You are on page 1of 16

CHAPTER 2

INTRODUCTION TO COLOSSAL
MAGNETORESISTANCE (CMR)

2.1 PHENOMENON OF CMR

2.2 STRUCTURE OF CMR MATERIALS

2.3 FEATURES OF CMR MATERIALS

2.4 REFERENCES
Chapter-2

Introduction to Colossal Magneto Resistance

Since the discovery of high temperature super conductivity in cuprate oxides by


Bednorz and Muller [1] in 1986 there has been renewed interest in the study of the
dynamics of systems having strong electron correlations near the mott transition. One of
the most remarkable features of these systems is that a strong spin charge coupling can
exist and the dynamics of carriers is expected to be strongly dependent on the spin
arrangement. The discovery [2-4] of "Colossal Magneto Resistance"(CMR) in pervoskite
based Rare earth Manganates is a spectacular example of strong interplay between itinerant
carriers and localized magnetic moments. This discovery generated wide interest in these
materials from the prospect of their technological applications in general and in magnetic
recording in particular.

2.1 Phenomenon of Colossal Magneto Resistance

Magneto Resistance is the relative change in the electrical resistance of a material


produced on the application of a magnetic field. It is generally defined by,

MR = [ / (O)] = [ (H) - (O)]/ (O) (1)

where (H) and (O) are the resistances or resistivities at a given temperature in the
presence and absence of a magnetic field, H, respectively. Magneto Resistance (MR) can
be negative or positive.

Most metals show a small MR (only a few percent). In non-magnetic pure metals
and alloys MR is generally positive and MR shows a quadratic dependence on H. MR can
be negative in magnetic materials because of the suppression of spin disorder by the
magnetic field. Suppression of quantum mechanical interference can give rise to negative
MR in highly resistive alloys. Large Magneto Resistance, referred to as Giant Magneto
Resistance (GMR), was first observed on the application of magnetic fields to atomically
engineered magnetic super lattices (e.g. Fe/Cr) [5] and in magnetic semiconductors [6].
Besides some of the bimetallic or multimetallic layers, comprising ferromagnetic and anti
ferromagnetic or non-magnetic metals, GMR is also found in ferromagnetic granules
dispersed in paramagnetic metal films (e.g. Co/Cu) [5]. GMR in layered and granular
magnetic materials arises from the ability of magnetic fields to change and control the

32
Chapter-2

scattering of conduction electrons through the modification of the electron-orbit and spin-
orbit interactions. GMR is a result of the reduction of the extra resistance due to the
scattering of electrons by the non-aligned ferromagnetic components in zero magnetic
field.

The discovery of negative GMR in rare-earth manganates,[7-9] Ln1-xAxMnO3


(Ln=rare-earth, A=a divalent cation such as an alkaline earth) with the perovskite structure,
has attracted wide attention. The magnitude of GMR in these materials can be very large,
close to 100% as per eq. (1). For this reason, many researchers prefer to call this colossal
magnetoresistance (CMR), as distinct from GMR in layered or granular metallic materials.
In metallic multi layers or granular alloys, the mechanism involves spin-polarized
transport. In the magnates also, spin polarized transport is responsible for the large
negative MR, but it is different from what happens in metallic multi layers.

A variety of manganese oxides, in the form of polycrystalline powders, thin films


and single crystals, have been investigated for CMR. These studies have provided valuable
insight into the CMR phenomenon in the manganates and have also revealed several other
fascinating features of these oxides.

2.2 Structure of rare earth manganates

LaMnO3 is an insulator with an orthorhombically distorted perovskite structure

(b>a>c/ 2 , Pbnm) and typically contains some Mn4+ as prepared by the usual solid state
reactions. LaMnO3 with a small proportion of Mn4+ (<5%) becomes anti ferromagnetically
ordered at low temperatures (TN~ 150 K). When the La3+ in LaMnO3 is progressively
substituted by a divalent cation as in La1-xAxMnO3 (A=Ca, Sr or Ba), it becomes
ferromagnetic with a well-defined Curie temperature, Tc, and metallic below Tc[9,10]. The
saturation moment is approximately 3.8 which is close to the theoretical estimate based
on localized spin-only moments, suggesting that the carriers are spin polarized. Below Tc,
the manganates exhibit metal-like conductivity. Figure 2.1 illustrates how the
ferromagnetism and the M-I transition occur around Tc in a typical La1-xCaxMnO3
composition[11].

