• Aucun résultat trouvé

From localization to superconductivity in granular niobium nitride thin films

N/A
N/A
Protected

Academic year: 2021

Partager "From localization to superconductivity in granular niobium nitride thin films"

Copied!
9
0
0

Texte intégral

(1)

HAL Id: jpa-00210756

https://hal.archives-ouvertes.fr/jpa-00210756

Submitted on 1 Jan 1988

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.

From localization to superconductivity in granular niobium nitride thin films

R. Cabanel, J. Chaussy, J. Mazuer, J.C. Villegier

To cite this version:

R. Cabanel, J. Chaussy, J. Mazuer, J.C. Villegier. From localization to superconductiv- ity in granular niobium nitride thin films. Journal de Physique, 1988, 49 (5), pp.795-802.

�10.1051/jphys:01988004905079500�. �jpa-00210756�

(2)

From localization to superconductivity in granular niobium nitride thin films (*)

R. Cabanel (**), J. Chaussy, J. Mazuer and J. C. Villegier (1)

Centre de Recherches sur les Très Basses Températures, C.N.R.S., BP 166 X, 38042 Grenoble Cedex, France (1) LETI, C.E.N.G., BP 85, 38041 Grenoble Cedex, France

(Requ le 3 décembre 1987, accepté le 26 janvier 1988)

Résumé.

2014

Nous avons étudié entre 1 et 300 K les variations de la résistivité de films de nitrure de niobium

préparés par pulvérisation cathodique réactive. En travaillant à des pressions d’azote de plus en plus élevées,

on augmente la concentration d’azote dans les couches qui deviennent de plus en plus poreuses. Ceci favorise leur oxydation ultérieure à l’ambiante. La concentration d’oxygène s’est révélée jouer un rôle primordial sur la

résistivité p. Des séries d’échantillons ont ainsi été obtenues avec des résistivités à l’ambiante variant de 14 03A9.cm à 1 m03A9.cm. Les caractéristiques p (T) évoluent régulièrement depuis un comportement du type de celui observé dans des semiconducteurs désordonnés jusqu’à celui du NbN supraconducteur classique. Les

échantillons sont très hétérogènes, avec une large distribution des barrières d’énergie entre les grains

conducteurs. Dans les échantillons de haute résistivité, la conduction a lieu par sauts activés thermiquement et Ln p ~ T- n avec 1/2 ~ n ~1/4. Pour les couches moins résistives, un chemin conducteur s’établit par

percolation à travers les barrières les plus basses et dès que T est supérieur à une certaine valeur critique T*, la conductivité devient linéaire en T. Il peut alors apparaître une transition supraconductrice même sur des

films dont le comportement à l’état normal est nettement du type semiconducteur. La plus haute résistivité observée avant transition a été de 60 m03A9.cm pour une résistivité à l’ambiante de 1 m03A9.cm, soit un rapport de 60. L’application d’un champ magnétique réglable de 0 à 4 T provoque la décroissance de la température critique selon la même loi linéaire que celle observée sur des supraconducteurs de type II homogènes.

Abstract.

2014

The variations with temperature T of the resistivities of reactively sputtered NbN films have been studied from 1 to 300 K. The nitrogen concentration in the films has been increased by using higher and higher nitrogen pressures during sputtering. The porosity of the films is also increased and this favours their post oxidation at ambient atmosphere. The oxygen concentration in the films revealed to play a prime part on the resistivities p. Series of samples were obtained with room temperature resistivities from 14 03A9.cm to 1 m03A9.cm.

