Vous êtes sur la page 1sur 5

Materials Letters 61 (2007) 4879 4883 www.elsevier.

com/locate/matlet

Grain boundary effects on the electrical and magnetic properties of Pr2/3Ba1/3MnO3 and La2/3Ca1/3MnO3 manganites
Neeraj Panwar a,b , Vikram Sen a , D.K. Pandya b , S.K. Agarwal a,
a

Superconductivity & Cryogenics Division, National Physical Laboratory, Dr. K.S. Krishnan Road, New Delhi-110012, India b Department of Physics, Indian Institute of Technology, Hauz Khas, New Delhi-110016, India Received 22 January 2007; accepted 17 March 2007 Available online 23 March 2007

Abstract Electrical and magnetic properties of orthorhombic Pr2/3Ba1/3MnO3 (PBMO) and La2/3Ca1/3MnO3 (LCMO) manganites with considerable difference in variance factors (2) are reported here. PBMO with higher variance exhibits distinct intrinsic (due to grains) and extrinsic (due to grain boundaries) transitions in the resistivity behaviour. Extrinsic effects, however, are not observed in the lower 2 LCMO system. Low field magnetoresistivity (LFMR) data also substantiate these results. Increase in the density of states obtained through Mott's 3-D variable range hopping mechanism in the paramagnetic insulating regime indicates the suppression of magnetic domain scattering with applied magnetic field. Ferromagnetic metallic regime below the extrinsic transition in PBMO seems to emanate from the electronmagnon scattering process. LFMR at 77 K also points towards the higher canting of spins in the vicinity of grain boundary regions in PBMO compared to that in LCMO. 2007 Elsevier B.V. All rights reserved.
PACS: 75.47.Gk; 71.30.+h Keywords: Manganites; Grain boundary effects; Low field magnetoresistivity (LFMR); Variable range hopping (VRH); Electronmagnon scattering

1. Introduction Perovskite manganites with the general formula L1 xAxMnO3 (where L is a trivalent rare-earth ion like La+ 3, Pr+ 3, Nd+ 3 etc. and A is the divalent alkaline earth ion like Ca+ 2, Sr+ 2, Ba+ 2 etc.) have triggered the attention worldwide due to the occurrence of colossal magnetoresistive (CMR) effect in the vicinity of insulatormetal (Tp) and paramagneticferromagnetic (TC) transitions [13]. Although such magnetoresistivity can be quantitatively explained by the Zener's theory of double exchange [4], more explanatory mechanisms involving the polaronic effects [5] and intrinsically inhomogeneous states [6], have been suggested to explain the observed high magnitude of magnetoresistivity. In these perovskites the physical properties have been noticed to strongly depend on both the electronic doping x and the rare Corresponding author. Tel.: +91 11 25742610 12x2239 2276; fax: +91 11 25852678. E-mail address: shyamagarwal2005@yahoo.co.in (S.K. Agarwal). 0167-577X/$ - see front matter 2007 Elsevier B.V. All rights reserved. doi:10.1016/j.matlet.2007.03.062

earth site mean ionic radius brAN [7,8] (where brAN is calculated for the nine-fold co-ordination from values tabulated in Ref. [9]). Higher CMR effects have generally been observed for x = 0.33 and for small values of brAN. Strong brAN dependence of Tp and TC has been observed in some manganites, however, it has also been established that additional parameters such as the A-site cation size mismatch 2 (also called variance or cationic disorder and defined by 2 = ixiri2 brAN2, where xi is the fractional substitution level of the ith rare-earth site species with ionic radius ri and brAN = ixiri) and oxygen stoichiometry also exert a strong influence [1015]. In some cases larger 2 affects the grain boundary properties as well. In this paper we report the effect of 2 on the grain boundary properties and the ensuing impact on various physical properties of Pr2/3Ba1/3MnO3 (PBMO) and La2/3Ca1/3MnO3 (LCMO) manganites where the former is having higher brAN (1.27 >2) and higher 2 (0.0187 2) and the later has relatively lower brAN (1.204 ) and 2 (0.000287 2 ) respectively. Significantly, PBMO material is also important from the fact that it does not follow the criteria of brAN vs. TC as proposed by Hwang et al. [16].

