• Aucun résultat trouvé

Phase transitions of spin polarized 3He : a thermodynamical nuclear orientation technique ?

N/A
N/A
Protected

Academic year: 2021

Partager "Phase transitions of spin polarized 3He : a thermodynamical nuclear orientation technique ?"

Copied!
13
0
0

Texte intégral

(1)

HAL Id: jpa-00209106

https://hal.archives-ouvertes.fr/jpa-00209106

Submitted on 1 Jan 1979

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.

Phase transitions of spin polarized 3He : a thermodynamical nuclear orientation technique ?

B. Castaing, P. Nozières

To cite this version:

B. Castaing, P. Nozières. Phase transitions of spin polarized 3He : a thermodynamical nuclear orienta- tion technique ?. Journal de Physique, 1979, 40 (3), pp.257-268. �10.1051/jphys:01979004003025700�.

�jpa-00209106�

(2)

Phase transitions of spin polarized 3He : a thermodynamical

nuclear orientation technique ?

B. Castaing and P. Nozières

Institut Laue-Langevin, 156X, 38042 Grenoble Cedex, France (Reçu le 24 août 1978, accepté le 24 novembre 1978)

Résumé.

2014

Sous bien des aspects, l’3He complètement polarisé se présente comme un nouveau corps quantique,

différent de l’3He normal. Cet article rappelle qu’on peut l’obtenir de façon purement thermodynamique, en partant du solide fortement polarisé par un champ magnétique à très basse température, à condition de travailler sur une

échelle de temps inférieure à T1, le temps de relaxation de l’aimantation. Les points cruciaux du problème

2014

dépen-

dance de l’énergie du liquide avec l’aimantation, valeur de T1

2014

sont longuement discutés, ainsi que les nouvelles

possibilités offertes. En particulier l’3He complètement polarisé se présente comme la meilleure chance d’observer

une phase de solide lacunaire telle que l’ont discutée A. Andreev et I. M. Lifshitz.

Abstract.

2014

Completely polarized 3He behaves in many respects very differently from normal 3He. In principle

it can be obtained by purely thermodynamical methods, starting from a strongly polarized solid (by a magnetic

field at very low temperature), as long as the overall experimental time is less than T1, the magnetic relaxation time.

We discuss the crucial points of such an experiment

2014

dependence of liquid 3He energy on magnetization, value

of T1

2014

and the new experimental possibilities it opens. In particular completely polarized 3He appears as the best chance to observe a vacancy solid phase, first described by A. Andreev and I. M. Lifshitz.

Classification Physics Abstracts

67.00

1. Introduction.

-

Nuclear spin relaxation in

liquid ’He is a compromise between two mechanisms.

(i) Intrinsic relaxation in the bulk, due to collisions of two particles via the dipole-dipole interactions.

Because of motional narrowing, the corresponding

relaxation rate is inversely proportional to the spin

diffusion coefficient D.

(ii) Extrinsic relaxation on the walls, either on magnetic impurities or indirectly because the walls

are coated with a layer of solid ’He where Tl is short.

Such a mechanism requires diffusion to the wall, and the corresponding rate is

-

D.

Thus, altogether

where A depends on the wall, B on the bulk.

Around 1 K, the intrinsic Tl is of order a few minutes ; unless special precautions are taken, wall relaxation usually predominates. It is however pos- sible to observe the intrinsic Tl by appropriate clean- ing of the walls [1].

Recently, it has been found that appropriate coating

of the walls could reduce A drastically. Barbé, Laloë and Brossel [2] showed that a solid hydrogen coat- ting practically stopped wall relaxation around 1 K

- a result confirmed by recent experiments of

Taber [3]. Neon coating was studied by Chapman

and Bloom, who argue that A is controlled by the

adsorbed layer of 3He. Even more simply preferential adsorption of ’He can provide an insulating layer [5]

which prevents formation of a further solid layer of

’He. From all these results, it appears possible to stop relaxation at the walls - at least at temperatures

> 0.1 K at which the diffusion coefficient is reasonably

small (when D --> oo, wall relaxations would always

take over).

If that is so, Tl is long, in the range 1-10 min. It is then possible to achieve non equilibrium states, with a very large spin polarization, which would be unattai- nable in any reasonable magnetic field. More speci- fically, F. Laloë and C. Lhuillier [6] have proposed to polarize the atoms by optical pumping in the vapour

phase, and to observe the thermodynamical properties

of such an artificially polarized system : transport properties, liquid vapour equilibrium, etc...

