• Aucun résultat trouvé

Negative temperature coefficient of resistivity in samples of Nb 3Ge irradiated by electrons or heavy ions

N/A
N/A
Protected

Academic year: 2021

Partager "Negative temperature coefficient of resistivity in samples of Nb 3Ge irradiated by electrons or heavy ions"

Copied!
11
0
0

Texte intégral

(1)

HAL Id: jpa-00209910

https://hal.archives-ouvertes.fr/jpa-00209910

Submitted on 1 Jan 1984

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.

Negative temperature coefficient of resistivity in samples of Nb 3Ge irradiated by electrons or heavy ions

F. Rullier-Albenque, L. Zuppiroli, F. Weiss

To cite this version:

F. Rullier-Albenque, L. Zuppiroli, F. Weiss. Negative temperature coefficient of resistivity in samples of Nb 3Ge irradiated by electrons or heavy ions. Journal de Physique, 1984, 45 (10), pp.1689-1698.

�10.1051/jphys:0198400450100168900�. �jpa-00209910�

(2)

1689

Negative temperature coefficient of resistivity in samples

of Nb3Ge irradiated by electrons or heavy ions

F. Rullier-Albenque, L. Zuppiroli

Section d’Etude des Solides Irradiés, Centre d’Etudes Nucléaires, B.P. 6, 92260 Fontenay-aux-Roses, France and F. Weiss

Laboratoire des Matériaux, ENSIEG, ER 155, B.P. 46, 38402 St Martin d’Hères, France (Reçu le 2 décembre 1983, révisé le 4 avril 1984, accepté le 21 juin 1984)

Résumé.

2014

Des échantillons de Nb3Ge ont été irradiés à basse température (T~ 20 K) par deux projectiles

aussi violemment différents que des électrons de 2,5 MeV et des ions lourds d’environ 100 MeV (fragments de

fission de l’uranium). La résistivité

a

été mesurée

en

fonction de la température avant et après irradiation. Pour les deux types de particules des valeurs négatives du coefficient de résistivité

avec

la température d03C1/dT sont mesurées

et corrélées à la résistivité résiduelle. Ces propriétés de transport anormales sont discutées et comparées

aux

résultats des récentes théories consacrées à

ce

sujet. Qualitativement, il semble clair que, s’agissant d’un système

où le libre parcours des électrons est de l’ordre de la distance interatomique, le changement de signe du coefficient

d03C1/dT est relié à un processus de localisation des électrons

avec

la participation d’interactions électron-phonon et

électron-électron. En d’autres termes, l’ecart entre les valeurs expérimentales et la théorie de Boltzmann du trans-

port dans un gaz d’électrons presque libres est dû à l’interférence entre les collisions sur les impuretés, les collisions

électron-phonon et (ou) les collisions électron-électron. Alors que la théorie de Belitz et Schirmacher

sur

l’effet tunnel induit par les défauts dans les métaux très désordonnés semble la plus appropriée pour expliquer les résultats d’irradiation

aux

ions lourds, bien qu’elle n’inclue que les interactions électron-phonon, les résultats des irra- diations

aux

électrons révèlent la nécessité d’inclure les interactions coulombiennes

comme

dans le modèle de Alts’huler et Aronov.

Abstract.

2014

Nb3Ge samples have been irradiated at low temperature (T ~ 20 K) by two drastically different

irradiation projectiles : 2.5 MeV electrons and ~ 100 MeV heavy ions (uranium fission fragments). The resistivity

has been measured

versus

temperature before and after these irradiations. For both projectiles negative temperature coefficients of resistivity d03C1/dT

were

measured and correlated with the residual resistivity p. These anomalous trans-

port properties

are

discussed and compared to the results of the recent theories

on

the subject. It is qualitatively

clear that the change of sign of the temperature coefficient of resistivity, occurring in

a

system where the electronic

mean

free path is of the order of the atomic distance, is indeed related to some localization process, helped by electron-phonon and electron-electron interactions. In other words, the deviation of the resistivity from the Boltz-

mann

nearly free electron behaviour is due to some interference effect between impurity scattering and electron-

phonon and/or electron-electron scattering. While the theory of defect induced tunnelling in strongly disordered

metals developed by Belitz and Schirmacher

seems

to be the most appropriate for the explanation of the heavy ion

irradiation results, although it includes electron-phonon interactions only, electron irradiation results reveal the

necessity of including the interplay with Coulomb interactions

as

in the model of Alt’shuler and Aronov.

J. Physique 45 (1984) 1689-1698 OCTOBRE 1984,

Classification

Physics Abstracts

61.80F201361.80J201372.15E

1. Introduction.

There has been a great deal of interest in the A-15

compounds such as V3Si, Nb3Sn or Nb3Ge because

of their various peculiar properties. First of all, they are among the compounds with the highest

superconducting transition temperatures. Secondly, they exhibit anomalies in various normal state pro-

perties. In particular, the fact that the A-15 resistivities do not increase linearly with temperature and struc- tural damage but rather tend to saturate at about 120 to 160 u.cm, corresponding roughly to a mean

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:0198400450100168900

(3)

free path I of the order of the interatomic distance a,

has attracted considerable attention [1, 2]. The

disordered character of these alloys is also reflected in the breakdown of Matthiessen’s rule, the tempe-

rature coefficient of resistivity (TCR) dp becoming

dT

even negative at low temperatures for sufficiently

disordered samples [3, 4]. This last point appears to be quite universal as emphasized by Mooij who

found such correlations between dlnp and the resisti- dT

vity p for a large class of materials [5]. The critical value of resistivity beyond which the TCR becomes

negative at room temperature was found by Mooij

to be approximatively 150 uQ . cm. It is striking to

notice that this value is precisely the one we have

mentioned for the A-15 saturation resistivity.