4880

N. Panwar et al. / Materials Letters 61 (2007) 48794883

2. Experimental techniques Both Pr2/3Ba1/3MnO3 and La2/3Ca1/3MnO3 materials have been synthesized in the polycrystalline form using the conventional solid state reaction route of taking stoichiometric ratios, grinding and calcining the powders (for homogeneity) at different temperatures between 900 C and 1100 C for 12 h with intermediate grindings. Finally, the powders were pressed in pellet form and sintered at 1260 C for 15 h. This sintering temperature of 1260 C was optimized for the single-phase formation [13]. X-ray diffractometer (Rigaku, CuK, = 1.54 ) was used for the determination of the phase purity of the materials. Four-probe method was employed for the measurement of temperature variation of the resistivity. Air drying silver epoxy was used for making the electrical contacts. Magnetic measurements were carried out in an AC susceptometer (Lakeshore Model ACS 7000) in the temperature range 300 K77 K at a fixed frequency of 111.1 Hz and an ac field of 80 A/m under the zero field cooled configuration. Scanning electron micrographs of the samples were taken using LEO SEM 440 operating at 5 kV. Magnetoresistivity (MR) data was obtained at 0.6 T magnetic field in the temperature range of 300 K77 K. MR has been defined using the relation: MR = [{R(0) R(H)} / R(0)] * 100, where R(H) and R(0) are the resistances of the sample with and without magnetic field respectively.
3. Results and discussion The single-phase nature of the synthesized PBMO and LCMO materials was revealed through their X-ray diffraction measurements. Both the samples possess the orthorhombic structure. The tolerance factor (t) also confirms the structure to be orthorhombic (as it falls within the specified limits (0.89 t 0.96). The temperature dependence of resistivity of Pr2/3Ba1/3MnO3 and La2/3Ca1/3MnO3 samples are shown in Fig. 1. The sample Pr2/3Ba1/3MnO3 shows a sharp transition (TP1) at 194 K followed by a broad transition like hump (TP2) at 160 K. However, La2/3Ca1/3MnO3 depicts only one transition at 265 K. In the polycrystalline sample the contribution to the resistivity originates from the two regions: grain and the grain boundary. Since the grain boundary is more chaotic than the core or the grain, the contribution of the grain boundary to the total resistivity in a polycrystalline sample therefore, always exceeds to that of the grain. The effects on the physical properties due to grain are termed as the intrinsic effects and those arising from the grain boundary as the extrinsic effects. In the ferromagnetic state, perovskites in general, behave like a metal in the electrical properties. In this sense, a polycrystalline sample or granular perovskite is a granular ferromagnet similar to granular transition metals. However, according to the low temperature transport properties observed the formation of the intergrain barrier may be a little different from that in granular transition metals. Since no magnetic material, which can be the potential barrier between the ferromagnetic grains exists in granular perovskites, the interface or the grain boundary between neighbouring grains should be taken into account as a barrier. In polycrystalline sample where the variance 2 is low e.g. LCMO (2 = 0.000287 2) insulatormetal (IM) transition is single and broader because of the grain and the grain boundary effects occurring simultaneously. But in samples where 2 is larger, due to the larger ionic size mismatch between ions present at the rare-earth site e.g. Pr2/3Ba1/3MnO3 (Pr+ 3 and Ba+ 2 sizes in

Fig. 1. Resistivity-temperature variation of Pr2/3Ba1/3MnO3 and La2/3Ca1/3 MnO3.