The present paper follows the reverse path : if the

relaxation time Tl is long, up and down spins are like

different chemical species, and one may apply to

them the usual concepts of phase equilibrium, frac-

tionate distillation, etc... Hence the possibility of polarizing the spins by purely thermal methods.

Let M be the magnetization, F(M) the free energy

Article published online by EDP Sciences and available at http://dx.doi.org/10.1051/jphys:01979004003025700

(3)

density. The system is subject to an external magnetic

field H,,,,. If M is allowed to relax, thermodynamical equilibrium corresponds to a minimum of the function G

=

F - MHeX,. We may define an internal magnetic

field

M will relax spontaneously until H is equal to Hext.

Let us now assume that the spin relaxation is blocked.

The relevant thermodynamical variable is M instead of Hext8 We may still define a field H through (1)

-

but it bears no relation to the physical applied field Hertz Rather, H represents the difference in chemical

potential between i and 1 spins (’)

(Jl(1

=

DFIDN,, being defined in the absence of Hext).

H is related to M by a magnetic equation of state, H(M). On a time scale « Tl, it may take any value.

Only over times > Ti will it relax to Hertz

Consider now the equilibrium of two phases, 1

and 2. Since the free energy is stationary under exchange of either i or 1 atoms between 1 and 2,

the chemical potentials are the same in the two phases,

and thus Hl

=

H2 = H - irrespective of the applied

field Hext’ which does not influence the equilibrium (its contribution to Jla is the same in both phases). The phase equilibrium is parametrized by (H, T), the magnetizations Mi and M2 being usually different.

In the (M, T) plane, the phase diagram has the usual

liquidus-solidus shape shown on figure 1 - a figure

which clearly displays the different behaviour on

times « or > Tl. Assume that we start with phase (1)

Fig. 1.

-

A typical phase diagram in a magnetic field. At a given temperature, phase (1) with magnetization MA is in equilibrium with phase (2) at magnetization MB. If one transforms (1) into (2) at

constant H, the points A and B do not move (spin relaxation making

for the change in M). If the transformation proceeds at constant M,

AB moves up to CD : the effective magnetic field increases in the transition.

(’) Throughout the paper, we set the magnetic moment of the 3He

atom equal to 1 : H is an energy, and the magnetization

at point A, and transform it to phase (2) at point B :

we thus deplete the magnetization of phase (1).

-

For a slow process (» Tl), relaxation is avai- lable to restore any missing magnetization. Since H

must remain equal to Hext’ MA and MB are unchanged

as the transformation proceeds : the points A and B

do not move.

-

For a fast process ( Tl), M must be conserved :

as the transformation proceeds, AB moves up, until it reaches CD when (1) is completely turned into (2).

The effective field H is thus increased by the phase transformation, well beyond its original equilibrium

value Hext.

As an example, consider solid ’He, at a temperature T much higher than the magnetic ordering tempera-

ture : it obeys a Curie law, and thus M1 = CH/T.

After melting, it becomes a Pauli paramagnet, with

magnetization M2

=

CHITF. Fast melting must con-

serve M. If H was equal to Hext at the start, it becomes

HTF/T at the end

-

a spectacular increase if T is

small !

In this paper, we develop these thermodynamical

arguments in some detail, and we propose experiments

which might allow the production of highly polarized 3He matter. In section 2 we first discuss the properties

of homogeneous liquid or solid 3He. The melting phase diagram of polarized 3He is described in sec-

tion 3. Section 4 deals briefly with other phase equili-

bria (liquid-vapour, or dissolution in 4He). In sec-

tion 5, we try to estimate the spin relaxation time Ti, a

central feature in our problem. Specific experimental proposals are made in section 6, together with specu- lations on new physical effects that could emerge at such high effective fields (such as the vacancy solid

proposed long ago by Andreev and Lifshitz [3]).

2. Properties of the homogeneous phases.

-

Solid 3He cristallizes in the b.c.c. lattice at moderate pres-

sures. Its molar volume is large (vS

=

24.2 cm3 jmole

on the melting curve at T = 0). Its compressibility

is unusually large, Ks - 6 x 10-3 bar-1. It behaves

as a Curie paramagnet down to TN

=

1 mK, at which temperature it turns into an antiferromagnetic state [8- 10]. The exchange energy J responsible for such a Néel

transition is probably due to rotational tunnelling

of a pair of atoms around each other, in the anisotropic

cage formed by the surrounding medium (2). Recent experimental data [9] suggest that a moderate magnetic field,

-

4 kG, can turn the antiferromagnetic phase

into a new state, apparently a weak ferromagnet [11].

