Open access peer-reviewed chapter

Low-Doped Regime Experiments in LaMnO3 Perovskites by Simultaneous Substitution on Both La and Mn Sites

Written By

Aminta Mendoza and Octavio Guzmán

Submitted: 19 July 2022 Reviewed: 23 August 2022 Published: 17 October 2022

DOI: 10.5772/intechopen.107309

From the Edited Volume

Recent Advances in Multifunctional Perovskite Materials

Edited by Poorva Sharma and Ashwini Kumar

Chapter metrics overview

77 Chapter Downloads

View Full Metrics

Abstract

Although extensive substitution studies exist on LaMnO3 perovskites, simultaneous substitution at the La and Mn sites in the low-doping regime is not as common and provides important insights into the subtle balance of various competing effects acting on the crystal structure and the magnetic properties. This chapter presents a study of the evolution of the crystal structure and magnetic properties by simultaneous substitution of magnetic and non-magnetic ions at the perovskite A and B sites, respectively. It will examine some of the ways in which experiments on the evolution of magnetic properties provide a suitable balance scenario between the competitive effects arising from doping at each site. The work discusses the evolution of the Curie-Weiss behavior and the formation of ferromagnetic (FM) clusters above the Curie temperature, whose structure is also dependent on doping.

Keywords

  • perovskite manganites
  • electron paramagnetic resonance
  • magnetic clusters
  • Jahn-Teller distortion
  • local magnetism

1. Introduction

Simultaneous substitution in polycrystalline LaMnO3 on the La3+ and Mn3+ ions by magnetic Dy3+ and non-magnetic Zn ions, respectively, offers an adequate scenario for studies on the evolution of structure and magnetic properties due to competing effects. Low-doped regime in La1 − xDyxMn1 − yZnyO3 manganites(La(Dy) Mn(Zn)O3)means the amount of substitution (x,y) is kept in the 0.00.1 regime for both sites. For both x=0, y=0, the parent compound LaMnO3 is an insulator composed of ferromagnetic (a,b)-planes of Mn3+ ions with orbitals configuration {t2g3,eg1}, total spin S=2, oriented in basal plane, but antiferromagnetically (AF) coupled along the c axis [1]. This antiferromagnetic ground state structure (A-type) due to the Jahn-Teller (J-T) distortion of the Mn3+ ions [2, 3, 4] tends to disappear by oxygen excess, and the LaMnO(3 + δ) manganite becomes ferromagnetic. The magnetic A-type structure occurs because of the different orientations of the eg1 orbitals within the (a,b) plane lead to FM superexchange, while covalent orientation of these orbitals along the c axis leads to AFM superexchange [5, 6, 7], which weakens the ferromagnetism within the (a,b) planes, but the AFM coupling along c axis remains unaltered. For LaMnO(3 + δ), the correlation between crystal structure and magnetism has been discussed in Ref. [8].

For y=0, the magnetic and transport properties of La1 − xDyxMnO3 compounds with nominal stoichiometry x=0.05,0.10 synthesized by sol-gel method have been studied by the authors in Ref. [9]. Although LaMnO3 and La1 − xDyxMnO3 compounds were synthesized with the same method, they observed that in La1 − xDyxMnO3 more La-site vacancies appear. This has been attributed to evaporation of La ions by chemical disorder. The La-site deficiencies, as mentioned, increase the FM double exchange (DE) interaction. Simultaneously, the ionic radius difference between La and Dy ions leads to local distortion, in addition to the J-T distortion caused by Mn3+ ions reducing the FM interaction between Mn3+ and Mn4+ ions. Furthermore, the large magnetic moment of Dy3+ (μeff10.83μB) randomly polarizes the magnetic Mn sublattice in favor of ferromagnetism. In slightly Dy-doped LaMnO3, the paramagnetic Dy ion can be represented as an impurity between the Mn-Mn spin pairs coupling to Mn ions through 3d and 4f electrons. At large Mn-Dy separation, the interaction between 3d and 4f electrons is smaller than (or just comparable) that between nearby Mn-Mn host ions [10, 11, 12, 13], confirmed that through weak coupling between magnetic impurities and their nearest neighbor host ions, the local AFM order in the host spin system will be enhanced. Through small Dy doping, the interactions between Dy impurities are determined by host Mn spins. On the contrary, when the Dy concentration increases, the interaction between the Dy impurities increases, and the lattice becomes much more disordered. Thus, Dy can introduce magnetic disorder that can be important to understand the presence of ferromagnetic clusters above the ferromagnetic ordering.

For x=0, a great deal of work has been conducted concerning the doping of the Mn site in LaMn1 − yZnyO3 manganites with y in the 0.050.4 range [14, 15, 16, 17]. Partial substitution of Mn by Zn2+ {3d10} nonmagnetic ions results in the simultaneous presence of Mn3+ and Mn4+ ions, triggering Zener’s DE interaction. Thus, the transformation of Mn3+ into Mn4+ reduces the antiferromagnetism of the A-type structure, inducing 3-dimensional ferromagnetism [15]. Moreover, the Zn dilution of the Mn sublattice locally weakens magnetism and induces disorder. At the same time, the magnetic La(Dy) sublattice interacts with the local field imposed by the ferromagnetic of the Mn sublattice. Depending on the magnetic nature of the La(Dy) sublattice, the interaction can be FM or AFM; in our case, ferromagnetism is obtained.