their nine-fold co-ordination are 1.179 and 1.47 respectively, 2 = 0.0187 2), the grain boundary effects are larger and separate out from the grain effects which is clear from T data of Fig. 1. Therefore, PBMO shows one sharp transition at 194 K (TP1) and the broader one at 160 K (TP2). The reason for larger grain boundary effects in PBMO sample is that due to the larger ionic size difference between Pr+ 3 and Ba+ 2 the lattice within the grain experiences a good deal of strain. The lattice seemingly unloads this strain to the grain surface or to the grain boundary. Consequently, the lattices at the grain boundary or in its vicinity are more distorted and would weaken the electron transfer probability from one Mn-site to the other. This results in the separation of the two transitions in PBMO. However, in LCMO the ionic size difference between the ions La+ 3 (1.216 ) and Ca+ 2 (1.18 ) is smaller resulting in the non-separation of the two transitions. Now the question arises why the broader transition TP2 occurs below TP1 in PBMO. The answer to this emerges from the Heisenberg theory of ferromagnetism [17]. The ferromagnetic transition can be expressed as TC = 2qJ / kB, where q is the co-ordination number of the ion, J is the exchange integral between the neighbouring atoms and kB is the Boltzmann constant. On one hand, the average co-ordination number (q) is lower at the grain boundary due to the presence of the dangling bonds and on the other hand, the overlapping between the neighbouring ions is lower due to the co-ordination number being lower and so is the exchange integral J. Thus TC or TP in the grain surface would certainly be lower to that within the core. In the PBMO sample due to the larger strain at the grain boundary these two effects (grain/grain boundary) separate out but not in LCMO where ionic size difference is not larger. Due to the reason that in the grains Mn+ 3/Mn+ 4 ions are parallel to each other and thus the double exchange mechanism is stronger in grain and hence larger TP but at the grain surface those ions are in chaotic order so TP for grain boundary is lower. It would be worth mentioning here that for the pristine sample PBMO increase in the sintering temperature leads only to the decrease in the peak resistivity with the transition at TP2 ( 160 K) remaining practically unchanged. Such a behaviour can be explained on the basis of the increase in the grain size with the increase in the sintering temperature resulting in the overall decrease in the disorder present there. Barnabe et al. [18] have used higher sintering temperature (1500 C) and reported that electrical resistivity at TP2 is lower than that at TP1, whereas it is higher in the present study. However, the value of TP2 is same in both cases. The higher

N. Panwar et al. / Materials Letters 61 (2007) 48794883

4881

temperature transition TP1 is also slightly higher ( 205 K) in case of Barnabe et al. [18]. The higher insulatormetal (IM) transition is related to the larger grain size. Further, it is observed that there is appearance of a re-entrant insulating behaviour below 35 K which has been ascribed to the localization of the carriers. Such localization could be due to the spin polarized tunneling in these samples but in PBMO due to the larger variance 2 an extra term is also added to the resistivity. Tunneling is made possible between the ferromagnetic grains via the grain boundary that is rather paramagnetic [19]. If the tunneling is of the spin polarized electrons between two grains of anti-parallel spins then to reverse it one has to add an extra term in the resistivity. Due to the strain in PBMO system the carriers are unable to overcome this strain (apparently due to insufficient energy) and get localized leading to the observed re-entrant insulating behaviour at low temperatures. Resistivity data (with and without the application of the magnetic field, Fig. 2) above TP (from room temperature up to TP) fits the 3-D variable range hopping model (VRH) = 0exp(T0 / T)1/4 as proposed by Mott [20] where T0 is a constant (16a3 / KBN(EF), is the inverse of the localization length, N(EF) being the density of states at the Fermi level). From the VRH fit (between 295 K and corresponding TP) we have calculated T0 and N(EF) (Table 1) both in the presence and absence of the magnetic field. It is observed that N(EF) increases with the application of the magnetic field. However, the increase is not significant in PBMO in comparison with LCMO because of the lesser effect of magnetic field