At the moment there exists many interpretations

of such a magnetic behaviour [12] : multiple

(2) Such a tunnelling is certainly assisted by a transient deforma- tion of the cage. Note that a modest decrease in molar volume raises the potential barrier : hence a drastic drop in the already small J, leading to the observed large magnetic Gruneisen constant :

d Log J

d Log y

20.

(4)

exchange [13], spin polarons around vacancies [14],

stabilization of the ferromagnetic state by zero point

vacancies [15], etc... Experiment will decide which of these models is correct (actually, they do not exclude

each other, and reality may partake of several). Here

we call attention to the model of Andreev et al. [15],

based on the spontaneous appearance of vacancies in

a magnetic field : this possibility will appear later in

our discussion.

In any case, we shall not be concerned with such low temperatures. We thus disregard completely the magnetic transition. As long as T exceeds a few mK

the magnetization is well represented by a Curie law.

Under typical conditions, T

=

5 mK, H

=

80 kG,

the spin polarization ms reaches 85 %.

Liquid ’He above a few mK is a standard Fermi

liquid, well described by the Landau theory at low enough temperature and magnetic field. The Landau

parameters m*/m, Fi, Fô are well known as a function of pressure [16]. We are especially concerned with the low field spin susceptibility, which we may write as

the spin temperature TF being defined as

The Landau theory applies as long as T, H TF.

The energy TF decreases from 0.54 K at vapour pressure, down to 0.30 K at melting pressure. It cha- racterizes the liquid features of the system : flow of

quasiparticles, etc... (which become harder and harder

as the density increases).

Besides the plastic flow typical of a liquid, ’He also displays acoustic oscillations, in which the atoms vibrate in the cages formed by the instantaneous

configuration of their neighbours. The corresponding

characteristic energy is that of a phonon with atomic

wave length

0D plays the same role as the Debye temperature in the solid

-

and indeed it is quite comparable : longitu-

dinal phonons are much the same in liquid and solid

’He. 0, increases with pressure, from 11 K at vapour pressure up to 25 K at melting pressure.

Traditionally, one considers liquid ’He as a nearly ferromagnetic system, to which one can apply the

usual paramagnon concept [17]. Indeed the exchange

enhancement factor, (1 + Fô)-1, is not small, - 3

throughout the relevant pressure range. The effect

is sizeable

-

yet not dramatic. What is really striking

is the huge ratio OD/TF, which reaches 80 near melting

pressure. From such a vantage point, one tends to

view liquid ’He as nearly solid. Thére exist two

distinct energy scales. On the fast scale, - n/ (JD,

the atoms vibrate in their own cages. The system is locally rigid, like a glass which, despite the lack of a crystal lattice, can resist a stress. On the much slower scale li/TF, the cages start drifting around each other, leading to the plastic deformation expected in a liquid.

According to (4), magnetic properties involve the slow time scale ni TF, That can be understood easily

on physical grounds. (2 XL) - 1 is just the coefficient of m2 in the usual Landau expansion of the ground state

energy

A finite XL is thus tantamount to an m-dependent Eo.

Now the magnetization dependence of Eo can only

arise from the exchange between the atoms of the liquid, which allows them to sense their relative spins (if they cannot exchange, they are de facto, if not de jure, distinguishable). Finally, hard core atoms can only exchange by rotating around each other. In the solid, the surrounding matter can at best give way

slightly, but it cannot participate in the rotation

-

hence the very small tunnelling probability. In the liquid, the vicinity can follow the rotating pair (back- flow), and thus J is much larger. Nevertheless, exchange involves a real motion of the fluid, hence the slow plastic time scale.

To put it shortly, the large measured susceptibility of liquid 3He can arise from two causes :

(i) The liquid is nearly ferromagnetic. Then the

coefficient a in (7) is accidentally very small

-

but the higher terms b, ..., may be much larger.

(ii) The liquid is nearly solid. The whole energy scale for magnetic properties is then - TF, limited

as it is by the slow exchange of two atoms (their rota-

tion being hindered by the nearby matter).

Fig. 2.

-

A sketch of Mi and MI as a function of m in liquid 3He

at zero temperature.

(5)

In order to answer the dilemma we must examine the behaviour of Eo(m) for large m. Let us for instance

plot the chemical potentials ,uT and ,u for up and down

spins as a function of m (Fig. 2). The initial slopes are

simply 1 = 2 T,13. The m2 term in the expansion of J1.