The resistivity variations with T regularly evolve from a disordered semiconductor like behaviour to the bad metal behaviour of the classical superconducting NbN. The samples are highly inhomogeneous with a wide

distribution of the energy barriers between conducting grains. In the high resistivity samples, thermally

activated hopping is present and Ln p ~ T-n with 1/2~ n ~1/4. In the less resistive samples, a percolating path is established through the lowest barriers and beyond some critical temperature T* the conductivity

becomes linear in T. A superconducting transition may then appear which could be observed even on samples

whose behaviour is markedly of the semiconductor type in the normal state. The highest resistivity before

transition was 60 m03A9.cm. This corresponds to a resistivity ratio of 60 with respect to the value at room temperature. Application of a magnetic field up to 4 T decreases the critical temperature in the same way as for

a homogeneous dirty type II superconductor.

Classification Physics Abstracts

73.60 - 74.70D - 81.15C

Introduction.

Niobium nitride studies extensively grew during the past twenty years, in close connection with the

development of new methods for thin-film depo-

sition. Among these methods, reactive sputtering

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:01988004905079500

(3)

796

certainly was the most successful for NbN [1]. Initial

interest at that time was the high critical properties

of the superconducting (SC) 5-NbN phase. Values of tc

=

15 K, Bc2

=

20 T and Jc

=

101° A. m-2 are now

currently achieved for the critical temperature field

at 4.2 K and current density at 4.2 K and 10 T [2]

respectively. Although research to improve these properties is still running [3, 4], current studies are

rather directed to the possible use of NbN films for

Josephson junctions based integrated circuits [5-7].

Superconducting properties actually are greatly con-

nected with the crystallographic structure, compo- sition and sputtering parameters of the films [8, 9].

For example, the highest critical temperatures are obtained on the 8-NbN phase, and superconductivity disappears as soon as the ratio [N ]/ [Nb ] of the nitrogen to the niobium atoms in the film reaches 1.5. At the same time, resistivity at ambient tempera-

ture drastically increases [8].

Much less attention has been paid to the normal phase and to the non-superconducting films of

niobium nitrides. Available results are concerning

either high ambient resistivity samples with an amorphous structure [10] or studies of the variations with temperature of the resistivity from ambient down to the SC transition [11]. In both cases, samples exhibit a semiconductor-like behaviour. A strong correlation has been found between the decrease of the critical temperature and the increase of the absolute value of the temperature coefficient of resistivity (TCR) which increases with the nitrogen

concentration in the reactive gas mixture in the

sputtering chamber [11]. These transport properties

have been proved to be reminiscent of a typical class

of materials, the so-called granular films, which are

mixtures of metal and insulators.

Granular films have been recognized for some

years as very convenient materials to study the metal

insulator transition (MIT) [12]. Moreover, when the

metal is a superconductor, an additional problem arises, i.e. the effect of localization on superconduc- tivity [13]. Interest of NbN as such a granular system then appears, inasmuch as it is possible to regularly modify the composition of the films, so that they

may regularly evolve between semi-metallic SC films and high resistivity films with a hopping type conduc- tivity. It is to be noted that these two limits only have

been studied up to now.

We report below a part of the work that we have undertaken to extend information on the transport properties of the NbN system [14]. First we briefly

describe our sputtering apparatus and analyse the

standard parameters of the film deposition. The

main parts played by the structure and post-oxi-

dation at ambient atmosphere of the films on their electronic properties are then emphasized. We ob-

tained a variety of samples whose room temperature resistivities (p RT ) lie between less than 1 mfLcm up

to p RT > 14 n. cm. The variations with the tempera-

ture of the resistivities are discussed in the frame of various hopping models. The last paragraph is specifically devoted to the more conducting films for

which a SC transition is appearing even in a highly semiconducting sample in the normal state. The ratio reaches 60 between the low temperature peak resistivity and that at room temperature ; this is, to

our knowledge, the highest reported value for a granular film.

Film deposition and characterization.

The films were produced by reactive d.c. magnetron sputtering. The apparatus was specially designed at

the Centre de Recherches sur les Tres Basses

Temperatures, in order to ensure a close control of the sputtering parameters. Since our aim was to increase the nitrogen concentration, we used a pure

nitrogen atmosphere instead of a mixture of nitrogen

and argon as is usual in depositing the classical SC NbN’s [9, 15]. As a result, we avoid any pollution

of films by argon. We have prepared a variety of

NbN films whose thicknesses vary between 0.45 and 2.6 fJ-m.

