• Aucun résultat trouvé

Aspects of the thermodynamics of metallic solutions

N/A
N/A
Protected

Academic year: 2021

Partager "Aspects of the thermodynamics of metallic solutions"

Copied!
11
0
0

Texte intégral

(1)

HAL Id: jpa-00236677

https://hal.archives-ouvertes.fr/jpa-00236677

Submitted on 1 Jan 1962

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.

Aspects of the thermodynamics of metallic solutions

O.J. Kleppa

To cite this version:

O.J. Kleppa. Aspects of the thermodynamics of metallic solutions. J. Phys. Radium, 1962, 23 (10),

pp.763-772. �10.1051/jphysrad:019620023010076300�. �jpa-00236677�

(2)

763.

ASPECTS OF THE THERMODYNAMICS OF METALLIC SOLUTIONS

By O. J. KLEPPA,

Institute for the Study of Metals and Department of Chemistry,

The University of Chicago, Chicago 37, Illinois.

Résumé,

2014

Un travail récent sur la thermodynamique des solutions métalliques a donné des

informations qui ne peuvent être obtenues à partir d’une analyse des différents diagrammes de phases seulement. Certains systèmes de solutions binaire sont discutés en vue d’illustrer l’influence

sur les propriétés thermodynamiques d’une différence de taille, d’électronégativité et la valence

caractéristique des deux corps en solution. Une attention particulière sera donnée à l’effet de valence sur l’enthalpie de mélange pour les solutions des métaux du groupe B et au problème général de l’entropie de mélange dans les systèmes des solutions métalliques.

Abstract.

2014

Recent work on the thermodynamics of metallic solutions has produced infor-

mation which cannot be obtained from an analysis of the various phase diagrams alone. Selected

binary solution systems will be discussed with a view towards illustrating the influence on the

excess thermodynamic properties of a difference in the size, the electronegativity and the charac- teristic valence of the two solution partners. Particular attention will be given to the effect of valence on the enthalpy of mixing in terminal solutions of group B metals, and to the general problem of the excess entropy of mixing in metallic solution systems.

LE JOURNAL DE PHYSIQUE ET LE RADIUM TOME 23. OCTOBRE 1962,

Since the turn of the century a great deal of information has become available on the thermo--

dynamic properties of metallic solutions. This work originally was stimulated through Gibbs’ dis-

covery of thelphase rule. Even today the most

extensive body of thermodynamic information on

alloy systems is contained in the literature on phase equilibria. However, it should be recognized that

this type of information, valuable as it is, leaves

much to be desired in any detailed discussion of solution thermodynamics. While the phase dia-

grams of course reflect the dependence of the total free energy of the system on composition and on temperature, they usually do not permit an unam- biguous numerical evaluation of the theoretically interesting excess thermodynamic quantities. In

the present discussion we shall in the main be con-

cerned with the (intégral) excess free energy, GB,

the excess enthalpy, HB, and the excess entropy of mixing, SB [1]. These quantities are related through the fundamental thermodynamic relation

GB = HB -TSB.

We shall not hère consider the various experi-

mental techniques used in detailed studies of the

thermodynamic properties of metallic solution [2].

Among these the most important are high tempe-

rature galvanic cell and vapor pressure methods for free energy déterminations, and high tempe-

rature reaction calorimetry for the mixing enthal- pies. The best entropy data usually are obtained through combination of equilibrium free energy values with éalorimetric enthalpies.

In considering data on metallic solutions repor- ted in thé literature, a word of caution is in order.

It is particularly important to keep in mind that

all solid state rate processes are exceedingly slow, compared to the time involved in making the usual thermodynamic measurements. Therefore, obser-

vations made on a given solid solution specimen

may not in fact relate to a state of thermodynamic equilibrium. Note also that the number of binary systems with extensive solid solubility is quite res-

tricted.

Some of the complications associated with studies of solid solutions are eliminated in work on liquid alloys. In fact, thermodynamic data on liquid

metallic solutions are in certain respects more extensive and also more reliable than corres-

ponding data on solid solutions. For this reason,

we shall in the present paper consider examples

selected both from solid and from liquid metallic

solution systems. Most of the available infor- mation relates to binary alloys involving Group B

metals. We shall restrict our discussion to this

type of system.