Advertisement

2. Crystalline structure

X-ray diffraction (XRD) patterns study of the samples of La1 xDyxMn1 yZnyO3 family (synthesized by solid reaction method) with x=0.0,0.05,0.1 and y=0.0,0.05,0.1 for samplesx=0,y=0.0 and x=0.0,y=0.05 exhibits a trigonal phase, while the other samples correspond to an orthorhombic phase. The effects on the crystal structure through simultaneous substitution on both La and Mn sites are complex. On the one hand, inclusion of Dy [18, 19, 20], with smaller ionic radius than La ions, and inclusion of Zn [16, 17], with a greater ionic radius than Mn ions, introduces distortions in the octahedra and, therefore, in the crystallographic structure. On the other hand, introduction of divalent Zn ions induces the change of Mn3+ for Mn4+ and the J-T distortion of the Mn3+ neighboring Zn2+ differs from one of the Mn3+ no-neighboring Zn2+ ions [17]. Figure 1 shows the XRD pattern forx=0 due to different Zn substitution: y=0 (bottom panel), y=0.05 (middle panel), and y=0.1 (top panel).

Figure 1.

X-ray diffraction patterns for La1 xDyxMn1 yZnyO3 (a) x=0.0,y=0.1; (b) x=0.0y=0.05; and (c) x=0y=0.0. Major Miller indices are indicated.

X-ray diffraction data were refined by the Rietvel method—Fullproff program [21]. Information, such as phase structure, identification of planes, cell volumes, and cell densities were obtained. Table 1 presents some crystallographic parameters corresponding to La1 xDyxMn1 yZnyO3 samples.

Samplex=0.0x = 0.05
y = 0.0y = 0.05y = 0.10y = 0.0y = 0.05y = 0.10
StructureTrigonal/R3COrthorhombic/Pbnm
Space group16762
Z64
a (Å)5.53635.54125.54595.53955.53545.5354
b (Å)5.53635.54125.50565.49655.50175.5072
c (Å)13.367013.36527.79957.77657.78637.7923
Cell volume (Å3)354.831355.400238.1236.775237.124237.542
Density (g/cm3)6.7856.8046.8076.8656.8406.828

Table 1.

Structural parameters for La1 − xDyxMn1 − yZnyO3 samples. Values in white cells are from Ref. [22].

Figure 2 shows a structural phase transition for x=0.0 samples, wheny>0.05 it is from triclinic to orthorhombic. This signals the high structural distortion introduced by Zn ions. Therefore, the sample x=0.0, y=0.1 relaxes the stress by symmetry breaking. As will be seen later, this behavior is also observed at Dy concentrations for x=0.05, 0.10 series, although for these samples, relaxation is achieved by octahedra tilting increases.

Figure 2.

Variation of unit cell parameter with Zn doping (0.0<y<0.1). (a) Cell parameter a (circle) and b (triangle); and (b) Cell parameter c (square). x=0.0 in all cases.

Use of 3D reconstruction tools of structures (like Visualization for Electronic Structural Analysis) allows obtaining interatomic distances, MnO and angles, θi in the MnOMn bonds (Figure 3). In the case of manganites, a distortion of structure is directly related to their magnetic properties. This phenomenon, known as the J-T effect, is a structural phase transition driven by the coupling between the orbital state and the vibronic configuration of the crystal lattice. The J-T coupling to the lattice manifests itself in changes in bond lengths MnO and θi angles, as well as in orbital order [3, 4].

Figure 3.

Unit cell of perovskite structure indicating the bond distances, Mn/ZnO1 and MnZnO1 and bond angles, θ1 and θ2. Atoms: O (red), Mn/Zn (black), and La/Dy (Green).

Regarding the measurements of structural distortion in perovskites with orthorhombic structure, two relations have been used to evaluate said distortion from the distances of the manganese with their neighboring oxygenMnOMn in the plane and with the apical oxygen ions. The octahedra distortion, Doct, can be calculated by [23, 24]:

Doct=10×i=1,6MnOiMnOaverageMnOaverage,Doct=1/6×i=1,6MnOiMnOaverageMnOaverage2E1

whereMnOi represents the bond distances between the manganese ion and each of the six oxygens of the octahedra that surround it. These distances will depend, among others, on the ionic radii and mainly on the doping of the ion in site B. Figure 2 shows these distances of the manganese MnO2 in the plane and MnO1 with the apical oxygens.

Octahedra distortion, from the energetic point of view known as J-T distortion, is evaluated through the relation [25]:

λJT=13×MnOiMnOaverage2E2

Although Eqs. (1) and (2) are different ways of evaluating the distortion, both are dimensionless; their range is between [0, 1], and their behavior is similar. In the manganites studied herein, an inverse relationship between distortions and the magnetic moment of individuals or clusters has been observed forT>Tc. Where does this structure-magnetism relationship originate? Clearly, the J-T effect results in spontaneous splitting of the energy levels, reducing the total energy of the Mn3+ also resulting in a spontaneous distortion of the octahedrons that increase their elastic energy at the cost of a reduction in energy of certain electron orbitals, thus, resulting in a net reduction in energy.