Details of the sputtering process and the prelimi-

nary characterization of the films are available in

previous papers [16, 17]. A more extensive report of the correlation between sputtering parameters, composition and resistivity of the films is in progress and will be published elsewhere [18]. We summarize below the typical sputtering conditions and our main conclusions.

For the room temperature resistivity p RT to be varied between the above mentioned limits, nitrogen

pressure has to be monitored from 1 to 10 Pa. The magnetron power density ranged between 8.5 to

28 W. cm-2 for a niobium target 6 cm in diameter.

X-ray diffraction patterns showed i) the existence on

all samples of a rather well crystallized &NbN phase

and ii) a large diffusion peak typical of an amorphous

second phase. The columnar structure of the film

appeared on SEM and TEM photographs. 100 to

200 nm in diameter columns are formed of grains

which stack to design cauliflowers. From the width of X-ray diffraction peak, the grain diameters are

found to be between 4 to 10 nm.

Diffraction patterns of transmitted electrons con-

firm that the small 8-NbN crystallites are oriented

with the [002] axis perpendicular to the substrate.

Careful analyses were performed on four samples

selected to overlap the full resistivity range. Results

are given in table I. Techniques of Rutherford Back

Scattering (RBS), Nuclear Reactions (NR) and Secondary Ion Mass Spectroscopy (SIMS) were used

to determine the atomic concentrations of niobium, nitrogen, carbon and oxygen in the films. A com-

parative study of the oxidation by means of Auger

(4)

Table I.

-

Sample characteristics.

electron spectroscopy [18, 19] lead us to the con-

clusion that oxygen concentration resulted from a

post-oxidation at ambient atmosphere of initially highly porous films. This concentration appeared to play a main part in both the TCR and the room

temperature resistivity which both increase almost

exponentially with the ratio [0]/[N] of the oxygen and nitrogen atomic concentrations.

To summarize, the studies of structure and compo- sition lead us to the conclusion that our granular

NbN films are made of metallic grains of the 5-NbN

phase embedded in an amorphous insulating phase.

For the most resistive samples, this second phase

most probably is amorphous Nb205 so that the general formula must be written in the form

NbN,,Oy. For the less resistive films, the amorphous phase is rather an inhomogeneous mixture of niobium and nitrogen atoms and voids.

Temperature variations of resistivities.

We have studied the variations of the resistivity of a

number of samples from room temperature down to 1 K. For films exhibiting high resistivities (p RT --

70 mO. cm) we could not explore the whole tempera-

ture range due to the off-limit values of the resist-

ances. Figure 1 shows a logarithmic plot of the resistivity ratio p(T)/pRT versus T-14 for some

Fig. 1.

-

Electrical resistivities of samples A to F nor-

malized with respect to their values at room temperature

versus T- 1/4. The lines. -. -. are used to separate approximately three regions where the conductivity ex-

hibits different behaviours.

samples illustrating the three typical behaviours that

we may distinguish for our films :

. when p RT > 10 mQ , cm, the conductivity cr is of

thermally activated hopping type (Region 1) ;

(5)

798

2022 when PRT’" 3 mO. cm, we will see that cr is linear versus T over a quite wide range of tempera-

ture (Region II) ;

2022 when P RT - 1 mfL cm, a SC transition occurs at T > 1 K (Region III).