Idealized models for metallic solutions.

-

During recent decades, impressive strides have been taken in the statistical mechanics of mole- cular solutions [3]. However, most of this work

has been based on models which bear very little resemblance to the metallic state. For example,

a characteristic feature of the important practical

solution theories (i.e. those that lend themselves to detailed numerical calculations) is the assumption

that the total cohesive energy may be approxi-

mated by a summation of pair interactions. It has become increasingly obvious in recent years

tliat this approach is wholly unsatisfactory for

metallic systems. Nevertheless, among- the sta-

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphysrad:019620023010076300

(3)

tistical theories the so-called lattice theories of solutions merit special comments. These have

contributed substantially toward our understan- ding of the order-disorder phenomenon, and pro- vide the basis for the usual theoretical estimates of

departures from random mixing in metallic solu- tions..

In the simplest version of the lattice theory of

solution it is assumed that the total cohesive energy arises from nearest neighbor interactions or

"

bonds. Each bond has a certain energy, v, the nature of which is unspecified. This energy is assumed to be independent of the other bonds

formed by the atoms in question, i.e. it is inde-

pendent of composition. Under this assumption,

the total internal energy of the alloy A-B is simply

Here nAB, ngg and nBB are the numbers of nearest

neighbor bonds, each of fixed energy vAB, vAA and vBB, respectively.

In statistical calculations based on this model

one adopts a latthe frame of reference, with a coor-

dination number z. It is then found that all the molar excess thermodynamic properties of the con-

sidered mixture can be expressed in terms of a

single interaction parameter

where No is Avogadro’s number.

For the purpose of the present brief discussion it is useful to consider this theory in two approxi-

mations [4].

a) Inthe" zeroth " approximation it is assumed

that À is sufficiently small (compared to RT) so

that the mixing of the two components is essen- tially random. In this (Bragg-Williams) case we

obtain

On general thermodynamic grounds we know

that for positive values of HE (and of À) this system

must segregate at low temperatures into an A-rich

and a B-rich phase. In the present approximation

the critical mixing temperature is 7’c = X12R.

In a similar way we arrive at the conclusion that for negative values of HE (and X) the equilibrium

state at low temperature must involve the for-

mation of an ordered phase AB. Again, in the con-

sidered approximation, the critical (Curie) point

for the ordering process is

b) In higher approximations of this theory it is

taken into account that for any non-zero value of the interaction parameter À there will be some departure from random mixing. For À > 0,

i.e. vAB > 1 /2 (VAA + VBB), a larger than random number of AA and BB bonds in predicted (" clus- tering "). Similarly, for X 0 one predicts some preference for the AB configuration at all finite

temperatures (" short range order ").

The theory permits estimates to be made of

these departures from random mixing, and of the ensuing loss in entropy. For moderate valu"es of À

we obtain for this configurational excess entropy

In view of this formula, it is indicated that above

the critical temperature (T > À/27?) these negative configurational entropy contributions should be

quite small (- 0. 1 cal/degree/mole or less). Often they may be neglected compared to other and

much larger thermal contributions (See below).

In recent years attempts have been made to make calculations of the thermodynamic properties

of metallic solutions on the basis of solid state

theory. So far, these theories have not approached

the solution problems from the more compre-

hensive point of view of molecular solution theory.

Instead they have focused their attention on some

particular problem which lends itself to theoretical treatment. The most successful of these theories is Friedel’s treatment of the relative valence effect in terminal solutions [5]. We shall have more to say about this below.

Selected examples of real metallie solutions.

-

It is well known from the extensive°work on binary phase diagrams, particularly by Hume-Rothery

and his school, that these diagrams reflect the working of at least three important " factors " the

"

size " factor, the " electro-chemical " (or better the " chemical affinity " or " electronegativity ")

factor and the " valence " factor. It is obvious that these factors are interrelated, and that to con-

sider them separately represents at best a crude approximation. Nevertheless, it serves as a useful starting point in a survey of the properties of real

metallic solutions.

