File: SECT05.WM of Tape: Various/ETH/s10-diss
(Source file text) 





			    -  80  -

@ka
	_________________________________________
	|	|	|	|	|	|
	|	|     -2|	|	|    fe	|
	|  Imp.	|A .10	| &DA/A  | &Dn/n  |  &DR   |
	|	| cu	|	|	|    o	|
	|	|	|	|	|	|

	_________________________________________
	|	|	|	|	|	|
	| Pure	| 2.364	|	|	|	|
	|	|	|	|	|	|
	|  Sn	| 2.515	|  .064	|  .002	|  .04	|
	|	|	|	|	|	|
	|  Bi	| 2.632	|  .113	|  .004	|  .06	|
	|	|	|	|	|	|
	|  Pb	| 2.524	|  .068	|  .002	|  .04	|
	|	|	|	|	|	|
	|  Tl	| 2.345	| -.008	| -.000	| -.00	|
	|	|	|	|	|	|
	|  Hg	| 2.248	| -.049	| -.002	| -.03	|
	|	|	|	|	|	|
	|  Cd	| 2.198	| -.070	| -.002	| -.04	|
	|	|	|	|	|	|

	_________________________________________


      Table 5.1: Rigid band changes in 1At.% alloys.

@ke
@ba
The calculation of the FS cross-section A was done with a 4-OPW

program. The changes in cross-section were calculated in the

rigid band approximation, i.e. the number of electrons per atom

was changed by taking into account the valency of the impurity.

The changes in c/a ratio, due to the alloying, were also taken

into account. The results are given in Table 5.1. The fourth

column gives the change &Dn/n = (&Dn\h\-&Dn\e\)/n in carrier

concentration between the second zone hole surface and the

third zone electron surface. The last column shows the

corresponding change of the Hall coefficient in the free

electron approximation, R\o\^fe^ = &Dn/(n^2^.e) [m^3^/As]. 
@be




			    -  69  -

@ka
V. HALL EFFECT MEASUREMENTS IN INDIUM.
______________________________________

1).Introduction.
________________
@ke
@ba
Our measurements of the Hall-effect in Indium were initiated by

the work of Cooper;Cotti;Rasmussen(1965). These authors studied

the influence of the size-effect on the low-field Hall effect

in Indium films. They obtained an important contribution of the

size-effect , i.e. the sign of the Hall effect could be made to

change from the usual positive sign to a negative sign in

sufficiently thin samples.
@be
@ba
These results were interpreted with the aid of anisotropic

relaxation times in connection with a multi-band scheme for a

polyvalent metal such as Indium. These strong lifetime effects

showed that it would be interesting to look for other possible

scattering mechanisms which would ,presumably, show analogous

effects. We decided to look at the effect of alloying, cold

working and, to some extent, of the temperature in both the

Hall effect and the magnetoresistance of rolled polycrystalline

specimens of Indium.
@be




			    -  70  -

@ka
2). Experimental setup.
_______________________

a). Cryogenics.
_______________
@ke
@ba
All measurements of the Hall effect were made in a standard

glass Dewar which fitted in a transverse electromagnet. Most

measurements were made in liquid Helium and the temperature was

determined with the aid of the vapour pressure. Some other

selected temperatures were also needed for the temperature

dependence measurements.The following temperatures were used:

20K, 77K, 210-300K. The first two temperatures were at the

normal pressure boiling points of liquid hydrogen and of liquid

nitrogen. The temperatures around 0C were made with a Freon 114

cooling setup. This cooling unit consisted of a standard refri-

gerator compressor followed by a Joule-Thomson expansion valve

and a vacuum pump which adjusted the pressure after the ex-

pansion. The magnetoresistance measurements were only done in

the liquid Helium temperature range.
@be
@ka
b). Magnetic Fields.
____________________
@ke
@ba
The magnet used for the Hall measurements was a conventional

iron core electromagnet with a maximal field of .8T in a gap of

100mm. The field-current relationship was calibrated with a

flux integrating method and the magnet was always saturated

prior to the runs so that the remanent field was the same. The

magnetoresistance measurements were made in a split-coil

Helmholtz superconducting magnet with a maximum field of 1.5T.

