• Aucun résultat trouvé

Anisotropy of transport properties in the Kondo compound CePt 2Si2 : experiments and theory

N/A
N/A
Protected

Academic year: 2021

Partager "Anisotropy of transport properties in the Kondo compound CePt 2Si2 : experiments and theory"

Copied!
14
0
0

Texte intégral

(1)

HAL Id: jpa-00211102

https://hal.archives-ouvertes.fr/jpa-00211102

Submitted on 1 Jan 1989

HAL is a multi-disciplinary open access archive for the deposit and dissemination of sci- entific research documents, whether they are pub- lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Anisotropy of transport properties in the Kondo compound CePt 2Si2 : experiments and theory

A.K. Bhattacharjee, B. Coqblin, M. Raki, L. Forro, C. Ayache, D. Schmitt

To cite this version:

A.K. Bhattacharjee, B. Coqblin, M. Raki, L. Forro, C. Ayache, et al.. Anisotropy of transport prop-

erties in the Kondo compound CePt 2Si2 : experiments and theory. Journal de Physique, 1989, 50

(18), pp.2781-2793. �10.1051/jphys:0198900500180278100�. �jpa-00211102�

(2)

Anisotropy of transport properties in the Kondo compound CePt2Si2 : experiments and theory

A. K. Bhattacharjee (1), B. Coqblin (1), M. Raki (2), L. Forro (2, *),

C. Ayache (2) and D. Schmitt (3)

(1) Laboratoire de Physique des Solides, Bât. 510, Université Paris-Sud, Centre d’Orsay,

91405 Orsay, France

(2) Service des Basses Températures, DRFG/CENG, 85 X, 38041 Grenoble Cedex, France

(3) Laboratoire Louis Néel, CNRS, 166X, 38042 Grenoble Cedex, France (Reçu le 29 mars 1989, accepté sous forme définitive le 19 mai 1989)

Résumé.

2014

Nous présentons ici des mesures de la résistivité électrique, du pouvoir thermoélectri- que et de la conductivité thermique d’un monocristal de CePt2Si2 de 1,5 à 300 K. Nous présentons

aussi un modèle théorique, basé sur l’Hamiltonien d’échange effectif et prenant en compte l’anisotropie du temps de relaxation des électrons de conduction, pour expliquer la forte anisotropie observée dans les propriétés de transport du composé Kondo du cérium CePt2Si2.

Abstract.

2014

Measurements of the electrical resistivity, thermoelectric power and thermal

conductivity from 1.5 to 300 K in single crystal CePt2Si2 are reported here. A theoretical model, based on the effective exchange Hamiltonian and taking into account the anisotropy of the

conduction electron relaxation time, is also presented to explain the strong anisotropy observed in the transport properties of the cerium Kondo compound CePt2Si2.

Classification

Physics Abstracts

72.15E - 72.15J

-

72.15Q

-

71.70C

-

75.20H

1. Introduction.

Cerium Kondo compounds have been extensively studied in the last years from both an

experimental and a theoretical point of view. At sufficiently high temperatures, i.e. for temperatures larger than the Kondo temperature Tk and the overall crystal-field splitting,

these compounds have a magnetic susceptibility which follows a Curie-Weiss law with a

magnetic moment corresponding roughly to the Ce 3, trivalent ion or to the 4f1 configuration

and show a decrease of the magnetic resistivity with increasing temperature, generally in Log T in a given température range. At very low temperatures compared to Tk, cerium

Kondo compounds are characterized by a heavy-fermion behaviour giving rise to enormous

values of the magnetic susceptibility and electronic specific heat term as experimentally

observed in CeA13 [1], CeCu6 [2], CeCu2Si2 [3], CePtSi [4], CePtln [5], CeInCu2 [6], CePd3B [7], CeCu4Ga [8], CeCu4Al [9] and many other cerium compounds. Three different

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:0198900500180278100

(3)

behaviours have been found at very low temperatures in these compounds : many of them,

such as CeAl2 [10], CeB6 [11], CeInAg2 [12] and roughly thirty other cerium compounds,

order magnetically, generally in an antiferromagnetic or modulated magnetic order. On the other hand, CeCu2Si2 becomes superconducting below roughly 0.6 K [3]. Finally, the other compounds, such as CeAl3 [1], CeCu6 [2], CeCu2Si2 [3], CeInCu2 [6], CeRu2Si2 [13] or CePt2Si2 [14] were initially thought to exhibit no long-range magnetic order and to become

non magnetic ; the problem is now more controversial since a very weak magnetic order has

been detected in some compounds such as CeAl3 [15] and moreover strong short-range magnetic correlations have been any way observed in most of these compounds.