These three factors, since they are recognized readily in the appearance of many phase diagrams,

are even more strongly reflected in the excess ther- modynamic properties. As might be expected,

one usually finds that a différence in atomic size

gives rise to positive contributions to H-, FE and SE Similarly, a difference in electronegativity gives

rise to negative contributions. Finally, it is found

that a difference in the characteristic valence of the two solution partners often produces an asymmetry in the excess thermodynamic functions, i.e. a pro-

nounced difference between the solutions of A in B and those of B in A. We shall illustrate these .

general observations by considering some selected

examples of binary solution systems.

(4)

765

a) SYSTEMS WITHOUT VALENCE EFFECTS. - The

alloys of silver and gold are sometimes uncritically

referred to as " ideal ". Thus this system has a

very simple phase diagram with a complete range

of solid solutions (see fig. 1), the lattice parameter

FIG. 1.

-

Phase diagram and excess thermodynamic data

for silver-gold alloys (ref. [7]).

departs by less than 0.15 % from Vegard’s rule,

and there is no convincing evidence for phase sepa-

’ration or for the existance of any ordered phase at

low temperatures.

It is well known that the atomic radii of gold

and silver are very similar. Thus the " size f actor " is very favorable. On the other hand,

these metals have a substantial différence in electro-

negativity (about 0.5 units on Gordy’s scale [6]).

Thus, chemical reasoning suggests that " ionic "

forms of the type Ag+ Au7- may be of some impor-

tance in bonding between these metals. We accor-

dingly anticipate negative departures from ideality.

The phase diagram, given in the upper part ’of figure 1, shows a very small liquidus-solidus sepa- ration, which points in the same direction.

Detailed thermodynamic measurements, the results of which have beeri summarized by White, Orr and Hultgren [7] confirm this, as illustrated in figure 1.

The rather large negative enthalpies of mixing suggest that this system might be expected to have

an ordered intermetallic phase at low temperature.

So far, this has not been observed.

It is noteworthy that this system also has a negative excess entropy of mixing, which amounts

to about 0.3 cal/degree g. atom in the middle of the

system. Although this value is not accurately known, it certainly is too large to be explained by departures from random mixing. Therefore, it is

indicated that the excess entropy in this system in

the main must be of thermal, i.e. in this case presu-

mably of vibrational origin. We have, under the assumption that we can eff ectively " freeze in "

the high temperature random configuration,

Here C§i represents the deviation of the heat capacity from a linear dependence on composition

( Kopj-Neumann rule). Note that non-zero values

of C,» of course contribute also to HB. However,

these contributions tend to be overshadowed by

the temperature independent terms.

,

The most significant thermal contributions to

,

the excess entropy usually arise at temperatures

well below the characteristic temperature of the alloy. But regrettably, we do not have low-tempe-

rature heat capacity data for silver-gold alloys

which permit a check of eqn. (3). On the other

hand, if our identification of the excess entropy as vibrational is correct, we should be able to turn the problem around and use the observed excess

entropies to estimate the Debye 6’s for the alloys.

For elevated temperatures (T » 6) We have on

the Debye model the following simple expression

for the vibrational excess entropy of the alloy

Here 0 is the Debye temperature for the alloy

and 6E

=

0 -[r0Au + (1-x) OAg]. The observed

negative excess entropies for gold-silver accordingly imply positive departures of 0 from a linear depen-

dence on composition (higher frequencies of vibra-

Fie. 2.

-

Young’s modulus for solid solutions in the

system silver-gold (ref. [8]).

(5)

766

tion). If we adopt values of 0 of 215° and 1800 for silver and gold, respectively, we estimate from the entropy data a 6 of 209° for the equi-atomic alloy.

This value is about 6 % larger than the mean of

the values for silver and gold. Our result is con- sistent with the data on Young’s modulus for this

system reported by Kôster and Rauscher [8], and presented in figure 2. This agreement demons- trates, as was originally suggested by Zener [9], that

data on the elastic constants for metallic solid solu- tions may be used to advantage in estimating vibra-

tional entropy contributions. Of course this holds

equally well for systems with and without valence effects.

As a second example of systems without valence effects let us consider copper-silver, for which the phase diagram and some other thermodynamic

data are presented in figure 3 [10]. Here the elec-

FIG. 3.

