You are on page 1of 45

Home Search Collections Journals About Contact us My IOPscience

The Hall effect in polycrystalline and powdered semiconductors

This content has been downloaded from IOPscience. Please scroll down to see the full text.

1980 Rep. Prog. Phys. 43 1263

(http://iopscience.iop.org/0034-4885/43/11/001)

View the table of contents for this issue, or go to the journal homepage for more

Download details:

IP Address: 134.99.128.41
This content was downloaded on 09/12/2013 at 17:13

Please note that terms and conditions apply.


Rep. Prog. Phys., Vol. 43,1980. Printed in Great Britain

The Hall effect in polycrystalline and powdered


semiconductors

J W ORTON and M J POWELL


Philips Research Laboratories, Redhill, Surrey RH1 SHA, UK

Abstract
T h e characterisation of the transport properties of semiconducting material
through a combination of Hall effect and resistivity measurements is as important in
understanding the electrical transport properties of polycrystalline and powdered
semiconductors as it is for single-crystal semiconductors. However, the interpretation
of these measurements for polycrystalline and powdered semiconductors is more
complicated due to the presence of grain boundaries and trapped interface charges
which lead to inter-grain band bending and potential barriers. I n recent years a
resurgence of interest in these materials, due to their potential for large-area device
applications, has led to a much better understanding of the influence of grain bound-
aries on these properties.
This review surveys the development of this subject from its origins in the early
1950s to the recent advances of the last few years, showing the extent to which a
considerable body of experimental results can now be understood in terms of simple
theoretical models.
We give a critical review of idealised two-phase geometrical models which, though
first considered nearly 30 years ago, still form the basis of the subject. These treat-
ments derive expressions for the resistivity and Hall coefficient of a composite material
in terms of the properties of its constituents. We show that these models can be
applied to the interpretation of transport measurements in polycrystalline films and
powder layers.
Important distinctions are made depending on whether the depletion layers
extend completely or partially through the grains, whether the Debye length is
greater or less than the grain size and whether the mean free path is greater or less
than the grain size. When the depletion region extends only partially through the
grain, the Hall effect measures the carrier concentration in the bulk of the grain while
the Hall mobility should be given by p= po exp (- + b / k T ) where +b is the band-
bending and ,uo is related to the grain diameter 1. When the depletion region extends
right through the grain, the carrier concentration measured may be very much lower
than the bulk doping level but is still close to the carrier concentration in the centre
of the grains. Band-bending is related to doping level N and interface trap density Nt.
It reaches a maximum value when N t = N l , and ,u therefore shows a minimum.
Consideration of the possibility of the mean free path being greater than the grain size

0034-4885/80/111263+45 $06.50 01980 The Institute of Physics


81
1264 J W Orton and M J Powell

leads to the conclusion that the Hall effect measurement would measure the number
of carriers capable of moving between grains, and not the number of carriers in the
bulk of the grains.
An important aspect of the conduction mechanism in these inhomogeneous semi-
conductors, not considered by the geometrical models, is percolation. We therefore
discuss the theoretical treatment of the Hall effect in percolative systems.
Experimental measurements on a wide range of materials are surveyed and the
results compared with theoretical predictions. Polycrystalline silicon films show
quantitative agreement, and in particular show the minimum in mobility as the inter-
grain band-bending is changed by doping. Other materials, including Ge, Te, CdS,
CdSe, InP, InSb, CuInSz and PbS, show some measure of agreement, particularly
in respect of the thermally activated mobility, but conclusions are more qualitative.
We also consider the modulation of conductivity and Hall coefficient by illumina-
tion and the adsorption and desorption of ambient gases. I n particular, experiments
on ZnO powder layers are described which show interesting effects due to the creation
of both surface accumulation layers, induced through strong uv illumination, and
surface depletion layers, induced through adsorption of oxygen. Appropriate modi-
fication of the composite geometrical model leads to a qualitative interpretation of the
experimental results.

This review was received in August 1980.


The Hall eflect in polycrystalline and powdered semiconductors 1265

Contents
Page
1. Introduction . . 1267
2. Theory of the Hall effect in inhomogeneous semiconductors . 1268
2.1. General considerations . . 1268
2.2. Geometrical models . . 1270
2.3. Application to polycrystalline materials and powders . 1274
2.4. Effect of grain size , . 1279
2.5. Preferred paths and percolation . . 1280
3. The Hall effect in polycrystalline thin films . . 1284
3.1. Introduction . . 1284
3.2. Effect of film thickness . . 1285
3.3. Effects due to grain boundaries . . 1286
3.4. Experimental results . . 1291
4. The Hall effect in zinc oxide powders . 1301
5. Conclusions . . 1303
References . . 1305
The Hall effect in polycrystalline and powdered semiconductors 1267

1. Introduction

The Hall effect has been used for many years as an aid to understanding the
electrical properties of single-crystal semiconductors (Putley 1960, Beer 1963, Blood
and Orton 1978). Its importance lies in its ability to measure free carrier density, at
least to within an accuracy of 10-ZO~o(limited by uncertainty in the Hall scattering
factor). This enables the experimenter to distinguish effects depending on carrier
density from those related to mobility.
T h e same technique has also been adopted by people studying the properties of
polycrystalline semiconductor films since the 1950s. As we shall see, it is even more
important in this case to distinguish carrier density variations from mobility varia-
tions because of the additional complications introduced by grain boundaries. I n
particular, we may cite the controversy as to whether photoconductivity in thin-film
materials such as PbS, CdS, or CdSe results from an increase of carrier density or
mobility (see 93.4.3). The answer to this question is almost certainly that either or
both may be important but that the precise details of why and when are still not
completely clear. Nevertheless, it is clear that measurements of Hall effect both in
the dark and under illumination are essential to this understanding.
The interpretation of Hall effect measurements on single crystals is straightforward,
apart from small uncertainties in scattering factor, so long as both carrier density and
mobility are uniform throughout the volume of material being sampled. Thus the
carrier density n and Hall mobility p~ of an n-type single crystal are given by