33
Chapter-2

Fig.2.1 Temperature variation of (a) magnetization, (b) resistivity (at H = 0 T and 6T)
and (c) magnetoresistance in La0.8Ca0.2MnO3 (from Mahendiran et al [11]).

The simultaneous observation of itinerant electron behavior and ferromagnetism in


the manganates is explained by Zener's double-exchange mechanism. The basic process in
3
this mechanism is the hopping of a d-hole from Mn4+ d 3 , t 23g eg0 , S = to Mn3+
3

(d 3
)
, t 23g e1g , S = 2 via the oxygen, so that the Mn4+ and Mn3+ ions change places. The t 23g

3
electrons of the Mn3+ ion are localized on the Mn site giving rise to a local spin of , but
2
the eg state, which is hybridized with the oxygen 2p state, can be localized or itinerant.
There is strong Hund's rule interaction between the eg and t 23g electrons. Good enough [12]

34
Chapter-2

has pointed out that the ferromagnetism is governed not only by double-exchange, but also
by the nature of the super-exchange interactions. It may be noted that the Mn3+-O-Mn4+
super exchange interaction is ferromagnetic while both the Mn3+-O-Mn3+ and Mn4+-O-
Mn4+ interactions are anti ferromagnetic.

The orthorhombic distortion of LaMnO3 decreases as La is progressively


substituted by a divalent ion and until it becomes rhombohedral or pseudo-cubic. The Mn-
O-Mn angle approaches 180o as the distortion decreases. When the size of the A-site
cations becomes small, orthorhombic distortion is favoured. The key factor responsible for
(
the distortion in LaMnO3 is the Jahn-Teller distortion due to Mn3+ ions t 23g e1g . )
). Across the M-I
Accordingly, LaMnO3 has three Mn-O distances (1.91, 1.96 and 2.19
transitions, the Jahn-Teller distortion increases in Ln1-xA-xMnO3, the distortion being
greater in the insulating regime[13]. Increasing static coherent MnO6 distortion is
associated with increasing insulating behaviour and decreasing Tc. The structural
parameters, specifically the oxygen thermal parameters, show a significant change across
the M-I transition. Clearly, Mn4+ ions decrease the Jahn-Teller distortion, increase the
itinerancy of electrons through double-exchange and increase the tolerance factor of the
perovskite structure (since Mn4+ is smaller than Mn3+).

2.3 FEATURES OF CMR MATERIALS


Temperature-composition phase diagrams for La1-xAxMnO3 have been worked out
with increasing x or Mn4+ content [9]. A composition of x ~ 0.3 is ideal to observe a good
M-I transition and negative Colossal Magneto Resistance (CMR). The magnitude of CMR
is maximum at the M-I transition (Fig.2.1) and scales with the field-induced
magnetization[8].
35
Chapter-2

CMR is found over a wide range of compositions La1-xCaxMnO3(x<0.5). However,


this material shows significant lattice effects in particular for low x. When x=0.2, a local
structural distortion occurs due to the formation of lattice polarons which arises from the
strong electron-lattice coupling[14]. Lattice polaron formation and local structural
distortion have important implications on the electron transport (high resistivity) in these
materials.

La1-xSrxMnO3 is somewhat different from La1-xCaxMnO3, the A-site ion size being
larger. The ferromagnetic metallic regime in La1-xSrxMnO3 extends over larger values of x
and the Curie temperatures are higher.