-

Phase diagram and excess thermodynamic data

for silver-copper alloys (ref. [10]).

tronegativity différence is zero on Gordy’s scale,

while the size factor is fairly large. Thus we expect and find substantial positive deviations

from ideality, both in the solid and in the liquid

state. Note that the (positive) excess fiée ener- gies are significantly larger in the solid than in the

liquid solutions. This behavior is characteristic of

simple alloy systems where there is a large diffe-

rence in size between the two solution partners. It

reflects the fact that the size factor is far more cri- tical in the solid than in the liquid state. On the

other hand, when the size difference is small, as in silver-gold (and in silver-palladium, see below), it

may be found that the solid solutions have lowér

excess free energies of mixing than the corres- ponding liquid alloys.

Beginning with the works of Pines [11] and

Lawson [12] several attempts have been made to calculate excess thermodynamic quantities for binary solid solutions from elastic theory, i.e. on

the basis of the " misfit " between solvent and solute atoms. On the whole, there is only fair agreement between theory and experiment in these

calculations.

b) HUME-ROTHERY TYPE SYSTEMS.

-

Histo-

rically, the most important systems exhibiting

valence effects are the alloys formed between the

mono-valent metals copper, silver and gold and

many metals of higher valence. Among these sys-

tems the simplest ones are those where both solu- tion partners belong to the same row in the periodic table, such as the alloys of copper with zinc, gal-

lium and germanium and the alloys of silver with cadmium, indium and tin. In these cases the size factors are moderate, and there is a relatively

small difference in electronegativity between the

two metals. Thus the setting is right for a display

of valence effects.

We give in figure 4 the equilibrium phase dia-

grams for the alloys of silver with cadmium, indium

and tin. These diagrams may be used to illustrate the well-known Hume-Rothery rules. However,

we shall restrict the present discussion to thermo-

dynamic information which cannot be derived from the phase diagrams alone. In particular, we shall

review the data on the enthalpies of mixing, which

are presented in figure 5 [13, 14]. These data apply for alloys which are stable at 450 OC, at

which temperature cadmium, indium and tin are

all liquid. This circumstance turns out to be a

fortunate one, since it gives an insight into the thermodynamics of these systems which can not readily be obtained through study of the solid

phases alone.

The most striking feature of the data in figure 5

is the remarkable difference in the properties of the

terminal solutions in these systems, i.e. between

the solutions of cadmium, indium and tin in silver,

on the one hand, and the solutions of silver in the

liquid multi-valent metals on the other.

From a purely thermodynamic point of view we

may characterize the terminai solutions by two sets of quantities, namely the liniiting slopes and

the limiting curvatures of the excess thermodynamic

functions. In the special case of the enthalpy of

mixing the limiting slope represents the change in

enthalpy associated with the transfer of a solute

(6)

767

FIG. 4.

-

Phase diagrams in the sys- tems silver-cadmium, silver-indium and silver-tin.

FiG. 5.

-

Excess enthalpy data at

450 DC for the systems silver-

cadmium, silver-indium and silver-

tin (ref. [13], [14]).

(7)

768

atom from its reference state (here the pure solid

.

or liquid solute) into the pure solvent. In general

this transfer will involve a very drastic change in

the environment of the solute atom. Therefore, it

is obvious that a realistic calculation of this quan-

tity is a very formidable task. This was attempted by Friedel on thé basis of a suitable thermo-

dynamic cycle [5]. However, in this cycle the heat

of solution appears as a différence between very

large numbers, and the results are of rather limited value.

More interest is attached to the limiting curva-

ture of the enthalpy of mixing. This quantity is a

measure of the interaction between the solute atoms in the matrix of the solvent, and was success-

fully calculated by Friedel [5]. In the present

paper we shall refrain from discussing the funda-

mental basis for this theory. The most important

conclusion reached is that if the complicating fac-

tors of a large différence in atomic size and of

strong chemical interaction between solute and solvent do not overshadow the valence effect, one might expect the limiting curvature of the enthalpy

of mixing to be determined by the difference in valence between solvent and solute. Thus, if the

solute is of higher valence than the solvent, there

should be an effective repulsion between the solute atoms, i.e. a positive curvature. If the situation is

reversed, there should be attraction and a negative

curvature. Furthermore, for différent solutes in the same solvent, it is predicted that there should be a rough proportionality between valence diffe-

rence and curvature.

