fe4O9 showing schematic crystal structure and magnetic spins in ab plane along c-axis where the competing exchange interactions between octahedral (Fe(O)) and tetrahedral (Fe(T)) coordinated Fe ions generates pentagon spin frustration. (b) Rietveld refinement of room temperature XRD patterns of BFO, BGFO0.5, BGFO1, BGFO1.5 and BGFO2. (c) The magnified pattern of (002) and (220) diffraction peaks.

The main aim of this manuscript is to investigate the dielectric, magnetic and magnetodielectric properties of Bi2(1−x)Gd2xFe4O9 (0 ≤ x ≤ 0.02) compounds. In the literature, it is well established that the doping of Gd (a rare-earth magnetic element) at Bi-site of BiFeO3 is very effective in increasing the magnetization.23,24 Also, Gd3+ ion is magnetically active and has larger effective magnetic moment (∼8 μB) as compared to Fe3+ ions (∼6 μB). Therefore, substitution of smaller ionic radii i.e. Gd3+ (1.05 Å) at Bi3+-site (1.17 Å) and the coupling between Gd3+ and Fe3+ ions would be interesting to search for emerging physical properties. In the present context, we report enhanced antiferromagnetic transition (TN) temperature along with the unambiguous evidence of ME coupling and enhanced magnetodielectric effect at room temperature in Gd3+ substituted Bi2Fe4O9 ceramics.

2. Experimental procedure

The polycrystalline samples of Bi2(1−x)Gd2xFe4O9 (0 ≤ x ≤ 0.02) were prepared using high purity oxides (>99.9%, Sigma Aldrich) namely Bi2O3, Gd2O3 and Fe2O3 by conventional solid state reaction route. The above raw materials taken in stoichiometric proportions were mixed thoroughly, grounded for 2 h and calcined at 800 °C for 12 h followed by a sintering at 850 °C for 10 h. The sintered samples were characterized by room temperature X-ray diffraction (XRD) measurement, RIGAKU JAPAN for detecting the phase formation. XRD was carried out using Cu-Kα radiation (λ = 1.54 Å) in a wide range of Bragg angles (10° ≤ 2θ ≤ 70°) with a step size of 0.002° and at a slow scan rate of 3° min−1. The room temperature Raman spectra were recorded using an Nd-Yag 532 nm laser in the back scattering geometry in a spectrometer (WiTec, alpha 200) equipped with a confocal microscope and CCD camera. The magnetization measurements were done using cryogen free 7T-SQUID-VSM, Magnetic Property Measurement System (Quantum Design, USA). The dielectric measurements were performed using high precession impedance analyser (Wayne Kerr 6500B) over a frequency range of 0.1 kHz to 100 kHz. The temperature variation from 15–320 K was attained using closed cycle refrigerator (Cryo industries, USA). The magnetodielectric measurements were carried out in presence of an external magnetic field (up to 1.3 T) using an electromagnet (GMW 5403) equipped with bipolar DC power supply. Finally, room temperature ferroelectric property (polarizations–electric field loops) of the samples were measured on a standard ferroelectric tester using a Radiant precision premier II unit.

3. Results and discussion

3.1 X-ray diffraction (XRD) study

Fig. 1(b) shows the room temperature Rietveld refined X-ray diffraction (XRD) patterns of parent (BFO) as well as the gadolinium substituted samples (Bi2(1−x)Gd2xFe4O9 (0 ≤ x ≤ 0.02)). The synthesized samples are abbreviated depending on x (Gd) concentration, such as BFO (x = 0), BGFO0.5 (x = 0.005), BGFO1 (x = 0.01), BGFO1.5 (x = 0.015) and BGFO2 (x = 0.02). All the XRD pattern reveals a well crystallized orthorhombic phase with space group ‘Pbam’ (JCPDS no. 76-1098). The Rietveld refinement was done using FULLPROF software25 to study the detailed crystal structure and the fitting parameters are listed in Table 1. Substitution of Gd3+ ions having lower ionic radius than that of Bi3+ ions results in reduction of lattice parameters and cell volume as seen from Table 1. A closer view of obtained XRD peaks around 29–32° in Fig. 1(c) shows slight shifting of the main peaks i.e. (002) and (220) to higher 2θ angles with increase in x concentration. This feature can also be attributed to lower ionic radii substitution thus, confirming substitution of Gd3+ ions at Bi3+ site. Also, the variation in tetrahedral and octahedral coordinated Fe–O bond lengths and O–Fe–O bond angles due to substitution are estimated and listed in Table 2. The comparative variation of Fe(T)–O as well as Fe(O)–O bond lengths in BFO and BGFO2 are schematically depicted in Fig. 2(a) and (b). Here, we observe a slight increase (∼0.2%) in average bond length of Fe–O due to Gd3+ substitution in BGFO2 samples with respect to BFO. This has given an insight to explore the effect of Gd3+ substitution on the bond lengths of strongly coupled tetrahedral coordinated Fe3+ ions i.e. Fe(T)–O as discussed in Fig. 1(a). Interestingly, we observed significant reduction in Fe(T)–O bond lengths of BGFO2 sample as compared to that of parent (BFO) sample. This feature is expected to play a vital role in governing the magnetic and magnetodielectric properties of substituted samples which will be discussed later.
Table 1 Rietveld refined X-ray diffraction parameters of Bi2(1−x)Gd2xFe4O9 (0 ≤ x ≤ 0.02)
Samples Lattice parameters χ2
a (Å) b (Å) c (Å) Vol (Å)
BFO 7.97350 8.44142 6.00277 404.0324 3.66
BGFO0.5 7.97316 8.44124 6.00272 404.0032 3.01
BGFO1 7.97279 8.44118 6.00238 403.9587 2.71
BGFO1.5 7.96758 8.43569 5.99818 403.1498 3.13
BGFO2 7.96737 8.43496 5.99978 403.2118 3.06