The J-T distortions obtained through Eq. (2) showed anomalous behavior (Figure 4): Zn increases in the range of 0<y<0.05 increases the J-T distortion and for y>0.05, a reduction in the J-T distortion is observed (Figure 4a). This is due to the difference between the ionic radii of Zn and Mn ions [16]. However, when x>0.05, the accumulated strain of the system is relaxed via octahedral tiltingsimilar to the introduction of dislocation when the maximum stress tolerated is reached in a crystal [22]. Tilting as a relaxation mechanism is observed in Figure 4b, where greater tilting corresponds to samples with lower J-T distortion. The same anomalous behavior is observed by calculating octahedra distortions (Eq. (2)). As will be seen ahead, the activation energy that allows the magnetic exchange is accompanied by a charge exchange facilitated by these lattice distortions.

Figure 4.

(a) Jahn-Teller distortion versus Zn composition for x = 0.10 samples (squares) and x = 0.05 samples (circles); and (b) Jahn-Teller distortion vs. Octahedra tilting, ω, by using 2ωi+θi=180° [22].

Local distortions affect directly the magnetic properties as a function of T, that is, exchange constants are proportional to the orthorhombic strain [8]. Also, magnetic moments and activation energies depend on distortion [22].

Advertisement

3. Local magnetism-electron paramagnetic resonance (EPR) analysis

Electron paramagnetic resonance is an important technique to study the microscopic nature of local interactions in magnetic materials and, particularly, in manganites [26], which in many cases show short-range interactions for T>Tc. This technique contrasts with the magnetization and susceptibility techniques that provide global information about the exchange mechanisms and the possible spontaneous creation of clusters at high temperatures.

The EPR signal corresponds to /dH, where χ = M/H is the susceptibility. In the paramagnetic region, where M is linear with H, the double integral of the EPR intensity is proportional to the magnetization. From EPR measurements, we can build the dependence of the inverse of the temperature-dependent susceptibility, χ1T.To carry out the EPR measurements, a Bruker ESP-300 spectrometer was used in the radiation X-band with 9.408 GHz frequency and in a temperature range from 10 to 290K. Figure 5a shows the resonant signals for manganite La0.9Dy0.1MnO3 in the temperature range 220<T<290K. The inset of Figure 5a is the integral of the EPR signal. It is observed that for high temperatures, the intensity of the EPR signal increases in all cases as the temperature decreases. For T > TC, it is found that the intensity of resonance line fits by the expression [28]:

Figure 5.

(a) EPR signal of La0.9Dy0.1MnO3 for different temperatures, Inset. Intensity obtained by integration of the EPR signal; (b) EPR signal intensity as a function of temperature, showing short-range interaction behavior for T>181K for the sample La0.9Dy0.1Mn0.95Zn0.05O3; and (c) resonant field vs. T for LaMn0.9Zn0.1O3 [27].

IPM=I0eE/kBTE3

where IPM is the intensity extracted from the resonance line, I0 is a fitting parameter, and E is an activation energy.

Figure 5c shows the variation of the resonant field with temperature. A change in the value of Hr is observed for T>Tc. This local signal evidences an FM phase in the PM region [27, 29]. This local magnetism is strongly dependent on doping and oxygen content, as Oseroff et al. [28] show in their work on collective spin dynamics above Tc in manganites (La1 − xCaxMnO3).

The peak-to-peak EPR linewidth, HPP, can also be used to confirm the presence of short-range interacting magnetic entities. In magnetic resonance, the EPR linewidth is related to the relaxation mechanisms of the magnetic units, whether individual spins or spin-coupled systems. A decrease in HPP as temperature decreases indicates PM behavior of the sample. However, in the case of La1xDxMnZnO manganites, an increase in HPP is observed as temperature decreases, indicating the presence of short-range interactions. In Figure 6a, the arrows indicate the corresponding critical temperatures for each sample. HPP increases as temperature decreases, indicating a greater range of the collective effects, approaching the FM phase. Tovar et al. [8] have found a direct dependence on temperature for the J-T distortion and the EPR linewidth.

Figure 6.

(a) Temperature dependence on the resonance linewidth; the corresponding TC values of the x=0.0, y=0.05 (red) and x=0.05, y=0.1 (black) samples are indicated; and (b) linewidth as a function of temperature or the x=0.0, y=0.05 sample; red line corresponds to the fitting line to obtain the activation energy value as Ea = 0.068 eV.

The energy transferred in the relaxation process is related to jumps of polarons thermally activated between the Mn4+ and Mn3+ states. Therefore, the jump rate of the charge carriers will determine the half-life of the spin state and, therefore, determines the EPR linewidth and the conductivity. The dependence of the EPR linewidth as a function of temperature in the paramagnetic region can be evaluated by [30]:

HPPT=ATeEakBTE4

Ea values showed on Table 2 are obtained by Eq. (4). Previous studies on relaxation modes in mixed-valence manganites have shown that the EPR predominant signal corresponds to Mn3+Mn4+ relaxation and the internal relaxation of the ions through the lattice. Shengelaya et al. [31] showed that the relaxation, R, of the Mn4+ (s) with the lattice can be negligible compared with the relaxation Mn3+ (σ) with the lattice (L), while the largest signal corresponds to the relaxation Mn3+Mn4+ and Mn4+Mn3+. Thus, a “Bottleneck” is formed, corresponding to the charge transfer between the two magnetic subsystems (3+ and 4+), while a slower relaxation process, Mn3+ with the network occurs. A schematic representation proposed is presented in Figure 7.