n = r/Re (1 1)
pH = Ru= Y p d (1.2)
where Y , R and U are the scattering factor (which lies between 1 and 2), Hall constant
and conductivity of the sample, while pd is its drift mobility.
I n the case of polycrystalline films it has been customary to apply these same
formulae, with the simplification that Y = 1 on the reasonable basis that other uncer-
tainties of interpretation are probably much greater than any uncertainty in Y . T h e
justification for this procedure has been based on simple two-phase geometrical
models of the Hall effect in inhomogeneous materials, which we shall consider in 92.
The essential point to appreciate here is that the electrical properties of polycrystalline
films are usually dominated by charge trapped at the inter-grain boundaries. This may
result in band-bending which sets up potential barriers to current flow (effectively
reducing carrier mobility) or to severe carrier depletion which reduces n considerably
below the bulk doping level N (or, in some cases, to a combination of these effects).
I n the case of severe depletion, where grain size is less than the Debye screening
length, n is approximately constant throughout a grain and therefore throughout the
whole film so the interpretation of the Hall effect appears straightforward. However,
in the case where depletion layer barriers exist adjacent to grain boundaries, the
carrier distribution is clearly inhomogeneous and the validity of the relation between
R and n in equation (1.1) is no longer obvious. This is the situation where the two-
phase models can be expected to apply, i.e. where the medium can be divided into
bulk regions of one conductivity interspersed with thin sheets of different conductivity.
1268 J W Orton and M J Powell

I n considering semiconducting powders, which are similar to polycrystalline films


in showing surface band-bending effects, we must also recognise a second type of
inhomogeneity arising from the geometry of grain packing. Current is constrained to
flow in non-uniform patterns for purely geometric reasons and the effective mobility
will be limited by such effects even in the absence of band-bending. Again, it is
possible to describe both conductivity and Hall effect in terms of an idealised geo-
metrical model which we outline in $2.3.
The last decade has seen a growing interest in the application of percolation and
effective medium theories to the description of electrical conduction in inhomogeneous
materials and we briefly review these in $2.5. The corresponding theories for the Hall
effect are less well established but we have attempted to summarise the present state
of knowledge. An important point we wish to emphasise is that both percolation and
effective medium theories apply to random mixtures of conducting phases whereas
the so-called geometrical models are highly ordered. I n other words, they apply to
fundamentally different problems.
While we discuss experimental results on polycrystalline films and powders very
largely in terms of the geometrical models it is important to realise that most real
systems probably show elements of both types of behaviour.

2. Theory of the Hall effect in inhomogeneous semiconductors


2.1, General considerations
One way of calculating the transport properties of an inhomogeneous medium is
to treat the medium as a composite of regions in which the single-crystal relationships
between Hall voltage, current density and magnetic field are assumed to hold. Each
contribution to the net Hall voltage can then be added by consideration of the
appropriate series and parallel contributions of an equivalent circuit.
Such an approach is limited to highly idealised geometrical models, but the use-
fulness of the results stems from the observation that, although idealised models are
necessary to perform the calculations, a number of features of the results appear to be
insensitive to the precise geometry considered. This encourages us to hope the
results will apply in a more general way to real systems.
The first paper to use this approach was that of Volger (1950). He considered the
composite medium shown in figure 1. Cubic grains of material, resistivity p l ,
mobility pl and carrier concentration nl are surrounded by inter-granular material,
resistivity p 2 , mobility p2 and carrier concentration n2. Further analysis of this model
of Volger (1950) and the application of the results to photo-Hall measurements have
been given by Bube and co-workers (Bube 1968, Blount et aZl970). However, these
papers contain an error in the analysis that was correctly pointed out by Lipskis et a1
(1971) and Heleskivi and Sal0 (1972).
Subsequently, Mathew and Mendelson (1974) considered the Hall effect in a
different, but related, idealised geometry. Instead of considering composite cubes,
they considered composite spheres. I n order to fill the whole of space it is necessary
to use spheres of all sizes, but with the ratio of diameters of the inner sphere to the
outer sphere remaining constant. I n the case of composite spheres the analysis is
possible in a more general way, but in the case where the inter-granular material is
thin and highly resistive, the results for the effective resistivity and effective Hall
constant are identical to those of the composite cubes. For the case where the inter-
The Hall efSect in polycrystalline and powdered semiconductors 1269

Figure 1. The Volger model of an inhomogeneous medium represented by composite cubes.