Table 2 Different bond lengths and bond angles of BFO and BGFO2 samples obtained from the Rietveld refinements of respective XRD patterns
Samples BFO BGFO2
Bond lengths (Å)
Fe(O)–O2 (×2) 1.956 2.064
Fe(O)–O3 (×2) 1.949 1.989
Fe(O)–O4 (×2) 2.048 2.045
Fe(T)–O1 1.837 1.808
Fe(T)–O2 (×2) 1.930 1.876
Fe(T)–O3 1.847 1.736
〈Fe–O〉a 1.945 1.949
[thin space (1/6-em)]
Bond angles (degree)
O3–Fe(O)–O2 87.51 92.85
O2–Fe(O)–O4 92.60 94.17
O3–Fe(O)–O4 97.31 96.72
O4–Fe(O)–O4 82.25 75.39
O2–Fe(O)–O2 175.07 178.21
O3–Fe(O)–O3 83.12 92.05
O1–Fe(T)–O2 113.25 114.71
O2–Fe(T)–O2 106.05 105.45
O3–Fe(T)–O2 108.18 104.26
O1–Fe(T)–O3 111.72 112.35



image file: c6ra24525k-f2.tif
Fig. 2 Fe(T)–O and Fe(O)–O bond lengths of (a) BFO and (b) BGFO2 samples.

3.2 Raman spectroscopy study

To have better insight to the vibrational, structural and lattice properties, the room temperature Raman spectra of parent (BFO) and its substituted samples (BGFO1 and BGFO2) were performed at wave number ranging from 100–1000 cm−1 and is illustrated in Fig. 3(a). Using group theoretical irreducible representation, Iliev et al. have estimated 42 Raman modes (Γ = 12Ag + 12B1g + 9B2g + 9B3g) for orthorhombic Bi2Fe4O9.26 But, due to the insufficient intensity arising from small polarizability of several modes, we could observe less number of modes as compared to above group theoretical calculations. As shown in Fig. 3, all the samples exhibits similar Raman vibrations, confirming the ‘Pbam’ symmetry with seven dominant Raman modes. All the observed Raman modes for parent and substituted samples are listed in Table 3. However, a close look at the spectra shows a slight shifting of major Raman modes to higher wave number (shown in dashed arrows) along with decrease in peak intensity with increase in Gd content. The reason for slight shifting of Raman modes could be attributed to the substitution of Gd atom (∼157 u) which is lighter than Bi atom (∼209 u) and lattice disorder in terms of contraction of unit cell volume as detected from XRD analysis. On the other hand, stronger force constants in general causes blue shift of Raman spectra lines.27 Moreover, decrease in peak intensity with increase in Gd3+ content may be due to substitutional disorder which has been recently argued to cause broadening of Raman modes and reduction in intensities.28 Although, the reduction in peak intensities in the present case is not very significant due to small amount of substitution. Another possible reason could be attributed to the decrease in dipole moment (arising due to Bi3+ lone s2 pair) in substituted samples leading to reduction in intensities. Here, we found that the experimentally obtained Raman modes are in good agreement with the predicted values.26 In fact, the low frequency modes (up to 167 cm−1) are assigned to Bi atoms, high frequency modes (above 262 cm−1) involve mainly oxygen-stretching motions while the Fe atoms actively participate in the modes between 152 cm−1 and 262 cm−1.29 Furthermore, we observed suppression of peak at ∼165 cm−1 while peaks at ∼360 cm−1 and ∼640 cm−1 stabilizes with increase in x content. In order to clearly assign the active Raman modes, we have deconvoluted Raman spectra of BGFO2 sample as shown in Fig. 3(b). The spectra is fitted to the sum of eight Lorentzian peaks with multiple Lorentzian oscillators of the form I = I0 + ∑AiΓi/[4(ωωi)2 + Γi2] where, i is the peak number, I0 is the background intensity, ωi is the centre frequency, Γi is the full width at half maxima (FWHM), and Ai is the area of ith peak.30 The deconvoluted Raman spectra of BGFO2 with seven major Raman modes at 204 cm−1 (Ag(3), B1g(3), B2g(3), B3g(3)), 278 cm−1 (B1g(4)), 325 cm−1 (Ag(6), B1g(5), B2g(6)), 364 cm−1 (Ag(7), B1g(6)), 429 cm−1 (Ag(8), B1g(7), B2g(7), B3g(6)), 552 cm−1 (Ag(9), B1g(8), B3g(8)) and 641 cm−1 (Ag(10), B1g(9)) are discernible from the Fig. 3(b).
image file: c6ra24525k-f3.tif
Fig. 3 (a) Room temperature Raman spectra of BFO, BGFO1 and BGFO2 samples and (b) deconvoluted Raman spectrum of BGFO2 fitted to the sum of eight Lorentzian peaks.
Table 3 Various active Raman modes observed for Bi2(1−x)Gd2xFe4O9 (x = 0, 0.01 and 0.02) samples recorded at room temperature
Active Raman modes Raman shifts (cm−1)
BFO BGFO1 BGFO2
Ag(3), B1g(3), B2g(3), B3g(3) 202 203 204
B1g(4) 276 277 278
Ag(6), B1g(5), B2g(6) 321 323 325
Ag(7), B1g(6) 359 362 364
Ag(8), B1g(7), B2g(7), B3g(6) 426 427 429
Ag(9), B1g(8), B3g(8) 549 551 552
Ag(10), B1g(9) 638 640 641
Ag(11) 724 725 727