The enthalpy data presented in figure 5 are in

reasonable agreement with these predictions.

Thus we find positive and increasing curvatures for the solid solution of cadmium, indium and tin in silver and, what is more remarkable, negative cur-

vatures for silver in liquid indium and tin. The

data for silver in liquid cadmium are somewhat

FIG. 6.

-

Plot of HEjx for cadmium and indium in silver (ref. [14]).

FIG. 7. - Plot of HE /x for zinc and gallium in copper

(ref. [15]).

uncertain. In this case it is probable that we have

a positive rather than a negative limiting curvature.

In order to attempt a more quantitative check

of the stated predictions we give in figure 6 a sui-

table graph for the solid solutions of cadmium and indium in silver. Note that the slope of HE lx

versus x is a measure of the curvature of the

enthalpy of mixing. The figure indicates that the

curvature for indium is about twice that for cad- mium. Very recent calorimetric work by Kleppa

and King [15] on the solid solutions of zinc and

gallium in copper shows comparable agreement, as

demonstrated in figure 7. However, it should be

recognized that the data in figures 6 and 7 do not

cover the very dilute range where the theory is

most applicable. Therefore, one should not over- emphasize this agreement with theory.

c) ALLLOYS OF GROUP II B METALS. -Many of

the Group B metals fall within a fairly restricted

range with respect to atomic size and electro- negativity. Therefore, pursuing the line of rea- soning advanced in the present paper, one might expect that the alloys formed between the divalent metals zinc, cadmium and mercury, on the one hand, and metals of higher valence, on the other,

should also display valence correlated solution pro-

perties. In this context it is interesting to note

the suggestion made by Raynor [16], that the very ’

(

(8)

169 low solid sofubilities of multivalent metals in zinc

and. cadmium might be related to certain details of

the band structure of these solvent metals. Apart

from this, there is little evidence in the phase dia-

grams for the importance of valence effects.

It turns out that a detailed analysis of the excess thermodynamic properties for some of the liquid alloys of zinc and cadmium is more revealing.

This is illustrated in figure 8, where we present

some of the integral excess thermodynamic quan-

Fic. 8.

-

Some excess thermodynamic quantities for liquid zinc-cadmium, zinc-indium and zinc-tin (ref. [17]).

tities for the liquid àlloys of zinc with cadmium,

indium and tin [17]. The data undoubtedly reflect

both a différence in size and a différence in valence between the solution partners. Nevertheless, the

correlation between excess* properties and valence différence is apparent.

A survey of the differential excess quantities for

various solutes in zinc is even more suggestive [181.

For this purpose we give in figure 9 the differential

excess enthalpies of liquid zinc as a solvent for a

wide range of other group B metals. In figure 10

we present a similar graph which shows the diffe- rential excess entropies plotted against électron

concentration. We see that for all the considered, muttivatent solutes the exeess entropies are posai-

tive-and,-except in"the case of bismuth, of compa- rable magnitude at the same electron concentra- tion. Similarly, the enthalpy data indicate a clear

corrélation between excess enthalpy and valence difference. However, it should be noted that the

"

valence effect 1) for solutions in liquid zinc (and cadmium) is opposite in sign to that predicted by

the Friedel theory [18 A).

.

We have suggested elsewhere [19]7,that the ther- modynamic properties in these terminal solutions may possibly be related to departures from free

electron behavior ni liquid. zinc and cadmium.

There is some support for this in the anomalous temperature dependenae of thé electrical conàuc-

tivity ouf thèse liquid metals. On the other hand,

(9)

770

FIG. 9.

-

Differential excess enthalpies for zinc’in zinc- rich liquid alloys (ref. [19]).

FiG. 10.

-

Differential excess entropies for zinc in zinc- rich liquid alloys (scale on left for Zn-Ag and Zn-In only. Other curves displaced by multiple of 0.1 cal/deg,

ref. [19]).

recent Hall coefficient measurements by Esch [20]

do not seem to support this interpretation.

Finally, it should be mentioned that the liquid alloys of mercury with metals of higher valence

show a much more complex pattern of thermo-

dynamic behavior than the corresponding zinc and

cadmium systems [21].