The magnetic and transport properties of cerium Kondo alloys and compounds have been

also extensively studied experimentally and theoretically in both the « high-temperature » and

« low-temperature » limits, i.e. for temperatures respectively larger and smaller than the Kondo temperature Tk. The transport properties, namely the electrical resistivity [16, 17], the magnetoresistivity [17], the Hall effect [18], the thermoelectric power [19] and the thermal

conductivity [20] of cerium Kondo alloys and compounds have been computed in detail by the Orsay group within a third-order calculation performed in the framework of the effective

exchange Hamiltonian [21] (the so-called « Coqblin-Schrieffer Hamiltonian »), which de-

scribes the large resonant scattering connected with the Kondo effect and takes into account the crystalline field effect of cerium. A good agreement with experimental data has been found in a broad temperature domain, i.e. for temperatures larger than Tk where the perturbation theory is valid, for most of the transport properties in many cerium systems.

First of all, the previous calculation [16] of the magnetic resistivity explains the experimental

behaviour of roughly forty cerium compounds at sufficiently high temperatures, i.e. an increase of the resistivity with increasing temperature, then a maximum corresponding roughly to the crystalline field splitting and finally a decrease in Log T at higher temperatures.

The « high-temperature » Log T decrease of the resistivity can be in fact considered now as a

signature of the Kondo effect in cerium compounds. This calculation has been extended to low temperatures [17] ; it yields a peak at the ordering temperature with a sharp decrease

below it and generally another slow decrease in approximately log T above it in the case of

compounds which order magnetically (such as CeAl2 or CeS). On the contrary, a continuous decrease with decreasing temperature below the crystal field maximum and a T2 behaviour at very low temperatures is generally obtained in compounds which become non magnetic (such

as CeAl3 or CeCu6) ; in CeCu2Si2, coherence effects which develop when Cerium atoms are

demagnetized, give rise to a second weak maximum before the low temperature decrease

[22].

The other transport properties have been computed within the same model in the « high temperature » limit. A negative magnetoresistivity has been found in the temperature domain around the Néel temperature, in good agreement with experiments in CeAl2 or CeB6 [17].

The temperature dependence of the thermoelectric power exhibits one or two, positive or negative, peaks which can reach large absolute values and are located at a temperature corresponding to a fraction 1/6 to 1/3 of the crystal-field splitting [19]. A qualitative agreement with experiments can be found in most cerium compounds. Moreover the calculations have

quantitatively accounted for thermoelectric power data in Ce1-xLaxAl3 alloys [19]. Finally,

the thermal conductivity has been recently computed in the same model [20] ; this theoretical

electronic contribution to the thermal conductivity behaves linearly at low and high

temperatures, while it exhibits an inflexion and even a maximum followed by a minimum at a

temperature approximately given by a fraction 1/3 to 1/2 of the crystal-field splitting. A

quantitative agreement is obtained with experimental data in CeAl2, CeCu2Si2 and

CePt2Si2 polycrystalline samples [20].

(4)

The purpose of thé présent paper is to study both experimentally and theoretically a new

feature observed in the transport properties of cerium Kondo compounds, namely, the anisotropy of the electrical resistivity, the thermoelectric power and the thermal conductivity

found in CePt2Si2 single crystals and to present a theoretical explanation of this anisotropy,

based on the effective exchange Hamiltonian of references [16, 21], by taking into account the anisotropy of the conduction electron relaxation time. Experimental evidence for anisotropy

effects has been previously found in the resistivity of CeCu2Si2 [23], CeA13 [24] and CePt2Si2 [25] single crystals and in the thermoelectric power of CeRu2Si2 single crystals [26].

The resistivity of CeCu2Si2 and CeAl3 is smaller along the c-direction than perpendicular to it,

in contrast to the case of CePt2Si2 [25].

In the present paper, we describe the experimental results on the transport properties of CePt2Si2 single crystals in the next section, the theoretical model in section 3 and the

comparison between experiment and theory in section 4.

2. Measurements of transport properties of single crystal CePt2Si2.

The ternary compound CePt2Si2 has been recently recognized as a new cerium Kondo

compound, which shows no sign of magnetic ordering or superconductivity down to 50 mK.