3.3 DC magnetization study

Fig. 4(a–c) illustrates temperature dependant DC magnetization plot of BFO, BGFO1 and BGFO2 under zero field cooled (ZFC) and field cooled warming (FCW) conditions performed at an external field of 1 kOe. A broad cusp around 250 K is noticed in the parent (BFO) sample, characteristically revealing the antiferromagnetic (AFM) behaviour and is consistent to previous reports,13,15,31 while for the substituted samples, the AFM transition (TN) is found to increase with substitution. To ascertain the exact AFM transition, the dχ/dT versus temperature is plotted which confirms the AFM transition (TN) of BFO = 250 K, BGFO1 = 252 K and BGFO2 = 256 K as shown in the insets of Fig. 4(a), (b-i) and (c-i) respectively. The variation of AFM transition (TN) with concentration (x) is depicted in Fig. 4(d). The origin of enhancement in the AFM transition (TN) upon Gd3+ substitution may be attributed to the Fe–Gd interaction. Moreover, as observed in the XRD analysis, increase in Gd3+ concentration leads to decrease in lattice parameters and reduction in strongly antiferromagnetically coupled Fe(T)–O bond lengths. This might be modifying the Fe(T)–O–Fe(T) super-exchange interaction in terms of the lattice distortion in the unit cell causing enhancement in TN. Also, BFO is known to be magnetically frustrated system showing a unique pentagon spin frustration and the degree of frustration (f) in a spin frustrated system is generally defined as f = |θCW|/TN, where |θCW| is the Curie–Weiss temperature.32 The value of f for BFO is determined to be ∼9 revealing a highly frustrated AFM spin in the material but a decrease in f parameter with increase in substitution can be clearly seen from Fig. 4(d). A clear quantitative picture of f parameter can be obtained by performing high temperature χT measurement in order to have a wide range of paramagnetic region, which is beyond the scope of present study. It is to be noted that ZFC–FCW of BFO exactly superimposes thus ruling out the possibility of magnetic glassiness or spin glass behaviour in the material. Whereas, a slight mismatch of ZFC–FCW curve is observed in case of BGFO1 samples. So as to confirm the possibility of magnetic glassiness, we have performed the ac susceptibility (χac) measurement of BGFO1 sample at selected frequencies as shown in inset of Fig. 4(d). Usually, for a spin glass system, χac versus temperature gives frequency dependant behaviour, but no such feature is seen for BGFO1 sample which implies absence of spin glass or magnetic glassiness in the materials. One of the interesting features in Gd substituted samples is that the magnetization sharply increases below ∼30 K which may be due to the ordering of Gd3+ magnetic moments.33 The enhanced magnetisation due to Gd substitution is also seen in isothermal MH plots.
image file: c6ra24525k-f4.tif
Fig. 4 (a–c) dc magnetic susceptibility (χ) versus temperature (80–325 K) of BFO, BGFO1 and BGFO2 samples under 1 kOe. Insets [(a-i), (b-i) & (c-i)] show the corresponding dχ/dT plots indicating the AFM transition (TN) and [(b-ii) & (c-ii)] shows χ versus T plot for a temperature range of 2–325 K respectively. (d) Variation of AFM transition (TN) and frustration parameter (f) with concentration (x) and its inset display χac versus T of BGFO1 sample at several static frequencies.