XL

may be found by integrating the Gibbs-Duhem

relationship

(s, v, m are quantities per particle). We thus obtain

near the origin, and for T

=

0 :

When m - 1, J.1t and y, go to well defined limits.

Since H = Mt - J.1 p the magnetization curve m(H)

has the form sketched on figure 3. The central issue is now the following :

-

either the extrapolation of the low m expansion

is qualitatively correct. The relevant energy scale remains - TF for any value of m. Then model (ii) is

correct,

-

or extrapolation to large m grossly underesti-

mates the actual variation of J.1t and y,. Then we have the same situation as in a second order phase transi-

tion. ’He is of type (i) (nearly ferromagnetic).

Fig. 3.

-

The magnetization as a function of H for liquid ’He at

zero temperature (full curve) and at finite T (dotted curve).

We believe that the former case holds : liquid 3He is fundamentally a nearly rigid system, and extrapolat- ing the low field susceptibility to large m is a reaso-

nable rough approximation. As it stands, this is an act

of faith. Only a microscopic calculation of Eo(m)

could provide an unambiguous proof. Such a calcula-

tion is very hard. Low density expansions are of no

avail for ’He : Oné must resort to molecular dyna-

mics [18]. Preliminary calculations by Levesque [19]

seem to support our contention

-

one must await

more detailed results before concluding. As of now,

it is better to rely on experiment : we shall see that the phase diagram of ’He in very high fields should pro- vide an unambiguous answer.

Let us assume that extrapolation of (9) to large m

is qualitatively correct. Then the variation of M, between m

=

0 and m

=

1 is

-

7p/3

=

0.2 K at

low pressure. This is to be compared to po itself

obtained from the heat of evaporation. At p ---> 0, po

= -

2.5 K. The magnetic variation of pi is thus

a small correction : liquid ’He remains firmly bound,

even when its spins are fully polarized. Such a conclu-

sion is markedly different from what one would expect for a free Fermi gas (in which spin alignment multiplies

the Fermi energy by 2213).

A few more words on volume variations of the

liquid. The molar volume vL is quite large, going from

37 cm’ at zero pressure down to 25.4 cm3 at melting

pressure (T

=

0). (Note the large compressibility.)

The change of volume on melting is small :

In large magnetic fields, one must worry about

magnetostriction. From (8), it follows that

The solid is a pure Curie magnet, and thus vs is field

independent. In the liquid, integration of (4) yields at

once

If we extrapolate this formula to m

=

1, we find a variation AVL - 0.1 cm’ near melting : magnetostric-

tion is a small correction (at vapour pressure, OvL is larger, - 0.8 cm’).

3. The liquid-solid equilibrium.

-

In zero field,

the melting curve has the shape shown on figure 4.

It displays a minimum around 0.3 K. As explained long ago by Pomeranchuk this minimum arises from the spin entropy of the solid, ss = Log 2, which

overcomes the entropy of the liquid sL

=

yT at low

Fig. 4.

-

The melting curve in zero field (full curve), in a small

field H TF (dotted curve) and in a large field H > TF (dot-dash

curve).

(6)

enough temperature. The liquid-solid equilibrium

curve obeys the Clausius Clapeyron relationship

In zero field, and neglecting the variation of (VL - vs),

we obtain the melting pressure

Hence the minimum in p(T), which occurs around

T = TF.

In finite fields, the melting pressure decreases. For a

given temperature, the variation Ap(H) is obtained

from(ll)

(Here again, we have neglected magnétostriction ; Ap is 0, since the solid is more susceptible than the liquid.) The integral in (13) is represented graphically

in figure 5. In moderate fields, H « TF the magnetiza-

tion of the liquid is negligible. The drop in melting

pressure is controlled by ms, and it occurs mostly for

T % H - hence the dotted curve in figure 4. For large fields H it TF, ms reaches saturation, and Ap is

instead controlled by mL. dp extends to higher tempe- ratures, and it eventually washes out the minimum in

the melting curve (dash-dot curve in Fig. 4). Indeed, for fully polarized matter, there is no spin entropy

left in the solid

-

hence’no Pomeranchuk minimum.

Fig. 5.

-

The magnetization curves of liquid and solid ’He : (a) at a small finite temperature, T TF, (b) at T

=

0. The lower-

ing of the melting pressure involves the shaded area.

The resulting melting pressure drop is sketched as

a function of H on figure 6, both at T = 0 (full curve)

and at finite T (dotted curve). For H « T, Ap starts quadratically

Fig. 6.