Table 1 N(EF)[H = 0.6 T] T0 (H = 0 T) T0 (H = 0.6 T) N(EF)[H = 0 T] (106 K) (106 K) (1020 eV 1 cm 3) (1020 eV 1 cm 3) LCMO 1.51 PBMO 1.96 1.24 1.94 13.4 10.4 16.4 10.5

on PBMO above TP. To estimate N(EF) we used the value of = 2.22 nm 1 calculated by Viret et al. [21]. It is noticed that the value of N(EF) increases (or corresponding T0 value decreases) on the application of the magnetic field which may be due to the suppression of the magnetic domain scattering by the magnetic field. In order to analyze the data in the metallic regime (below 265 K in LCMO and 160 K in PBMO), the following equations were fitted q q0 q1 T 2 q q0 q2 T 2:5 q q0 q1 T 2 q3 T 4:5 1 2 3

where the temperature independent part 0 is the resistivity due to domain, grain boundary and other temperature independent scattering mechanisms. 1T2 term in Eqs. (1) and (3) represents the electrical resistivity due to the electronelectron scattering process and is generally dominant up to 100 K. 2T2.5 is the term arising due to electronmagnon scattering process. On the other hand, the term 3T4.5 is a combination of electronelectron, electronmagnon and electron phonon scattering processes. We find that in the metallic regime, conductivity data for both the samples best fit the equation = 0 + 2T2.5 both in the presence and absence of the magnetic field. Therefore metallic regime can be attributed to the electronmagnon scattering processes, which further demonstrates that the metallic regime is in the ferromagnetic phase. However, 2 is much larger for PBMO than that of LCMO which further implies the prominent role of the grain boundary in PBMO by making the sample lesser ferromagnetic as compared with LCMO after the transition at TC. The best-fit parameters (the linear correlation coefficient R2 is maximum for 2T2.5) obtained from the fitting of the low temperature metallic regime of the resistivity data with Eq. (2) are shown in Table 2. It is observed that 0 decreases significantly with magnetic field, but the influence of the field on term 2 is small. As the magnetic field increases, the size of the domain boundary decreases and 0 becomes smaller. The slight decrease of 2 with field may be due to the suppression of spin fluctuations in the field. The electronmagnon scattering process ( = 0 + 2T2.5) has also been invoked earlier to fit the electrical resistivity data in the metallic region of the manganites [22]. The MR behaviour (Fig. 2) shows sharp peaks at 194 K and 262 K for the samples PBMO and LCMO respectively and below these transitions it again starts increasing reflecting the role of grain boundaries. The susceptibility ( T) measurements (Fig. 3a) for both PBMO and LCMO samples show transitions from the paramagneticferromagnetic (PMFM) state at 194 K and 267 K respectively.

Table 2 ( = 0 + 2T2.5) R2 0 ( cm) 2 ( cm K 2.5) Pr2/3Ba1/3MnO3 H=0 0.999 0.06062 3.1 10 6 H = 0.6 T 0.998 0.05251 2.51 10 6 La2/3Ca1/3MnO3 H=0 0.999 3.478 10 3 2.04 10 8 H = 0.6 T 0.999 2.823 10 3 1.98 10 8

Fig. 2. Resistivity-temperature variation with and without magnetic field and MR behaviour of Pr2/3Ba1/3MnO3 (a) and La2/3Ca1/3MnO3 (b).

4882

N. Panwar et al. / Materials Letters 61 (2007) 48794883

Fig. 3. a. Plot for T of Pr2/3Ba1/3 MnO3 and La2/3Ca1/3MnO3. b. Plot for T of Pr2/3Ba1/3 MnO3 and La2/3Ca1/3MnO3.

Fig. 4. Scanning electron micrographs of Pr2/3Ba1/3MnO3 and La2/3Ca1/3MnO3.

These transitions being close to IM transitions confirm that electrical and magnetic properties are indeed coupled in manganites [23]. The magnetic transitions have been calculated by differentiating the T curves and measuring the temperature where d / dT is minimum. After TC, for the sample LCMO is almost constant which shows that it is a long-range ordered ferromagnetic sample. However, decreases drastically for PBMO sample indicating the cluster glass behaviour and short-range ordered ferromagnet. The imaginary part (Fig. 3b) of the susceptibility ( = + i ) also peaks near the IM transitions for both the samples and the signal of is about two orders of magnitude smaller than as have been reported earlier also [24]. Since is a measure of the heat loss in the sample so it should correspond to the electrical transition which is the case here. This also tells about the better homogeneity of the samples. The T curve also reflects only the higher temperature transitions in both the samples as the field is lower and unable to make the surface spins parallel so no signature of the second transition is seen. However, as T represents the magnetic behaviour of the sample below 160 K where grain boundary effects are prominent there is competition between the core Mn ions (ferromagnetically ordered) and surface Mn ions (which are rather paramagnetic) and the result is destruction of the long range order in PBMO, making the sample cluster glass. The SEM pictures (Fig. 4) of the samples also show clear grains and the average grain sizes for PBMO and LCMO are 3 m and 5 m respectively. The grain sizes are smaller reflecting the role of grain boundaries in the transport properties. The MR curves (Fig. 2) also indicate the higher temperature transition as the field is low it only reduces the grain resistivity noticeably and not of the grain boundary because of chaos there. The MR vs. temperature curve (0.6 T)