Fig. 5(a–c) shows MH curves of BFO, BGFO1 and BGFO2 samples at selected temperatures up to a maximum field of 60 kOe. A linear behaviour of magnetisation can be seen for BFO sample for all the recorded temperatures typically revealing its AFM/PM characteristics, as discussed earlier. Interestingly, the substituted samples also shows a linear magnetisation which signifies AFM/PM nature being intact for 30 K ≤ T ≤ 320 K but for T < 30 K, magnetisation shows a non-linear curve implying an onset of weak FM in the samples. Eventually, the presence of weak FM ordering in the substituted samples were evaluated from the Arrott plots (M2 versus H/M) based on the Weiss-molecular field theory.34,35 Fig. 5(d) shows the Arrott plots of BFO (inset), BGFO1 and BGFO2 samples at 5 K. A consistent linear increase for BFO is observed whereas BGFO1 and BGFO2 show a non-linear increase of the curves. Also, the linear extrapolation of the high field region of M2H/M plot of both the samples yields a negative intercept on the M2 axis which specifies the crossover from AFM/PM to weak FM ordering.36 Further, to strengthen the indications of the growing weak FM interactions along with AFM/PM, we have analysed the MH loop K using the following expressions consisting of both FM and AFM/PM parts:37

 
image file: c6ra24525k-t1.tif(1)
where, MSFM is FM saturation magnetisation, MRFM is remanant magnetisation, Hci is intrinsic coercivity, χ is magnetic susceptibility for AFM/PM part, M is the observed magnetisation and H is the applied magnetic field. Here, first term represents FM component and the second term is related to the linear component representing AFM/PM contribution. Inset of Fig. 5(b) and (c) exhibits the fitting of MH curve of BGFO1 and BGFO2 samples respectively at 5 K using the above eqn (1). From the fitting, the values of MSFM, MRFM and Hci is determined to be ∼2.7 emu g−1, ∼0.02 emu g−1 and ∼575 Oe for BGFO1 while ∼3.6 emu g−1, ∼0.7 emu g−1 and ∼1035 Oe for BGFO2. Also, AFM/PM susceptibility (χ) increases monotonically from ∼3.3 × 10−5 emu g−1 Oe−1 to ∼5 × 10−5 emu g−1 Oe−1, as x increases from x = 0.01 to 0.02 respectively.


image file: c6ra24525k-f5.tif
Fig. 5 The isothermal MH curves of (a) BFO, (b) BGFO1 and (c) BGFO2 samples at several fixed temperatures up to 60 kOe. Inset (b and c) illustrates experimental MH curve of BGFO1 and BGFO2 samples respectively at 5 K with the fitted (solid line) and theoretically simulated AFM/PM and FM part (dotted lines). (d) Arrott plots of BFO, BGFO1 and BGFO2 samples at 5 K.

It is to be noted that value of magnetisation increases with increase in x content and a sudden rise in magnetisation is more prominently seen towards the lowest temperature i.e. T < 30 K (insets of Fig. 4(b-ii and c-ii)). This can be explained by considering two sub-lattice mode model.30,38 In the present studied compound, we have Fe as well as Gd sub-lattice which have their own temperature dependence of magnetisation. Usually, Gd being a rare earth element, orders at preferably lower temperature.39 Generally, AFM material comprising of two spin sub-lattices have zero net magnetic moment due to complete compensation of spins between these two sub-lattices. However, if any resultant small magnetic moment arises in AFM materials, then it could be possibly due to uncompensated spins in the two sub-lattices. In our case, substitution of Gd3+ causes enhancement in magnetisation below T ∼ 30 K because Gd3+ having large effective magnetic moment (∼8 μB) than Fe3+ (∼5.9 μB) orders FM at preferably low temperature and coupling between Gd3+ and Fe3+ ions (f–d coupling) creates a possibility of enhanced magnetization.