-

The lowering of the melting pressure as a function of field at T

=

0 (full curve) and a finite T (dotted curve).

and it is very small. For H > T, it becomes linear, going to a finite limit when H » TF. The result is especially simple if T = 0. Then

We thus know the beginning of the curve of figure 6.

In estimating the limit Opmax for high fields, we meet again the dilemma of the preceding section. If the fluid

is not close to a ferromagnetic instability, extrapola-

tion of (15) to m

=

1 is not grossly unreasonable. We thus find

In such a case, the melting pressure remains positive

even for an infinite H. The phase diagram of fully polarized ’He has the shape shown on figure 7a.

If instead the large XL is due to the proximity of a ferromagnetic instability, the m(H) curve flattens, and

Fig. 7.

-

Possible phase diagrams for fully polarized 3He :

(a) If I1Pm is small, a liquid phase persists down to T

=

0. (b) If

A/?m is large, a triple point is restored. (c) For incomplete polariza-

tion, case (b) should yield two triple points t, and t2.

(7)

it takes a much larger field to saturate m. The shaded

area in figure 5 increases, and f1p Max is much larger.

A correction by a factor 5 would make the melting

pressure negative. Then the melting curve meets the evaporation curve, and we recover the usual phase diagram with a triple point shown on figure 7b. Note

that in such a case, there must exist a range of magnetic

fields for which the minimum of the melting pressure is

negative, white/?(0) is still positive. The phase diagram

should then display the unusual shape of figure 7c, with two triple points.

It thus appears that a study of the phase diagram

should answer unambiguously the question raised

in section 2

-

namely the behaviour of strongly polarized liquid ’He. Of course, if we could achieve

a field H - TF, there would be no need for a phase diagram

-

we could study the curve ML(H) directly.

Unfortunately, 0.3 K - 4 MG ! The reason for study- ing melting is that one generates effective fields of that order of magnitude by working faster than the relaxa- tion time Tl.

Our argument is illustrated in figure 8, the equivalent

of a liquidus-solidus curve. For a given temperature T,

we plot the magnetizations mL and ms of the two

phases which are in equilibrium at a pressure p.

Figures 8a and 8b correspond respectively to the phase diagrams of figure 7a and 7b. Consider for instance

case (a), and start from a solid at point B. The first drops of liquid correspond to point A. In the absence

of spin relaxation, AB moves down as melting pro- ceeds. Melting is completed when the pressure reaches pc; by that time the effective magnetic field is much

higher than Hext. If we stop at a pressure intermediate between PA and pc, melting is incomplète ; subse- quently, it is triggered by spin relaxation, which reduces H and moves CD back toward AB. (Note

the unusual situation in which melting is completed only thanks to magnetic relaxation.)

Fig. 8.

-

The liquidus-solidus diagram for melting of polarized

’He, as a function of pressure for a given T. Cases (a) and (b) cor- respond to the phase diagrams of figures 7a and 7b.

In case (b), melting cannot be completed if the initial solid magnetization exceeds the critical value mc, however small the pressure. There again, it is only spin relaxation which will allow full melting.

Melting the solid at constant m increases the field H

(the liquid goes from A to C). Conversely, freezing the liquid by a pressure increase decreases H, the solid

going from B to E in figure 8a. Subsequently, spin

relaxation will take the solid back to B - but in a

transient regime, the solid is underpolarized - a

feature which bears on the efficiency of Pomeranchuk

cooling.

All these figures have been plotted at constant T, conceptually the simplest case. Actually, the heat

transfers at such low temperatures are extremely difficult, because of the Kapitza resistance at the walls.

Consequently, any transformation will by necessity

be adiabatic rather than isothermal. Let us assume

that the transformation is slow (and thus reversible)

for every internal variable but m. Then, the picture

should be modified in two ways. First the liquidus and

solidus curves of figure 8 should be drawn at constant

entropy S, instead of constant T. Second, we should

worry about entropy production (and concurrent temperature increase) during the process of spin

relaxation.

Strictly speaking, since there are now two conserved

extensive quantities in the transformation, M and S,

we should plot liquidus-solidus surface in the (p, M, S)

space. For a given initial condition however

-

say

point A in figure 8a

-

the temperature is a well defined function of the percentage of liquid phase

-

and we can thus plot equilibrium curves in the (p, m) plane

-

keeping in mind that these curves would

change for another initial state A. The corresponding diagram is qualitatively the same as in figure 8. Let us

start for instance from a solid at very low temperature To ; its initial entropy is purely magnetic, Ss(ms).