Presently the A-15 superconductor with the highest

transition temperature, Nb3Ge, is prepared by sputter- ing, coevaporation or chemical vapour deposition.

These elaboration methods allow the A-15 Nb3Ge

metastable phase to be stabilized and provides films

with transition temperatures which can reach 23 K.

The transition temperatures, Tc, residual resistivities and temperature coefficients have been shown to be

strongly correlated to each other and to all depend

on the global degree of disorder of the material [3, 4, 6].

But the precise nature of this disorder is not at present clear. One may wonder if the disorder is concentrated

mainly at the atomic scale such as disorder in amor-

phous alloys (metallic glasses) or if it is related to an

inhomogeneity of any kind which could depend on

the « metallurgical )) qualities of the films. In the present paper we try to discuss this problem in

connection with a set of new radiation damage experiments in Nb3Ge films. We have tried to change

the level and the spatial distribution of damage by using two kinds of very different irradiation projec-

tiles : 2.5 MeV fast electrons which produce mainly

isolated point defects and ~ 100 MeV heavy ions (uranium fission fragments) which deposit high energy densities in the form of electronic excitation and atomic collisions. Studies of the effects of electron, neutron, a-particle and heavy-ion irradiation on the electrical resistivity have been reported by other investigators [3, 4, 6, 7]. One of the originalities of

the present work lies in irradiating the same material by projectiles which are expected to produce very different damage distributions.

The outline of this paper is as follows. In section 2

we describe the experimental techniques and present

our results. In section 3 the different models which have been proposed to explain the anomalous beha- viour of resistivity are reviewed and experimental

results are discussed in connection with the theoretical

predictions and the defect structure in irradiated

Nb3Ge samples is analysed. Section 4 contains our

conclusion.

2. Experiments.

2.1 SAMPLES. - All the Nb3Ge samples used in this work have been prepared by chemical vapour depo-

sition. The elaboration techniques are described in detail elsewhere [8, 9]. Except for one (Nb3Ge-9)

the Nb3Ge films (of thickness from 1.5 to 4.5 J.1m)

are deposited on 300 J.1m thick sapphire substrate.

The resistivities of these samples were measured by

a standard four points method. The estimated error

in the absolute values of resistivity is about 20 % primarily due to an imprecise knowledge of the sample thickness. For the last sample (Nb3Ge-9) depositions are made on both sides of a steel foil [9].

In this case, the sample is about 20 tim thick. This

relatively small thickness is very helpful for electron irradiation because high electron beam current can

then be used without producing important heating

of the sample. On the other hand the resistivity

determination is more difficult. At each measurement

point, we have to measure both the resistance of the

Nb3Ge-9 sample (Nb3Ge on steel) and the resis- tance of a steel sample of the same nature. In this way

we were able to obtain the absolute Nb3Ge resistivity

with an estimated error of 40 %.

The initial sample specifications are given in table I.

The different critical temperature Tc and resistivity p

(25 K) values are the results of different growth condi-

tions.

2. 2 IRRADIATIONS.

2. 2 .1 Electrons.

-

The electron irradiation has been carried out in the Vinkac low temperature facility.

This device, described elsewhere [10], consists of a liquid hydrogen cryogenerator coupled to a Van der Graaff accelerator, allowing irradiation between 18 and 25 K.

In addition, it is possible to isolate the irradiation cryostat from the cryogenerator and to warm up the

sample by a linear increase of the temperature between 20 to 300 K with heating rates from 0.1 to 10 Kelvin

per minute. The sample Nb3Ge-9 and a steel sam- ple were irradiated together with 2.5 MeV electrons up to a dose of 4 x 102° e/CM2 corresponding to an

electron flux of about 5 x 1014 e/cm2/s. The irradia-

tion temperature never increased above 22 K. The

sample resistivity was measured continuously during

irradiation. After given irradiation times, the irradia- tions were interrupted and the resistivity versus tem-

perature curves p(T) were determined between 18 and 60 K after annealings at 300 or 60 K.

Electron irradiations are known to produce isolat-

ed point defects. In a diatomic compound such as Nb3Ge it is possible to determine their number by using a calculation due to Lesueur [12]. Knowing

the two respective displacement threshold energies [14] this calculation gives the fraction of atoms displac-

ed (displacements per atom (dpa)) for each atomic

species. The outline of this calculation and its appli-

cation to Nb3Ge are indicated in the appendix. It is

(4)

1691

Table I.

-

Sample specifications.

shown that an electron flux of 5 x 1014 e/cm2 . s

corresponds to a damage production rate of about

5 x 10- 8 dpa/s.

2. 2. 2 Ions.