Let us summarize firstly the main known properties of tetragonal CePt2Si2. The magnetic resistivity of polycrystalline CePt2Si2 increases regularly up to a maximum at 76 K and then decreases logarithmically down to room temperature ; at low temperatures below roughly

15 K, the resistivity follows a law given by AT 2 [27]. The magnetic susceptibility of CePt2Si2 follows a Curie-Weiss law above 150 K with a magnetic moment close to that of Ce3+, then goes through a maximum around 60 K and is almost constant below 20 K reaching

a value of 3.6 x 10-3 emu/mole at 1.5 K, after correction of impurity effects [14, 27].

Measurements of the specific heat C yield an extrapolated value of C/T down to 0 K equal to

80 [14]-86 [27] mJ/mole K2 and a maximum of C/T close to 120 mJ/mole K2 at 2 K. Both the low temperature value of the magnetic susceptibility and the A coefficient of the low temperature T2 law of the resistivity decrease rapidly with pressure [27].

Recent measurements on CePt2Si2 single crystals have shown the strong effect of anisotropy

on the magnetization [28] and electrical resistivity [29]. The low temperature magnetization is larger along the (100) direction than along the (001) direction and the observed strong decrease of the initial susceptibility is associated with the possible existence of a metamagnetic

transition at about 30 kOe in the basal plane [28]. These results differ from the magnetization

measurements in CeRu2Si2 which show clearly an important metamagnetic transition occurring along the c-direction at about 80 kOe [13]. The analysis of magnetization and

inelastic neutron scattering measurements in single crystals [30], consistent with resistivity and specific heat data, has led to an evaluation of the crystal-field splitting scheme : the ground

state is the doublet ± 1/2 and the two doublets ± 3/2 and ± 5/2 lie at approximately 80 and

230 K above it.

We report here measurements of the transport properties of CePt2Si2 single crystals along

the c-direction and along one direction (110) in the basal plane of the tegragonal structure.

Resistivity measurements were already presented in a short paper [25] and we report here

new data on the thermoelectric power and the thermal conductivity.

The two samples used in the present study were cut from the same single crystal ; the effect

of heat treatment is reported elsewhere [31]. For the present measurements, the samples were equiped with an Au-Fe/Chromel thermocouple for the determination of the thermal gradients

while the thermopower was obtained by using copper wires. Experiments were carried out by

sweeping continuously the heating power through the sample in both directions following a

procedure described in reference [31].

(5)

In figure 1, we reproduce the results for the resistivity [25], which shows stronger values for the (001) direction over the whole investigated temperature range. The thermoelectric power is shown in figure 2 and the anisotropy appears to be again very strong. Along the (110) direction, the thermoelectric power is always positive ; at 300 K its value is 5.4 tV/K and it

increases logarithmically with decreasing temperature down to 50-55 K where it reaches the maximum value of 31 f.L V /K. At lower temperatures, the thermoelectric power decreases

monotonically with decreasing temperature down to zero. At 300 K, the thermoelectric power along the (001) direction is negative and its absolute value, 12.3 f.L V/K, is higher than in

the former case. Around and below the room temperature, the thermoelectric power varies in

a logarithmic manner ; it crosses zero near 180 K and it reaches a maximum of 8.8 f.LV/K

around 85 K. Below this maximum, it exhibits an oscillatory behaviour, crossing zero around

56 K and 22 K and going through a minimum of - 7 f.L V /K at 35 K. At 4 K, it joins the value

of the first orientation (110) after another maximum at 8.8 K. The thermoelectric power of

CePt2Si2 presents, therefore, a strong anisotropy with larger values for the component perpendicular to the c-axis ; the same tendency is observed in CeRu2Si2 [26] but the thermopower curves have a different temperature dependence in the two cases.

Fig.l.

Fig. 2.

Fig. 1.

-

Experimental plots of the electrical resistivity along the (001) and (110) directions for a single crystal CePt2S’2 compound-.

Fig. 2.

-

Experimental plots of the thermoelectric power along the (001) and (110) directions for a

single crystal CePt2Si2 compound.

Figure 3 gives the two curves of the thermal conductivity K versus temperature for the two directions (110) and (001) and once again we can observe a strong anisotropy which increases

steadily from 4 to 300 K. The thermal conductivity is larger along the (110) direction, as expected from the electrical resistivity data. The temperature dependence of both components of K shows a smooth oscillation separating two quasi-linear regions ; the inflexion point of the

two curves is located approximately at 50 K. Our previous investigation on a polycrystal [32]

has shown a similar behaviour which appears to be typical of the thermal conductivity for

most cerium Kondo compounds. Inset of figure 3 shows the Wiedemann-Franz ratio

L = Kp /T versus temperature. Between room temperature and roughly 150 K, the

Wiedemann-Franz ratio L is constant for both components and is a little larger than the

Sommerfeld value Lo

=

2.45 x 10- 8 WOK- 2 of the Lorenz number ; in fact, this discrepancy

(6)

Fig. 3.