Electron transport and magnetic properties of the x=0.3 compositions of the Ca and
Sr compositions have been investigated in thin films as well as in bulk materials[15].
While the magnetization decreases as T2, the resistivity increases proportional to T2. A
study of La0-7Sr0.3MnO3 by neutron scattering has shown that the spin dynamics and the
magnetic critical scattering are surprisingly comparable to those of typical metallic
ferromagnets[16]. In La1-xSrxMnO3 (x=0.17), the local spin moments and charge carriers
couple strongly to changes in the structure and the structure can be switched by the
application of a magnetic field depending on the temperature. The magnetic field which
causes this cross-over from the orthorhombic to the rhombohedral structure is rather low
(around 2T), showing thereby that a very small energy difference exists between these two
structures [17]. Evidence for magnetic polarons at T>Tc has been found in
localized magnetic clusters which grow under magnetic
(LaY)0.67Ca0.33MnO3, with 12
fields[18].

Hydrostatic pressure stabilizes the ferromagnetic metallic state in La1-xSrxMnO3


(i.e. increases the Tc). Hydrostatic pressure enhances the Mn-O-Mn transfer integral
through a change in the Mn-O-Mn angle. A similar effect can be brought about by
increasing the radius of the A-site cation. There is a direct relationship between Tc and the
average radius of the A-site cation, <rA>, the Tc increasing with <rA>[19,20]. A small <rA>
gives rise to a distortion of the MnO6 octahedra by bending the Mn-O-Mn bond and in turn
causing the narrowing of the eg bandwidth to a greater extent. By plotting the Tc against
<rA>, one obtains a phase diagram separating the ferromagnetic metal and paramagnetic

36
Chapter-2

insulator regimes [19,21]. In Figure 2.2, a phase diagram obtained in this manner by
Mahendiran et al [11] is given. In the left-hand bottom comer, we see the ferromagnetic
insulator regime. CMR generally decreases with increase in <rA>, just as the peak
resistivity at the M-I transition. Enhanced CMR could arise from a reduced mobility of the
doping holes and an increase in the coupling between the localized and itinerant electrons.

Fig.2. 2 Variation of the ferromagnetic Tc or the insulator-metal transition temperature, Tim, in


Ln1-xAxMnO3 with the average radius of the A-site cations (from Mahesh et al[17]).

The effect of particle size on the electron transport and magnetic properties of
polycrystalline La0.7Ca0.3MnO3 has been investigated [22]. Although Tc decreases with the
decreasing particle size, the magnetoresistance is insensitive to particle size. The M-I
transition becomes broader when the particle size is small, but the percentage MR is nearly
constant over a wide temperature range. The low-temperature MR is significantly
dependent on the grain size. For larger grain size, the MR at the lowest temperature
decreases eventually becoming zero as in good single crystals. Samples of manganates
with different particle sizes prepared with different heat treatments often show large
differences between the ferromagnetic Tc and the temperature of the M-I transition.
Clearly, grain boundaries contribute substantially to magneto resistance in polycrystalline
materials, in particular at T Tc.

The effect of dimensionality on CMR and related properties have been studied by
examining the (SrO) (La1-xSrxMnO3)n family[23]. The n=l member which is two
dimensional with the K2NiF4 structure is an insulator, while the n= member is the three-

37
Chapter-2

dimensional perovskite showing CMR and other properties. The n=2 member exhibits the
sharpest I-M transition and high MR. A recent study [20] of La1.4Sr1.6Mn2O7 has shown
that the current perpendicular to the MnO2 planes is large, particularly close to Tc and the
barrier to transport provided by the intervening insulating layers is removed by the
magnetic field[24]. Interlayer transport seems to involve tunnelling.

Studies of the electrical, magnetic and other properties of La1-xAxMnO3 systems


have revealed certain unusual features with respect to the charge carriers in these oxides.