-

In order to irradiate the films with

heavy ions we have used sources of fission fragments containing 90 % of 2 3 5 uJ. The samples together with

uranium or uranium dioxyde sources were irradiated with thermal neutrons producing the fission of 2 3 5 U nuclei. The neutron irradiations of these « sandwiches))

were performed in the Vinka low temperature facility

of the Triton reactor of Fontenay-aux-Roses. This

device allows neutron irradiation at liquid hydrogen temperature [11]. The thermal neutron flux was

about 7 x 1011 n/cm2/s corresponding to a fission fragment flux of order of 4 x 1010 FF/CM2/S for

a bulk uranium source. During irradiation the tem-

perature always remained below 35 K.

Uranium fission fragments are expected to deposit high densities of energy both in electronic excitation and atomic collisions. One can calculate the latter contribution and evaluate the number of atomic dis-

placements (see appendix). For the thermal neutron flux used in our experiment (7 x 1011 th.n./cm2 . s)

the damage production rate is found to be of the

order of 7 x 10-6 dpa/s.

2. 3 EXPERIMENTAL RESULTS.

2. 3.1 Electrons.

-

Figure 1 shows the evolution of

Fig. 1.

-

Evolution of resistivity

versus

temperature

curves

of the Nb3Ge-9 sample under electron irradiation. See table I for the sample description. The arrows indicate the

resistivity minima.

(5)

the resistivity versus temperature curves p(T) for

different electron doses. Although it is not very visible in this figure, a resistivity minimum is observed after

a dose of 2.3 x 102° e/cm2 (corresponding to

~

0.02 dpa). It appears at about 40 K and is weakly

shifted towards higher temperature with irradiation.

On figure 2 the negative TCR is shown with a more

appropriate scale; this particular curve obtained

at the end of irradiation after annealing at 300 K is reproduced together with the curve before irradiation.

Fig. 2.

-

Comparison between the shapes of p(T)

curves

before and after irradiation for the Nb3Ge sample irradiated

with 2.5 MeV electrons.

During this electron irradiation the critical tem-

perature Tc

c

has decreased at a rate Ap (Ap Ap AP is the corresponding resistivity increase) equal to

-

0.1 K/pf2. cm, which is the value that we have

already reported for previous electron and neutron irradiation [13].

2.3.2 Fission fragments.

-

Eight different Nb3Ge samples ( 1 to 8) were irradiated with fission fragments.

Figure 3 shows some typical p(T) curves obtained

before and after irradiation. These curves are simi-

Fig. 3.

-

Some of typical p(T)

curves

for as-grown and fission fragment irradiated Nb3Ge samples. See table I

for the sample descriptions.

lar to those reported after a-particle [3] or heavy-ion [4]

irradiations. The most damaged samples exhibit a negative TCR in the whole temperature range investi-

gated (20-300 K). In these cases the p(T) curves are

very similar to that observed for amorphous Nb3Ge [14].

2. 3 . 3 Comparison between electron and fission frag-

ment irradiations.

-

The most striking resdlt of the

present study is the appearance of a negative TCR

after both electron and fission fragment irradiations.

But whereas the negative TCR is extending up to room

temperature after fission fragment irradiation, it

appears only at low temperatures (T 50 K) after

electron irradiation.

Figure 4, where the temperature coefficient of

resistivity values are plotted versus the resistivity

values at room temperature for all the samples, shows clearly that there is a strong correlation between these two quantities : the larger the resistivity p, the lower the temperature coefficient dp which

dT becomes negative at a resistivity of the order of

150 uQ2. cm.

In order to compare irradiation effects at low temperature we have represented the same plot at

25 K on figure 5. It can been seen on this figure that

Fig. 4.

-

Correlation between temperature coefficient of resistivity (TCR) and resistivity values at 280 K for all the samples (as-grown, electron

or

fission fragment irra- diated).

*

and o symbols refer respectively to electron-

and fission fragment-irradiated samples.

(6)

1693

Fig. 5.

-

Correlation between temperature coefficient of

resistivity (TCR) and resistivity values at 25 K for all the

samples (as-grown, electron

or

fission fragment irradiated).

*

and . symbols refer respectively to electron- and fission

fragment-irradiated samples.

a correlation between dp (25 K) and p (25 K) is

again visible but the values for ion and electron irradiations are not situated on the same correlation

curves. At 25 K the change from a positive to a nega- tive TCR occurs for 60 p 80 uQ. cm for elec-

tron irradiation and for 120 p 160 uQ. cm for

ion irradiation.

3. Discussion.

3.1 THE DIFFERENT MODELS EXPLAINING NEGATIVE TEMPERATURE COEFFICIENTS OF RESISTIVITY.

-

The observation of a negative coefficient of resistivity

and its correlation with the resistivity has.stimulated

a lot of theoretical work. The numerous theories can

be classified in the five following categories.

1) The extended Ziman theories [15-18] have been remarkably successful to explain the resistivity versus temperature curve of metal glasses [19-21]. Like in the

theory of the resistivity of liquid metals [15], the con- ducting electrons are considered to be quasi-free

and weakly scattered due to the random position of

atoms (pseudopotentials). The resistivity can be

related to the pair distribution function (or, in the k-space, to the X-ray interference function). The posi-

tion of the Fermi wave vector kF with respect to the

peak position of the X-ray interference function deter- mines the sign of the temperature coefficient. The main

success of this model was to explain the change in

the temperature coefficient of resistivity when the

Fermi wave vector is shifted by changing the compo-

sition of glassy metallic alloys [19, 20].