-

Experimental plots of the thermal conductivity along the (001) and (110) directions for a single crystal CePt2Si2 compound. Inset gives the reduced Lorenz number along the two directions.

with Lo above 150 K is slight and not far from the uncertainty on the absolute values. But, below 150 K, the ratios defined along the two directions have opposite behaviours. That

corresponding to the (001) direction increases significantly and reaches a plateau between 4

and 50 K corresponding to a value of roughly 1.5 Lo. On the contrary, the ratio L along the (110) direction decreases steadily down to roughly 0.7 Lo at 4 K, after crossing Lo at 35 K.

Thus, experimental measurements of the electrical resistivity, thermal conductivity and

thermoelectric power show a strong anisotropy between the directions (110) and (001) [33]

and we will present a model able to explain it in the next section.

3. Theoretical model.

We shall compute here the three transport coefficients in single crystals, i.e. the electrical

resistivity p, the thermal conductivity K and the thermoelectric power S along the c-direction and along a direction in the basal plane of a tegragonal or hexagonal crystal structure, as in

CePt2Si2 or CeAl3 respectively. In the present paper, we do not consider the anisotropy within

the basal plane ; this effect will be described elsewhere [34].

There are certainly several origins for the anisotropy, but we will show here that the effective exchange Hamiltonian of reference [21], including crystal-field effects as in

reference [16], yields a large anisotropy in transport properties of cerium Kondo compounds.

A first calculation using the self-consistent ladder approximation by applying Abrikosov’s

pseudo fermion technique has been recently performed by Kashiba et al. [35] in order to

compute numerically the electrical resistivity. Here, we present the main points of the

derivation of the third-order calculation of the transport properties, as in references [16, 19, 20], but now in the case of a single crystal. Detailed calculations will be published elsewhere [34].

The effective resonant-scattering Hamiltonian can be written as :

(7)

with the usual notations and definitions of references [16, 21] : ctM represents the creation operator of a conduction electron of energy Ek in the partial-wave state M (defined by the quantum numbers Q = 3, s = 1/2, j

=

5/2), while cù represents the creation operator of a localized 4f electron in the state M of the Ce atom. M denotes, as usual, the quantum number characterizing an eigenstate of energy EM in presence of the crystalline field (CF), i.e. either any state of the doublet fy or the quartet T 8 for the cubic symmetry (as in CeAl2), or any

eigenvalue ±1/2, ± 3/2, ± 5/2 for the hexagonal or tetragonal symmetry (as in CePt2Si2). The energies Ek and EM are defined with respect to the Fermi energy. The exchange integrals are given by :

defined with a cut-off D. In equation (2), Vu is the hybridization parameter between 4f and conduction electrons. The VMM parameters, which represent pure direct scattering, are all

taken equal to a common value V.

Our previous calculations [16, 19, 20] of the transport properties have been always performed for polycrystals and we have there approximated the relaxation time of a

conduction electron by an isotropic average over the different k directions of the conduction electrons. In the present case of a single crystal, we compute the transport properties along

different crystallographic directions i (defined by the direction of the electrical field or that of the temperature gradient) and we must calculate the relaxation time Tku for a conduction electron plane wave of wavevector k and spin (T, which turns out to be highly anisotropic.

The classical formulae for the electrical conductivity a

=

1 /p , thermal conductivity

K and the thermoelectric power S can then be written, for each direction i, as :

where the integrals Kn (written in the following Kn for each direction i) are defined by :

where the conduction-electron energy Ek is defined by (1) and f k is the Fermi-Dirac distribution function. The relaxation time Tku is given here by :

In the expression (7), Cm’ (f2 k) represents the weight of the partial wave 1 kM> inside the plane wave 1 ko, > [21]. The coefficients Cm’ (f2k) are given by [21] :

where Yf({1k) are the spherical harmonics of order 3.

(8)

The partial wave relaxation time TkM is given by :

RkM and SkM are, respectively, the second- and third-order terms and they are given by the expressions (38) and (39) of reference [16] without summation on M, i.e. :

In the expression (9), m, vo and c are respectively the mass of the conduction electron, the

considered volume and the cerium concentration.

As usual in the calculation of the transport properties, we need to invert the relaxation times in order to compute the integrals Kn and we use the third-order perturbation approximation which implies that the third-order terms are much smaller than the second- order terms. Thus, we can write :

We will consider now the k-dependence of both LeM RkM and LeM SkM. Let us explain

M M

how we proceed : the average value RkM of RkM can be written as :

where Rk = E Rkm is obtained from the expression (10).