The same calibration method was used.
@be




			    -  71  -

@ka
c) Measurements.
________________
@ke
@ba
The Hall effect was measured with a DC technique. The sample

had three electrodes for measuring the Hall voltage; two of

these electrodes were connected to a potentiometer of very low

resistance which was situated at the same temperature as the

sample. This potentiometer was used to compensate for the vol-

tage due to the misalignement of the Hall-electrodes and was

made out of a material similar to the sample, thus the effect

of magnetoresistance on the offset was largely compensated.
@be
@ba
This compensated voltage was then fed to a null-compensating

bridge. The null detection was done with a "Tinsley" galvano-

meter amplifier and a final standard galvanometer. The sample

and compensation currents could be switched simultaneously in

order to obtain a better null criterion.
@be
@ba
Any remaining misalignement voltage was eliminated by reversing

the magnetic field for every data-point. Our setup had an ul-

timate sensitivity of approximately 2nV in the range of sample

currents used (.5-5A). The resistivities were measured with the

same circuit connected to two other electrodes on the sample.
@be
@ba
The incremental magnetoresistance was measured by having two

wires of the same material constitute the two arms of a

Wheatstone bridge, one arm being in the magnetic field. The

off-null voltage of the bridge, which was proportional to

&D&r/&r, was directly fed to an X-Y recorder with a sensitivity

of 50&mV.
@be




			    -  72  -

@ka
3). Alloys and samples.
_______________________
@ke
@ba
The alloys were prepared from 99.999% pure Indium of Johnson

and Matthey Co. and 99.99% pure solutes of different sources.

They were molten and thoroughly mixed in sealed glass ampoules,

usually under vacuum, but, in the case of solutes with high

vapour pressure under helium gas. We first prepared master

alloys of a relatively high concentration (~5At.%), which were

then further diluted to the required atomic percent concen-

tration. The characteristics of the different alloys were

already discussed in sect.II: we did not make any alloys with

Mg,Zn,Li for the reasons indicated therein. The ingots so

obtained were then tempered at 100^o^C for several days and at

room temperature for several weeks.
@be
@ba
As we wanted to measure only polycrystalline substances, the

Hall samples were rolled to the required thickness (.1-.25mm as

dictated by size effect considerations) and the magnetoresis-

tance samples were extruded in the form of wires of .6mm. The

rolled Hall film was then cut to size on a template and its

thickness determined from its weight. The dimensions were: 1cm

wide by 4cm long with electrodes of .5mm by 5mm projecting from

the sample. This shape should reduce the systematic errors due

to the shape of the film to less than 1%.(see Hurd(1972) p.184)

For the measurements of cold working, the sample was twisted

through a three pronged fork before being cooled down. During

the experiment this fork was then moved from the top to the

bottom of the sample and in this way it was deformed twice

through an angle of approximately 30^o^.
@be




			    -  73  -

@ka
4). Results.
____________
@ke
@ba
We will now give the results of our measurements and give a

short discussion of several special details. The results can be

divided in three different topics.
@be
@ka
a) Residual resistivity results.
________________________________
@ke
@ba
In Fig 5.1 we give the residual resistivity (RRR = &r\4.2\/

&r\300\) of the indium films of .1mm thickness which we

investigated. Also given are the size effect results (Samples 5

and 6) of Cooper et al.(1965) and the results on cold worked

(CW) samples. The scales given for the last two effects are

arbitrary and have no connection with the At.% impurity scale.

The cold work RRR's make a jump at the first deformation and

then only increase slowly with further cold work cycles; this

probably comes from the fact that in the first deformation a

crystallite size corresponding to the 30^o^ bending is

established and this size is not changed much thereafter.
@be
@ba
The value indicated for 99.999% pure indium is already size-

effect limited at the thickness of .1mm, because, as Cooper et

al.(1965) have shown, the material used had a bulk mean free

path of ~.27mm. We used an extra thick sample (.25mm) for the

Hall effect result on pure indium; this film had a RRR of

9.10^-5^.
@be




			    -  74  -

@ba
As can be seen from Fig. 5.1, there were no significant devia-

tions of a linear relationship between the At.% concentration

and the RRR, up to the highest concentration measured (1-

2At.%). The numerical values will be discussed further in a

later paragraph.
@be
@fa













		FIGURE 5.1

















      Figure 5.1: Residual resistivities of Indium alloys
                  ( the results on cold-working and size-
                    effect are also indicated )

@fe




			    -  75  -

@ka
b) Magnetoresistance.
_____________________
@ke
@ba
Some representative results are shown in Fig. 5.2 , where

&D&r/&r\o\ is plotted against the reduced field H\red\ =

B&r\&J\/&r\o\. These results were used, along with the

transverse Hall voltages, to compute the Hall angle &j. They

also show the deviations from Kohler's rule which are to be

expected if the relaxation time distribution is changing from

sample to sample. (&J\D\=100K, &r\&J\=2.91 &m&O-cm)
@be
@fa













		FIGURE 5.2















        Figure 5.2: Magnetoresistance in Indium alloys.
@fe




			    -  76  -

@fa













		FIGURE 5.3






















           Figure 5.3: Hall angle in Indium alloys.
@fe
@ka
c) Hall effect.
_______________
@ke
@ba
In Fig. 5.3 we have plotted tan(Hall-angle)/ Red.Field in

function of the reduced field instead of the Hall coefficient

R\o\, for several reasons: The Hall angle is defined by,
@be
@ea
                      j\y\     E\Hall\         E\Hall\(B)
(5.1)    tan( &j )  =  --  =  -----  =  --------------
                      j\x\     E\long\     &r\long\(B) j\long\
@ee