3.4 Temperature dependant dielectric study

Fig. 6(a–c) depicts the temperature dependant (150–300 K) dielectric permittivity (ε′) in the main plot for BFO, BGFO1 and BGFO2 samples at several static frequencies. The insets of the respective plots show the variation of tan loss and dε′/dT in the same temperature window. For all the above samples, the values of ε′ decreases with the decrease in temperature i.e. ∼27 (BFO), ∼23 (BGFO2) at room temperature to ∼25.5 (BFO), ∼21.5 (BGFO2) at 150 K. The dielectric plot of parent as well as modified samples can be divided in two regions namely the low temperature almost linearly rising dispersion-less region (T < 180 K) and dispersedly rising region for T > 180 K. From the dielectric data, we observe a slight oppression (shown in circle) in the studied samples and to exactly trace the dielectric transition or peak, derivative of ε′ at 12 kHz of all the samples are shown in the respective insets. The dε′/dT plot shows a sharp peak at 250 K, 252 K and 256 K for BFO, BGFO1 and BGFO2 respectively, which also happens to be the AFM transition (TN). Along with, a change in slope around TN was seen in the respective tan loss plots taken at 12 kHz. This strongly indicates a plausible coupling of magnetic and electric order parameters in our samples. Moreover, Fig. 6(d) shows temperature and field dependant tan loss plot of parent and BGFO2 samples which indicate that with the application of magnetic field, tan loss factor decreases and a clear distinction of broad shoulders around TN can be seen. This further provides the evidence of ME coupling in BFO as well as in BGFO1 and BGFO2.
image file: c6ra24525k-f6.tif
Fig. 6 (a–c) Temperature dependant ε′ of BFO, BGFO1 and BGFO2 samples and respective insets show tan loss and dε′/dT of the same. (d) tan loss factor is plotted as a function of temperature at 0 T and 1.3 T for BFO (upper panel) and BGFO2 (lower panel). Dotted circles and arrows indicate the TN while dotted lines are guide to the eye.

3.5 Magnetodielectric study

Strong evidence of ME coupling encouraged us to observe the influence of magnetic field on the dielectric property and investigate the magnetodielectricity in our samples. The magnetodielectric (MD) effect is expressed as:40
 
image file: c6ra24525k-t2.tif(2)
where, ε′(H) and ε′(0) are dielectric permittivity with applied field (1.3 T) and zero field respectively. Fig. 7(a) shows temperature variation of MD% at 12 kHz for BFO (inset), BGFO1 and BGFO2. All the samples exhibited negative MD effect along with a contrasting behaviour in case of parent and substituted samples. The parent BFO sample shows almost negligible MD effect at low temperatures but as temperature approaches TN (250 K), a sudden rise in negative MD is seen which continues till 300 K and the MD% value becomes ∼−0.3%. On the contrary, BGFO1 and BGFO2 shows an abrupt change in MD values, one at lower temperature i.e. ∼50 K (marked in dotted circle) which might be due to the ordering of Gd3+ as discussed above, and the other around magnetic transition (TN, marked by an arrow). In BGFO1 and BGFO2, as temperature approaches TN, the magnitude of MD% increases with rise in temperature and reaches room temperature with values −0.31% (BGFO1) and −0.39% (BGFO2). This shows a remarkable enhancement of ∼7% and ∼30% in MD% for BGFO1 and BGFO2 samples respectively at 300 K as compared to that of BFO sample.

image file: c6ra24525k-f7.tif
Fig. 7 (a) Temperature dependent MD% of BFO (inset), BGFO1 and BGFO2 samples at 12 kHz. Dotted circle indicates Gd3+ ordering. (b) Frequency variation of MD% under H = 1.3 T of BFO, BGFO1 and BGFO2. The dotted lines show the power law fitting of the form MD% ∝ exp(ω)n. (c) Field variation of MD% of BFO, BGFO1 and BGFO2. The solid lines are fitted using the power law of the form MD% ∝ (H)m. (d) Variation of MD% with M2 for BFO (inset), BGFO1 and BGFO2 samples.

Now, we shall discuss the variation of MD% with frequency (0.1–100 kHz) at room temperature as shown in Fig. 7(b). A negative MD effect is seen for BFO, BGFO1 and BGFO2 samples. The MD% for BFO shows a maximum value of ∼−6% at 0.1 kHz whereas the substituted samples shows a very less value (<35%), but with increase in frequency MD% approaches to zero varying non-linearly in the log scale of frequency. Further, to analyse the frequency dependant behaviour, our MD plot was fitted using power law of the form, MD% ∝ exp(ω)n, distinctly for lower frequency regions (<10 kHz) and higher frequency regions (>10 kHz). The low frequency process is vulnerable to space-charge polarizations, leakage currents, dc conductivity etc., whereas high frequency process may be regarded as the intrinsic contribution. The exponent n ∼ −0.2 shows a reasonably good fit for low frequency regions while n ∼ −0.5 fits well for higher frequency regions for all the samples. Thus, the MD result suggests that both the lower and higher frequency processes are responsible for observed MD in these samples.