M

However, in the expression (12), if we replace RkM by its average value Rk/ (2 j + 1 ), we

obtain :

The isotropic approximation of référence [16] implies taking an average value of

CM equal to 2 j 1+ 1 ’ 2 ’+1 i.e. equal to - . ô However, the correct average given by (14) is found to

be 1, which leads to a discrepancy of factor 3 between the coefficients of the electrical 2

resistivity and thermal conductivity obtained below and those in références [16] and [20]

respectively.

(9)

Then, we write Rkm as RkM + (Rkm - RkM) and we expand the first term of (12) up to the first order in (RkM - RkM) / Rk. This corresponds to an expansion in (J MM’ /91)2 since RkM is of order 912 and (RkM - Rx,,) of order 1 jMM@ 12 this expansion is valid in the present

case where 1 ’U 1 is larger than 1 J MM’ 1.

.

This expansion was not considered in our previous

short paper [25]. Finally, we have replaced RkM by its average value Rkm in the denominator of the second term of (12) to be consistent with the third-order approximation.

Thus :

The integrals Kri given by (6) become equal to :

where mi is the effective mass along the i-direction, ki the i th component of k and

kF the Fermi wave number.

In the expression (16), the integral over f2k gives different results for the different directions. In order to perform the integration over f2k, we write ki and kx as a function of

spherical harmonics, as follows :

and, after performing the integration of the products of spherical harmonics, we get different results for the integrals Kn and Kn and finally we obtain different expressions for the transport properties along the z- and x-directions.

We present the results here only in the case of three doublets with M values equal to

± 1/2, ± 3/2, ± 5/2 ; thus, according to (10) and (11), RkM

=

Rk - M and SkM

=

Sk - M. In this

case, there is no anisotropy in the basal plane and the transport coefficients are equal to each

other along the x-and y-axes. The situation corresponds to the case of CePt2Si2 which has a

± 1/2 ground state and two excited states ± 3/2 and + 5/2 [28, 30].

The expressions (16) for Kn become, therefore :

with :

(10)

and

Using the different energy integrals given in references [16, 19, 20] for the three transport properties, we can finally write them down for the z- and x-directions. We have seen in the

previous calculations [16] that the so-called « fk

=

1/2 » approximation yields results very close to the exact ones and that it simplifies very much the calculations. Thus, we will use it

here and, when f k in the expression (9) takes the f k

=

1/2 value at the Fermi energy,

Rk and A k(2) become independent of k and will be called respectively R and A (2) in the

following. If we call also :

the resistivity p along the i-direction is given by :

The numerical coefficient of the resistivity is here three times larger than the coefficient

previously derived in reference [16], as already explained. However, apart from this numerical discrepancy, the first two terms of (25) give exactly the previous calculation of reference [16], while the last terms represent the anisotropy of the resistivity.

Using the same approximations as for the electrical resistivity, the thermoelectric power

Si along the i-direction is given by :

The first term of (26) gives the same result as previously derived in reference [19] with exactly

the same coefficient, while the second term yields the anisotropy of the thermoelectric power.

Finally, the thermal conductivity K’ along the i-direction is given by :

where we find again the thermal conductivity previously derived in reference [20], apart from the numerical factor of 3, and the two anisotropy terms in A (2) and A (3).

Thus, the formulae (25), (26), and (27) give the expressions of the electrical resistivity pi, thermoelectric power Si and thermal conductivity Ki along the i-direction and we will use

them in the next section for the specific case of CePt2Si2. Let us finally remark that the electrical resistivity behaves as mf and the thermal conductivity as llm?, while the

thermoelectric power is independent of the effective mass mi along the i-direction.

4. Application to CePt2Si2 compound.

We shall apply now the preceding theoretical results to the experimental measurements

presented in section 2. In fact we have computed only the electronic contributions to the

transport properties, via « impurity » scattering. The phonon contribution must be also

(11)

considered for heat transport, as well as the phonon scattering of electrons. In the case of the electrical resistivity, the usual procedure is to take the phonon scattering contribution equal to

the resistivity of the equivalent Lanthanum compound. But, in the present case of CePt2Si2, no resistivity is available in single crystal LaPt2Si2. Moreover, the electrical resistivities of polycrystalline LaPt2Si2 and CePt2Si2 show an increase with temperature which is much larger than that presented in figure 1. So, it is at present impossible to have a direct

estimation of the phonon contribution to the electrical resistivity. The problem of separating

the phonon and electronic contributions is even more difficult for the thermal conductivity

and thermoelectric power.