Zn content yDy x=0.00Dy x=0.05Dy x=0.1
Activation Energy EaeVActivation Energy EaeVActivation Energy EaeV
0.000.0294(8)0.0656(7)0.0596(1)
0.050.0683(9)0.0555(9)0.0584(5)
0.100.0487(7)0.0477(6)0.0624(6)

Table 2.

Activation energies as a function of Zn doping (y). White cells from Ref. [27].

Figure 7.

A block diagram showing the energy flow paths for the Mn4+ and Mn3+ spin subsystems and the lattice. The relaxation rates RσS,R represent relaxation between the subsystems. The thickness of the arrows is a measure of the magnitude of the particular relaxation rate [31].

Furthermore, the intensity of the EPR signal is proportional to the static magnetic susceptibility, χe, it is possible to explain the bottleneck regime due to the coupling of the spins (s) and (σ) as:

χS=χSo1+λ´χσo1λ´2χσoχSoE5

where χSo and χσo are the ion susceptibilities without exchange of Mn4+andMn3+, respectively, and λ is the dimensionless coupling constant.

λ´=zJngsgσμB2E6

with n being the number of spins per cm3 and gs,gσ being the g-factors of the Mn4+ and Mn3+ ions, respectively. For TTC, the EPR signal intensity drops faster than predicted by Eq. (5), associated with a transition from the bottleneck to an isothermal regime: RσLRσs, where the relaxation, Rσs, is T independent [31].

For manganites, the contribution of each coexistent phase has been evaluated by means of the EPR technique. In this case, FM clusters also contribute to the magnetization of the material, so that the total magnetization is the result of the PM and FM contributions above TC by [32]:

MHT=xTχPMTH+1xTMFMHTE7

where x and 1x are the fractions of PM and FM signals, respectively. The intensity of ESR lines are:

IPMTxTχPMTandIFMT1xTMFMTE8

For TTc, the EPR signal consists only of a PM line; then xTTc = 1, and MHT/H=χPM, while xT<Tc = 0 and MHT/H = χFM.As Eq. (8) shows, it is possible to find the fraction of the FM phase in the samples by subtracting the EPR intensity from the magnetic susceptibility (Figure 8). The inset shows the fraction as a function of temperature.

Figure 8.

(°) IPMT, (solid line) M/HT at 3.5 kG and (•) IFMT vs. T for La0.75Ca0.25MnO3. Adapted from Ref. [32].

Dormann and Jaccarino [33] proposed the following Huber approximation for a coupled system (clusters) in the PM state:

HppTχsT/χEPRTHppE9

where χsT=C/T is the susceptibility of the individual ions M3+ and M4+; Hpp corresponds to spin-only interactions (Hpp=2600G reported by Causa et al. [34] for perovskites with A = La) and χEPRT the PM signal of the coupled system.

C=NAμeff2μB2/3kBE10
Advertisement

4. Macroscopic magnetism: MTH

Thereafter, we will refer to one of the most-used macroscopic techniques to characterize magnetic materials: direct measurement of magnetization as a function of temperature and applied field. This technique offers vast information regarding the type of transition [35, 36, 37], transition temperature [38, 39], magnitude of the magnetic moments [22, 38, 40, 41], and the possible presence of clusters in the PM region, which is quite characteristic of manganites. Next, we will present results on the use of this technique in manganites.

4.1 Type of transition

To determine the nature of the FM-PM phase transition (first- or second-order), it is beneficial to use Arrott’s plot [42]: μ0H/M vs. M2. According to the Banerjee criterion curves, showing positive or negative slopes without inflection points are characteristic of second- or first-order transitions, respectively [35, 43]. Figure 9 shows M vs. T and μ0H/M vs. M2 for La2/3Sr1/3MnO3 [43]. In Figure 9b, in Arrot’s plot for μ0H/M vs. M2, the slope is always positive, indicating the second order of the phase transition for this sample.

Figure 9.

(a) Temperature dependence on FC magnetization for La2/3Sr1/3MnO3 under an applied magnetic field of H=5Oe; and (b) Arrott’s plot for μ0H/M vs. M2 for La2/3Sr1/3MnO3 [36].

4.2 Critical temperature and magnetic moment

Magnetization vs. T is used to find the Curie temperature, Tc, or Neel temperature, TN, in manganites; also, this technic provides information about the presence of magnetic clusters in PM phase. Figure 10a presents the curves M vs. T for the LaMn0.95Zn0.05O3 sample, where TC=185K is obtained from dM/dT. The effect of doping on the Tc value is generally one of the first factors to be evaluated in manganites. For low Dy doping in La1 xDyxMn1 yZnyO3, two competing effects have been observed: firstly, the large magnetic moment of Dy favors FM by increasing the value of Tc, and secondly, the small ionic radius of Dy (close to 50% of the La ionic radius) distorts the lattice, destroying the FM of the MnOMn chains, which leads to a reduced Tc value. As seen in Figure 10b, Zn determines which of the two effects predominates; at low Zn concentrations, the crystallographic distortion by Dy predominates and for concentrations of Zn>0.05, breaking of the MnOMnchains reduces the effect of lattice distortions by Dy and allows the effect due to the magnetic moment of Dy.