To elucidate the above feature, field dependant MD% is done for BFO, BGFO1 and BGFO2 samples at room temperature. Fig. 7(c) shows variation of MD% versus field at 50 kHz in order to nullify the effect of interfacial space-charge artifacts.41 The change in MD% observed with a change in the applied magnetic field indicates strong coupling of electric and magnetic order parameters. We observe a negative MD% in all the cases with values ∼−0.03%, ∼−0.06% and ∼−0.12% for BFO, BGFO1 and BGFO2 samples at an applied field of 1.3 T. The field dependence of MD% is found to obey power law of the form MD% ∝ (H)m and the value of ‘m’ is found to lie in the range of 0.45–0.5 for all the studied samples. The value of ‘m’ obtained is very close to that of the value of ‘n’ in the frequency dependant MD plot of the intrinsic region (>10 kHz), thus confirming a similar dependence of MD on frequency as well as the applied magnetic field. Also, from Fig. 7(c), it is clear that the relationship between MD% and H is nearly quadratic in nature and can be explained in the framework of the Ginzburg–Landau free energy theory for the second order phase transition. This free energy expression can be used to determine the magnetic ordering effect on dielectric susceptibility. It was concluded that the magnetic ordering effect on a dielectric constant will be proportional to γM2 (i.e., MD% ∼ γM2).42 Fig. 7(d) shows the linear variation of MD% with the square of magnetization (M2) for BFO, BGFO1 and BGFO2 samples which clearly emphasises that the MD effect emerges from the magnetoelectric coupling term ‘γP2M2’ obtained from free energy expression. From linear fitting, the absolute values of γ was calculated to be ∼0.35 (emu g−1)−2, ∼0.7 (emu g−1)−2 and ∼1.55 (emu g−1)−2 for BFO, BGFO1 and BGFO2 respectively.

3.6 Ferroelectric study

Since, BFO and Gd3+ substituted samples show room temperature MD effect, we expect the ferroelectric hysteresis behaviour for the same. Also, BFO has been reported to exhibit room temperature ferroelectric behaviour.16,19 Fig. 8 shows PE hysteresis loops of BFO, BGFO1 and BGFO2 samples at room temperature. No saturation in hysteresis curve could be obtained for all the samples up to the maximum applied electric field of ±40 kV cm−1 which implies partial reversal of the polarisation. A clear yet unsaturated hysteresis curve is observed for BFO while the highest substituted sample (BGFO2) shows a slim PE loop. In case of BFO, it has been reported that a lone s2 pair of electrons of the Bi3+ ions hybridizes with 2p- orbital of O2− ions and forms a localised lobe which leads to the non-centrosymmetric distortions and causes ferroelectricity.19 But with gradual increase in x content, we obtain slim PE loops and decrease in remanant polarisation (2Pr) and coercive field (2Ec). This shows that the substitution has a significant effect on the ferroelectric property of the system which suppresses the Bi3+ ions lone s2 pair effect. Moreover, the slim PE loop in case of BGFO2 suggest a comparatively less insulating behaviour of the sample and this fact is also revealed from the increased value of tan loss in BGFO2 as compared to BFO. Though the main plot of Fig. 8 shows a slim PE loop but a smaller hysteresis in PE loop can be seen for BGFO2 sample as illustrated in inset of Fig. 8. For BFO remanant polarisation (2Pr) and coercive electric field (2Ec) are found to be 0.119 μC cm−2 and 16.89 kV cm−1 respectively while the values are 0.05 μC cm−2 (2Pr) and 5.83 kV cm−1 (2Ec) for BGFO1 and 0.009 μC cm−2 (2Pr) and 1.17 kV cm−1 (2Ec) for BGFO2.
image file: c6ra24525k-f8.tif
Fig. 8 Room temperature polarization hysteresis (PE) loops of BFO, BGFO1 and BGFO2 samples. Inset shows magnified PE loop of the above samples up to a range of ±4.5 kV cm−1.