			    -  77  -

@ba
This angle is the only quantity which can be directly compared

to calculations of j\y\ and j\x\. (j\y\ and j\x\ are not acces-

sible in the usual experimental situations.) In low fields, the

quantity plotted in Fig. 5.3 is directly related to the low

field Hall coefficient R\o\:
@be
@ea
         tan(&j)  &r\o\      E\Hall\(B)      R\o\
(5.2)    ----------  =  ----------  =  --
            B &r\&J\        j\long\ B &r\&J\     &r\&J\
@ee
@ba
This quantity is also known under the name Hall mobility &m\H\.

In intermediate field ranges it seems to change in an almost

linear way and this fact will allow us to extrapolate to zero

field. The data shows a departure from linearity at low fields;

this is partly a real effect (one expects the Hall mobility to

saturate eventually) and partly is caused by the loss of

precision at low fields. The high field value of pure indium is

not shown, but we obtained values of 14.9-15.2.10^-11^m^3^/As.

This saturation value agrees well with other results (Hurd

(1972) p.312; the theoretical value for one electron is

15.9.10^-11^m^3^/As). One can also see that Kohler's rule is

not obeyed. This is to be expected if the relaxation times are

changing.
@be
@ba
The zero-field extrapolated values of R\o\ are given in Fig.

5.4 in function of the RRR's. These values were obtained by

extrapolating the Hall angle to zero field , disregarding the

small deviations from linearity at low fields. The two size-

effect points are from Cooper et al.(1965).
@be




			    -  78  -

@fa













		FIGURE 5.4
















          Figure 5.4: Extrapolated low field R\o\'s.

@fe
@ba
The points can be roughly separated in three groups: Ga, Cd,

Tl, Size saturate at about -5.10^-11^, Pb at -2.5.10^-11^ and

Hg, Sn, Bi, cold-work stay at +1.10^-11^m^3^/As. In Fig. 5.5 we

show some selected data on the temperature dependence of the

Hall coefficient. For comparison the values of Al are also

shown. (from Shiozaki; Sato(1967))
@be
@ba
The symbols appearing in Fig.5.4 and some of the symbols in

Fig.5.5 are defined in Fig.5.1.
@be




			    -  79  -

@fa













		FIGURE 5.5











         Figure 5.5: Temperature dependence of R\o\.

                     (the numbers in this figure refer to the

                      At% concentration of the impurity.)
@fe
@ka
5). Rigid band effects on the Hall effect.
__________________________________________
@ke
@ba
At the values quoted, the concentration of the impurities is

always less than 1At.%. One might ask oneself if these concen-

trations are big enough to have a direct influence on the

electron concentration and thereby change the Hall coefficient.

In order to check this influence we have computed changes in

the [110] cross-section A of the third zone for additions of

1At.% impurities. This is shown in Table 5.1.
@be




			    -  80  -

@ka
	_________________________________________
	|	|	|	|	|	|
	|	|     -2|	|	|    fe	|
	|  Imp.	|A .10	| &DA/A  | &Dn/n  |  &DR   |
	|	| cu	|	|	|    o	|
	|	|	|	|	|	|

	_________________________________________
	|	|	|	|	|	|
	| Pure	| 2.364	|	|	|	|
	|	|	|	|	|	|
	|  Sn	| 2.515	|  .064	|  .002	|  .04	|
	|	|	|	|	|	|
	|  Bi	| 2.632	|  .113	|  .004	|  .06	|
	|	|	|	|	|	|
	|  Pb	| 2.524	|  .068	|  .002	|  .04	|
	|	|	|	|	|	|
	|  Tl	| 2.345	| -.008	| -.000	| -.00	|
	|	|	|	|	|	|
	|  Hg	| 2.248	| -.049	| -.002	| -.03	|
	|	|	|	|	|	|
	|  Cd	| 2.198	| -.070	| -.002	| -.04	|
	|	|	|	|	|	|

	_________________________________________


      Table 5.1: Rigid band changes in 1At.% alloys.

@ke
@ba
The calculation of the FS cross-section A was done with a 4-OPW

program. The changes in cross-section were calculated in the

rigid band approximation, i.e. the number of electrons per atom

was changed by taking into account the valency of the impurity.