All samples exhibit a negative temperature coef- ficient of resistivity TCR

=

(1/p ) (dp /dT) between

1 and 300 K. Therefore, when we refer to Ander-

son’s criterium [21] for the MIT, all films are on the insulating side. In agreement with this classification,

the low temperature conductivity is always lower

than the Mott minimum metallic conductivity [22]

(T mio ~ 0.025 e2/ha leading to P max

=

if mio ~

6 mfL cm for a lattice parameter a

=

4 A.

1. HOPPING CONDUCTIVITY (REGION I).

-

All the

curves in region I exhibit over some range from the lowest temperature a linear variation versus T- 1/4 in the logarithmic plot of figure 1. The temperature dependences are accounted for by the now classical

Mott law [22] of variable range hopping a = A T- 1/2 exp - (ToIT)n. n depends on the dimen-

sionality D of the film n

=

1/4 at 3D and 1/3 at 2D.

Mott’s analysis is referring to a disordered homo- geneous medium with localized electrons. A more convenient representation of granular metals might

be that of the model of Ambegoakar, Halperin and

Langer (AHL) [23]. Here the medium is viewed as a

resistance network where regions of high conductivi- ty are connected by less conductive resistors through

which electrons are able to move by phonon assisted tunneling. The same T- 1/4 law is obtained as in Mott’s analysis. It is to be noted that both models consider a constant density of states N (E) at the

Fermi level EF and neglect electron interactions.

When Coulomb interactions between electrons

are considered, it has been shown that N (EF)

decreases and the conductivity becomes oc exp -

(Tol T)1/2 [24]. The value n

=

1/2 has been previously reported for NbN films [10]. In order to distinguish

between the various values for n, we attempt a more detailed analysis of our results by plotting

Ln p (T) / P RT as a function of T - n with n = 1/2, 1/3

and 1/4 (Fig. 2). For the specimens with p RT ~

10 mil. cm, the best fit clearly is obtained when

n

=

1/4, but the linear variation is limited within 2.5 and 35 K (Fig. 2a). For the highest resistivity samples (Fig. 2c, d) the differences is not striking at all. The temperature range investigated is too narrow to

allow for a final choice. Nevertheless, n seems to

evolve from 1/4 for P RT ’" 10 mO. cm to 1/2 for P RT :> 10 O. cm. Such an evolution has been pre-

viously reported for granular Al [25] where a linear

variation versus T- 1/2 could be observed over more

Fig. 2.

-

Electrical resistivities of samples from region I versus T-n with n

=

1/2, 3/7, 1/3, 1/4. Different scales are used

from a) to d).

(6)

than two decades of temperature. It is also typical of

the cermets films [1].

From 1/4 to 1/2, n crosses over the value 1/3 which

unambiguously appears as the best fit for specimens

with p RT ~ 100 mn. cm (Fig. 2b). In terms of Mott’s

model this would correspond to a 2D hopping conductivity. Such an interpretation is quite unrealis-

tic for our sample whose thickness is - 700 nm. An

explanation can be found through a phenomenologi-

cal approach which takes into account a possible

variation of N (E) at the Fermi level. Pollak, by using percolation theory [26] and E. M. Hamilton, by extending the original Mott calculation [27] both proved independently that, when the density of

states is a power function of energy E near the Fermi level

then results for the 3D-conductivity u

No and p being some constants.

When we apply this model to our results, we are

led to the conclusion that an increase of the volumic fraction of the amorphous phase in our films induces the break and broadening of a potential wall for the

conduction electrons at the Fermi level. So when p

goes from 0 to 1/2 and up to 2, then n takes the

values 1/4, 1/3 and 1/2 respectively and p RT increases from - 10 tao - 102 and - 104 mn. cm.

A new model has been recently proposed which

considers the fractal feature of the medium [28]. The

authors deduce a value n

=

3/7 i. e. not very different from 1/2 for samples near the percolation threshold

where states are strongly localized. This could be the

case of our specimens with p RT > 75 mfY. cm. From table I we are able to calculate the metallic concen-

tration in sample B (p RT

=

75 mQ . cm) under rough hypothesis for the metal insulator arrangement in the film : we assume the metal to be NbN and the insulator Nb205. We find a NbN ratio equal to 20 %,

which is not so far from the geometrical critical percolation threshold pc

=

15 % [13]. In the range of

applicability of this model, To - d-3 is composition independent (d is the grain diameter). We have previously pointed out the difficulty to conclude for

n. Yet we may use our experiments to determine To by means of least squares fits. Applying this

method to several samples having + 5 mn. cm - P RT 14 000 mn. cm leads to a regular increase of

To from 104 K to 2.55 x 104 K when n = 1/2. With

n

=

3/7, the same calculations give higher but still

realistic values : 3.25 x 104 K To M 9.25 x 104 K.