(d) ALLOYS INVOLVING TRANSITION METALS ; THE

SILVER-PALLADIUM SYSTEM. --At the present time

the bulk of the detailed thermodynamic infor-

mation on metallic solutions pertains to alloys for-

med by non-transition metals. Therefore, there

is as yet no basis for any general systematic dis-

cussion of the solution properties of transition metal

alloys. On the other hand, reliable thermodyna-

mic data have become available in recent years for

a number of binaries involving transition metals.

In some cases these exhibit features not found in the examples considered above, as we shall illustrate

by considering the. silver-palladium system.

We present first in the upper part on figure 11

the accepted phase diagram for this system. In

any discussion based only on the phase diagram,

this systems would appear to be very simple, and it might be compared, for example, to silver-gold.

Thus, we estimate from the course of the liquidus

curve that GE(s)

-

G’(1) for silver-palladium

FIG. 11.

-

Phase diagram and excess thermodynamic data

for silver-palladium alloys (refs. [23], [24]).

should be about -150 cal/g; atom in the middle of the system. The comparable figure for silver- gold is about

-

50 cal/g. atom [22]. On the other hand, the larger liquidus-solidus separation indi-

cates a more positive (or less negative) departure

from ideality than in silver-gold.

It was only on the completion of the recent

e. m. f. study by Pratt [23], and the calorimetric work by Chan, Anderson, Orr and Hultgren [24]

that it became generally recognized that the ther- modynamic properties of silver-palladium are really quite complex. This will be noted from a

brief look at the thermodynamic data given in the

lower part of figure 11. It is particularly note- worthy that silver-palladium has a very large nega- tive excess entropy of mixing, which for equi-

atomic alloys amounts to about

-

1.8 cal/degree

g. atom at 1 000 OK. This figure should be com- pared with the ideal entropy of mixing of + 1.38

cal/degree. Thus we find that at elevated tempe-

ratures the formation of an essentially random solid

solution of composition Ago.5Pdo.5 actually is asso-

ciated with a net reduction in entropy. This some-

what surprising result shows that the unfüled d- shell in this case gives rise to large thermodynamic

effects which are either absent, or are present only

(10)

771 to a very limited extent, in the types. of systems

discussed so far. Presumably, these eff ects are of electronic and magnetic as well as of vibrational

origin. To a first approximation it probably is justified to consider these three effects separately,

and they will of course contribute to all the various

excess thermodynamic properties of the mixture.

However, they are most readily recognized in the entropy of mixing, and we shall confine our dis- cussion here to this quantity.

Numerical estimates of the vibrational and elec- tronic contributions to the high temperature excess entropy of silver-palladium alloys can be made

from the helium-range heat capacity data of Hoare

and Yates [25]. Values of 6 and y obtained in their work are given in figure 12, along with values

FIG. 12. - 03B8 and y for silver-palladium alloys (refs. [25], [26]).

of 0 derived from elastic constant measurements

by Hoare, Matthews and Walling [26].

By use of eqn. (4) above we estimate from these data the vibrational contribution to the excess

entropy for an equi-atomic alloy to fall between

--- 0.8 cal/degree g. atom (from elastic constants)

and 2013’1.2 cal/degree (from heat capacities).

If the electronic heat capacities depend linearly

on the absolute temperature we have for the elec- tronic excess entropies

where Y-e

=

y

-

[XAg YAg + XPd ypd]. On the assumption that this applies up to 1. 000 OK, we estimage Sli for a 50-50 alloy to be about

-

0.8 cal/

degree g. atom. Actually, available high tempe- rature heat capacity data for pure transition metals indicate that the electronic contributions to the heat capacity at elevated temperatures probably

are somewhat smaller than yT. Therefore, this

estimate of Sé may be high numerically.

Finally, we shall consider very briefly the ma- gnetic contributions to the entropy of mixing. In général such contributions will depend on the

nature of the magnetic properties of the compo- nents. In silver-palladium pure silver is diama-

gnetic, palladium is paramagnetic, and the alloys

are paramagnetic for xpd > 0.5 [26]. Clearly the d-electrons, presumably localized on the palladium atoms, are spin-paired at high silver contents, with

an ensuing loss in entropy.