The changes in c/a ratio, due to the alloying, were also taken

into account. The results are given in Table 5.1. The fourth

column gives the change &Dn/n = (&Dn\h\-&Dn\e\)/n in carrier

concentration between the second zone hole surface and the

third zone electron surface. The last column shows the

corresponding change of the Hall coefficient in the free

electron approximation, R\o\^fe^ = &Dn/(n^2^.e) [m^3^/As]. 
@be




			    -  81  -

@ba
We see that these rigid band changes are very small and even if

one assumes that this rough evaluation is wrong by a factor 10,

our observed R\o\'s cannot be explained by the electron-hole

changes. (Even the systematics is wrong!). We are thus led to

conclude that anisotropic relaxation times must be responsible

for the changes in R\o\ on alloying.
@be




			    -  82  -

@ka
6). Discussion.
_______________

a). Resistivity:
________________
@ke
@ba
We will now discuss our results in the light of the results

obtained earlier in sect. IV. Eq.4.3 obtained there, was an

expression for the inverse relaxation time representing a

lifetime broadening. This means that all scattering events are

weighted equally and this expression is appropriate for

calculating Dingle temperatures (see sect.VI). In the context

of resistivity calculations one must start by solving the

Boltzmann equation. As shown in sect.IV, eq.4.6 will now take

the form: (see also Sorbello(1974)
@be
@ea
           1                v\z\(k')&t\z\(k')
(5.3)    -----  =  $I  [1 -  ------------] P\kk'\ dk'
         &t\z\(k)              v\z\(k) &t\z\(k)
@ee
@ba
In the limit of a spherical Fermi surface and isotropic

scattering this leads to (see e.g. Ziman(1964)) an expression

containing only the scattering potential and a term which takes

into account the loss of velocity in the direction of current

flow. By inserting eq. 3.31 in eq. 5.3 one obtains,
@be
@ea
                    1
         1     6&pZ
(5.4)    -  =  ---  $I  |<k+q|V|k>|^2^ x^3^ dx
         &t      E\F\
                    0
@ee




			    -  83  -

@ba
here x=q/2k\F\ and all quantities are in atomic units. One can

also express the resistivity directly with the aid of the

formula, &r=m\e\/(ne^2^&t), and then the constant in eq. 5.4

will be: 6&p&O\o\/E\F\ and the resistivity &r will be expressed

in atomic units (as defined in sect. III.1).
@be
@ba
This same expression for the resistivity can be written in

function of the phase shifts we calculated in sect.III and one

can show that the correct formula is:
@be
@ea
                      l=4
                4 &p
(5.5)   &r^au^  =  ---   $S    l sin^2^ (&d\l-1\ - &d\l\)
                Zk\F\
                      l=1
@ee
@ba
In calculating the resistivities with these formulas and our

scattering potentials (or phase shifts) we first got values

which were sensibly lower than the experimental values of the

indium alloys. This can be understood by looking at the scatte-

ring potentials (see Fig.3.3) and noting that they are usually

peaked near x=0. It is just this region which is ignored by the

x^3^ term in eq.5.4. Because these formulas were derived for a

free electron surface, it is not astonishing that they do not

work in a metal like indium which has a rather strongly dis-

torted Fermi surface.
@be




			    -  84  -

@ba
As we noted in sect.IV we should do the full calculation in the

framework of an OPW band structure. This is a rather costly and

complicated calculation and we have taken the approximate way

out. Because the trivalent metals have a Fermi surface con-

sisting of several bands, the scattering angle necessary to

annihilate practically all the current contribution of a state

&f\k\ is quite small (approximately 30^o^ i.e. 1/3 of the angle

in the free electron case). This fact led us to separate the

Fermi surface in 3 bands each with a relaxation time given by

eq.4.3. Thus, we say that the intraband scattering events are

angle independent and that interband scattering is neglected.
@be
@ba
In order to evaluate the resistivity we recast the eq.4.11

expression for the conductivity,
@be
@ea
                  1
(5.6)   &s\xx\  =  ----  $I   &t v dS
                12&p^3^
@ee                    FS
@ba
this expression is in atomic units and if one inserts compu-

tational units under the integral and defines the relaxation

time &t as (see eq. 4.3):
@be
@ea
                     l\m\
           1
(5.7)    -----  =   $S    (2l+1) F\l\(k) sin^2^ (&d\l\)
         &t\o\(k)
                    l=0
@ee