After completion of adiabatic melting, that entropy

must be recovered in the liquid, whose temperature is such that

In zero field, Tl would just be the temperature of the minimum in the melting curve (which precisely cor- responds to a zero entropy change). In a large field, Ss is reduced ; if we neglect the dependence of SL on magnetization, Ti should be smaller - say 0.1 to

0.2 K. The pressure drop pmaX required to complete melting will consequently be higher than in the iso- thermal case (heating and magnetization act in the

same direction). Note that the temperature increase

on adiabatic melting is just the reverse of Pome-

ranchuk cooling by adiabatic freezing.

Let us finally consider the heat dissipation due to spin relaxation after the fast melting at constant m

has been completed. Let S(Je, M) be the entropy at

constant pressure, Je being the enthalpy. When M

varies by ôM, the magnetic work provided to the fluid

is ô W

=

Hext ôM. The entropy production in an

adiabatic process is thus

(8)

Hence the irreversible entropy production due to

relaxation from ml to m2 (per particle) :

From (19), one may infer the temperature change

(note that in a slow process, H remains equal to Hext : then àSi,,

=

0). Actually one may get it directly by an energy conservation argument (treating the

relaxation as a magnetic Joule Kelvin process). The

total magnetic work furnished in the transformation is Hext(m2 - mi) ; in the absence of heat exchange,

one must have

Consider first melting, in which ml » m2’ The last term of (20) is negligible (equivalently, H » Hext in (19)). Since

it follows that (T2 - Tl) is of order Fp - 0.3 K. In

the reverse case of freezing, M2 » ml ; then the last

term of (20) is instead dominant. The enthalpy of the

solid has the form

where OD is the Debye temperature (Jeg is m indepen-

dent if we neglect exchange). In that case,

Thus the final temperature T2 should be

Irreversible behaviour in a magnetic field makes Pomeranchuk cooling very inefficient.

4. Other phase equilibria of polarized ’He.

-

The’

1

Clausius Clapeyron equation (11), can also be applied

to the liquid-gas equilibrium. Since the gas is again

a Curie paramagnet, more susceptible than the liquid,

the vapour curve will move upward - as the magnetic

field is increased. The displacement, however, should

not be spectacular.

First of all, it is known that the critical point is not

sensitive to the statistics. There exist two types of quantum effects :

-

Effects due to the finite spread of thermal wave

packets, which average the effective interaction bet-

ween atoms. Such effects are important

-

but they

are the same for bosons or fermions.

-

Effects due to the symmetrization of the wave function, through the exchange of two atoms. There the statistics is crucial.

The latter corrections are known to be small near

Tc

-

it is just for that reason that de Boer et al. were

so successful in predicting the critical point of 3He

from that of ’He by a mere scaling of the mass. Now,

the spin structure of the fluid can only enter in the exchange correction. If the latter are small, the critical

point should be essentially independent of magnetiza-

tion.

At low temperature, quantum effects become

important in the liquid, and the vapour pressure curve is affected by the magnetic field. Rather than using (11),

we note that the up and down spin vapour pressure of a

perfect gas at temperature T are proportional to

(from which it follows Pt/p¡

=

e HIT : the vapour is a

Curie paramagnet). As soon as H > T, the vapour is

fully polarized and its pressure is controlled by ,ut.

If we extrapolate the low field behaviour (9), the change of y T on going from mL = 0 to mL = 1 is

-

TF/3 1" 0.1 K, much smaller than Mo = - 2.5 K :

the liquid vapour curve is not changed drastically.

If the extrapolation were quite wrong, one could conceive that 1À, would become > 0 when m = 1 : the liquid would cease to be found. (It would take

a minimum pressure to make it stable.) We need

not consider that case : for such a large variation of y,, the melting curve would meet the evaporation

curve well before y, changes sign.

A similar discussion applies to the equilibrium

between polarized ’He and its dilute solution in ’He.

In fact, at low temperature, the 4He substrate acts as a

vacuum, and the dilute solution is like a dense gas.

At low temperatures, both phases are Pauli parama- gnets, but the dilute solution has a lower Fermi tem-

perature and is thus more susceptible. In a moderate magnetic field the magnetizations of the dense and

dilute phases are in the inverse ratio of their spin

temperatures

At constant magnetization, the effective field H should,

1

thus decrease upon dissolution of pure 3He in 4He.