shows higher and constant value of MR after TP1 for PBMO sample while that of LCMO shows a lower MR and an increasing trend below the transition TP. MR vs. H variation at 77 K (Fig. 5) shows the behaviour of low field magnetoresistivity (LFMR) due to spin polarized tunneling for both the samples (with a sudden rise under low magnetic field and then saturation at higher magnetic field value) [25]. But MR of PBMO remains unsaturated which means that it requires a larger field to get the saturation for this particular sample as compared to LCMO where MR gets saturated. This also indicates that canting of spins at the grain boundary region is higher in PBMO

Fig. 5. MR variation with applied magnetic field of Pr2/3Ba1/3 MnO3 and La2/3 Ca1/3MnO3 at 77 K.

N. Panwar et al. / Materials Letters 61 (2007) 48794883

4883

(larger strain) sample than LCMO (lesser strain). The role of oxygen content is very crucial on the electrical as well as the magnetic properties. The reduced oxygen content will increase the Mn+ 3 concentration and can result in two transitions in the resistivitytemperature behaviour [26]. TP in our LCMO sample is 265 K and matches well with others [2729] and it shows IM transition which confirms that our sample is not oxygen deficient. Since both PBMO and LCMO samples were synthesized under the same conditions so the second transition in PBMO due to the reduced oxygen content is ruled out. Barnabe et al. [18] have synthesized PBMO and checked the oxygen content to be 3 and the material shows two IM transitions.

4. Conclusions Electrical and magnetic properties of orthorhombic (PBMO) and (LCMO) manganites are reported here. These materials have been chosen because of the considerable difference in their variance (2) values. PBMO exhibits two types of transitions in the resistivity-temperature behaviour, characteristic of both intrinsic (due to grains) and extrinsic (due to grain boundaries) situations whereas only intrinsic effects are observed in LCMO. Low field magnetoresistivity (LFMR) data also substantiate these results. Increase in the density of states (and the decrease in the T0) obtained through Mott's 3-D variable range hopping mechanism in the paramagnetic insulating regime (above intrinsic transition) is attributed to the suppression of magnetic domain scattering with applied field. Ferromagnetic metallic regime below the extrinsic transition in PBMO seems to emanate from the electronmagnon scattering process. LFMR at 77 K also points towards the higher canting of spins in the vicinity of grain boundary regions in PBMO compared to that in LCMO. Acknowledgements The authors express their gratitude to the Director, NPL for his keen interest in the present work. Assistance from the scanning electron microscopy section, NPL, New Delhi is gratefully acknowledged. Two of us (NP and VS) are thankful to the CSIR, New Delhi for the grant of Senior Research Fellowships. References
[1] R. von Helmolt, J. Wecker, B. Holzapfel, L. Schultz, K. Samwer, Phys. Rev. Lett. 71 (1993) 2331. [2] H.L. Ju, C. Kwon, Q. Li, R.L. Greene, T. Venkatesan, Appl. Phys. Lett. 65 (1994) 2117.