			    -  85  -

@ba
we get for our 3 band model,
@be
@ea
                 24&p
(5.8)    &r\xx\  =  ---  1/ [ &t\1\v\1\S\1\ + &t\2\v\2\S\2\ + &t\3\v\3\S\3\ ]
                  k\F\
@ee
@ba
We use the values for v and S in the different bands as given

in the last paragraph in Table 4.2. If we insert these band

parameters in eq. 5.8 and use the &t's obtained from eq. 5.7,

we obtain theoretical values for the resistivity of 1At%

solutions of the different impurities under consideration. We

compare these theoretical values with our experimental results

in Table 5.2.
@be
@ba
In the same framework, we have also calculated the resistivi-

ties of our selection of impurities in aluminum. These results

are compared, in Table 5.3, to experimental data taken from

Blatt (1968). For a further discussion of the resistivities of

the samples used in our de Haas- van Alphen experiments, see

Wejgaard (1975).
@be




			    -  86  -

@ka
		_________________________________
		|	|	|	|	|
		|	|   	|   fe	|   mb	|
	        | Solute|  &r    |  &r    |  &r    |
	        |       |   exp |   calc|   calc|
		|	|	|	|	|

		_________________________________
		|	|	|	|	|
		|  Hg	|  .17	|  .08	|  .25	|
		|	|	|	|	|
		|  Cd	|  .33	|  .08	|  .35	|
		|	|	|	|	|
		|  Tl	|  .24	|  .04	|  .05	|
		|	|	|	|	|
		|  Ga	|  .20	|  .07	|  .13	|
		|	|	|	|	|
		|  Pb	|  .57	|  .16	|  .43	|
		|	|	|	|	|
		|  Sn	|  .50	|  .11	|  .53	|
		|	|	|	|	|
		|  Bi	| 1.38	|  .49	| 1.59	|
		|	|	|	|	|

		_________________________________


     Table 5.2: Resistivities for Indium alloys (&m&O-cm/At%).

                &r\exp\ is taken from Fig. 5.1 with

                &r\300\= 9.01 &m&O-cm       

                &r^fe^\calc\ is from eq. 5.4 or 5.5.

                &r^mb^\calc\ is from the multi-band eq. 5.8.
@ke
@ba
We see that the agreement is encouraging, in a simple model

like ours, for the charged impurities. The homovalent solutes

Tl and Ga do not give the correct values of &r. There can be

different reasons for this; the scattering potential is mainly

given by the distortions of the lattice and, as we discussed in

paragraph 3.3.b, this is the most doubtful part of the calcu-

lation. Another reason might be that these scatterers have a

strong s-like character and that the 3-band model breaks down.

We have tried to use the Ashcroft potential of paragraph 3.2.b

for these solutes, but the results were not sensibly different

and we shall not give these results here.
@be




			    -  87  -

@ka
		________________________________
		|	|	|	|	|
		|	|	|   fe	|   mb	|
		| Solute|  &r    |  &r    |  &r    |
		|	|   exp	|   calc|   calc|
		|	|	|	|	|

		_________________________________
		|	|	|	|	|
		|   Cd	|  .50	|  .07	|  .14	|
		|	|	|	|	|
		|   Zn	|  .22	|  .08	|  .27	|
		|	|	|	|	|
		|   Mg	|  .45	|  .15	|  .55	|
		|	|	|	|	|
		|   Ga	|  .30	|  .05	|  .05	|
		|	|	|	|	|
		|   Ge	|  .80	|  .24	|  .43	|
		|	|	|	|	|
		|   Si	|  .70	|  .22	|  .59	|
		|	|	|	|	|

		_________________________________


        Table 5.3: Resistivities for aluminum alloys.

                   in units of &m&O-cm/At%.

                   The experimental data is from Blatt(1968)

@ke
@ba
The same general remarks, as given for Indium, apply to the

results of the Aluminum impurities. Here, also, the worst

agreement is observed for the homovalent solutes.
@be




			    -  88  -

@ka
b). Low field Hall coefficient.
_______________________________
@ke
@ba
In sect.IV we obtained the expressions 4.11-13 for the con-

ductivity tensor. For a typical transverse arrangement we can

take H along the z-axis and put E\z\=0, J\y\= J\z\= 0 and we

get in second order in H:
@be
@ea
(5.9)   &s\xx\  =  &s^o^\xx\ + &s^2^\xx\ H^2^,  &s\xy\  =  &s^1^\xy\ H,  &s\zz\  =  0
@ee
@ba
In order to compare these quantities with experiment one has to