4. Conclusion

We have investigated structural, magnetic, magnetodielectric and ferroelectric behaviour of Bi2(1−x)Gd2xFe4O9 (0 ≤ x ≤ 0.02) ceramics. XRD confirms the formation of well crystalized samples of orthorhombic phase along with the decrease in lattice parameters while Raman spectra depicts slight shifting of Raman modes to higher wave number due to partial substitution of Gd3+ at Bi3+-site. The temperature dependence of magnetization and dχ/dT plot shows a sharp AFM (TN) transition which is found to increase with Gd3+ substitution i.e. from 250 K (BFO) to 256 K (BGFO2). The values of magnetization are found to increase rapidly towards lower temperature (T < 30 K) and is explained in terms of two sub-lattice mode model. Also, the magnetic measurements disclose the onset of weak FM along with AFM phase in the system for T < 30 K. Moreover, evidence of plausible magnetodielectric coupling in all the samples are observed from the concomitant anomaly realized in the temperature dependant dielectric and tan loss data around TN. At room temperature, enhanced magnetodielectricity (MD) of ∼7% and ∼30% are determined for x = 0.01 and 0.02 respectively with respect to parent (BFO) sample. Further, frequency and field dependant MD at 300 K is studied to establish a strong coupling of dielectric and magnetic order parameters and both are found to obey power law of the form ∝exp(ω)n and ∝(H)m respectively. Thus, we expect that the above striking features would make it a worthy candidate for various multifunctional device applications.

Acknowledgements

AKS acknowledges the Board of Research in Nuclear Science (BRNS), Mumbai (Sanction No. 2012/37P/40/BRNS/2145), UGC-DAE-CSR Mumbai (Sanction No. CRS-M-187, 225) and Department of Science and Technology (DST), New Delhi (Sanction No. SR/FTP/PS-187/2011) for funding. AKS acknowledges Dr P. Kumar for PE loop measurement. Lastly, SRM is thankful to the BRNS for the financial assistance.