Figure 10.

(a) M vs. T curves for x=0.00,y=0.05show magnetization measured at field cooling (open) and the zero-field cooling (close), at H=0.001T. Inset. The derivative of magnetization dM/dT vs. T [27]; (b) Tc vs. Zn composition. Circles for 0.05Dy series and squares for 0.10Dy series [22]; and (c) χ1vs.T for x=0.00,y=0.05. Linear fit (red) at high temperatures. CW temperatures, θCW, is indicated [27].

In manganites, it has been observed that the inverse of the susceptibility χ1 vs. T does not satisfy the Curie-Weiss law. However, over a wide range of temperatures (TTc), χ1T is linear and can be described by the Curie-Weiss model (Figure 10c); the Curie constant (Eq. (10)) can be used to estimate the effective magnetic moments, as well as the doping effect on the magnetic moment. Overall, for all the samples, it has been found that the experimental effective moments are higher than the theoretical values [38, 39, 41, 44], evidence of the presence of clusters in the material. For the La1 xDyxMn1 yZnyO3 manganite, the theoretical m values are expressed by [22]:

μeffth=0.862yμeffthMn3+2+0.12+yμeffthMn4+2+xμeffthDy3+212E11

according to the stoichiometric formula.

To visualize the cluster behavior throughout the nonlinear PM region, we calculated the values of μeffexp=2.83CTμB for T>TC from the difference of χ1T/T=1/CT [38]. Figure 11b shows μeffexp vs. T for LaMn1 yZnyO3 that μeff increases when T decreases in the temperature range corresponding toaχ1T that has a positive curvature. This suggests an increase in the strength of the exchange coupling with T.

Figure 11.

(a) Theoretical and experimental effective magnetic moments at TTC for LaMn1 yZnyO3; and (b) experimental effective magnetic moments as a function of temperature on La0.95Dy0.05Mn1 yZnyO3 for y=0.0 (black), y=0.05 (red), y=0.01 (blue) [22] (b).

4.3 Coupled moments in a mean-field approximation

The Mean-field theory of coupled moment pairs in an effective molecular field approximation, Be, has been discussed in the literature [28, 34, 45]. The molecular field constants, as well as the magnetic susceptibilities that depend highly on the three main crystal axes direction, result in a slight deviation from the Curie Weiss (C-W) law.

The effective field, Be=2z1JM/Ng2μB2 on the coupled moment pair corresponds to a molecular field coefficient, λ=2z1J/Ng2μB2. λ differs from the Weiss model in that z is replaced by (z1). In the Constant-coupling approximation (CCA) [34, 46, 47], this local field “aligns” the magnetic moments of some Mn3+ and Mn4+, resulting in FM clusters (S1+S2)- where S1 and S2 correspond to the Mn3+ and Mn4+ spins, respectively, which conform the cluster unit. This can be modeled, even in the paramagnetic region, by a Heisenberg-type isotropic interaction between pairs of Mn3+ and Mn4+ ions, subjected to the action of the effective field:

H=2JS1S2+gμBS1+S2H+BeE12

where H is the external field. Because the chosen pair is arbitrary, they must all have the same magnetic moment as every other pair. This condition requires that [47]:

½NgμBSz´=ME13

with Sz´=S1+S2. From Eq. (12) [10], the inverse of susceptibility is obtained as:

χ1=T12zJ/kB/CE14

Expanding Eq. (14) around jjC, where j=J/kBT is the reduced exchange constant, we obtain an expression for the inverse of the susceptibility in terms of Tc:

χ1=TTc/ρCE15

with ρ=2Jc(2z1Jc+z4. This expression reproduces a deviation from the linear C-W behavior for FM interactions.

A Heisenberg model over all points of the magnetic lattice ij is insoluble because said magnetic solid has on the order of 1022 magnetic moments,

H=ijJijSiSjE16

If we consider only first neighbors in the FM Heisenberg Hamiltonian, it can be rewritten as [48]:

H=2Jn,n´SnSn´gμBHnSnzE17

From a power series expansion in j=J/kBT [49], the susceptibility is obtained as:

χ=Ng2μB2SS+13kBTl=0aljlE18

with N being the number of sites in the sample, gμBS the magnetic moment associated with the spin, S at each lattice point, z the coordination number, and j being a dimensionless value: j>0 for FM and j<0 for AFM. Finally, Eq. (18) is rewritten as [48]:

χ01=TC1τ4/3fτE19

with τ=TC/T, fτ=1b1τ1b2τ/1a1τ1a2τ1a3τ and ai,bi parameters obtained by the Padé approximants that only depend on the crystal structure and on the S value. ai,bi parameters for z=6 are presented in Table 3.

Sa1a2a3b1b2
3/2−1/21,9677−1/1.62282−1/0.38910−1/1.743001/0.38911
2−1/51.0423−1/1.73724*−1/1.89133*

Table 3.

ai,bi coefficients obtained for Z = 6, from Padé approximants.

Corresponding to conjugate values reported by Gammel et al. [48] (* a3=b2).