obtain the corresponding resistivity expressions by calculating

the expression &r\ce\= [ J.E ] / [ J^2^ ] and we get,
@be
@ea
                     &s\xy\         &s^1^\xy\ H                   &s^1^\xy\
(5.10)  tan( &j )  =  ---  =  ---------------- ,  R\o\  =  -------
                     &s\xx\     &s^o^\xx\  +  &s^2^\xx\ H^2^           [&s^o^\xx\]^2^
@ee
@ba
for the Hall angle and coefficient respectivily. It should be

noted that the sign of &s^2^\xx\ is implicitly negative due to

the twofold differentiation of a trigonometric function in

eq.4.13. Thus, one would expect the quantity tan(&j)/H\red\,

plotted in Fig.5.3, to vary as R\o\(1+|&s^2^\xx\|.H^2^+...) in

the limit of low fields. This seems not to be the case, unless

the quadratic behaviour is located at very low fields.
@be
@ba
The expression tan(&j) = &o\H\.&t for the free-electron Hall

angle, can be used along with eq.5.2 to give the relation

&o\H\&t = R\fe\.H\red\/&r\&J\, where R\fe\ is the free electron

Hall constant and H\red\ our reduced field variable.
@be




			    -  89  -

@ba
Numerically, this gives in the case of indium, &o\H\&t =~

2.10^-3^.H\red\. This means that in a reduced field H\red\ =~

500 T we expect the transition from the low- to high-field

behaviour of the free electron parts of the metal in question.

This was the case for our high field values, where we observed

a complete saturation of R\o\ at H\red\ = 1000 T.
@be
@ba
We see from our experimental curves in Fig.5.3, that if there

is a quadratic behaviour at all, it must be limited to field

values H\red\ <^~^ 5 T. This could be explained by assigning an

effective lowest cyclotron mass m* =~ .01 or (see sect.IV.2.a)

a curvature radius of the smallest part of the FS of &r <^~^

&r\fe\/100. Such a small curvature radius is hard to understand

in the light of the results obtained in Table 4.2. On the other

hand it must be remembered that the low field condition is

really given by &o\H\&t << m*/m\e\, which might explain the

discrepancy of a factor of 10. 
@be
@ba
For the following numerical discussion we limit ourselves to

our simple model of the Fermi surface. If we consider each band

to be of a cylindrical shape, it is easy to see that the cur-

vature radii &r drop out of eq.4.12 and the two first terms

reduce to a sin(&f)^2^, which has an average value of 1/2. For

our 3 bands we get the following simple expression,
@be




			    -  90  -

@ea
                         i=3
                    e^3^
(5.11)   &s^1^\xy\  =  -----   $S  <&t\i\^2^> <v\pi\^2^>  &f\i\ h\i\
                  8&p^3^h^2^
                         i=1
@ee
@ba
If we now use eq.5.10 to express the Hall coefficient one gets

in computational units:
@be
@ea
                  3                         3
             9&O\o\
(5.12)  R\o\ = ---  $S  [ &t\i\^2^ v\pi\^2^ &f\i\ h\i\ ] / [ $S  ( &t\i\ v\pi\ S\i\) ]^2^
              e
                  1                         1
@ee
@ba
Here we have dropped the averages and the constant in front of

the sum is = 1.43.10^-9^ for In and = 0.92.10^-9^ for Al, when

R\o\ is expressed in [m^3^/As].
@be
@ba
We can now insert the FS model parameters obtained in Table

4.2. The relaxation times used are given by the dimensionless

formula 5.7 with the <F\l\> characteristic of each band taken

from Table 4.1. We give the results for Indium and Aluminum in

Table 5.4. The experimental values of the low field Hall coef-

ficient are taken from Fig.5.4 at the RRR value of ~7.10^-3^

where the impurities dominate the phonon scattering. The few

experimental values for Aluminum were taken from Boening et

al.(1975), for comparison purposes. 
@be
@ba
At a first inspection of the data in Table 5.4, we see that the

order of magnitude of the predicted R\o\ is correct. We noticed

during the computation of these values that the results are

extremely sensitive to changes in the scattering potential.
@be




			    -  91  -

@ka
	_________________________________________
	|	|	|	|	|	|
	| 	|    Indium	|   Aluminum	|
	| Solute|	|	|	|	|
        |       | R\calc\ | R\exp\  | R\calc\ | R\exp\  |
        |       |       |       |       |       |
	|	|	|	|	|	|