granular material is thin but highly conducting the results are identical for the effec-
tive resistivity and differ by only a small numerical factor for the effective Hall constant.
These results are significant in that they allow us to postulate the applicability of the
results to any medium composed of grains separated by a uniformly thin inter-
granular material.
The next important point to consider is the effect of depletion barriers in the regions
adjacent to grain boundaries. I n fact, in a polycrystalline film the grains will be in
such intimate contact that there is no inter-granular material as such. However, the
depletion layer can be treated similarly provided it behaves as an ohmic conductor.
Conduction parallel to the barrier can be represented as ohmic with a carrier
density reduced by a factor of order exp ( - +b/kT), where ~$b is the band-bending.
Normal to the barrier, current flow is by thermionic emission, which is also ohmic for
small applied voltages.
The standard treatment of the transport properties of polycrystalline films is
based on an often quoted paper by Petritz (1956). He showed the effective resistivity
to be controlled by the barriers and hence the temperature dependence to show an
activation energy equal to the barrier height?. However, in order to go one step
further and associate this activation energy with the mobility rather than the carrier
concentration Petritz cited Volger (1950) as evidence that a Hall effect experiment
would measure the carrier concentration in the grains in polycrystalline films.
Although more recently Snejdar and Jerhot (1976) and Jerhot and Snejdar (1978)
have considered a generalised model of the type shown in figure 1 and included
potential barriers between the grains and inter-granular material, they do in fact still
cite Petritz (1956) as justification for associating the activated process with mobility.
Consequently, we see that the results from the idealised geometrical models first
considered 30 years ago play an important role in the interpretation of the Hall effect
in polycrystalline materials and therefore we consider them in more detail in 92.2.
The derivation of these results depends on the assumption that a polycrystalline
material can be represented as a composite of homogeneous regions in which the
The literature on this subject contains many references to barrier height when what is
intended is band-bending and not the barrier height as measured from the Fermi level.
Wherever we refer to it in this review it should be construed as band-bending, i.e. the energy
difference between the conduction band edge at the grain boundary and in the bulk.
1270 J W Orton and M J Powell

conduction mechanism is ohmic as in a single crystal. However, if the grain size is


sufficiently small it is possible for the mean free path of the carriers to be of the same
order as or greater than the grain size. This regime is not catered for by the simple
geometrical models and requires a quite different approach, which is discussed in
$2.4.
Another criticism which can be made of these models is that they consider
uniform sized grains and uniformly thin inter-granular material whereas in real
materials we can expect there to be a variation in these properties. I n general, the
behaviour of a medium cannot be represented in terms of an idealised model with an
average grain size and average inter-grain resistance, since the current carriers will
tend to follow the path of least resistance. This general phenomenon is known as
percolation and the study of electrical conductivity in percolating systems has been
the subject of a number of papers (see, for example, Kirkpatrick 1973, Pike and Seager
1974, Shklovskii and Efros 1976).
I n the classical percolation problem, regions of different conductivity are assumed
to fill space completely randomly, and the conductivity is calculated as a function of
the volume fraction of each component. If there are two components, one of which is
conducting and the other non-conducting, then there is a discontinuous transition
between an overall conducting and overall non-conducting state when about 15% of
the total volume is occupied by conducting material. The transition point is known
as the percolation threshold.
I n addition to the behaviour of the conductivity near threshold, it is interesting to
know how other transport phenomena behave, in particular the Hall effect, and we
discuss this problem in $2.5.

2.2. Geometrical models


I n this subsection we derive expressions for the effective Hall constant and effective
conductivity for the idealised composite medium shown in figure 1. I n figure 2 we
draw a section perpendicular to the magnetic field and sketch the field lines for the
current density (J)and electric field ( E ) in the regions 1 to 4. The approach we have
adopted is basically similar to that of Lipskis et aZ(l971) and Heleskivi and Sal0 (1972).
One of the main complications and reasons for the error in the analysis of Bube
(1968) and Blount et aZ (1970) was an incorrect assumption concerning the field in
region 2. We can see from figure 2 that the electric field (E) in region 2 is the same
as in region 1 and the current density in region 3 the same as in region 1. This
description, however, is only valid if region 3 is thin and hence the behaviour in
region 4 is an edge effect, which can be neglected or conveniently grouped with
region 3. I n the opposite limit, that region 1 is much thinner than region 3, region 1
itself becomes the anomalous edge-effect region. It is not possible to derive general
results that cover the complete range of 11/12 ratios in any simple way, so we shall
confine our attention to the case 12 < ZI. However, the results of the composite sphere
model (Mathew and Mendelson 1974) are appropriate to the entire range of 11/12
ratios.
For the purpose of calculating the effective resistivity, the model of figure 1 can be
represented by the equivalent circuit of figure 3. The resistance rl, ~2 and YQ are the
resistances of regions 1, 2 and 3, respectively, and il, iz and i3 are the total currents
flowing in these regions. (Note that there are two r~ resistors representing the thin
slabs of inter-grain material perpendicular and parallel to the magnetic field.)
The Hall efJect in polycrystalline and powdered semiconductors 1271

ibi Field k ) Current


Figure 2. A section perpendicular to the magnetic field of the composite medium of figure 1,
showing (a) the designated regions, (b) the electric-field lines (-) and (c) the
lines of current density (----).

If the resistivities of the regions are p l and p2 and the applied voltage that produces
the currents is AV, per 'unit cell' then we have

Figure 3. The equivalent circuit for the model of figure 1 for calculating the effective resistivity.
The subscripts refer to the regions designated in figure 2.
1272 J W Orton and M J Powell

By writing the effective resistivity as

and substituting P = 12/11 and a= plIp2 we find for the effective resistivity

For the Hall effect, we consider that in each of the regions 1 , 2 and 3, Hall voltages
V I , V Zand V3 are generated where V,cci, (n= 1,2, 3). Then, in order to find the
total Hall voltage that would be measured experimentally we must sum these contri-
butions by using an appropriate equivalent circuit.
We believe that the circuit shown in figure 4 represents the correct choice. (Note
that VZ'is the voltage generated in the region 2 that is perpendicular to the magnetic
field.) Since there is negligible current flowing in the slabs perpendicular to the mag-
netic field and since ~2 is large we can ignore the limb containing V2' and Y Z and the
equivalent circuit reduces to the one considered by Bube (1968), Blount et aZ(l970)
and Heleskivi and Sal0 (1972). Lipskis et aZ(l971) employed a much more complicated
circuit to take into account the randomly staggered arrangement of grains which they
considered, but it can easily be shown that such a circuit may be simplified and is
exactly equivalent to the one used here.
T h e main feature of this equivalent circuit is that the relatively large Hall voltage
produced in region 3 is shorted by the more conducting region 1. I n terms of the
equivalent circuit picture, a circulating current is produced. An alternative way of
seeing that the Hall voltage produced in region 3 will be shorted is to consider the
current and field lines under the application of a magnetic field. I n figure 5 we can see
that the relatively large Hall angle in region 3 is accommodated by the current flowing
at an angle to the faces of the cubes, and thus a negligible Hall voltage will be meas-
ured perpendicular to the direction of net current flow. Such a picture is analogous