With the coefficients indicated in Table 3, Eq. (19) shows a deviation from the straight lines predicted from the C-W theory (Figure 12a). The blue line represents experimental data for La0.9Dy0.1MnO3, where clusters are present at T>TC and J is unknown. For general Heisenberg Hamiltonians, where more than one relevant exchange constant Ji(Ji =1, 2, 3, 4) is required, a high-temperature expansion (HTE) has been developed [50, 51]. Figure 12b shows χ1vs.T experimental and fitting data for La0.95Dy0.05Mn0.9Zn0.1O3. The J value obtained is J/KB=15K. The effect of Dy and Zn doping is evident on J (see Figure 12c) [27].

Figure 12.

χ1vs.T (a) using Eq. (19) for z=6, s=3/2 (black line), S=2 (red line) and the experimental values [22] for La0.9Dy0.1MnO3 (blue line); (b) La0.95Dy0.05Mn0.9Zn0.1O3 [22]. Black line, experimental data; red line, fit to HTE algorithm; dashed line, C-W model. J exchange constant values for both series from HTE algorithms; and (c) J values vs. y for La1 xDyxMn1 yZnyO3 [27].

Advertisement

Acknowledgments

The authors thank the BC foundation for its support in measuring the magnetic properties of the manganites and thank José Fernando López Toro for authorizing the use of diverse results from his Ph.D. thesis to illustrate the analysis presented herein. This publication was partially supported by the Science Faculty and physics department of the Universidad Nacional de Colombia.

Advertisement

Conflict of interest

The authors declare no conflict of interest.