	_________________________________________
	|	|	|	|	|	|
	|  Li	| -2.3	|  ---	|  ---	|  ---	|
	|	|	|	|	|	|
	|  Hg	| +3.1	| +0.5	|  ---	|  ---	|
	|	|	|	|	|	|
	|  Cd	| -1.1	| -5.0	| +0.9	|  ---	|
	|	|	|	|	|	|
	|  Zn	| -0.6	|  ---	| -2.5	| -1.2	|
	|	|	|	|	|	|
	|  Mg	| -3.0	|  ---	| +1.7	| -0.5	|
	|	|	|	|	|	|
	|  Tl	| +0.5	| -5.0	|  ---	|  ---	|
	|	|	|	|	|	|
	|  Ga	| -2.4	| -5.0	| +5.9	|  ---	|
	|	|	|	|	|	|
	|  Pb	| -3.6	| -2.5	|  ---	|  ---	|
	|	|	|	|	|	|
	|  Sn	| -0.7	| +0.7	|  ---	|  ---	|
	|	|	|	|	|	|
	|  Ge	|  ---	|  ---	| +6.9	| +1.8	|
	|	|	|	|	|	|
	|  Si	|  ---	|  ---	| +4.2	|  ---	|
	|	|	|	|	|	|
	|  Bi	| -1.7	| +0.7	|  ---	|  ---	|
	|	|	|	|	|	|

	_________________________________________


       Table 5.4: Comparison of calculated and experimental

                  values of R\o\ in Indium and Aluminum.

                  ( in units of 10^-11^ m^3^/As)
@ke
@ba
Some values were calculated with the aid of the simple Ashcroft

potential (see sect.III) and, although the scattering poten-

tials look quite similar, quite substantial changes in the Hall

coefficient resulted. We do not include these results here

because they would only confuse the issue and, surely, the Shaw

potentials are closer to the reality.
@be




			    -  92  -

@ba
On the average the R\o\ for In are more negative than those for

Al. This is just the opposite behaviour to the one one would

expect by looking only at the Fermi surfaces. Remember that the

third zone of In should contribute much less to the R\o\ than

the Al "monster". It seems that the influence of the strong

differences in s-p character, reflected in the values of F\l\,

is responsible for this abnormal behaviour.
@be
@ba
The calculated values for Tl and Bi in In, have the worst fit

to the experimental values. Tl in In was also the worst can-

didate for the resistivity calculation and it seems plausible

to assume that some aspect of the potential calculation is in

error; most probably the distortion effect, because homovalent

impurities are prone to these kind of errors. Bi has a valency

difference of 2 with In and it's s-phase shift is already quite

large, which might partially invalidate the Born-approximation.
@be
@ba
The agreement for the Al solutes is not too good, the calcu-

lated values are too large in absolute value, whereas the trend

in polarity seems to be better. It is possible that due to the

limited solubility in Al the experimental values did not attain

their saturation level and that the phonon contribution to the

Hall coefficient is still predominant.
@be




			    -  93  -

@ka
c). Temperature dependence of the Hall coefficient.
___________________________________________________
@ke
@ba
The effect of temperature on the Hall effect in selected alloys

and pure indium and aluminum was shown in Fig.5.5 and we shall

now discuss these results here.
@be
@ba
Let us first note that the Hall coefficient of In has a rather

strong variation between it's Debye temmperature &J\D\ of 100K

and the melting point of 157C, where it even makes an abrupt

jump to the value of three free electrons. It is very likely

that this change comes from the change in c/a ratio as shown in

Fig.2.4. We saw there that the c/a ratio stayed approximately

constant up to the temperature of 150K. At higher temperatures

this ratio diminishes quite rapidly toward the cubic ratio of

1. This temperature range coincides with the range where R\o\

is dropping rapidly. If we assume that the final jump in R\o\

corresponds to the change in c/a ratio from the extrapolated

value of 1.069 to 1, than, assuming the change in R\o\ is

linear in c/a-1, we can assign a value of ~-.7.10^-11^ to the

high temperature R\o\, had there been no change in c/a ratio.

Then we can see that the temperature dependence of Al and In

are very similar, the only difference being a constant shift

toward the positive side of ~5.10^-11^m^3^/As for In.
@be




			    -  94  -

@ba
Now, in a more quantitative description of the influence of the

phonons on the value of R\o\, we must emphasize a very prac-

tical property concerning the calculation of R\o\. As we saw in

eq.5.12 the expression giving R\o\ contains the factors &t^2^

both in the numerator and the denominator. This means that the

calculation is reduced to purely geometric problem weighted

with the relative influence of the relaxation time. This is

very important in discussing temperature effects, because, as

is well known, the assignement of absolute values to e.g.

resistivity calculations, is quite difficult.
@be
@ba
The most difficult part in the discussion of the influence of

phonon scattering is the fact that, in contrast to impurity

scattering, the phonon carries a momentum q of it's own which

leads to so called "Umklapp" scattering. This Umklapp scat-

tering only occurs if a reciprocal lattice vector g can be

found so that k\initial\ + q = g + k\final\. This condition is

always satisfied on those parts of the FS where a Brillouin

zone intersects it, but when these parts are reassembled in a

reduced zone scheme as shown for instance in Fig.4.1, it can be

seen that all these Umklapp processes are still intraband

effects in our definition of these bands and in the limit of

low temperatures.
@be




			    -  95  -

@ba
The limit of low temperatures is given by the fact that all

phonon q's should be smaller than q\min\ as defined in Fig.4.1.