b
Figure 4. The equivalent circuit for the model of figure 1 for calculating the effective Hall
coefficient.
The Hall effect in polycrystalline and powdered semiconductors 1273

--e-->---

Figure 5. Schematic diagram showing the electric field (-) and current density (---)
under the application of a magnetic field. The Hall angle in region 3 is exaggerated.

to the conventional view of geometrical magneto-resistance in thin slabs (Beer 1963).


This picture of the Hall voltage generated in region 3 being shorted out is a crucial
step in the argument and we must be convinced that the correct equivalent circuit is
used as the final results depend critically on this choice.
We write the Hall constants in regions with resistivity p l and p 2 as RI and R2,
respectively, and the net Hall voltage appearing across a 'unit cell' in a magnetic field
B as AVH. Using the approximation that r2Bq we can write

and we can define the effective Hall constant by

Now
Y1
AVH=Vi+ V2- -__ ( V i - V3). (2.6)
y1+ y 2
Substituting leads to

Writing a= p 1 I p 2 and 13 = 12/11 as before, and using again the approximation that
r2Bq (i.e. a/3(1+/3)< 1 ) gives us finally
R" =Ri(1+p) + 2R24(1+ p). (2.8)
1274 J W Orton and M J Powell

This result is identical to that obtained by Lipskis et a1 (1971), but the second
term differs by a factor of two from the result of Heleskivi and Sal0 (1972).
We can summarise by writing down the result for the effective resistivity p",
effective Hall constant R" and effective mobility p* = R"/p*, with the further approxi-
mation that (1+ ,B) E 1:
p" =Pl + p2p (2.9)
R" = RI + 2 R 2 4 (2.10)
(2.11)

These results are identical to those derived by Mathew and Mendelson (1974) for
the composite sphere geometry under the same conditions of a thin, more resistive,
inter-grain region. T h e agreement suggests that the results should be of general
validity, independent of precise geometry.
It is also of interest to consider the case where the thin inter-granular region is
more conducting. This will be of interest when we discuss the effect of accumulation
layers in $52.3 and 4.
If we consider our model of figure 1, when p2< pl, then all the current will flow
along the inter-granular regions. The effective resistivity is merely increased by the
cross-sectional area, i.e.
p" = p2/2p. (2.12)
A Hall voltage is developed in both the slabs perpendicular to and parallel to the
magnetic field. T h e slabs perpendicular to the magnetic field are continuous in the
direction perpendicular to both field and current and the largest Hall voltage is
generated here.
Writing the effective Hall constant as before leads to the simple result

i.e.
(2.13)
These expressions for p" and R" can be compared with the corresponding expressions
for the composite sphere geometry (Mathew and Mendelson 1974) under the condi-
tions that p 2 Q p l and 12<11, namely
P" = P 2 P P (2.14)
Rx = 3R2ISP. (2.15)
Again we see that the composite sphere geometry gives an identical result for resis-
tivity and a similar result for the Hall coefficient, a factor of 8 appearing instead of 3.

2.3. Application to polycrystalline materials and powders


We now apply the results of the previous subsection to the case where the inter-
grain region is a back-to-back Schottky barrier and current flows between grains by
thermionic emission.
T h e current flowing in region 3 over the barrier is

(2.16)
The Hall effect in polycrystalline and powdered semiconductors 1275

where v b is the voltage drop across the first of the pair of back-to-back Schottky
barriers. For thermionic emission io is given by
io = &zlqEexp ( - $b/k T ) (11 + 12)2 (2.17)
where nl is the carrier concentration in the grains (region l), E is the thermal velocity
= (8kT/nm")1/2 and $b is the barrier height.
For a back-to-back pair of Schottky barriers and a small applied voltage, V <kT/q,
the voltage drop across the barrier is proportional to the current and equally divided
between the two barriers. We can therefore define an effective resistance for the
barrier r3 by
(2.18)

We can then define a quantity pb which has the dimensions of mobility:


(2.19)
where
(2 * 20)
and we have written (11 + 12) = 1, the overall size of the grains.
Thus
(2.21)

Writing r3 E p p 2 / 1 and substituting in equation (2.9) then gives


1
p"=p1+---. (2.22)
nlqPb
Notice that equation (2.18) could be interpreted either in terms of a constant
carrier density associated with a thermally activated mobility or conversely in terms
of a constant mobility and thermally activated carrier density.
Petritz, in his original paper (Petritz 1956) on the mobility of polycrystalline
materials, referred to Volger (1950) as justification for assuming R" = R I (i.e. a Hall
experiment measures nl, the carrier density in the bulk of the grain) and that the
mobility must therefore be activated. However, we should really go back to the
equivalent circuit (figure 4) and satisfy ourselves that this is correct. Under these
circumstances the resistance 1 2 will be very small, and the voltage V 2 will also be
small since very little current can flow parallel to the potential barrier. We assume
that the effective carrier density n2 is less than nl by a factor exp ( - $b/kT) and also
write p 2 = p 1 which appears to be a good approximation for a homogeneous
material. Any modifications to scattering mechanisms as a result of the band-bending,
such as extra Coulomb or grain boundary scattering as considered, for example, by
Bednarezyk and Bednarezyk (1979), will reduce p 2 compared to p1. Notice that p2
is not p b nor is it po, the prefactor to the exponential in the expression for pb; it is
the bulk mobility in the region of the potential barrier.
Therefore, if we assume p 2 = p1 and write Rzcx = RI then equation (2.8) becomes
(2.23)
Thus
(2.24)
1276 J W Orton and M J Powell
and p* is given by
1 - 1 +-.1
--- (2.25)
Px pb