This problem has been discussed in a recent paper by Oriani and Murphy [27]. Adapting an approach advanced by Weiss and Tauer [28], these

authors assume that in the diamagnetic range the

magnetic contribution to the entropy of mixing

may be represented by - xpd R In (03BCpa + 1).

Here 03BCpd is the effective atomic moment of palla-

dium (1.44 Bohr magnetons). Assuming that the alloy is fully diamagnetic at xid == 0.4, they suggests that the magnetic excess entropy in the paramagnetic range will also vary linearly with composition (going to zero at pure palladium). In

this manner they arrive at an estimate of about - 0 . 7 cal/degree g. atom for the magnetic contri-

bution to the excess entropy in Ago.5Pdo.5.

The sum of the quoted vibrational, electronic

and magnetic entropy contributions amounts to

--

2.3 to --2.7 cal/degree g. atom. Àlthough

this result is numerically somewhat larger than the experimental value, it is of comparable magnitude.

Acknowledgements.

-

This work has been sup-

ported by the Office of Naval Research under contract No. Nori-2121 with the University of Chicago.

REFERENCES [1] The integral excess thermodynamic quantities are de-

fined through the relations YE = 0394 Y 2014 0394 Yideal

where 0394 Y is the molar change in the function Y on mixing. Note that

0394Gideal = RT(x In x + (1 2014 x) In (1

2014

x)).

Here x and (1 2014 x) are the mole fractions of the

two components. Similarly, for the entropy of mixing

0394Sideal = 2014 R(x In x + (1 2014 x) In (1 2014 x)),

i.e. the entropy of random mixing. On the other hand we have for enthalpy, internal energy, volume and heat capacity

0394Hideal

=

0394Eideal = 0394Videal

=

0394Cp ideal

=

0.

(11)

772

[2] Detailed discussions of these methods are presented in

the National Physical Laboratory Symposium No. 9,

The Physical Chemistry of Metallic Solutions and Intermetallic Compounds. H. M. S. O., London,

1959.

[3] See e.g. PRIGOGINE (I.), The Molecular Theory of Solutions, North Holland Publishing Company, Amsterdam, 1957.

[4] An authoritative discussion of the lattice theories of solution is given in E. A. Guggenheim’s

"

Mixtures ",

Oxford University Press, Oxford, 1952.

[5] FRIEDEL (J.), Adv. in Physics, 1954, 3, 446.

[6] GORDY (W.) and THOMAS (W. J. O.), J. Chem. Physics, 1956, 24, 439.

[7] WHITE (J. L.), ORR (R. L.) and HULTGREN (R.), Acta Met., 1957, 5, 747.

[8] KÖSTER (W.) and RAUSCHER (W.), Z. Meta!lk., 1948, 39, 111.

[9] ZENER (C.), in Thermodynamics in Physical Metal- lurgy, A. S. M., Cleveland, 1950.

[10] HULTGREN (R.), Private communication.

[11] PINES (B. J.), J. Physics, U. S,B S. R., 1940, 3, 309.

[12] LAWSON (A. W.), J. Chem. Physics, 1947, 15, 831.

[13] KLEPPA (O. J.), Acta Met., 1955, 3, 255.

[14] KLEPPA (O. J.), J. Phys. Chem., 1956, 60, 846.

[15] KLEPPA (O.,J.) and KING (R. C.), To Acta Met (in press).

[16] RAYNOR (G. V.), Progress in Metal Physics, 1949, 1, 1.

[17] KLEPPA (O. J.), KAPLAN (M.) and THALMAYER (C. E.),

J. Phys. Chem., 1961, 65, 843.

[18] In the dilute range, the differential excess quantities

of the solvent are related in a simple manner to the

curvature of the integral, excess quantity. (See also

ref. [21].)

[18A] Note added in proof. Very recent theoretical work

by Blandin and Deplante reported during this col- loquium, represents an improvement on the earlier Friedel theory, and appears to account for this change

in sign.

[19] KLEPPA (O J.) and THALMAYER (C. E.), J. Phys.

Chem., 1959, 63, 1953.

[20] WILSON (E. G.), Private communication.

[21] KLEPPA (O J.), Acta Met., 1960, 8, 435.