Following To - d- 3 this means that the grain diame-

ter regularly decreases when the metallic NbN ratio decreases. The reduction coefficient does not fall below 0.7, which seems quite acceptable.

2. CONDUCTIVITIES WITH LINEAR DEPENDENCE ver- sus TEMPERATURE (REGION II).

-

When the am-

bient resistivity lies between 1 and 10 mQ.cm the

logarithmic plot of figure 1 is no longer useful and

we rather have to discuss our result by plotting the conductivity ratio versus temperature in linear scales,

as shown for two samples in figure 3. A critical temperature T* can then be defined above which the temperature variation of a can be written :

Fig. 3.

-

Electrical conductivity of samples E and F

normalized with respect to their values at room tempera-

ture versus T.

We observed this behaviour on several samples

with p RT ~ 3 mO. cm. T* ranged from 10 to 80 K,

the constant u was either positive or negative and

the coefficient b varied between 1 and 4

(0. cm) - K-1, depending on the sample. Such a

linear variation has been previously reported for

GeAu granular films [29]. AHL [23] have suggested

that a conductivity proportional to T might follow

from their analysis when, in the resistance network,

a small number of high conductivity regions exist

which are randomly connected through the whole

film length. We expect the film to be highly in- homogeneous. In our case, T* could be interpreted

as the temperature at which a high conductivity path

is established throughout the sample by connecting

various conductive regions through small energy barriers of order kT*.

This interpretation emphasize the part played by

the barriers in our films. For low resistivity samples,

these barriers remain weak over the whole film.

When the ratio of the insulating phase increases,

some barriers become higher and begin to live apart

some high conductivity paths ; there is a wide

distribution of the barrier heights : this is the upper

linear conductivity regime.

(7)

800

When we go on increasing the insulator concen-

tration, all the high conductivity grains are insulated

from one another by large barriers and the hopping regime is reached.

At temperatures lower than T*, the slopes of the

or ( T) curves decrease and over a decade of tempera- ture, the conductivity is described by :

From the graphs of figure 4, for samples E and F, we deduced q

=

0.86 and q

=

0.75 respectively. Starting

Fig. 4.

-

The temperature dependent term of the conduc-

tivity of samples E and F versus T.

from the scaling theory of localization [30], Imry [31]

shows that inelastic scattering may govern the con-

ductivity when the inelastic mean free path lin is

greater than some localization length, here the grain

size. The mean time between inelastic events T;n increases when T goes down to zero like

T - 2 q, so that cr oc- 1 in oc T - 112 oc Tq. ’The value

q

=

0.75 is typical of inelastic electron-electron interactions [32]. Such variations with some power of T for a (T ) have been reported for NbN films [33]

which exhibit q values of order 1.5 related to

electron-phonon interactions.

The values (Tõ that we have calculated for our

samples E and F are 380 mfl.cm and 175 mn.cm

respectively. In the scaling theory for localization in

a 3D system [30] the maximum metallic resistivity is proportional to the length scale of the sample. For granular metals this length is the metallic grain size d

rather than the lattice parameter of Mott’s model

[22]. Taking d - 50 A for our films leads to p max ~ 40mn.cm. Thus samples E and F would be in- sulators at T

=

0 K, and the preceding analysis is by

no means pertinent. Actually we have to recall that

our films are highly inhomogeneous and the conduc-

tivity can be greater on preferential paths, as em- phasized above.