The manganates exhibit high resistivities even in the metallic state, particularly at
low temperatures[25].The values of resistivities are considerably higher than Mott's value
of maximum metallic resistivity in many of these materials[11] the resistivity reaching
constant values as high as 103-104 cm at low temperatures. A clear picture of the nature
of carriers responsible for the high resistivity is yet to emerge. The cause of the observed
temperature dependence of resistivity at low temperatures is also not entirely clear.
Another unusual feature of the manganates is related to the occurrence of large MR in a
small field at relatively high temperatures. The scale of Zeeman energy associated with a
field of 1 T of these spin systems is approximately 5 K. This is much too small compared
to the scale of thermal energy (200-300 K) where one sees such a large MR.

Optical conductivity measurements on La0.825Sr0.125MnO3 in the range 0-10 eV


have shown a band of 1.5 eV and the spectral weight is transferred from this band with
decreasing temperature. The 1.5 eV band is due to interband transitions between the
exchange split spin polarized eg bands. At T < Tc, intraband transitions in the eg band
dominate the spectrum [26]. A recent study of tunnelling density of states by low
temperature STM in La0.8Ca0.2MnO3 has shown that at T < Tc, a large number of states
become available near EF [27]. The specific heat of the metallic samples La0.7Ba0.3MnO3
and La0.8Ca0.2MnO3 give a value of g (for the linear term) in the region of 4-6 mJ mole-1 K-
2
[28]. From the cubic term, the Debye temperature is approximately 400 K. The term is
significantly enhanced from the free electron and the band values, the enhancement arising
from correlation effects and electron-lattice interaction. This value is comparable to that
in a number of similar metallic ABO3 perovskites and is not unusual and does not show
any indication of depletion of the density of states at the Fermi level. The Hall coefficient

38
Chapter-2

has been measured in epitaxial films of La0.67Ca0.33MnO3 and La0.67Sr0.33MnO3 grown by


MOCVD and the carrier concentration estimated from the Hall effect is more than that
calculated with the assumption that each Mn4+ contributes one hole and that all the holes
are mobile[15]. For the Sr-doped sample, it is 2.1 holes per Mn and for Ca-doped samples
it is 0.9 holes per Mn.

Double-exchange seems necessary for the CMR of the manganate system as


recently shown by a study of the LnMn1-xCrxO3 system where Cr3+ replaces Mn[8,29]. The
basic question that remains is why these manganates are insulating when T > Tc. The
anomalous resistance has been explained by Millis et al[30] considering electron-phonon
interaction arising from the Jahn-Teller splitting of the Mn d levels. Jahn-Teller distortion
localizes conduction electrons as polarons. As the temperature decreases, the transfer
integral, tij, increases and the ratio of the Jahn-Teller energy to tij decreases. The coupling
varies with temperature and composition and increases across the metal-insulator transition
[31]. The large effect of local lattice distortion and crystal structure symmetry on electron
transport and magnetism show that one cannot explain the unusual behaviour without
taking the lattice into account. The observed Tc (approximately 250-400 K) in most
manganates is considerably lower than that estimated from the measured spin-wave
stiffness constant [32] implying that the ferromagnetic exchange arising from the double-
exchange mechanism is not the only factor that affects Tc. The importance of electron-
lattice interaction is also indicated by the observation of a large oxygen isotope shift of Tc
in La0.8Ca0.2MnO3[33]. Besides lattice effects, electron correlation yields a large
distribution of spectral weights [34]. The electrical resistivity of the rare-earth manganates
(with small bandwidth) across the M-I transition is indeed difficult to understand
especially at low temperatures based on double-exchange or electron-phonon interaction
alone. We also do not understand why CMR actually occurs in these oxides.

39
Chapter-2

Fig2.3 Temperature variation of (a) magnetization, (b) lattice parameters and


(c) resistivity in Nd0.5Sr0.5MnO3 (from Kuwahara et al[31]).