However, in spite of these difficulties, according to previous analyses of the transport

properties, we think that the Kondo scattering electronic contribution is the dominant one

and on this basis we derive here theoretical curves for the three transport coefficients with the

same set of parameters. Let us discuss now the values used here for the different parameters.

First, we take thé crystal-field scheme recently derived by magnetization and inelastic neutron

scattering irr single crystal CePt2Si2 [28, 30] : the ground state is the 1/2 doublet and the two

doublets ± 3/2 and ± 5/2 lie at respectively 80 and 230 K above. The exchange parameter

JMM for the ground state doublet ± 1/2 is chosen here equal to a typical value Jl,

= -

0.1 eV

and we take also the same parameters as previously used for cerium Kondo compounds [16] :

the hybridization parameter Vu entering (2) is equal to Vu

=

0.07 eV, the density of states of

the conduction band at the Fermi energy for one spin direction n (EF) = 2.2 statesleV.at., the

cut-off D

=

850 K, the number of conduction electrons per atom z = 3, the atomic volume vo

=

234 atomic units and kF

=

0.72 atomic unit. Finally, the direct potential parameter cU and the two effective masses mx and mz are chosen in the present work, in order to obtain

an optimal fit of the anisotropic transport properties of CePt2Si2. First, we take

CU = - 0.26 eV, in agreement with previous results, in particular for the thermal conductivity [20]. Then, the two effective masses have been chosen as follows : if we adjust the theoretical

curves to the experimental ones for the electrical resistivity at its experimental maximum value, i.e. 60 K along the z-direction and 90 K along the x-direction, we get the two values

mi = 21.3 and mx

=

8.2 in atomic units ; if, on the contrary, we make this adjustement for the

thermal conductivity at 100 K, we get the two values mi = 10.8 and mx

=

4.7. Thus, we take

here the mean values for mi i.e. mi = 16 and mx

=

6.5, (or mz

=

4 and mx

=

2.55 in atomic

units).

Figures 4, 5 and 6 give the theoretical plots versus temperature of respectively the electrical resistivity, thermoelectric power and thermal conductivity. When comparing figures 1 and 4

on the one hand and figures 3 and 6 on the other, we see a relatively good agreement between theory and experiment for both the electrical resistivity and the thermal conductivity. In the

first case (p ), the shape of the curves, the magnitude, the ratio of the values along x and z,

which is indeed fixed by the choice of mx/mz, and finally the respective positions of the

maxima are similar in theory and experiment, except at very low temperatures where coherence effects develop and single-impurity third-order perturbation calculations are no

longer valid. In the second case (K), there is also a good agreement for the ratio and the shape

of the two curves, in particular with two inflexions at approximately the right temperatures and a more pronounced inflexion along the Oz-axis ; this good agreement is clearly connected

with our choice of mx/mz and CU.

On the other hand, the agreement between 2 and 5 for the thermoelectric power appears to

be relatively good for Sx and poor for Sz, but we must note that the theoretical thermoelectric

power does not depend on the adjustable effective masses mx and mz. For Sx, the shape, the

position of the maximum at roughly 50 K and the positive values obtained by theory agree

with experiment and there is a disagreement only with respect to the magnitude of

(12)

Fig. 4.

Fig. 5.

Fig. 4.

-

Theoretical plots of the electrical resistivity along the z 01 c) direction and along the direction x or y perpendicular to it. The following parameters have been used : III

= -

0.1 eV for the ground state ;

the ground state is the ± 1/2 doublet and the two excited doublets ± 3/2 and ± 5/2 lie at respectively 80

and 230 K ; V kf

=

0.07 eV ; n(EF)

=

2.2 states/eV. at. for one spin direction ; D

=

850 K ; cU=-0.26eV; z

=

3 electron/atom ; v o

=

234 atomic units ; kF

=

0.72 atomic unit ; mz

=

4 and mx

=

2.55 atomic units.

Fig. 5.

-

Theoretical plots of the thermoelectric power along the z 01 c) direction and along the direction

x or y. The parameters are the same as those used in figure 4.

Fig. 6.

-

Theoretical plots of the thermal conductivity along the z 01 c) direction and along the direction

x or y. The parameters are the same as those used in figure 4.

Sx. On the contrary, we obtain only two extrema instead of three for Sz, but Sz is generally negative and much smaller than Sx, in agreement with experiment.