[22] KLEPPA (O. J.), Acta Met., 1960, 8, 804.

[23] PRATT (J. N.) Trans. Faraday Soc. 1960, 56, 975.

[24] CHAN (J. P.), ANDERSON (P. D.), ORR (R. L.) and

HULTGREN (R.), 4th Tech. Report, Mineral Research

Laboratory, Berkeley, Calif.,1959.

[25 HOARE (F. E.) and YATES (B.), Proc. Roy. Soc., 1957, A 240, 42.

[26] HOARE (F. E.), MATTHEWS (J. C.) and WALLING (J. C.),

Proc. Roy. Soc., 1953, A 216, 502.

[27] ORIANI (R. A.) and MURPHY (W. K.j, Acta Met., 1962, 10, 879.

[28] WEISS (R. J.) and TAUER (K. J.), J. Phys. Chem.

Solids, 1958, 4, 135.

DIFFUSION STUDIES OF VACANCIES AND IMPURITIES

By DAVID LAZARUS, ,

Department of Physics, University of Illinois, Urbana, Illinois, U. S. A.

Résumé.

2014

Les défauts ponctuels dans les métaux ont d’abord été introduits pour expliquer les phépomènes de diffusion, et le succès des modèles est généralement mesuré par le succès dans la correlation des résultats des mesures de diffusion. Dans cet article, on passe en revue l’utilisa- tion de la diffusion comme instrument d’étude des imperfections, et on cherche à définir les limites de la validité des modèles théoriques à la lumière des études expérimentales de la variation, en fonction de la température, de la pression et de la masse, de la diffusion dans un ensemble de métaux purs et de solutions solides.

Abstract.

2014

Point defects in metals were first introduced to explain diffusional phenomena,

and the success of the models is generally measured by the success in correlating results of diffusion measurements. In this paper, the use of diffusion as a tool to study imperfections will be reviewed, and an attempt made to assess the limits of validity of theoretical models in the light of experi-

mental studies of the temperature, pressure, and mass dependence of diffusion in a variety of

pure metals and solid solutions.

LE JOURNAL DE PHYSIQUE ET LE RADIUM TOME 23, OCTOBRE 1962,

1. Introduction.

-

Diffusional phenomena are

basic to reactions in metallic systems. Diffusion limits the rate of phase transformations, solubility,

creep, grain growth, and recrystallization. Dif-

fusion rates dictate whether a material will be useful in a given environment, as in high tempe- rature reactors, under high flux conditions, or com- pletely useless, as in corrosive atmospheres. Tech-

nical interest in the field, therefore, has always been lzigh.

From a purely scientific viewpoint, the most important problems have been associated with deli-

neating specific mechanisms for diffusion which

permit the observed large flues of matter without perturbing the essentially perfect lattice structure.

Of the many mechanisms suggested to explain dif- fusion, the concept of mobile point defects, parti- cularly interstitials and vacancies, has proven most viable. Since point defects were essentially

invented " to explain diffusion, it is perhaps appropriate to consider how diffusional measu-

rements have been useful as, a tool for studying point defects in various systems,.

In homogeneous systems, the diffusion coefficient,

Références

Documents relatifs

• complete, immersed minimal surfaces M which are strongly symmetric, have finite total curvature, embedded ends, and such that M + is homeomorphic to a planar domain.. There are

A necessary and sufficient condition seems to leave little hope of being able to apply the method to provide an existence theorem for absolutely continuous minimizers under

- We prove the existence of a minimizer for a multidimen- sional variational problem related to the Mumford-Shah approach to computer vision.. Key words :

I will suggest that the minimalist approach (often tacitly) assumes that word meanings have two essential features: first, that words in general make a stable

In the context of solving the lattice equation we have obtained many types of fractional solutions of the nonlinear lattice equation by using fractional transformations method;

On Figure 6, we plotted the variations of the derivative y’(x) of the grain groove profile for Mullins approximation as a function of the distance x from the grain

In the Arctic tundra, snow is believed to protect lemmings from mammalian predators during winter. We hypothesised that 1) snow quality (depth and hardness)

Keywords : Critical and Generalized Korteweg-de Vries equations, wave operator, large data, solitons, linear-non linear interaction... For such a solution, one has conservation of