Many of the rare earth manganates exhibit charge-ordering. Charge-ordering of the


Mn3+ and Mn4+ ions in La1-xAxMnO3 is associated with dramatic changes in properties[9].
Charge-ordering in these materials is strongly dependent on the average radius of the
A-site cations, <rA>. The ferromagnetic Tc increases with increase in <rA> while charge-
ordering is favoured by small <rA>. Charge-ordering and double-exchange are thus
competing interactions. When <rA> is sufficiently large, as in La0.5Sr0.5MnO3 (<rA>=l.26
), the manganates become ferromagnetic and undergo a metal-insulator transition around

), is
Tc as described earlier. Nd0.5Sr0.5MnO3 with a slightly smaller <rA> (1.236
ferromagnetic and metallic at 250 K (Tc), but transforms to a charge-ordered (CO) anti

40
Chapter-2

ferromagnetic state at 150 K [35]. The charge-ordering transition temperature, TCO, is the
same as TN. There is a significant specific heat anomaly at TCO (TN) and the resistivity
increases markedly due to the metal-insulator (FM-CO) transition as shown in Figure 2.3.
)
This situation is to be contrasted with that of Nd0.5Ca0.5MnO3 (rA>=l. 17 which
exhibits a TCO of 250 K in the paramagnetic state exhibiting a small heat capacity
anomaly[36]. Nd0.5Ca0.5MnO3 is an insulator at all temperatures. Y0.5Ca0.5MnO3 with a
also gets charge-ordered in the paramagnetic state (TCO, 240 K) and
<rA> of 1.13
becomes anti ferromagnetic at 140 K

Fig. 2.4 Schematic diagram showing the different types of behavior of the manganates,
Ln0.5A0.5MnO3, depending on <rA>. Region A corresponds to the manganates which show
ferromagnetism and insulator metal transitions at Tc while regions B and D correspond to
Nd0.5Sr0.5MnO3 and Y0.5Ca0.5MnO3, respectively. Region C would show complex magnetic
and related properties. FMM, ferromagnetic metal; PMI, paramagnetic insulator, AFMI,
antiferromagnetic insulator, CO, charge-ordered state.

In Figure 2.4 we show a simple diagram where we delineate the different types of
behavior of the manganates depending on the <rA>. The charge-ordered state of
Nd0.5Sr0.5MnO3 is melted by the application of magnetic fields, rendering the material
metallic. On the other hand, a magnetic field of 6 T has no effect on the charge-ordered
insulating state of Y0.5Ca0.5MnO3. The charge-ordered states in Nd0.5Sr0.5MnO3 and
Nd0.5(Y0.5)Ca0.5MnO3 are clearly of different kinds, the difference being caused by the
difference in <rA>[37]. The two types of CO states can be understood qualitatively in terms
of the variation of the exchange couplings JFM and JAFM with <rA>, noting that the single-

41
Chapter-2

ion Jahn-Teller energy does not vary with <rA>[38]. Another important feature of the
charge-ordered manganates is the presence of alternate Mn3+ and Mn4+ ions and hence
alternating long and short Mn-O distances [36-38]. Charge-order stripes are often seen
because of this feature [39].

It was mentioned that the CO state is melted into a ferromagnetic metallic state in
some of the manganates, on application of magnetic field. Application of electric fields and
irradiation by appropriate lasers also has a similar effect. Application of internal or external
pressure and appropriate cation substitution in the Mn3+ site also melts the CO state[40,41].
These features of the charge-ordered manganates may have potential technological
applications

Although CMR in the rare earth manganates has many unique features, it should be
noted that CMR also occurs in other materials, Tl2Mn2O7, CrFe2S4 and nonstoichiometric
EuO being examples [9] clearly, many other oxidic materials will exhibit CMR. However,
the main problem facing applications is that CMR should occur at ordinary temperatures
and at small magnetic fields. Composites involving the rare earth manganates and other
materials hold some promise in this direction. The electronics and information technology
industry seems to be hopeful of using GMR or CMR materials, possibly the bimetallic in
the near future. There is much to investigate with regard to charge ordering and related
properties of the manganates. In particular, region C of the diagram in Figure
2.4 is particularly fascinating as shown recently by a study of the unusual CO-FMM
transition and other aspects of one of the members [42]. The role of strain due to mismatch
in the size of A-site cations [43] is getting to be recognized to be important in determining
TC and TCO. There is considerable scope for theoretical approaches and models directed
towards understanding CMR and charge-ordering in manganates.