If p l < pb, then we expect p" = p1, the bulk single-crystal mobility, but if ~ 1 pb %
then the effective mobility is given by p b . Notice that the occasionally quoted result
p* = p1 exp ( - $b/kT) is incorrect. If the mobility is thermally activated the pre-
exponential factor is given by equation (2.20). A useful rule of thumb is obtained by
taking m" = 0-3m which leads to a room-temperature relationship ~ O1061 Z (1 in m,
po in m2 V-1 s-1). As we shall see in $3, there is experimental evidence for the linear
relationship between po and 1.
T h e temperature dependence of po, though slight, is sufficient to modify lg p*
against 1/T plots when $b is small ( < 0.05 ev), and strictly $b should be obtained
from a plot of 1g (p*T112) against 1/T. Finally, because p b and p1 vary in opposite
senses with temperature it may be possible to reverse the inequality p1> p b simply by
raising the temperature, giving rise to an unexpectedly complex temperature variation
of px (Mankarious 1964).
We now wish to consider the problem of powder layers and how we can apply the
ideas of the preceding subsection to this situation. There is a general problem in
interpreting the Hall effect in powders, in that the whole of space is not filled with
conducting material. T h e packing density for randomly close-packed spheres is
~ 6 0 %(Finney 1977), and much lower packing densities are found for some sub-
micron-sized powder layers (Clarke et a1 1978). I n this situation, is the measured
Hall coefficient still equal to that of the grain material?
The answer to this question depends on how the non-conducting phase (i.e. the
space between particle grains) is spatially arranged. If it occupies space entirely at
random, then the result would be given by the effective medium and percolation
approaches of $2.5. However, this is not the most important feature. I t is clear that
the conductivity in a powder layer will be limited by the size and nature of the inter-
grain contacts, so any theory of conductivity must include contact radius as a para-
meter. T o do this we consider an idealised model where the arrangement of the
non-conducting phase is completely non-random. This consists of a simple cubic
array of spheres in contact (see figure 6). We choose a simple cubic array rather than
a close-packed array since the mean number of contacts of a randomly packed array
of spheres is about 6 (Powell 1980a). Using such a geometry and assuming the resist-
ance of each grain to arise from current crowding at the contact, i.e. rz p$a where
a is the contact radius, it is easy to show that the effective resistivity is given by

1
P"Zz-, PI (2.26)

where 1is now the grain diameter.


T h e effective Hall constant can be similarly estimated by summing the Hall
voltages generated in each sphere. The current density distribution in a single sphere,
when the current is flowing between diametrically opposed 'point' contacts, can be
calculated using Legendre polynomials (J P Whelan, private communication). I t
leads to the simple relation
R" 2d2
= -- Rl 0.9 RI (2.27)
77
The Hall eflect an polycrystalline and powdered semiconductors 1277

I I

Figure6. Idealised model of a powder, represented as a regular cubic array of spherical


particles.

i.e. the effective Hall constant is, to a fair approximation, the Hall constant of the
grains.
T h e main consequence of the granular structure of a powder, therefore, is an
increased resistivity associated with the contacts. This will also be true if we include
potential barriers or an inter-granular phase at the contacts. For example, if inter-
grain potential barriers are present, we can include the effect of the contact by writing
(2.28)
RXx R1 (2.29)
l N 1 1
--,-+-- (2.30)
pX 8Pl YPb
where, for very thin barriers, y = 4a2/12 and 6 = 2aJl.
Finally in this discussion of the application of simple geometrical models to
polycrystalline and powdered materials, we should mention the possibility of surface
accumulation rather than depletion. There is good experimental evidence for the
existence of strong accumulation layers on surfaces of ZnO under certain circum-
stances. So, for example, in ZnO powders this can lead to conduction taking place
predominantly through the surface regions of each grain.
For this case, too, we can develop a simple geometrical model for the Hall coeffi-
cient and resistivity using the cubic array of spheres shown in figure 6. We now
assume the current through each grain to be carried wholly within an outer conducting
shell of thickness 12/2. Reference to figure 7, which represents a cross section through
the centre of the sphere perpendicular to the direction of current flow, shows that the
Hall voltage between points P and Q is given by the integral:

(2.31)
where I2 is the current flowing through the sphere.
1278 J W Orton and M J Powell

Figure 7. Section through the centre of a composite sphere normal to the direction of current
flow, illustrating the calculation of Hall voltage between points P and Q. Current
Iz normal to paper.

Summing these voltages across the array of spheres results in the following
expression for the effective Hall coefficient :

(2.32)

The resistance yg of a single sphere is determined for this case by the resistance of
the shell rather than by contact resistance (provided 12 < a ) and is given by

(2.33)
where
A = In (cot 1/72) (2.34)
and i,h is the ratio of contact radius to grain radius, i.e. $/2= ail. This result leads to
an expression for the effective resistivity of the array:
2A
p*=- pz (2.35)
4
and finally, combining equations (2.32) and (2,35),
pLQ
= R*/p" = A-1p2. (2.36)
The parameter A is a slowly varying function of i / ~and for sub-micron grains held
together by van der Waals forces takes a value of about 3 (J W Orton unpublished).
Thus, to within small factors of order 3 these results are in agreement with those
derived for composite cube or sphere models in 52.2 (equations (2.12)-(2.15)) and
support the contention that both resistivity and Hall constant are determined by the
The Hull efJect in polycrystalline and powdered semiconductors 1279

surface region. I n 94 we shall illustrate this point with the results of measurements
made on ZnO powder-binder layers.