This diffraction model was used by Bieger et al. [22]

to explain the shape of the resistivity curves of some

A-15 compounds. In our case, it is worth mentioning

that the p(T) curves for the most heavily irradiated Nb3Ge samples (Fig. 3) are very similar to those of many other metallic glasses [14, 19-21].

There are two serious criticisms, in principle, concerning the application of the extended Ziman

theories to transition metal glassy alloys and to A-15 compounds. First these models are all based on a

nearly free electron approach. One may wonder if

they can properly describe electronic systems for which a tight binding approach should be much more

appropriate. Secondly, most of the Ziman extended theories assume weak coherent scattering of electrons

despite the fact that the mean free paths are of the

order of one atomic distance and that the resistivities

are approaching the « maximum metallic resistivity »,

which indicates that these alloys are not far from the

regime of localization. These remarks have inspired

some of the following alternative models.

2) Imry evoked the possible role of incipient

Anderson localization in describing the resistivities of highly damaged A-15 alloys and more generally

in explaining the universality of the Mooij correlation in highly disordered metals [24]. His interesting specu- lations are based on the recent ideas of Thouless [25]

and the scaling theory of localization of Abrahams

et al. [26] which predict that the zero temperature resistance of a disordered electronic system depends

on its length scale in a universal manner. Extensions

of these theories to higher temperatures are somewhat controversial [27, 50]. They are based on the compari-

son of the different lengths characteristic of the elec- tron propagation (critical localization length ç, elastic

mean free path le, size of the sample L, inelastic mean

free path li the temperature dependence of which plays

the main role in the resistivity variation). Moreover,

localization theories are, basically, independent elec-

tron theories and they cannot be applied crudely

to a system in which a particularly strong electron-

phonon interaction and a substantial electron-elec- tron interaction are expected to occur.

It is worth mentioning here that localization toge- ther with Coulomb interactions have been envisaged by Anderson et al. [31] to be responsible for the « uni-

versal » degradation of the transition temperature

T, in high-T, superconductors like A-15 compounds.

They argue that the anomalous behaviour of the electron diffusion constant associated with localiza- tion leads to an increase of the repulsive interaction between Cooper pair electrons, which results in a

decrease of Tc. Once the resistivity exceeds some

(7)

critical value p,, Tc is found to decrease as a universal

function of p/pc.

3) The electron-phonon interaction is the main

concern of Girvin and Jonson [23] and Belitz and Schirmacher [49] in their study of the resistivity of

disordered metals. In normal metals the additional disorder associated with phonons causes the mobility

to decrease with temperature. The new feature which appears in the high resistivity regime is the fact that the phonons can increase rather than decrease the electron mobility. A new contribution called defect induced tunnelling appears in the conductivity when

the Boltzmann hypothesis of independent scattering

events fails. Jonson and Girvin developed a theory

in the case where the usual adiabatic approximation

breaks down. Although their numerical calculations

give a qualitative agreement with the Mooij correla- tion, they do not give the correct numbers : this theory

would be more appropriate for the description of

metals with resistivities of the order of 1 000 pQ . cm

than for metals with resistivities of the order of 100 gfl . cm. Belitz and Schirmacher have developed

very similar ideas with a different formalism. We will show that their results can account rather accu-

rately for the heavy ion irradiation results shown in the present paper.

4) Whereas the theories of the previous category

concentrate on the electron-phonon interaction, Alt’-

shuler and Aronov [32, 33] proposed a description of

disordered materials which deals with interacting

electrons in the presence of weak impurity scattering.

In three dimensional systems, the interaction model

predicts that 1 /p varies as 1/ T at low temperatures [32].

Such a dependence has been observed, for the low temperature electrical resistivity of semi-metallic bis- muth in form of evaporated films and even of polyme-

ric sulfur nitride (SN)x [34].

5) Several other theories have been proposed to explain the anomalous transport properties of disorder- ed metals. In the two level tunnelling model [28, 29, 51], the negative TCR arises from a Kondo type exchange between the conduction electrons and two

equivalent atomic positions separated by a low energy barrier. This model has been applied to amorphous Nb3Ge by Tsuei [14].

In a general way, a negative TCR can be observed each time the electrons have to cross a distribution of barriers even if the distances between these barriers

are larger than atomic distances and are more related to the macroscopic granularity of the samples. This

introduces a new class of models which, to our know- ledge, have not yet been applied to Nb3Ge. Let us point out that a theory of disordered materials gene-

rally characterized by large conducting regions sepa- rated by small insulating barriers has been proposed recently by Sheng [30].

3.2 CORRELATIONS BETWEEN THE TEMPERATURE COEF- FICIENT OF RESISTIVITY dpld T AND THE RESISTIVITY.

-

Figures 4 and 5 clearly show that there is a strong corre- lation between the temperature coefficient of resis-

tivity -72013 and the resistivity p, not only at room tem-

perature (Mooij plot) but also at low temperatures.

This correlation is better in the former case than in the latter where two different curves can be distinguished depending on the irradiation type. Thus the kind of defects created does influence the behaviour of the

resistivity. It is remarkable anyway that the change

of sign of the TCR occurs around 100 u. cm for both

cases, that this value is close to the saturation value of the resistivity at high temperature and is not so far from values obtained at the same temperature in other systems with quite different electronic proper- ties [48].