42
Chapter-2

Aim of present studies


In order to understand the effect of A-site lattice mismatch in ABO3 pervoskite
oxides due to the substitution of smaller Yttrium at Lanthanum site in La0.7Ba0.3MnO3 and
La0.7Sr0.3MnO3 compounds, systematic studies have been undertaken in the present
investigation
(i) Preparation of La0.7-xYxBa0.3MnO3 and La0.7-xYxSr0.3MnO3 (x=0.0-1.0)
compounds by solid state ceramic technique.
(ii) X-ray diffraction (XRD) Powder X-ray diffraction technique has been used for the
determination of phase purity of the samples.

(iii) Electrical resistance of the samples was measured as a function of temperature in


the presence of zero, 1T, 4T, and 7T magnetic fields to determine the insulator-
metal transition temperatures and also magnetoresistance of the samples as a
function of temperature.
(iv) DC Magnetic susceptibility of the samples was measured as a function of
temperature (5K-300K) in the presence of 5T applied magnetic field to determine
the ferro-paramagnetic transition temperature TC. Magnetic isotherms at 20K with
increasing and decreasing magnetic fields upto 5T were also recorded to study the
variation in saturation magnetization with Yttrium substitution in these compounds.

43
Chapter-2

References

1. J.G. Bednorz and K.A. Muller, Z. Phys.,B64, 189 (1986).


2. K. Chabara, T. Ohno, M. Kasai and Y. Kozono, Appl. Phys. Lett. 63, 1700 (1993).
3. R. Von Helmolt, J. Wocker, B. Holzapfel, Schultz and K. Samwer, Phys. Rev. Lett.
71, 2331 (1993).
4. S. Jin , T.H. Tiefel, M. Mc Cormack, R.A. Fortmacht R. Ramesh and L.H. Chen.,
Science, 264, 413 (1994).
5. P.M. Levy, Solid State Phys., 47, 367 (1994); P.M. Levy and S. Zhang, J. Magn.
Magn. Mater., 151, 315 (1995).
6. N.B. Brandt and V.V. Moschalkov, Adv. Phys., 33, 193 (1984).
7. K. Chahara, T.Ohno, M.Kasai and Y.Kozono Appl. Phys. Lett., 63, 1990 (1993);
M. McCormack et al., Appl. Phys. Lett., 64, 3045 (1994); R. Mahesh,
R.Mahendiran, A.K.Raychaudhary and C.N.R.Rao, J. Solid State Chem.,114,
297(1995).
8. A. Urushibara, Y.Moritomo, T.Arima, A.Asamitsu, G.Kido and Y.Tokura,Phys.
Rev., B51, 14103 (1995).
9. C.N.R. Rao, A.K.Cheetam and R.Mahesh Chem. Mater., 8, 2421 (1995); C.N.R.
Rao and A.K. Cheetham, Adv. Mater., 9, 1009 (1997) and the references therein.
10. G.H. looker and J.H. van Santen, Physica, 16, 377 (1950).
11. R. Mahendiran, S.K.Tiwary, A.K.Raychaudhari, T.V.RamaKrishnan, R.Mahesh,
N.Rangavittal and C.N.R.Rao, Phys. Rev., B53, 3348 (1996).
12. J.B. Good enough, Progr. Solid State Chem., 5, 149 (1971).
13. J.L. Garcia-Munoz, M.Suaaid, J.Fontcuberta and J.Rodriguez, Phys. Rev., B55, 34
(1997).
14. S.J.L.Brillinge, R.G.DiFrancesco, G.H.Kwei, J.J.Neumeier and J.D.Thompson,
Phys. Rev. Lett., 77, 715 (1996).
15. J.G. Snyder, R.Hiskes, S.Dicarolis, M.R.beasley and T.H.Geballe, Phys. Rev., B53,
14434 (1996).
16. M.C. Martin, G.Shirane, Y.Endoh, K.Hirota, Y.Moritomo and Y.Tokura, Phys.
Rev., B53, 14285 (1996).