[3] R. Mahendiran, S.K. Tiwari, A.K. Raychaudhuri, T.V. Ramakrishnan, R. Mahesh, N. Rangavittal, C.N.R. Rao, Phys. Rev., B 53 (1996) 3348. [4] C. Zener, Phys. Rev. 81 (1951) 440. [5] A.J. Millis, Nature 392 (1998) 147. [6] C. Sen, G. Alvarez, E. Dagotto, Phys. Rev., B 70 (2004) 064428. [7] L. Pi, M. Hervieu, A. Maignan, C. Martin, B. Raveau, Solid State Commun. 126 (2003) 229. [8] H. Jain, A.K. Raychaudhuri, Y.M. Mukovskii, D. Shulyatev, Solid State Commun. 138 (2006) 318. [9] R.D. Shannon, Acta Crystallogr., A 32 (1976) 751. [10] L.M. Rodriguez-Martinez, J.P. Attfield, Phys. Rev., B 54 (1996) 15622. [11] E. Surad, F. Fauth, C. Martin, A. Maignan, F. Millange, L. Keller, J. Mag. Magn. Mater. 264 (2003) 221. [12] G. Venkataiah, V. Prasad, P.V. Reddy, J. Alloys Compd. 429 (2007) 1. [13] N. Panwar, S.K. Agarwal, G.L. Bhalla, D. Kaur, D.K. Pandya. Int. J. Mod. Phys. B (June, 2006) Acceptance No. jpb061484. [14] V. Sen, N. Panwar, G.L. Bhalla, S.K. Agarwal. J. Alloys Compd. (in press), doi:10.1016/j.jallcom.2006.04.078. [15] W. Zhong, W. Chen, C.T. Au, Y.W. Du, J. Mag. Magn. Mater. 261 (2003) 238. [16] H.Y. Hwang, S.-W. Cheong, P.G. Radaelli, M. Marezio, B. Batlogg, Phys. Rev. Lett. 75 (1995) 914. [17] W. Heisenberg, Z. Phys. 49 (1928) 619. [18] A. Barnabe, F. Millange, A. Maignan, M. Hervieu, B. Raveau, G. Van Tendeloo, P. Laffez, Chem. Mater. 10 (1998) 252. [19] N. Zhang, F. Wang, W. Zhong, W. Ding, J. Phys.: Condens. Matter. 11 (1999) 2625. [20] N.F. Mott, MetalInsulator Transitions, Taylor & Francis, London, 1990. [21] M. Viret, L. Ranno, J.M.D. Coey, Phys. Rev., B 55 (1997) 8067. [22] L. Pi, L. Zhang, Y. Zhang, Phys. Rev., B 61 (2000) 8917; R. Ang, Y.P. Sun, J. Yang, X.B. Zhu, W.H. Song, J. Appl. Phys. 100 (2006) 073706; J.M. De Teresa, M.R. Ibarra, J. Blasco, J. Garcia, C. Marquina, P.A. Algarabel, Z. Arnold, K. Kamenev, C. Ritter, R. von Helmolt, Phys. Rev., B 54 (1996) 1187. [23] F.M. Araujo-Moreira, M. Rajeshwari, A. Goyal, K. Ghosh, V. Smolyaninova, T. Venkatesan, C.J. Lobb, R.L. Greene, Appl. Phys. Lett. 73 (1998) 3456. [24] A.J. Duyneveldt, Review on ac-Susceptibilty Studies in Solid State Magnetism and references therein, Application Note (Lake Shore Cryotronics, Westerville, OH, 1996. [25] H.Y. Hwang, S-W. Cheong, N.P. Ong, B. Batlogg, Phys. Rev. Lett. 77 (1996) 2041. [26] L. Malavasi, M.C. Mozzati, C.B. Azzoni, G. Chiodelli, G. Flor, Solid State Commun. 123 (2002) 32. [27] R. Mahendiran, R. Mahesh, A.K. Raychaudhuri, C.N.R. Rao, Solid State Commun. 94 (1995) 515. [28] G. Li, S.-J. Feng, F. Liu, Y. Yang, R.-K. Zheng, T. Qian, X.-Y. Guo, X.-G. Li, Eur. Phys. J. B32 (2003) 5. [29] L. Seetha Lakshmi, K. Dorr, K. Nenkov, V. Sridharan, V.S. Sastry, K.-H. Muller, J. Mag. Magn. Mater. 290291 (2005) 924.

Vous aimerez peut-être aussi