2.4. Eflect of grain size


I n the previous subsections we made the implicit assumption that the mean free
path for electron scattering (A) is short compared to the grain size 1. This is a necessary
condition for being able to define the grain resistance and for expressing it in terms
of a bulk mobility p1. If this is not the case, then we have to reconsider the mechanism
of carrier transport between grains, and the consequence in terms of the mobility.
It is important therefore to consider the likely magnitude of A.
For our present purpose we neglect the energy dependence of the relaxation time
T and write
h = CT = plm*E/q (2.37)
which gives h = 6.5 x 10-7 (m*/m)lIzp1, for h in m and p in m2 V-1 s-1. Thus h varies
from about 6 nm for CdS to 450 nm for n-type InSb, and illustrates the possibility
of h being greater than a grain size for a high mobility material (I has frequently been
measured in the range 10-100 nm).
The significance of this result is considerable. I n particular, consideration of the
transport mechanism leads to the observation that the Hall effect may not measure
the bulk carrier density n l , but a very much smaller quantity nl exp ( - + b / k T ) , and
consequently the Hall mobility will no longer be thermally activated. This represents
a significant departure from the theory of 992.2 and 2.3.
I n figure 8 we illustrate the energy band diagram of a granular n-type semi-
conductor. I n the absence of scattering within a single grain an electron with energy
greater than +b may cross several grains before colliding with a phonon. Electrons
with energies less than +b are confined with the potential wells which occur between
grain boundaries and are unable to take part in current transport. Only those carriers
with E > +b respond to an applied magnetic field so the quantity measured by a Hall
effect experiment is not n 1 but is now given approximately by nl exp ( - +b/kT).
The Hall mobility of these carriers is no longer activated but is limited by other
scattering processes. (Note that it is possible to define a mobility only over a distance
of several mean free paths and therefore over many grain diameters.) The situation
is somewhat analogous to the behaviour of a single crystal at low temperatures where
carrier freeze-out becomes significant. Those electrons trapped on donor sites take
no part in conduction and a Hall effect measurement records only the remaining free
carriers in the conduction band. The carrier density is thermally activated but not
the Hall mobility.

I \
Trapped electrons

Figure 8. Schematic representation of the conduction band in an n-type polycrystalline film


for the case where the electron mean free path exceeds the grain dimension 1.
82
1280 W Orton and M J Powell

It is interesting to compare the expected results of measuring Hall and drift


mobilities for the two cases, i.e. when the mean free path is either larger or smaller
than the grain size.
I n both situations an injected pulse of charge carriers will be trapped behind inter-
grain barriers, and excitation via a phonon interaction is necessary to enable them to
surmount the barrier. Therefore, in both interactions we expect a thermally activated
drift mobility but the Hall mobility is activated only for the situation where X < 1.
Detailed calculations which demonstrate this effect for InSb have been made by
Bednarezyk and Bednarezyk (1979) who also consider the additional scattering due to
grain boundaries. They treat each grain as a charged scattering centre and use a
modified Conwell-Weisskopf ionised impurity formula to calculate the mobility. T h e
validity of using a point-charge Coulombic potential may be questioned, but we can
make a reasonable estimate of grain-boundary-limited mobility without considering
the details of the interaction, simply by equating the effective mean free path to the
grain radius 112, i.e.
(2.38)

This is closely similar to the pre-exponential term ( P O ) for the mobility determined
by thermionic emission between grains (equation (2.20)).

2.5. Preferred paths andpevcolation


T h e current through an inhomogeneous semiconductor depends both on the
nature of the inhomogeneities and their spatial arrangement. If there is some com-
ponent of the inhomogeneity which is distributed spatially randomly, then the electric
current will follow preferred paths, determined by a percolation process.
Within the general framework of percolation theory there are a number of related
percolation problems which can arise (separately or together) in inhomogeneous
semiconductors. T h e classical problem, referred to in $2.1, that arises when two
regions of widely differing conductivity fill space randomly by volume is known as the
continuous percolation problem (Zallen and Scher 1971, Webman et a1 1976). This
can arise in semiconductors, for example, as a result of inhomogeneous bulk doping.
I n the type of inhomogeneous materials we have considered in previous sub-
sections, i.e. polycrystalline and powdered semiconductors, the regions of different
conductivity are not distributed randomly, e.g. the inter-granular or barrier regions
always occur between grains and the current has to flow through them. I t is true that
the current will still flow along preferred paths of least resistance, which exist because
of irregularities in grain size, shape and contact distribution, but the conductivity is
dominated by the barriers even though the total volume fraction occupied by the
barrier region is low. However, there is the possibility in this situation that a related
percolation problem can arise.
Consider a powder layer where the effective resistivity of the layer is dominated by
the grain-to-grain contact resistances. If the value of the contact resistance can take
a range of values due to fluctuations in the barrier height and/or' variations in the
contact area, then the current will flow along paths which offer minimum resistance.
It is not difficult to envisage mechanisms that could lead to a variation in barrier
height, e.g. a spatial variation of the surface state densities which cause the barriers
The Hall eflect in polycrystalline and powdered semiconductors 1281