The systems of interest are dirty metals with mean

free paths of the order of one atomic distance. This indicates that they are not far from the regime of loca-

lization. A priori, the models of category 2 in sec- tion 3.1 are the most attractive for explaining the properties of interest but we have already mentioned

that they are not able to account for the shapes of

the p(T) curves that have been observed in heavily

irradiated samples (Fig. 3). The theories able to

explain the strong deviation from the Boltzmann behaviour leading to a negative TCR up to room

temperature are the models of categories 1, 3 and 4 in section 3.1. All of them explain the negative TCR by some interference effect between two scattering

events.

In the Ziman extended model, the origin of the negative TCR has to be found in the interference bet-

ween waves diffracted by two adjacent centres of the amorphous material. But because A-15 alloys are

narrow d-band metals with bandwidths of approxi- matively 0.3 eV it is reasonable to exclude the extended Ziman theories to explain the temperature variations of the resistivity.

In the Alt’shuler and Aronov interaction model the interference between impurity scattering and

Coulomb scattering is responsible for the anomalous behaviour of the resistivity : the electron impurity scattering process is influenced by the presence of the

surrounding interacting electrons and in turn Cou- lomb interactions between electrons are enhanced when these particules are substantially slowed down

by frequent collisions with impurities. We have tried

to fit the low temperature parts of the 1 curves to the

p

T1/2 law of the interaction model. In the case of the electron irradiation (curves similar to this of Fig. 2)

the signal to noise ratio is approximately 5. Thus

numerous temperature increasing functions can be

fitted to 1/p including a T 1/2 variation. In the case of ion irradiations the parts of the curves with negative

TCR are far more extended but the fit is less good

(8)

1695

below 40 K. It is better above 60 K in a region where

it has less sense.

Finally, the interference between impurity and phonon scattering is also very important for under- standing the deviations from the Boltzmann theory

as mentioned by Girvin and Jonson who have em-

phasized that in the dirty limit the electrons are diffusing so slowly that they are sensitive to the time dependence of the phonon amplitudes. Unfortuna- tely, their numerical calculations imply that when

the TCR changes sign, the resistivity is by one order

of magnitude larger than the experimental value.

A theory of phonon controlled conductivity in the high resistivity metals has been developed more recently by Belitz and Schirmacher [49] on the basis

of similar concepts. They obtained for the temperature dependent conductivity

-

Mo and Lo are static contributions due to disorder.

--

MT is the generalized phonon scattering rate

which reduces to the nearly free electron expression

when the mean free path is large compared to the

Fermi wavelength ÀF.

-

LT is the new term due to the interference pro-

cess that they call phonon controlled tunnelling rate.

The random tunnelling process increases with temperature as does the scattering rate but the former is proportional to the conductivity thus allowing for negative TCR. When the time scale for the electronic motion is much smaller than that of the phonons (adiabatic approximation) the previous formula is

a generalisation of the Ziman theory for amorphous

metals. Thus it is not surprising that such a formalism

can account for the conductivity of the heavy-ion

irradiated samples of figure 3 (negative coefficients from 20 to 300 K). Figure 6 shows that this model

can indeed account for the resistivity curves obtained

after heavy ion irradiation with Mott’s maximum metallic resistivity value of the order of 700 gf2.cm.

Even if the initial concept of maximum metallic

resistivity is in contradiction with the scaling theory [26] in which the conductivity vanishes continuously,

it has been shown [49], and this experimental study

also confirms, that pm is the relevant resistivity scale

near the Anderson transition. It is worth mentioning

that this theory is also able to explain the resistivity

saturation at high temperature at a value of about 150 gfl. cm.

However the small negative dp parts of the curves

dT

for the electron irradiated samples can probably not

be explained in the same way and some Coulomb interactions halve probably to be included as men-

tioned in the semi-quantitative analysis of Kaveh

and Mott [50].

Fig. 6.

-

Comparison of two p(T ) curves measured after

heavy ion irradiations with the model of Belitz and Schir- macher [49]. Dots are experimental data presented on figure 3 (see table I for sample description). Symbols (*)

are

the fits using curves P PM = f T 0 of the figure I (a) in reference [49]. Here, pm is the value of Mott’s

«

maximal

resistivity

»

pm

=

3 n’hle’kf and 0 is the Debye tempe-

rature equal to 302 K for N b3Ge [52]. Best fits for p(T)

curves

6 and 9 are obtained using respectively the fourth

and fifth (from bottom) 1!.... curves of Belitz et al. with

a

PM

value of pM of the order of 700 gf2.

cm.

The following section is an attempt to determine in a more microscopic way which are the defects

responsible of these macroscopic effects on the resis- tivity and to compare electron and fission fragment

irradiation.

3.3 DEFECT STRUCTURE IN IRRADIATED Nb3Ge SAM-

PLES.

-

In an ordered diatomic compound AB,

irradiation produces point defects either isolated

(electrons) or in cascades (neutron or heavy ions)

and antisite defects, i.e. A atoms occupying B sites

and vice versa. This atomic interchange results in a

decrease of the long range order parameter S. The existence of antisite defects has been demonstrated in neutron irradiated A-15 compounds using neutron

and X-ray diffraction techniques [35, 36]. Due to the

structural instability of high-T, A-15 alloys, the

presence of defects can result in local atomic rearran-

gements. Testardi et al. [37, 38] have suggested that

the major effect of point or antisite defects is to

generate microscopic strains occurring over dimen-

sions of the order of some unit cells.