Assuming a Debye model we find that q\D\ = 1 c.u. and q\min\/

q\D\=0.05. This corresponds to a temperature of .05 &J\D\ = 5K

for both Al and In (compare Figs. 2.2 and 2.6). Thus, in our

liquid helium temperature range, it is a good approximation to

ignore Umklapp processes and treat the scattering as a normal

quasi-elastic event. In order to calculate the corresponding

phase shifts one still has to introduce the Bose statistics of

the phonon distribution. This leads to a replacement of the

usual matrix element <k+q|w|k> by an expression proportional

to,
@be
@ea
                            q/c(T)                   T
(5.13)      <k+q|w|k> ----------------- ,  c(T) = q\D\ -
                      exp[ q/c(T) ] - 1              &J\D\
@ee
@ba
here c(T) is a cutoff variable which has the value of 20 in our

low temperature range. We see that this effective matrix

element is very sharply peaked at q=0 and has a width of of

approximately .04 in q/(2k\F\). The formula 3.45, giving the

phase shifts, being in some sense a Fourier transform in sphe-

rical coordinates, we find that the phase shifts resulting are

all equal up to at least l=4.
@be
@ba
In the opposite case of very high temperatures the Bose factor

in 5.13 is reduced to a constant and the scattering of the

phonons is the same as those produced by vacancies of the pure

metal. We assume that the neglect of Umklapp processes is not

too serious because all the normal process q vectors are

already interconnecting the totality of the FS.
@be




			    -  96  -

@ba
Invoking the argument given on the top of the preceding page we

use the uniform value of 1 for the phase shifts at low tem-

peratures and the values given under Al in Table 3.8 for the

high temperatures. We can now calculate the R\o\'s in Al and In

and compare them with experiment in Table 5.5.
@be
@ka
		_________________________________
		|		|		|
		|   Low temp.	|   High temp.	|
		|		|		|
		|	|	|	|	|
                | R\calc\ | R\exp\  | R\calc\ | R\exp\  |
                |       |       |       |       |
		|	|	|	|	|

		_________________________________
		|	|	|	|	|
		|	|	|	|	|
	   Al	|  -0.3	|  -1.5	|  -2.3	|  -4.5	|
	   --   |	|	|	|	|
		|	|	|	|	|
	   In	|  +2.2	|  +2.5	|  +0.2	|  -0.7	|
	   --   |	|	|	|	|
		|	|	|	|	|

		_________________________________


        Table 5.5: Temperature dependence of R\o\ in

                   Al and In. [10^-11^m^3^/As]
@ke
@ba
We see that the results are encouraging in the case of indium

and less so in the case of aluminum. The positive trend in the

calculated values of Al was already noticeable in the impurity

results. It may be that some aspect of the FS was neglected and

that the 3-band parameters are somewhat biased.
@be
@ba
The results on the temperature dependence of the impure speci-

mens are clearly a manifestation of the transition of an

impurity dominated regime to a phonon dominated one.
@be




			    -  97  -

@ba
The low concentration specimen with Cd has an RRR which is

already very close to the RRR of pure In and its temperature

dependence is very strong, whereas the specimen with Pb has an

RRR of .063 and phonons with a higher q vector are needed to

bring R\o\ back to the pure In case.
@be
@ba
Up to here we have not yet discussed the results of cold work

and size effect. We left these aspects out on purpose because

they are difficult to handle in the framework of atomic

scattering potentials.
@be
@ba
The cold work introduces additional grain boudaries in the

normally large grained samples. These grain boundaries are very

large objects compared to the electron wavelength, when we

assume that their principal scattering effect is due to their

associated strain field. This leads to the conclusion that they

should behave in a similar way as the low temperature long

wavelength phonons. The experimental result that the value of

R\o\ stays roughly constant when going from the pure to the

cold worked samples (see Fig.5.4), confirms this.
@be
@ba
It is difficult to see how the size effect can be incorporated

in our model. Let us assume that the scattering is isotropic on

the sample boundary and that this leads to a predominant s-like

phase shift. Our previous results showed that the impurities

with large s-like phase shifts tended to have a negative R\o\.

Thus, isotropic surface scattering might explain the negative

sign of R\o\ for the size effect experiments.
@be