The superconducting samples.

For sample E, the decrease of the conductivity when

T is decreasing is sharply stopped at Tc ~ 2 K where

a SC transition arises. We observed such a transition

at T > 1 K on almost all samples with p RT 1 mH.cm. Generally the resistivity slowly increased

from room temperature to liquid helium tempera- ture, as shown in figure 5 for sample D. Sample E

was sputtered under the same conditions, but it was

further annealed at 350 °C and - 10-4 Pa for 12 h.

Fig. 5. - The variations versus temperature of the resis- tivities of samples E and D. A transition to a superconduct- ing state occurs at T, 1.5 K (sample D) and Tc ~

2 K (sample E).

The resulting resistivity behaviour is quite interest- ing : just before transition, at 2.5 K, the resistivity

reaches 60 mfl.cm, the highest value we may observe

on these samples. It corresponds to a ratio

p (2.5 K)/ P300 ~ 60, which is greater, by almost one

order of magnitude, than the currently reported

results on granular materials [11, 34-36]. The value

60 mfl.cm is much larger than the critical resistivity

Pc

=

(h/e2) d ~ 3 mn. cm associated with the grain

size d

=

7 nm for specimen E. This means that this

size is larger than the critical length dc for which the superconducting gap d is equal to the separation

between energy levels of a single grain at Fermi

energy :

For NbN, taking N (EF) - 1.4 x 10" eV-1 cm-’ [37]

and 4 - 3 meV [6] we find de == 3 nm. Our sample

then illustrates the large-grain model of Imry and Strongin [38] where SC may extend over the whole

film, even if the localization is governing the conduc- tion. The sharp transition at a well defined tempera-

ture Tc well below the critical temperature of bulk material would mean that the normal state resistivity

of the grains is very small compared to that of the

barriers between grains, in agreement with our

(8)

conclusions on the microscopic structure of the

films. 7c is the temperature at which the Josephson coupling energy between a sufficient number of

grains over reach the energy barrier.

Our study of the variation of the upper critical field B,2 versus temperature near T, also confirms this idea (Fig. 6). Bc2(T) is linear in T as is the case

Fig. 6.

-

Influence of an applied magnetic field from 0 to 3.67 T on the low temperature resistivity of sample E. The

insert shows the linear decrease of the upper critical field

Be2 when T is increased.

for a dirty type II SC. This behaviour has been

previously reported for NbN films [39] having

0.2 mil.cm p RT 2.5 mQ . cm but with a resistivity

ratio which does not exceed 3. It disagrees with what

is generally observed in high resistivity granular

films made of a mixture of metal and insulator such

as InGe or PbGe [40]. From the GLAG theory the slope of Bc2(T) near Tc for a homogeneous dirty type II SC is proportional to the normal state resis-

tivity p,, and the electronic coefficient of the specific

heat y :

The numerical factor is for a result in T. K-1 with y in mJ . cm- 3 . K- 2 and p in >Q . cm. From the insert of

figure 6 we deduce dBc2/dT == 5 T. K-1. Assuming

y

=

0.32 x 10- 3 mJ . cm- 3 . K- 2, as for bulk NbN

[39] we find Pn ~ 350 um . cm which is much lower than the film resistivity, as expected from the analysis just above, and which is physically accept- able as the resistivity of the NbN grains, compared

to the bulk resistivity of this material [39].

Conclusion.