In the case of fission fragment irradiation, the large

energy deposited by electronic excitation can be

thought to strongly influence the damage production.

Moreover, the observation that p(T) curves for sam- ples irradiated at the highest doses are very similar to that for amorphous Nb3Ge could be interpreted

in terms of local amorphization. Let us point out that

a mechanism by thermal spikes has been proposed by Lesueur [39] to explain the Pd80S’20 amorphiza-

tion under fission fragment irradiation. In this assump-

(9)

tion, fission fragment irradiated Nb3Ge samples

are considered as binary mixtures composed of small amorphous regions in an undamaged matrix, and electrical resistivity can be expressed within the framework of the effective medium theory [40].

Attempts to fit experimental resistivity production

curves with this interpretation have been unsuccess-

ful showing that a’ more homogeneous model is required to explain these resistivity results.

It is possible to relate the long range order para- meter S to the resistivity by using the relationship

of Muto [47]

where PD and po are the electrical resistivities for the disordered and the ordered states, respectively. More-

over, a functional dependence of S with irradiation based on probability arguments has been derived by

Aronin as

where A is the replacement effectiveness per atomic

displacement c (dpa).

In the case of fission fragment irradiated Nb3Ge samples, a linear dependence between S (deduced from resistivity measurements) and the number of atomic displacements, is observed once the dose exceeds

about 0.04 dpa. Experimental value of A is found to

be N 6. The same value of A has been reported by

Brown et al. [42] upon measuring the r sistivity of

a Nb3Ge sample irradiated by fission

f gment.

Then we reach the conclusion th t amorphous

behaviour observed in the most radiation-damaged samples is not due to a local amorphization induced by irradiation but results from a very large accumula-

tion of antisite defects. This interpretation is consis- tent with the experimental X-ray results of Cox

et al. [36] in neutron irradiated Nb3Ge samples;

they found a relation between the transition tem-

perature and the long-range order parameter up to

a dose of - 0.1 dpa and observed a loss of crystalli- nity after a dose of N 0.2 dpa.

In the case of electron irradiation, antisite defects

probably play also a role in the damage mechanisms,

but it is not possible to determine their contribution from resistivity measurements.

Antisite defects can be created during electron

irradiation by replacement collision sequences. In A-15 compounds production of large amounts of

antisite defects in such sequences is anyhow unlikely

because the 102 > atomic rows along which repla-

cements produce disorder, are « broken » between the ABA sequences (see Fig. 7), whereas the very

probable collision sequences along 100 ) do not produce antisites. A detailed study of electron irra- diation effects in V3Si samples [41] suggested that

antisite defects are produced at about the same con-

centration levels as are point defects in electron irradiated A-15 compounds.

Fig. 7.

-

Schematic sketch illustrating in the A-15 struc- ture the 102 > directions along which replacement col-

lisions

can

produce disorder.

Thus the main difference between fission fragment-

and electron-irradiated Nb3Ge samples at a given

number of dpa appears to be a much larger concen-

tration of antisite defects in the former (1 ) than in the latter. One can see here the indication that the effect of antisite defects is mainly to affect the electron-

phonon interaction while point defects tend to inter- fere with the electron-electron interactions.

4. Conclusion.

We have studied the influence of two very different types of irradiation (electrons and fission fragments)

on the transport properties of Nb3Ge samples. In

both cases a change of sign of TCR dP was observed

dT

and a correlation between dp and the residual resis- dT

tivity was established either at high or at low tempera-

tures. The problem is to decide if these correlations

are universal in the sense of Mooij. In the detail of the

values, they are not universal and the type of disorder has some influence ; but in all cases the sign of the

TCR changes at values of the order of 100 uQ. cm

which is a fraction of about 0.15 of the maximum metallic resistivity pm initially suggested by Mott (in three dimensions Pm ~ hale 2 = 600 u cm when

a

=

3 A). The scaling theories [26] have shown that there is no maximum metallic resistivity in 3 D.

It seems clear anyway that pm is the relevant resis-

tivity scale near the Anderson transition [49] though

the conductivity vanishes continuously. In the same

way the Mooij criterium give the correct numbers for

most of the changes of sign of the TCR, even if there

is no real reason for this correlation to be valid in

detail, at any temperature.

Among all the models proposed to explain the

anomalous transport properties, the most appro-

priate seem to us those which are related to some inter-

(1 ) In this case disordering by thermal spikes is probably

the most efficient mechanism.

(10)

1697

ference effects between impurity scattering and elec- tron-phonon or electron-electron scattering. In order

to distinguish between these two models, further experimental work is needed. In particular magne- toresistance measurements should be very helpful.

Acknowledgments.

It is a pleasure to thank Dr. Y. Quere for many valua- ble suggestions and comments. We wish also to thank

S. Paidassi who provided some samples used hire

and J. Dural and R. Blot for their technical assis- tance during the neutron and electron irradiations.

Several useful comments from the referees are grate- fully acknowledged.

Appendix

CALCULATION OF THE NUMBER OF ATOMIC DISPLACE- MENTS.

1) Electrons.

-

Lesueur has recently proposed

an analytic calculation of the number of atomic

displacements in polyatomic compounds [12]. The principal limitation of his calculation is that the electronic stopping is neglected but, concerning elec-

tron irradiation, this approximation is quite valid.