or the effect of depletion regions causing greater or lesser pinch-off in the contact
region between sintered powder particles (Beekmans 1978).
A variation of the barrier height will lead to a wide variation in contact resistance,
and the wider the range of contact resistances then the more significant will be the
percolation. I n the limit of a very wide range of contact resistances the overall effec-
tive resistivity will be determined by the critical resistance at the percolation threshold
level. We can see this by following the procedure of Ambegoakar et a1 (1971). T h e
contact resistances together form a network. If we imagine rebuilding the network
by replacing all the resistances in order of increasing value, then there is zero con-
ductivity until we have replaced a critical number of resistances corresponding to the
percolation threshold. T h e value of resistance at this point will dominate the con-
ductivity since resistances of a higher value are already shunted and resistances of a
lower value have a negligible resistance in comparison. Seagar and Pike (1974) have
investigated this procedure in a number of model percolation systems. T h e situation
is clearly related to the simple bond percolation problem (Kirkpatrick 1973), in which
a network of sites is connected by bonds, which can be considered to be either con-
ducting or non-conducting. T h e overall conductivity is calculated as a function of
the fraction of conducting bonds and found to be zero below a critical concentration
p,, the percolation threshold, and given by a power law just above threshold :
U= .o(p -pc)t. (2.39)
Below the threshold, there are only isolated groups of conducting bonds, and at
the threshold there is one infinite cluster of connected conducting bonds together with
a number of separate isolated clusters of conducting bonds. T h e conductivity
increases much more slowly, above threshold, than the percolation probability P ( p )
(the probability of a bond being conducting and connected to the infinite cluster),
since much of the infinite cluster, although connected, will consist of dead-ends and
not contribute to the conduction.
T h e value of p , depends on the connectivity of the network but the power index
depends only on-the dimensionality (t = 1.65 in 3D and t= 1.1 in ZD) (Kirkpatrick
1978).
There is a related percolation problem, known as the site percolation problem,
where the sites rather than the bonds are either conducting or blocking. Again the
conductivity is given by equation (2.39), but p is now the fraction of sites. T h e value
of t is the same as for the bond percolation problem and pc depends on the con-
nectivity of the network (Powell 1980b).
T h e relationship between percolation problems on networks and the continuous
percolation problem has been the subject of considerable discussion (Scher and Zallen
1970, Zallen and Scher 1971, Webman et a2 1975,1976), but one aspect of this relation-
ship is important when we consider the Hall effect and that is the role of correlation.
Webman et a1 (1975,1976) showed that by taking the simple bond percolation problem
on a regular lattice and increasing the degree of correlation between neighbouring
bonds, i.e. the tendency for a neighbouring bond to be of the same type, then the
percolation threshold decreased and tended toward a constant value for a network
with correlation up to more than about nine bond lengths. At this point the percola-
tion has become continuous and the threshold value is the continuous percolation
threshold.
At larger fractions of conducting species, much greater than the percolation
threshold, the conductivity is well described by the effective medium theory. This is
1282 J W Orton and M J Powell

a self-consistent embedding approximation invented by Bruggeman (1935) and


christened later by Landauer (1952). The method calculates the properties of a
homogeneous 'effective' medium in which the real inhomogeneities are embedded in
such a way that the spatial average fluctuation of the electric field is zero. This
approach is not limited to small concentrations or small fluctuations, but the exact
realm of validity is very difficult to ascertain.
T h e effective medium theory was generalised to include the magneto-conductivity
tensor a, and hence the Hall coefficient, by Cohen and Jortner (1973). For a random
mixture of two components 01, R I , p1 occupying a volume fraction C and 02, Rz,p2
occupying a volume fraction (1 - C), the effective conductivity and Hall constant are
given by
rJ" =f.1 (2.40)

(2.41)
where
x = u2/01 (2.42)
Y = P2IP.1 (2.43)
(2.44)
and
(2.45)
When one component is non-conducting, i.e. x = 0, these expressions simplify to

and
(2 47)
I

+
The conductivity goes to zero at a volume fraction of (this can be compared with the
percolation threshold of 0.15 for continuous percolation in three dimensions), but the
Hall coefficient never differs from the Hall coefficient of the conducting medium by
more than a factor of 2 (i.e. when C= Q, R" = 2R1).
A number of papers have modified the effective medium theory (Davidson and
Tinkham 1976, Granquist and Hunderi 1978) and it has been shown that the percola-
tion threshold can be moved toward the 0.15 value (Granquist and Hunderi 1978),
but neither of these papers have considered the Hall coefficient.
T h e original effective medium theory gives results different from percolation
theory near the percolation threshold, and consequently it is of interest to consider
the behaviour of the Hall coefficient in this region. Kirkpatrick (1973), in line with
earlier work (Eggarter and Cohen 1970), suggested that the Hall coefficient should be
proportional to the inverse of the percolation probability,
R" = P-l(p) RI. (2.48)
This implies that the 'dead-ends' occurring in the infinite cluster contribute to the
Hall coefficient, even though they do not carry current or appear on the surface of a
sample, a picture which seems unrealistic.
More recently, a number of workers (Straley 1976, 1977, 1978, Levinshtein et a1
1976a, Webman et a1 1977b) have developed a scaling theory to describe the critical
The Hall effect in polyciytalline and powdeved semiconductors 1283