References

  1. 1. Wollan EO, Koehler WC. Neutron Diffraction Study of the Magnetic Properties of the Series of Perovskite-Type Compounds [(1-x)La, xCa]MnO3. Physics Review. 1955;100:545. DOI: 10.1103/PhysRev.100.545
  2. 2. Bersuker I. The Jah-Teller Effect. Cambridge: Cambridge University Press; 2006. Chapter 8. pp. 479-597. DOI: 10.1017/CBO9780511524769
  3. 3. Van Aken BB, Jurchescu OD, Meetsma A, Tomioka Y, Tokura Y, Palstra TTM. Orbital-Order-Induced Metal-Insulator Transition in La1-xCaxMnO3. Physical Review Letters. 2003;90:066403. DOI: 10.1103/PhysRevLett.90.066403
  4. 4. Qiu X, Proffen T, Mitchell JF, Billinge SJL. Orbital Correlations in the Pseudocubic O and Rhombohedral R Phases of LaMnO3. Physical Review Letters. 2005;94:177203. DOI: 10.1103/PhysRevLett.94.177203
  5. 5. Goodenough JB. Magnetism and the chemical bond. New York, NY: Interscience Publishers; 1963. ASIN: B002TT366A
  6. 6. Booth CH, Bridges F, Snyder GJ, Geballe TH. Evidence of magnetization-dependent polaron distortion in La1-xAxMnO3, A=Ca, Pb. Physical Review B. 1996;54:15606. DOI: 10.1103/PhysRevB.54.R15606
  7. 7. Billinge SJL, DiFrancesco RG, Kwei GH, Neumeier JJ, Thompson JD. Direct Observation of Lattice Polaron Formation in the Local Structure of La1-xCaxMnO3. Physical Review Letters. 1996;77:715. DOI: 10.1103/PhysRevLett.77.715
  8. 8. Tovar M, Alejandro G, Butera A, Caneiro A, Causa MT, Prado F, et al. ESR and magnetization in Jahn-Teller-distorted LaMnO3-δ: Correlation with crystal structure. Physical Review B. 1999;69:10199. DOI: 10.1103/PhysRevB.60.10199
  9. 9. Liu H, Zhang H, Li Y, Chen Y, Chen L, Dong X, et al. Magnetism and Resistances of Slightly Dy Doped LaMnO3 Solid Solutions. Journal of Superconductivity and Novel Magnetism. 2012;25:1049-1054. DOI: 10.1007/s10948-011-1348-5
  10. 10. Fabrèges X, Mirebeau I, Bonville P, Petit S, Lebras-Jasmin G, Forget A, et al. Magnetic order in YbMnO3 studied by neutron diffraction and Mössbauer spectroscopy. Physical Review B. 2008;78:214422. DOI: 10.1103/PhysRevB.78:214442
  11. 11. Wehrenfennig C, Meier D, Lottermoser T, Lonkai T, Hoffmann J-U, Aliouane N, et al. Incompatible magnetic order in multiferroic hexagonal DyMnO3. Physical Review B. 2010;82:100414. DOI: 10.1103/PhysRevB.82.100414
  12. 12. Lüscher A, Sushkov OP. Long-range dynamics of magnetic impurities coupled to a two-dimensional Heisenberg antiferromagnet. Physics Review. 2005;71:064414. DOI: 10.1103/PhysRevB.71.064414
  13. 13. Engel J, Wesse S. From enhanced to reduced quantum antiferromagnetism by tuning a magnetic impurity. Physical Review B. 2009;80:094404. DOI: 10.1103/PhysRevB.80.094404
  14. 14. Ghosh K, Ogale SB, Ramesh R, Greene RL, Venkatesan T, Gapchup KM, et al. Transition-element doping effects in La0.7Ca0.3MnO3. Physical Review B. 1999;59:533. DOI: 10.1103/PhysRevB.59.533
  15. 15. Hébert S, Martin C, Maignan A, Retoux R, Hervieu M, Nguyen N, et al. Induced ferromagnetism in LaMnO3 by Mn-site substitution:The major role of Mn mixed valency. Physical Review B. 2002;65:104420. DOI: 10.1103/PhysRevB.65.104420
  16. 16. Álvarez-Serrano I, Pico C, Veiga ML. Structural characterization, electric and magnetic behavior of Zn-doped manganites. Solid State Sciences. 2004;6:1321. DOI: 10.1016/j.solidstatesciences.2004.07.022
  17. 17. Tang S, Yue S, Zhang Y. Jahn-Teller distortion induced by Mg/Zn substitution on Mn sites in the perovskite manganites. Physics Letters A. 2003;319:530-538. DOI: 10.1016/j.physleta.2003.10.061
  18. 18. Mitra C, Raychaudhuri P, Dhar SK, Nigam AK, Pinto R, Pattalwar SM. Evolution of transport and magnetic properties with dysprosium doping in La0.7-xDyxSr0.3MnO3 (x= 0-0.4). . Journal of Magnetism and Magnetic Materials. 1999;192:130. DOI: 10.1016/S0304-8853(98)00388-6
  19. 19. Terai T, Kakeshita T, FuKuda T, Saburi T, Takamoto N, Kindo K, et al. Electronic and magnetic properties of (La-Dy)0.7Ca0.3MnO3. Physical Review B. 1998;58:14908. DOI: 10.1103/PhysRevB.58.14908
  20. 20. Xu S, Tong W, Fan J, Gao J, Zha C, Zhang Y. Influence of doped Dy on magnetic and electronic properties in La0.67-xDyxSr0.33MnO3. Journal of Magnetism and Magnetic Materials. 2005;288:92-105. DOI: 10.1016/j.jmmm.2004.08.022
  21. 21. Rodríguez-Carvajal J. Recent advances in magnetic structure determination neutron powder diffraction. Physica B. 1993;192:55-69. DOI: 10.1016/0921-4526(93)90108-I
  22. 22. López-Toro JF, Lezama L, Salazar D, Mendoza A. Influence of Non-magnetic Dilution on the Magnetic Properties of La1-xDyxMn1-yZnyO3 Perovskites at High Temperature. Physica Status Solidi A: Applications and Materials Science. 2021;219(1-8):2100513. DOI: 10.1002/pssa.202100513
  23. 23. Alonso JA, Martínez-Lopez MJ, Casais MT, Fernández-Díaz MT. Evolution of the Jahn-Teller Distortion of MnO6 Octahedra in RMnO3 Perovskites (R = Pr, Nd, Dy, Tb, Ho, Er, Y): A Neutron Diffraction Study. Inorganic Chemistry. 2000;39:917-923. DOI: 10.1021/ic990921e
  24. 24. Rodríguez-Carvajal J, Hennion M, Moussa F, Moudden AH, Pinsard L, Revcolevschi A. Neutron-diffraction study of the Jahn-Teller transition in stoichiometric LaMnO3. Physical Review B. 1998;57:R3189. DOI: 10.1103/PhysRevB.57.R3189
  25. 25. Blasco J, Ritter C, García J, de Teresa JM, Pérez-Cacho J, Ibarra R. Structural and magnetic study of Tb1-xCaxMnO3 perovskites. Physical Review B. 2000;62:5609. DOI: 10.1103/PhysRevB.62.5609