Different assumptions of this calculation (see Ref.

[12]) result in very simple expressions; the number of j atoms displaced by a primary of type k moving with

an energy Ek can be expressed at low energy in the form

where Ckj is a constant only depending on the k - j

collision and Ej is the threshold energy to displace a j atom. Let us notice here that a j atom can displace another j atom provided that its energy is at least equal to 2 Ed ; but, to displace a k-atom, its energy has to be at least equal to - E§ where £kj

=

4 Mk Mj

2 .

has to be at least equal to - Ajk where Åkj

=

( 4MkM. J )2 . Mk + Mj

In the case of Nb3Ge one finds

The total cross-sections to displace Nb atoms can

be then written with respect to atomic concentrations

as :

and a similar expression for dd e.

Emb (resp. Eme ) represents the maximal energy transferred to a Nb (resp. Ge) primary atom and

d 6 (resp.(-rp) ) the differential cross section

dE IVb p V/Ge/

for collisions between electrons and Nb (resp. Ge)

atoms.

In a previous experiment we have determined the

respective threshold energies in Nb3Ge. By using

numerical results tabulated by Oen [45] it is straight-

forward to calculate these integrals for 2.5 MeV elec-

trons. The results are :

resulting in a total dpa production rate of 5 x 10-8 dpa/s for an electron flux of 5 x 1014 e/cm2/s.

2) uranium fission fragments.

2013

Energetic heavy

2) Uranium fission fragments.

-

Energetic heavy

ions like uranium fission fragments lose most of

their energy in the form of electronic excitation. In this case an analytic determination of the energy transferred to lattice, and then of the number of atomic displacements, is not possible. In a mono-

atomic compound, one can calculate the damage

created by fission fragment sources [44] but such a

calculation is much more complicated for diatomic

compound. On the other hand a diatomic compound

like Nb3Ge can be shown to be assimilated with a monoatomic compound for calculations of energy losses. In this way it is possible to determine the energy transferred to lattice by a fission fragment

source. A typical «damage profile » in Nb3Ge is

shown in figure 8. Knowing the total energy trans-

Fig. 8.

-

Damage profile per average fission fragment in

a

Nb3Ge target for

a

thick uranium source.

ferred to

lattice E E = dEn it is possible to

estimate the number of atomic displacements by

Ed represents here the average threshold displace-

ment energy that we have determined from the atomic percentage of Nb and Ge and found to be equal to

30 eV. With a thermal neutron flux of 7 x 1011 n/cm2/s,

one finds a damage production rate of about 7 x

10-6 dpa/s for a thick uranium source.

(11)

References

[1] FISK, Z., WEBB, G. W., Phys. Rev. Lett. 36 (1976) 1084.

[2] GURVITCH, M., Phys. Rev. B 24 (1981) 7404.

[3] TESTARDI, L. R., POATE, J. M., LEVINSTEIN, M. J., Phys. Rev. B 15 (1977) 2570.

[4] MÜLLER, P., ISCHENKO, G., ADRIAN. H.. BIEGER, J.,

LEHMANN, L., HAASE, E. L., Superconductivity in d and f-band metals, D. H. Douglass, ed. (Plenum, NY) 1980, p. 369.

[5] MooIJ, J. H., Phys. Status Solidi (a) 17 (1973) 128.

[6] GURVITCH, M., GHOSH, A. K., LUTZ, H., STRONGIN, M., Phys. Rev. B 22 (1980) 128.

[7] CATON, R., VISVANATHAN, R., Phys. Rev. B 25 (1982)

179.

[8] PAÏDASSI, S., Thesis-Grenoble (1979).

[9] WEISS, F., Thesis-Grenoble (1983).

[10] DURAL, J., Rapport CEA-R 5091 (1980).

[11] CONTE, R. R., Adv. Cryogen. Eng. 12 (1967) 673.

[12] LESUEUR, D., Philos. Mag. A 44 (1981) 905.

[13] RULLIER, F., QUERE, Y. and PAÏDASSI, S., J. Physique 41 (1980) 515.

GELY, M. H., RULLIER-ALBENQUE, F., QUERE, Y., DUNLOP, A., J. Nucl. Mat. 103 (1982) 745.

[14] TSUEI, C. C. Solid State Commun. 27 (1978) 691.

[15] ZIMAN, J. M., Philos. Mag. 6 (1961) 1013.

[16] EVANS, R., GREENWOOD, D. A. and LLOYD, P., Phys.

Lett. 35A (1977) 57.

[17] NAGEL, S. R. and TANC, J., Phys. Rev. Lett. 35 (1975)

380.

[18] COTE, P. J. and MEISEL, L. V., Phys. Rev. Lett. 39 (1977)

102.

[19] SINHA, A. K., Phys. Rev. B 1 (1970) 4541.

[20] BOUCHER, B., J. Non Cryst. Solids 7 (1972) 277.

[21] COTE, P. J., Solid State Commun. 18 (1976) 1311.

[22] BIEGER, J., ADRIAN, H., MÜLLER, P., SAEMANN- ISCHENKO, G., and HAASE, E. L., Solid State Commun. 36 (1980) 979.

[23] GIRVIN, S. M., JONSON, M., Phys. Rev. B22 (1980) 3583.