By ensuring a close control of the sputtering par- ameters, we have been able to prepare series of NbN films with various nitrogen concentrations. Careful

analysis of the structure and composition of the films lead us to the conclusion that they always are made

of metallic grains of the ô-NbN phase separated by

an amorphous insulating phase. We have shown

that, when the nitrogen pressure in the sputtering

chamber is increased, the resulting films undergo

oxidation at ambient atmosphere. The fixed oxygen atoms play a prime part in establishing the energy barriers for electrons between grains, so that the

room temperature resistivity of the films is well correlated to the ratio between oxygen and nitrogen

atoms. The variations with the temperature of the resistivities regularly evolve from a disordered semi- conductor-like behaviour to the bad-metal behaviour of the classical SC NbN. The samples are highly inhomogeneous with a wide distribution of the energy barriers between conducting grains. In the

less resistive samples, a percolating path is estab-

lished through the lowest barriers and beyond some

critical temperature T* the conductivity becomes

linear in T. For such samples, inelastic scattering can prevail at low temperatures. Increasing the concen-

tration of the &NbN phase in the films allows for a

semiconductor to superconductor transition to occur

at Tc by Josephson’s coupling of grains along a percolating path. A value as high as 60 has been observed between the low and room temperature resistivities of a previously annealed film. When a

magnetic field is applied, T, is depressed in the same

way as it would be for a homogeneous dirty type II SC. Finally, the granular NbN system has been shown to be very interesting for the study of

MIT at the same level as other metal insulator mixtures such as Al-Ge, Pb-Ge or Al-A1203, Sn-Sn02 which have been more widely studied.

Acknowledgments.

The authors wish to thank Dr. R. Rammal for

enlightening discussions and very fruitful comments

on the analysis of the results. The help of J. Geneste

during sample elaboration is gratefully acknow- ledged.

References

[1] GAVALER, J. R., HULM, J. K., JANOCKO, M. A., JONES, C. K., J. Vac. Sci. Technol. 6 (1969) 177.

[2] GAVALER, J. R., SANTHANAM, A. T., BRAGINSKI,

A. I., ASHKIN, M., JANOCKO, M. A., IEEE Trans. Mag. 17 (1981) 573.

[3] ASHKIN, M., GAVALER, J. R., GREGGI, J., DE- CROUX, M., J. Appl. Phys. 55 (1984) 1044.

[4] VAN DOVER, R. B, BACON, D. D., SINCLAIR, W. R., J. Vac. Sci. Technol. A 2 (1984) 1257.

[5] CUKAUSKAS, E. J., CARTER, W. L., Adv. Cryog.

Eng. 32 (1986) 643 (Plenum Press, New York).

[6] VILLEGIER, J. C., VIEUX ROCHAZ, L., GONICHE, M., RENARD, P., VABRE, M., IEEE Trans.

Mag. 2 (1985) 498.

(9)

802

[7] CHEN, L., GREN, S. M., LUO, H. L., J. Low Temp.

Phys. 64 (1986) 145.

[8] SPITZ, J., CHEVALLIER, J., AUBERT, A., J. Less Com. Met. 35 (1974) 181.

[9] THAKOOR, S., LAMB, J. L., THAKOOR, P. and KHANNA, S. K., J. Appl. Phys. 58 (1985) 4643.

[10] CHEVALIER, J., BAIXERAS, J., ANDRO, P., J. Phys.

42 (1981) 405.

[11] JONES, H. C., Appl. Phys. Lett. 27 (1975) 471.

[12] ABELES, B., Appl. Solid State Sci. 6 (Raymond Wolf

Academic Press, N.Y.) 1976 1.

[13] DEUTSCHER, G., PALEVSKI, A., ROSENBAUM, R., Localization-Interaction Transport Phenomena,

Eds. B. Kramer, G. Bergman, Y. Ruynseraede,

Solid State Sci. (Springer, Verlag) 61 (1985) 108.

[14] CABANEL, R., Thesis, Institut National

Polytechnique de Grenoble (1986).

[15] BACON, D. D., ENGLISH, A. T., KAKAHARA, S., PETERS, F. G., SCHREIBER, H., SINCLAIR, W. R., VAN DOVER, R. B., J. Appl. Phys. 54 (1983) 6509.

[16] CABANEL, R., JOUBERT, J. C., CHAUSSY, J., MAZUER, J., Ext. Abstract VII, Int. Conf. of Solid Compounds of Transition Elements, Vie-

nne (9-13 avril 1985), P2 A4.