44
Chapter-2

17. A. Asamitsu, Y.moritomo, R.Kumai, Y.Tomioka and Y.Tokura, Phys. Rev., B54,
1716 (1996).
18. J.M. DeTeresa et al., Nature, 386, 256 (1997).
19. H.Y.Hwang, S-W.Cheong, P.G.Radaelli, M.Marezio and B.Batlogg, Phys. Rev.
Lett., 75, 914 (1995).
20. J. Fontcuberta, B.Martinez, A.Seffar, S.Pinol, J.L.Garcia-Muoz and X.Obradors,
Phys. Rev. Lett., 76, 1122 (1996).
21. R. Mahesh, K.R.Kannan and C.N.R.Rao, J. Solid State Chem., 114, 294 (1995).
22. R. Mahesh et al., Appl. Phys. Lett., 68, 2291 (1996).
23. Y. Moritomo, A.Asamitsu, H.Kuwahara and Y.Tokura, Nature, 380, 141 (1996).
24. T. Kimura et al., Science, 274, 1698 (1996).
25. J.M.D. Coey, M.Viret and L.Ranno, Phys. Rev. Lett., 75, 3910 (1995).
26. Y. Okimoto, T.Katsufuji, T.Ishikawa, T.Arima and Y.Tokura, Phys. Rev., B55,
4206 (1997).
27. A. Biswas and A.K. Raychaudhuri, J. Phys. Condens. Matter, 8, L739 (1996).
28. J.J. Hamilton, E.L.Kentley, H.L.Ju, A.K.Raychaudhari, V.N.Smolyaninova and
R.L.Greene, Phys. Rev., B54, 14926 (1996).
29. R. Gundakaram, A.Arulraj, P.V.Vanitha, C.N.R.Rao, N.Gayathri,
A.K.Raychaudhari and A.K.Cheetam, J. Solid State Chem., 127, 354 (1996).
30. A.J. Millis, P.B.Littlehood and B.I.Shraiman, Phys. Rev. Lett., 74, 5144 (1995).
31. H. Rder, J.Zang and A.R.Bishop, Phys. Rev. Lett., 76, 1356 (1996).
32. J.W. Lynn, R.W.Erwin, J.A.Borchers, Q.Huang and A.Santoro, Phys. Rev. Lett.,
76, 4046 (1996).
33. G. Zhao, K.Conder, H.Keller and K.A.Mller, Nature, 381, 676 (1996).
34. S. Satpathy, S.Zoran and R.V.Fillip, Phys. Rev. Lett., 76, 960 (1996).
35. H. Kuwahara, Y.Tomoika, A.Asamitsu, Y.Moritomo and Y.Tokura, Science, 270,
961 (1995).
36. T.Vogt, A.K.Cheetam, R.Mahendiran, A.K.Raychaudhari, R.Mahesh and
C.N.R.Rao, Phys. Rev., B54, 15303 (1996).
37. C.N.R. Rao and A.K. Cheetham, Science, 276,911 (1997).

45
Chapter-2

38. N. Kumar and C.N.R. Rao, J. Solid State Chem., 129, 363 (1997).
39. S. Mori, C.h.Chen and S.W.Cheong, Nature, 392, 473 (1998).
40. C.N.R. Rao et al., J. Solid State Chem., 135, 169 (1998); B. Raveau et al., J. Solid
State Chem., 117,424 (1995).
41. B. Raveau, A.Maignan, C.Martin, J. Solid State Chem., 130, 162 (1997); also
Appl. Phys. Lett., 71, 3907 (1947).
42. A. Arulraj, A.Biswas, A.K.Raychaudhari, C.N.R.rao, P.M.Woodwork, T.Vogt,
D.E.Cox and A.K.Cheetam, Phys. Rev., B57, 8115 (1998).
43. A. Maignon et al., Phil. Trans. Roy. Soc. London, A356, 1635 (1998), Phys.
Stat.Sol. (a) 143 (1994) 101.

46

You might also like