References

  1. N. A. Hill, J. Phys. Chem. B, 2000, 104(29), 6694–6709 CrossRef CAS.
  2. W. Erenstein, N. D. Mathur and J. F. Scott, Nature, 2006, 442, 759–765 CrossRef PubMed.
  3. D. Khomskii, Physics, 2009, 2, 20 CrossRef.
  4. N. A. Spaldin, S.-W. Cheong and R. Ramesh, Phys. Today, 2010, 63, 38 CrossRef.
  5. H. J. Zhao, W. Ren, Y. Yang, J. Íñiguez, X. M. Chen and L. Bellaiche, Nat. Commun., 2014, 5, 4021 CAS.
  6. K. Dey, A. Indra, D. De, S. Majumdar and S. Giri, ACS Appl. Mater. Interfaces, 2016, 8, 12901–12907 CAS.
  7. C. Ederer and N. A. Spaldin, Phys. Rev. B: Condens. Matter Mater. Phys., 2005, 71, 060401 CrossRef.
  8. S. N. Tripathy, D. K. Pradhan, K. K. Mishra, S. Sen, R. Palai, M. Paulch, J. F. Scott, R. S. Katiyar and D. K. Pradhan, J. Appl. Phys., 2015, 117, 144103 CrossRef.
  9. R. A. M. Gotardo, L. F. Cótica, I. A. Santos, M. Olzon-Dyonisio, S. D. Souza, D. Garcia, J. A. Eiras and A. A. Coelho, Appl. Phys. A, 2013, 111, 563–567 CrossRef CAS.
  10. A. K. Singh, S. Patnaik, S. D. Kaushik and V. Siruguri, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 81, 184406 CrossRef.
  11. J. Yong Park, J. Hwan Park, Y. Kyu Jeong and H. M. Jang, Appl. Phys. Lett., 2007, 91, 152903 CrossRef.
  12. N. Niizeki and M. Wachi, Z. Kristallogr., 1968, 127, 173–187 CrossRef CAS.
  13. S. R. Mohapatra, B. Sahu, T. Badapanda, M. S. Pattanaik, S. D. Kaushik and A. K. Singh, J. Mater. Sci.: Mater. Electron., 2016, 27, 3645–3652 CrossRef CAS.
  14. I. Rousochatzakis, A. M. Lauchli and R. Moessner, Phys. Rev. B: Condens. Matter Mater. Phys., 2012, 85, 104415 CrossRef.
  15. A. K. Singh, S. D. Kaushik, B. Kumar, P. K. Mishra, A. Venimadhav, V. Siruguri and S. Patnaik, Appl. Phys. Lett., 2008, 92, 132910 CrossRef.
  16. S. R. Mohapatra, B. Sahu, M. Chandrasekhar, P. Kumar, S. D. Kaushik, S. Rath and A. K. Singh, Ceram. Int., 2016, 42, 12352 CrossRef CAS.
  17. S. Sun, W. Wang, L. Zhang and M. Shang, J. Phys. Chem. C, 2009, 113, 12826–12831 CAS.
  18. E. Ressouche, V. Simonet, B. Canals, M. Gospodinov and V. Skumryev, Phys. Rev. Lett., 2009, 103, 267204 CrossRef CAS PubMed.
  19. Z. M. Tian, Y. Qiu, S. L. Yuan, M. S. Wu, S. X. Huo and H. N. Duan, J. Appl. Phys., 2010, 108, 064110 CrossRef.
  20. C. M. Raghavan, J. W. Kim, J. Y. Choi, J.-W. Kim and S. S. Kim, Ceram. Int., 2014, 40, 14165–14170 CrossRef.
  21. S. Huang, L. R. Shi, Z. M. Tian, S. L. Yuan, C. M. Zhu, G. S. Gong and Y. Qiu, J. Solid State Chem., 2015, 227, 79–86 CrossRef CAS.
  22. M. G. Rozova, V. V. Grigoriev, I. A. Bobrikov, D. S. Filimonov, K. V. Zakharov, O. S. Volkova, A. N. Vasiliev, E. V. Antipov, A. A. Tsirlin and A. M. Abakumov, Dalton Trans., 2016, 45, 1192–1200 RSC.
  23. R. Guo, L. Fang, W. Dong, F. Zheng and M. Shen, J. Phys. Chem. C, 2010, 114, 21390–21396 CAS.
  24. P. Uniyal and K. L. Yadav, Mater. Lett., 2008, 62, 2858–2861 CrossRef CAS.
  25. H. M. Rietveld, J. Appl. Crystallogr., 1969, 2, 65–67 CrossRef CAS.
  26. M. N. Iliev, A. P. Litvinchuk, V. G. Hadjiev, M. M. Gospodinov, V. Skumryev and E. Ressouche, Phys. Rev. B: Condens. Matter Mater. Phys., 2010, 81, 024302 CrossRef.
  27. A. Ubaldini and M. M. Carnasciali, J. Alloys Compd., 2008, 454, 374–378 CrossRef CAS.
  28. K. K. Mishra, A. K. Arora, S. N. Tripathy and D. Pradhan, J. Appl. Phys., 2012, 112, 073521 CrossRef.
  29. P. Hermet, M. Goffinet, J. Kreisel and P. Ghosez, Phys. Rev. B: Condens. Matter Mater. Phys., 2007, 75, 220102 CrossRef.
  30. S. Chauhan, M. Kumar, S. Chhoker, S. C. Katyal and M. Singh, RSC Adv., 2016, 6, 43080–43090 RSC.
  31. A. A. Zatsiupa, L. A. Bashkirov, I. O. Troyanchuk, G. S. Petrov, A. I. Galyas, L. S. Lobanovskii, S. V. Trukhanov and I. M. Sirota, Inorg. Mater., 2013, 49, 616–620 CrossRef CAS.
  32. A. P. Ramirez, Annu. Rev. Mater. Sci., 1994, 24, 453–480 CrossRef CAS.
  33. V. A. Khomchenko, V. V. Shvartsman, P. Borisov, W. Kleemann, D. A. Kiselev, I. K. Bdikin, J. M. Vieira and A. L. Kholkin, Acta Mater., 2009, 57, 5137–5145 CrossRef CAS.
  34. S. Blundell, Magnetism in condensed matter, Oxford University press Inc., New York, 1st edn, 2001 Search PubMed.
  35. P. Kumar, N. Shankhwar, A. Srinivasan and M. Kar, J. Appl. Phys., 2015, 117, 194103 CrossRef.
  36. P. Tong, B. Kim, D. Kwon, T. Qian, S.-I. Lee, S.-W. Cheong and B. G. Kim, Phys. Rev. B: Condens. Matter Mater. Phys., 2008, 77, 184432 CrossRef.
  37. S. Duhalde, M. F. Vignolo and F. Golmar, Phys. Rev. B: Condens. Matter Mater. Phys., 2005, 72, 161313 CrossRef.
  38. T.-J. Park, G. C. Papaefthymiou, A. J. Viescas, A. R. Moodenbaugh and S. S. Wong, Nano Lett., 2007, 7, 766–772 CrossRef CAS PubMed.
  39. P. Suresh, P. D. Babu and S. Srinath, Ceram. Int., 2016, 42, 4176–4184 CrossRef CAS.
  40. P. Uniyal and K. L. Yadav, J. Phys.: Condens. Matter, 2009, 21, 405901 CrossRef CAS PubMed.
  41. D. M. Evans, M. Alexe, A. Schilling, A. Kumar, D. Sanchez, N. Ortega, R. S. Katiyar, J. F. Scott and J. M. Gregg, Adv. Mater., 2015, 27, 6068–6073 CrossRef CAS PubMed.
  42. T. Kimura, S. Kawamoto, I. Yamada, M. Azuma, M. Takano and Y. Tokura, Phys. Rev. B: Condens. Matter Mater. Phys., 2003, 67, 180401 CrossRef.

This journal is © The Royal Society of Chemistry 2016
Click here to see how this site uses Cookies. View our privacy policy here.