[17] CABANEL, R., CHAUSSY, J., JOUBERT, J. C., MAZUER, J. reference [14] p. 227.

[18] CABANEL, R., CHAUSSY, J., GENESTE, J., MAZUER, J., VILLEGIER, J. C. (to be published).

[19] BLANC, E., LAUROZ, C., SEGAUD, J. P., Contrat GR-705.089/CENG.

[20] CABANEL, R., BARRAL, G., CHAUSSY, J., DE- LABOUGLISE, G., JOUBERT, J. C., MAZUER, J., submitted to J. Electrochem. Soc.

[21] ANDERSON, P. W., Phys. Rev. 109 (1958) 1492.

[22] MOTT, N. F., Philos. Mag. 26 (1972) 1015.

[23] AMBEGAOKAR, B., HALPERIN, B. I., LANGER, J. S., Phys. Rev. B 4 (1971) 2612.

[24] EFROS, A. L., SHKLOVSKII, B. I., J. Phys. C 8 (1975)

L 49.

[25] CHUI, R., DEUTSCHER, G., LINDENFELD,, P., MCLEAN, W. L., Phys. Rev. B 23 (1981) 6172.

[26] POLLAK, M., J. Non Cryst. Sol. 8-10 (1972) 486.

[27] HAMILTON, E. M., Philos. Mag. 26 (1972) 1043.

[28] DEUTSCHER, G., LEVY, Y., SOUILLARD, B., Europhys. Lett. 4 (1987) 577.

[29] DODSON, B. W., McMILLAN, W. L., MOCHEL, J. M., DYNES, R. C., Phys. Rev. Lett. 46 (1981)

46.

[30] ABRAHAMS, E., ANDERSON, P. W., LICCIARDELLO, D. D., RAMAKRISHNAN, T. V., Phys. Rev. Lett.

42 (1979) 673.

[31] IMRY, Y., Percolation, Localization, Superconduc- tivity, Eds. A. M. Goldman and S. A. Wolf

(Plenum Press, N.Y.) 1984, p. 189.

[32] PROBER, D. E., reference [31] p. 231.

[33] DOUSSELIN, G., Private communication.

[34] GUBSER, D. U., WOLF, S. A., FRANCAVILLA, T. L., FELDMAN, J. L., Inhomogeneous superconduc-

tors 1979, AIP Conf. Proc. , Eds. Du Gubser,

T. L. Francavilla, J. R. Leibowitz and S. A.

Wolf 58 (1980) 159.

[35] SHAPIRA, Y., DEUTSCHER, G., Phys. Rev. B 27 (1983) 4463.

[36] VAN HAESENDONCK, C., BRUYNSERAEDE, Y., Phys.

Rev. B 33 (1986) 1684.

[37] TOTH, L. E., Trans. Mat. Carbides and Nitrides

(Acad. Press., N.Y.) 1971.

[38] IMRY, Y., STRONGIN, M., Phys. Rev. B24 (1981)

6253.

[39] JUANC, J. Y., RUDMAN, D. A., VAN DOVER, R. B., SINCLAIR, W. R., BACON, D. D., Adv. Cryog.

Eng. 32 (1985) 2.

[40] DEUTSCHER, G., Percolation-Localization-Supercon- ductivity, Eds. A. M. Goldman and S. A. Wolf

(Plenum Press, N.Y.) 1984, p. 95.

Références

Documents relatifs

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

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

After annealing our samples well above Tx the resistance minima became extremely shallow (no- te the large scale in figure 1 ) and are shifted to very low temperatures so that

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

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des

We will show that despite the presence of high industrial electromagnetic noise, we managed to recover a reliable 3D resistivity image of the geothermal field but

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

i\lthough the effects of heat stress on pig performance are weil descrihed in the litera\ ure. little is known about how am bien\ temperature could aflèct the ir ability to