In conclusion, we have obtained a semi-quantitative agreement for the three transport properties along two directions with the same set of parameters. In particular, the positions of

the maxima for p and Sx and of the inflexions for K, as well as the relative values of

Sz and Sx, indicate clearly that the values used here for the crystal-field scheme and in

(13)

particular for the overall splitting are certainly correct for CePt2Si2. The values used here for

JMM, and ’U, which play obviously a very important role for determining the shape of the

curves, are quite reasonable and in good agreement with previous determinations [16, 19, 20].

Finally, we have taken mZ

=

4 and mx

=

2.55, which allow a good interpretation of the anisotropy of p and K. At this point, these values are rather phenomenological ; to our knowledge, no direct study of the Fermi surface has been carried out and we have not really

considered the anisotropy of the Fermi surface in the integrals.

5. Conclusion.

We have presented in the present paper a complete set of measurements for the transport properties of single crystal CePt2Si2 along the c-axis of the tetragonal structure and perpendicular to it. The observed anisotropy is quite large in CePt2Si2, as previously shown

for the resistivity of CeAl3 [24] and the thermoelectric power of CeRu2Si2 [26]. On the other

hand, we have presented an explanation of this anisotropy in terms of the anisotropic

relaxation time resulting from the Coqblin-Schrieffer Hamiltonian (1) appropriate for cerium impurities. The most striking point concerns the theoretical calculation of the thermoelectric power which shows a strongly anisotropic behaviour, with even values of different signs according to the considered directions, and this anisotropy is due only to the anisotropic

character of the Hamiltonian (1) since the thermoelectric power does not depend on the

effective masses. We must add that this intrinsic effect of the anisotropic Hamiltonian (1)

exists also obviously for the electrical and thermal conduction, although there is also an

anisotropy effect phenomenologically described by a ratio mz/mx taken here different from 1.

Further studies of transport properties in single crystals would be necessary to better understand the anisotropy in cerium Kondo compounds.

References

[1] ANDRES K., GRAEBNER J. E. and OTT H. R., Phys. Rev. Lett. 35 (1975) 1779.

[2] OTT H. R., RUDIGIER H., FISK Z., WILLIS J. O. and STEWART G. R., Solid State Commun. 53

(1985) 235 ;

FISHER R. A., LACY S. E., MARCENAT C., OLSEN J. A., PHILLIPPS N. E., FLOUQUET J., AMATO A. and JACCARD D., Jpn. J. Appl. Phys. Suppl. 26-3 26 (1987) 1257.

[3] STEGLICH F., AARTS J., BREDL C. D., LIEKE W., MESCHEDE D., FRANZ W. and SHÄFER H., Phys. Rev. Lett. 43 (1979) 1892.

[4] LEE W. H. and SHELTON R. N., Phys. Rev. B 35 (1987) 5369 ;

SERA M., SATOH T. and KASUYA T., Jpn. J. Appl. Phys. Suppl. 26-3 26 (1987) 551.

[5] FUJITA T., SATOH K., MAENO Y., UWATOKO Y. and FUJII H., J. Magn. Magn. Mater. 76-77 (1988)

133.

[6] LAHIOUEL R., PIERRE J., SIAUD E., GALERA R. M., BESNUS M. J., KAPPLER J. P. and MURANI

A. P., Z. Phys. B 67 (1987) 185.

[7] SERENI J., NIEVA G., KAPPLER J. P., BESNUS M. J. and MEYER A., J. Phys. F 16 (1986) 435.

[8] BAUER E., PILLMAYR N., GRATZ E., GIGNOUX D., SCHMITT D., WINZER K. and KOHLMANN J., J. Magn. Magn. Mater. 71 (1988) 311.

[9] BAUER E., GRATZ E. and PILLMAYR N., Solid State Commun. 62 (1987) 271.

[10] BARBARA B., BOUCHERLE J. X., BUEVOZ J. L., ROSSIGNOL M. F. and SCHWEIZER J., J. Phys.

Colloq. France 40 (1979) C5-321.

[11] VETTIER C., BURLET P. and ROSSAT-MIGNOT J., J. Magn. Magn. Mater. 63-64 (1987) 18.

[12] GALERA R. M., PIERRE J., SIAUD E. and MURANI A. P., J. Less Common Met. 97 (1984) 151.

(14)

[13] HAEN P., FLOUQUET J., LAPIERRE F., LEJAY P. and REMENYI G., J. Low Temp. Phys. 67 (1987)

391.

[14] GIGNOUX D., SCHMITT D., ZERGUINE M., AYACHE C. and BONJOUR E., Phys. Lett. A 117 (1986)

145.