  26. 26. Santiago-Teodoro M, Hernández-Cruz L, Montiel-Sánchez H, Álvarez-Lucio G, Flores-González MA, Legorreta-García F. Synthesis, Microstructure and EPR of CaMnO3 and EuxCa1-xMnO3 Manganite, Obtained by Coprecipitation. Journal of the Mexican Chemical Society. 2011;55(4):204-207. DOI: 2073-4352/11/5/473
  27. 27. López JF. PhD Thesis, Universidad Nacional de Colombia; 2022
  28. 28. Oseroff SB, Torikachvili M, Singley J, Ali S, Cheong S-W, Schultz S. Evidence for collective spin dynamics above the ordering temperature in La1-xCaxMnO3. Physical Review B. 1996;53(10):6521. DOI: 10.1103/PhysRevB.53.6521
  29. 29. Eremina RM, Fazlizhanov II, Yatsyk IV, Sharipov KR, Pyataev AV, Zavoisky EK, et al. Phase separation in paramagnetic Eu0.6La0.4-xSrxMnO3. Physical Review B. 2011;84:064410. DOI: 10.1103/PhysRevB.84.064410
  30. 30. Yang J, Rong X, Suter D, Sun YP. Electron paramagnetic resonance investigation of the electron-doped manganite La1-xTexMnO3 (0.1≤x≤0.2). Physical Chemistry Chemical Physics. 2011;13:16343-16348. DOI: 10.1039/c1cp21807g
  31. 31. Shengelaya A, Zhao G-M y, Keller H, Müller KA. EPR Evidence of Jahn-Teller Polaron Formation in La1-xCaxMnO3+y. Physical Review Letters. 1996;77(26):5296. DOI: 10.1103/PhysRevLett.77.5296
  32. 32. Ccahuana DL, Winkler E, Prado F, Butera A, Ramos CA, Causa MT, et al. Magnetic phase coexistence in CMR manganites: ESR evidence. Physica B. 2004;354:55-58. DOI: 10.1016/j.physb.2004.09.020
  33. 33. Dormann E, Jaccarino V. High temperature EPR line widths in MnO and MnS. Physics Letters. 1974;48A:81. DOI: 10.1016/0375-9601(74)90409-5
  34. 34. Causa MT, Tovar M, Caneiro A, Prado F, Ibañez G, Ramos CA, et al. High-temperature spin dynamics in CMR manganites: ESR and magnetization. Physical Review B. 1998;58(6):3233. DOI: 10.1103/PhysRevB.58.3233
  35. 35. Banerjee BK. On a generalised approach to first and second order magnetic transitions. Physics Letters. 1964;12:16-17. DOI: 10.1016/0031-9163(64)91158-8
  36. 36. Chebaane M, Bellouz R, Oumezzine M, Hlil EK, Fouzri A. Copper-doped lanthanum manganite La0.65Ce0.05Sr0.3Mn1-xCuxO3 influence on structural, magnetic and magnetocaloric effects. RSC Advances. 2018;8:7186-7195. DOI: 10.1039/C7RA13244A
  37. 37. Manh TV, Shinde KP, Nanto D, Lin H, Pham Y, Razaq DS, et al. Critical behavior and magnetocaloric effect in La0.7Ba0.25Nd0.05Mn1-xCuxO3. AIP Advances. 2019;9:035345. DOI: 10.1063/1.5079842
  38. 38. Khammassi F, Lopez JF, Chérif W, Mendoza A, Lanceros-Mendez S, Salazar D, et al. Short-range magnetic behavior in manganites La0.93K0.07Mn1-xCuxO3 (0.0≤x≤0.09) above the Curie temperature. Journal of Physics D. 2021;54:175001. DOI: 10.1088/1361-6463/abde6b
  39. 39. Issaoui F, Bejar M, Dhahri E, Bekri M, Lachkar P, Hill EK. Crystal, spin glass, Griffiths phases and magneticaloric properties of the Sr1.5Nd0.5MnO4 compound. Physics B. 2013;414:42-49. DOI: 10.1016/j.physb.2012.12.039
  40. 40. Vergara J, Ortega-Hertogs RJ, Madurga V, Sapiña F, El-Fadli Z, Martínez E, et al. Effect of disorder produced by cationic vacancies at the B sites on the electronic properties of mixed valence manganites. Physical Review B. 1999;60:1127. DOI: 10.1103/PhysRevB.60.1127
  41. 41. Kamel R, Tozri A, Dhahri E, Hlil EK. Anomalous behavior above the Curie temperature in (Nd1-xGdx)0.55Sr0.45MnO3 (x=0, 0.1, 0.3 and 0.5). RSC Advances. 2019;9:27541-27548. DOI: 10.1039/c9ra03303c
  42. 42. Standley HE. Introduction to phase transitions and critical phenomena. London: Oxford university Press; 1971
  43. 43. Mira J, Rivas J, Rivadulla F, Vázquez-Vázquez C, A. López-Quintela MA. Change from first- to second-order magnetic phase transition in La2/3(Ca, Sr)1/3 MnO3 perovskites. Physical Review B. 1999;60:2998. DOI: 10.1103/PhysRevB.60.2998
  44. 44. Banik S, Banu N, Das I. Evolution from non-Griffiths phase to Griffiths phase: Giant enhancement of magnetoresistance in nanocrystalline (La0.4Y0.6)0.7Ca0.3MnO3 compound. Journal of Alloys and Compounds. 2018;745:753-760. DOI: 10.1016/j.jallcom.2018.02.243
  45. 45. Oguchi T. A Theory of Antiferromagnetism, II. Progress in Theoretical Physics. 1955;13(2):148-160. DOI: 10.1143/PTP.13.148
  46. 46. Allub R, Alascio B. Effect of Disorder on the Magnetic and Transport Properties of La1-xSrxMn03. Solid State Communications. 1996;99:9613-9617. DOI: 10.1016/0038-1098(96)00337-7
  47. 47. Smart JS. Effective field theories of magnetism. Philadelphis: Saunders; 1966
  48. 48. Gammel J, Marshall W, Morgan L. An Application of Pade Approximants to Heisenberg Ferromagnetism and Antiferromagnetism. Proceedings of the Royal Society of London A. 1963;275:257-270. DOI: 10.1098/rspa.1963.0169
  49. 49. Rushbrooke GS, Wood PJ. On the Curie points and high temperature susceptibilities of Heisenberg model ferromagnetics. Molecular Physics: An International Journal at the Interface Between Chemistry and Physics. 1958;1(3):257-283. DOI: 10.1080/00268975800100321
  50. 50. Schmidt H-J, Lohmann A, Richter J. Eighth-order high-temperature expansion for general Heisenberg Hamiltonians. Physical Review B. 2011;84:104443. DOI: 10.1103/PhysRevB.84.104443
  51. 51. Lohmann A, Schmidt H-J, Richter J. Tenth-order high-temperature expansion for the susceptibility and the specific heat of spin-s Heisenberg models with arbitrary exchange patterns: Application to pyrochlore and kagome magnets. Physical Review B. 2014;89:014415. DOI: 10.1103/PhysRevB.89.014415

Written By

Aminta Mendoza and Octavio Guzmán

Submitted: 19 July 2022 Reviewed: 23 August 2022 Published: 17 October 2022