Conversely, precipitating 3He from the dilute solu- tion will increase H. Finally, we note that the zero temperature solubility will increase for polarized 3He (like the vapour pressure). It is a straightforward thermodynamical problem to calculate the change in solubility, etc..., in a magnetic field : we shall not do it

here.

5. The relaxation time Tl.

-

AU this discussion

relies on the assumption that Ti is long enough that

experiments can be carried out on a faster scale. It is

(9)

thus of importance to assess its order of magnitude

-

and also to make sure that high polarizations will not

reduce it unduly.

Experimentally, the observed Tl is a complicated

mixture of extrinsic relaxation at the walls of the sys- tem (or on the powder used to cool the fluid), and of

intrinsic relaxation in the bulk due to dipole-dipole coupling between the nuclear spins. Hence the large

scatter in the literature. Here, we shall assume that extrinsic relaxation can be suppressed completely,

and we shall focus our attention on the intrinsic mechanism.

Since we are mostly concerned with melting, we

first consider the Tl of the liquid (which is also easier to understand). Its low field behaviour [20] is sketched

on figure 9. Ti has a minimum at a temperature

-

0.5 K. The height of the minimum depends on

pressure

-

roughly 5 min. at zero pressure, closer to 2 min. at melting pressure. The data are somewhat uncertain and they represent a lower bound. At lower temperatures, Ti increases as 1/T2. The latter increase is easily understood in the framework of the Landau

theory of Fermi liquids. Relaxation is due to scattering

of two quasiparticles in a layer of width T around

the Fermi level. The phase space for such a collision is of order T2 - hence a transition probability of that

order. The same argument holds for the lifetime r,

of a quasiparticle (controlling transport properties).

The only difference lies in the nature of the interaction.

Fig. 9.

-

A sketch of the intrinsic relaxation time of liquid ’He as

a function of temperature.

The scalar potential between two atoms contributes to Te

-

but not to Ti, since the corresponding colli-

sions are spin conserving. The only way to relax spin

conservation is to invoke the dipole interaction.

Since the phase space is the same for the two pro- cesses, the only difference lies in the matrix elements.

As a result

where Hd is the dipolar interaction and t the scattering

matrix. Let Wd be the precession frequency in the

average dipolar field of two neighbours (- 104) ;

since the interaction is of intermediate strength, ton - TF. Thus

At 0.3 K, te is of order 10- i l s, yielding Tl - 10’

second, in qualitative agreement with the observed value.

The above estimate holds as long as H « T. In the opposite limit H > T, the phase space available to a spin flip collision is much larger. From the energy distribution of up and down spins sketched in figure 10, it is clear that the particles that can take part in spin

relaxation extend over a range H, instead of T. The situation is similar to the lifetime of a quasiparticle

with e » T. In the latter case, detailed calculation shows that the transition probability is proportional

to (n’ T2 + e’). We can thus expect that in a field H » T, the relaxation Tl will be the same as in zero

field at a temperature T’

=

77/Tr. Since the maximum

field is of order TF, T’ should be a few tenths of a K, and the corresponding relaxation time should roughly correspond to the minimum of figure 9, whatever the actual temperature T. We conclude that Tl will remain

very long

-

several minutes

-

even in highly pola-

rized liquid 3He.

Fig. 10. - The energy distribution of up and down spins in

polarized ’He.

We now turn to the solid. For fairly high tempera- tures, T > 1 K, relaxation is supposed to result from

a modulation of the dipolar interaction by the random

motion of vacancies. A recent review of the problem

has been given by Landesman [21]. The corresponding Tl is given by the following formula

Here w is the precession frequency in the applied magnetic field Hext (the only one which is relevant in the energy conservation); iL is the hopping time of

vacancies near a given site (which characterizes the

frequency of the dipolar coupling modulation). M2

is essentially the square of the dipolar precession

frequency wd. (M2

=

5 x 108 s- 2.) iL depends critically

(10)

on the number of vacancies, and thus Tl varies rapidly with temperature. For a given Hext, Tl has a

minimum when W’tL

=

1.2, at which point T1 = W/ M 2’

For an external field of 80 kG, W

=

1.5 x 109, and

the minimum T i is roughly 3 s. Such a lower bound is

certainly very pessimistic, as at low temperature the number of vacancies is very small, making iL very long.

At lower temperature, relaxation occurs through

the exchange between nuclear spins, and Tl is roughly temperature independent. The theory of this regime

is not clearly understood. Experimentally, the corres- ponding Tl increases rapidly with the Larmor fre-

quency w, but decreases when one gets closer to the melting curve. In 25 kG and at melting pressure, Tl

is of order 10-30 s ; extrapolating to 80 kG, one should

find T1 ~ 5 min. It thus appears that at very low temperature the relaxation time of the solid is quite long.