[15] BARTH S., OTT H. R., GYGAX F. N., HITTI B., LIPPELT E., SCHENCK A., BAINES C. , VAN DEN

BRANDT B., KONTER T. and MANGO S., Phys. Rev. Lett. 59 (1987) 2991 ;

NAKAMURA H., KITAOKA Y., ASAYAMA K. and FLOUQUET J., J. Phys. Soc. Jpn 57 (1988) 2644.

[16] CORNUT B. and COQBLIN B., Phys. Rev. B 5 (1972) 4541.

[17] LASSAILLY Y., BHATTACHARJEE A. K. and COQBLIN B., Phys. Rev. B 31 (1985) 7424.

[18] FERT A., J. Phys. F. 3 (1973) F2126.

[19] BHATTACHARJEE A. K. and COQBLIN B., Phys. Rev. B 13 (1976) 3441.

[20] BHATTACHARJEE A. K. and COQBLIN B., Phys. Rev. B 38 (1988) 338.

[21] COQBLIN B. and SCHRIEFFER J. R., Phys. Rev. 185 (1969) 847.

[22] FRANZ W., GRIESSEL A., STEGLICH F. and WOHLLEBEN D., Z Phys. B 31 (1978) 7.

[23] ASSMUS W., HERRMANN M., RAUCHSCHWALBE U., RIEGEL S., LIEKE W., SPILLE H., HORN S., WEBER G., STEGLICH F. and CORDIER G., Phys. Rev. Lett. 52 (1984) 469.

[24] JACCARD D. , CIBIN R. , JORDA J. L. and FLOUQUET J. , Jpn. J. Appl. Phys. Suppl. 26-3 26 (1987)

517.

[25] MARSOLAIS R. M., AYACHE C. , SCHMITT D., BHATTACHARJEE A. K. and COQBLIN B., J. Magn.

Magn. Mater. 76-77 (1988) 269.

[26] AMATO A. , JACCARD D., SIERRO J., LAPIERRE F. , HAEN P., LEJAY P. and FLOUQUET J., J.

Magn. Magn. Mater. 76-77 (1988) 263.

[27] AYACHE C. , BEILLE J. , BONJOUR E., CALEMCZUK R. , CREUZET G., GIGNOUX D., NAJIB A., SCHMITT D., VOIRON J. and ZERGUINE M., J. Magn. Magn. Mater. 63-64 (1987) 329.

[28] GIGNOUX D., SCHMITT D. and ZERGUINE M., Phys. Rev. B 37 (1988) 9882.

[29] MARSOLAIS R. , AYACHE C. , BONJOUR E., CALEMCZUK R. , COUACH M. , LOCATELLI M. , SALCE B., GIGNOUX D., SCHMITT D. and ZERGUINE M. , Jpn. J. Appl. Phys. Suppl. 26-3 26 (1987)

2101.

[30] ZERGUINE M., Thesis, Grenoble (1988).

[31] RAKI M., Thesis, Orsay (1989).

[32] AYACHE C. , RAKI M. , SALCE B., SCHMITT D. , BHATTACHARJEE A. K. and COQBLIN B., J. Phys.

Colloq. France 49 (1989) C8-791.

[33] MARSOLAIS R. M. , AYACHE C. and SCHMITT D., to be published.

[34] BHATTACHARJEE A. K. and COQBLIN B., to be published.

[35] KASHIBA S., MAEKAWA S., TAKAHASHI S. and TACHIKI M., J. Phys. Soc. Jpn. 55 (1986) 1341.

Références

Documents relatifs

This line, interpreted in terms of a crossover between two different Kondo regimes, crosses the critical line where magnetic order takes place.. On the basis of our

2014 Using the theory of Lacroix-Lyon-Caen et al., who have calculated the magnetic susceptibility of a cerium Kondo system in the Anderson model, including the

rature paramagnetic properties of many intermediate valence compounds appear to exhibit strong para- magnons feature of nearly magnetic Fermi liquids. We have

- We have calculated, in the second Born approximation, the conduction electron scattering rate due to impurity pairs with long range magnetic coupling in a non magnetic

Finally, one has to mention, that very recently Meaden [5] suggested that the occurrence of the break point in the log-log plot of the da/dT curve and the anomaly in

An important pararheter in the subsequent inter- pretation ofp data (Sec. V) is the c/a ratio of the pseudo- binary compounds obtained by partial replacement of Ce in

- Temperature dependence of the magnetization and reci- procal magnetic susceptibility for CeMnCuSi,(a) and field depen- dence of magnetization for CeMnCuSi, 1 and

- Relativc reversible permeabilities pk at different magnetic polarizations versus temperature of Ni-Mn-Co ferrites.. ( x = 0.027) determined from