[24] IMRY, Y., Phys. Rev. Lett. 44 (1980) 469.

[25] THOULESS, J., Phys. Rev. Lett. 39 (1977) 1167.

[26] ANDERSON, P. W., ABRAHAMS, E. and RAMAKRISHNAN,

T. V., Phys. Rev. Lett. 43 (1979) 718.

[27] GIORDANO, N., Phys. Rev. B 22 (1980) 5635.

[28] COCHRANE, R. W., HARRIS, R. H., STRÖM-OHLSON,

J. O. and ZUCKERMANN, M. J., Phys. Rev. Lett. 35

(1975) 676.

[29] ANDERSON, P. W., HALPERIN, B. I. and VARMA, C. M., Philos. Mag. 25 (1971) 1.

[30] SHENG, P., Phys. Rev. B 21 (1980) 2180.

[31] ANDERSON, P. W. MUTTALIB, K. A., RAMAKRISHNAN, T. V., Phys. Rev. B 28 (1983) 117.

[32] ALT’SHULER, B. L., ARONOV, A. G., Zh. Eksp. Teor. Fiz

77 (1979) 2028 (Sov. JETP 50 (1979) 968).

[33] ALT’SHULER, B. L., KHMEL’NITSKII, D., LARKIN, A. I., LEE, P. A., Phys. Rev. B 22 (1980) 5142.

[34] BRIMLOW, G. M., PRIESTLEY, M. G., Solid State Com-

mun.

45 (1983) 1063.

[35] SWEEDLER, A. R., COX, D. E., MOEHLECKE, S., J. Nucl.

Mat. 72 (1978) 50.

[36] COX, D. E., SWEEDLER, A. R., MOEHLECKE, S., NEW-

KIRK, L. R., VALENCIA, F. A., Superconductivity

in d and f-band-metals, D. H. Douglas, ed. (Plenum, NY) 1980, p. 335.

[37] POATE, J. M., TESTARDI, L. R., STORM, A. R., AUGUSTY-

NIAK, W. M., Phys. Rev. Lett. 35 (1975) 1290.

[38] DYNES, R. C., POATE, J. M., TESTARDI, L. R., STORM, A. R., HAMMOND, R. H., IEEE Trans. Mag.

Mag. 13 (1977) 640.

[39] LESUEUR, D., Rapport CEA-R 4502 (1973).

[40] LANDAUER, R., J. Appl. Phys. 23 (1952) 779.

[41] RULLIER-ALBENQUE, F., Thesis, to be published (1984).

[42] BROWN, B. S., BIRTCHER, R. C., KAMPWIRTH, R. T., BLEWITT, T. H., J. Nucl. Mat. 72 (1978) 76.

[43] MÜLLER, P., WALLNER, G., WECK, W., RULLIER- ALBENQUE, F., J. Nucl. Mat. 108-109 (1982) 585.

[44] RULLIER-ALBENQUE, F., QUERE, Y., Phys. Lett. 81A (1981) 232. The threshold energy values

are erro-

neous

in this paper. The exact values are

ENbd

=

33.7 eV and EGed

=

22.4 eV.

[45] OEN, O. S., Report ORNL-4897 (1973).

[46] LETEUTRE, J., POUCHOU, J. L., SOULLARD, J., ZUPPIROLI, L., J. Nucl. Mat. 54 (1974) 254.

[47] MUTO, R., Inst. Phys. Chem. Res. Sci. Tokyo 20 (1936)

99.

[48] ZUPPIROLI, L., BANDHYOPADHYAY, A., BOUFFARD, S.

and BERNARD, C., Open questions in the field of

localization in low dimensional conductors, to be published in 1984.

[49] BELITZ, D. and SCHIRMACHER, W., J. Phys. C. : Solid

State Phys. 16 (1983) 913.

[50] KAVEH, M. and MOTT, N. F., J. Phys. C. : Solid State

Phys. 15 (1982) L-707.

[51] IGASAKI, Y. and MITSUHASHI, H., J. Appl. Phys. 54 (1983) 836.

[52] STEWART, G. R., NEWKIRK, L. R. and VALENCIA, F. A.,

Solid State Commun. 26 (1978) 417.

Références

Documents relatifs

Resistivity measurements suggest that the metal stoechiometry and/or atomic disorder within the octahedral chains are key factors governing.. the transport

The samples show a negative temperature coefficient of the electrical resistivity (resistance anomaly) like the Kondo effect of SnMn, reported by Buckel et alJ.

solution of the phonon Boltzmann equation for the case of mixed phonon-electron and phonon-dislo- cation scattering.. A wide range of reasonable dislocation densities

By taking into account the effective electron mass in the HAP model and the empty core radius in the ECP model, an accu- rate fit of experimental data can be achieved which

Prediction of the bed friction coefficient using either high resolution bathymetric data or granulometry samples. N. Rouen, CNRS, Normandie Université, SFR SCALE) 2 ,

Faber Ziman theory are in reasonable agreement with the measured resistivity, but the temperature coefficient is four times smaller than the

For many of these alloys, particularly those without metalloid components such as YFe /17/ ZrCu 1181 and TiZrBe 16,191, there is never an extended range li- near in temperature but

These factors cause ties of transition metals amorphous alloys resistance to rise with temperature, All (metal glasses) suggested that in the low the mentioned mechanisms, both