Actually, one should be careful in applying these

results to a solid in the process of melting. Very likely, the solid is full of vacancies, and the former regime of vacancy modulation might extend to much

lower temperature. Thus it is not unconceivable that the minimum value of Ti (a few seconds) might apply

even at T N 0.1 K. That is unlikely, though : if there

are many vacancies, they move quickly, andrl is short, which again makes Tl long.

In any case, the Ti of the solid is not as critical as

that of the liquid, since the solid spends a much

shorter time off equilibrium. Consider an ideal case

in which the liquid solid interface moves at velocity u.

As the solid melts, it enriches in up spins, since the liquid is less polarized. If magnetization could diffuse

infinitely fast, this increase in ms would distribute in the whole sample. With a finite spin diffusion coeffi- cient Ds, the increase in ms remains localized near the interface, in a diffusion length A

=

Dslu. The corres- ponding magnetization profile near the interface is sketched on figure 11. The liquid has the same magne- tization as the bulk solid, the effective magnetic field H building up only in the surface sheath of width Â.

With a typical DS = 3 x 10- 8 cm2/s, and an inter-

face velocity 10 - 3 cm/s, the surface sheath is very

narrow (3 000 Â). The time a given atom of the solid spends in the surface sheath (i.e. off equilibrium) is

With the above values, im

=

0.03 s. The only require-

ment is that the solid Tl be longer thanr,,,

-

a condi-

tion which should be easily satisfied. It may be that

near melting Ds is much larger due to the large number

of vacancies ; then we could pull on u somewhat, and anyhow there is a long way to go before r. reaches Ti.

Finally, we might worry about relaxation right at

the interface, where the crystal structure breaks down.

There is no theory available there

-

but a minimum

condition for such a relaxation would be that the time

an atom spends on the last crystalline plane be larger

Fig. 11.

-

A sketch of the magnetization pattern near a moving liquid solid interface (melting case).

than the precession period in the dipolar field. Irrespec-

tive of any narrowing mechanism, one should at least

have

(where a is the crystal spacing). For u

=

10- 3, this

condition is not met : interface relaxation should not be important.

One may also worry about fluctuations of the

demagnetizing field in homogeneous systems, with solid and liquid interspersed into a snow like structure.

Such a situation certainly occurs in practice. These

fluctuations are certainly sufficient to kill T2. However, in a very large field Hext, they should not hurt Tl very

much, as for macroscopic droplets their frequency

spectrum is not adapted to the Larmor frequency w.

In conclusion, it seems that relaxation in the solid should not be the limiting factor in our problem. Of

course, it is hard to conclude, as we may have forgot-

ten an important effect. Still, one fact is clear : a large applied magnetic field Hext always helps in making Tl long.

6. Possible experiments. - The simplest expe- riment one can think of is simply to polarize liquid 3He by adiabatic melting of the solid. It amounts to working

a Pomeranchuk cell in reverse in a magnetic field. 3He

is solidified under pressure, brought to very, low temperature

-

say 5 mK - and placed in a large magnetic field

-

say H

=

80 kG. We take it for

granted that the solid is initially in thermal equili-

brium - that might in fact take quite a long time.

The pressure is then suddenly dropped below the melting curve (there is no heat exchange because of the Kapitza resistance). If melting is completed on a

time « Tl, the resulting liquid retains the same polarization as the initial solid, 2 TF/3 T bigger than

its equilibrium polarization Meq in the applied field.

The magnetization will subsequently relax to Meq

in a time T,. One should first try to observe that relaxation. That can be done for instance by applying

a 900 NMR pulse, and watching the decay of the spin

Références

Documents relatifs

We propose to explain the structure with transitions from a fourfold valence band state with degeneracy lifted by the magnetic field to a spin-polarized singlet state..

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

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

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

In this report we present results of a theo- retical evaluation of liquid structure func- tion and one-particle momentum distribution for spin-polarized and normal 3 ~ e

Finally, if we assume we have high intensity and can measure polarization, will the beams remain polarized while in collision in a storage ring? The field due to

2014 The hyperfine coupling in metallic cobalt and in cobalt alloys has been determined both by nuclear orientation and by low temperature specific heats.. The results are

Abstract.- We present here some experimental aspects concerning the production of highly polarized liquid 3 ~ e : previous polarization of the solid, relaxation times in