behaviour near the percolation threshold, analogous to the behaviour of thermo-


dynamic quantities in the theory of critical phenomena. A correlation length can be
defined which becomes infinite at the percolation threshold, and near the percolation
threshold is given by a power law
LK(p-p,)-. (2.49)
Several other quantities can be defined, including the conductivity and the per-
colation probability, which are also given by power laws, and relationships between
the indices are inferred. For example, the relationship between the conductivity
index t and the correlation length index v has been proposed as (de Gennes 1976,
Skal and Shklovskii 1975)
t=(d-2)v+ 1 (2*50(a>)
or alternatively (Levinshtein et a l 1976b) as
t = (d- 1) (2 5 V ) )
where d is the dimensionality.
Skal and Shklovskii (1975), Levinshtein et aZ(1976b) and Shklovskii (1977) have
developed the scaling arguments to include the Hall coefficient. Writing the Hall
coefficient as
R K (p -pc)-g (2.51)
they deduce the relation between the indicesg and v to be
g=v(d-Z). (2.52)
Thus in two dimensions the Hall coefficient is predicted to be constant in agreement
with effective medium theory (Adkins 1979), but in three dimensions the Hall coeffi-
cient is predicted to diverge at the percolation threshold.
There are some possible objections to the scaling arguments as applied to the Hall
coefficient, which have been discussed by Orton and Powell (1978), and some un-
resolved inconsistencies in scaling relationships such as (2,50(a)) and (2.50(b))
(Straley 1980). Computer calculations enable independent determinations of t and v
to be made and the indices have been calculated several times by different workers,
by a number of different methods, and a consensus gives in three dimensions v = 0.85
( -t. 0.05) and t = 1.65 ( k 0.05) (Kirkpatrick 1978).
T h e corresponding independent numerical simulation of the Hall coefficient on a
computer that would enable g to be calculated is more difficult. However, Webman
et a1 (1977a) have successfully simulated the Hall coefficient using a simple cubic
lattice with correlated bonds. I n order to generalise the finite difference approxima-
tion of the continuity equation for the magneto-conductivity tensor, correlation
between neighbouring bonds is essential. This is understandable since only a con-
tinuous percolation problem can have a meaningful Hall coefficient. For random
mixtures of a conducting and a non-conducting component, they found two interesting
results. Firstly, for C > 0.4, the calculated Hall coefficient is in excellent agreement
with effective medium theory, in the same way that numerical simulations of the
conductivity agree with effective medium theory in this range. Secondly, in the
range 0.25 > C > 0.15, the effective Hall constant showed a pronounced increase as
C -+ 0.15, the percolation threshold. Although their results are not detailed enough
to enable g to be calculated (this would involve some very long calculations) they do
indicate a divergent behaviour at threshold.
1284 J W Orton and M J Powell

It is interesting to see if there is any experimental evidence of divergence of the


Hall coefficient at the percolation threshold, but, unfortunately, a comparison with
experimental results is extremely difficult for two reasons. Firstly, it is difficult to
know whether some microscopically inhomogeneous systems are true representations
of the percolation problem, and secondly, since the conductivity falls more rapidly
than the expected increase in the Hall constant, the mobility tends to zero and conse-
quently experimental measurements become very difficult.
However, we can point to experiments on metal-ammonia solutions (Nasby and
Thompson 1970), porous copper (Goldin and Juretschke 1958) and model experi-
ments using stacks of resistive card (Levinshtein et al 1976b), all of which show an
increase in the Hall coefficient as the percolation threshold is approached, but the
evidence is far from substantial and an increase of the Hall coefficient of even one
order of magnitude has not been reported. Experiments on mixtures of conducting
ZnO and insulating ZnS powders were also inconclusive (Clarke et at 1978).
The true role of percolation in the transport properties of inhomogeneous semi-
conductors is very difficult to ascertain. While it is possible to invent model systems
displaying purely classical percolation phenomena, few systems of practical interest
fall into this category. Nevertheless, percolation is likely to play some part in the
transport of materials such as polycrystalline films and powder semiconductor layers.
It seems necessary therefore that a full treatment of the transport properties should
combine the approach of the geometrical models, describing the non-random aspects,
with percolation theory or effective medium theory, describing the random processes.
An area where percolation effects seem likely to be important is discussed in 93.4.3,
where we consider barrier modulation effects.

3. The Hall effect in polycrystalline thin films

3.1. Introduction
Polycrystalline thin films have become of considerable importance in the electrical
and electronics industries, particularly where there is a requirement for cheap, large-
area semiconductor devices. Perhaps the most significant recent developments have
been concerned with solar cells, where CdS and poly-silicon show promise as materials
for terrestrial applications, but thin films have many other uses. Examples include
thin-film transistors, photoconductors, electroluminescent panels, transparent con-
ductors and the use of poly-silicon in integrated circuits. It has therefore become of
importance to gain a fuller understanding of conduction mechanisms in these films
and the combination of Hall effect and resistivity measurements plays a major role.
A detailed description of the technology and general properties of the films lies
outside the scope of this review but in the interest of clarity we provide a brief outline.
A number of preparative methods have been employed, the most common being
vacuum evaporation of the material onto glass or quartz substrates (though a wide
range of other substrate materials has been used in particular cases). Typical base
pressures in the evaporator are z 10-7 Torr and substrate temperatures between
room temperature and 350C. Resistive heating, electron bombardment or flash
heating have all been employed, depending on the nature of the evaporant. Other
deposition methods include sputtering, chemical vapour deposition and spray pyroly-
sis which has recently been given considerable attention for growing films of CdS and
The Hall effect in polycrystalline and powdered semiconductors 1285

CdSe. I t frequently proves beneficial to anneal the as-grown film in an inert atmo-
sphere at temperatures up to 500C.
T h e use of non-crystalline substrates precludes the growth of single-crystal films.
Electron microscopy reveals the presence of a granular structure with crystallite
diameters usually in the range 5-500 nm though some workers have reported diameters
of several microns. Grain sizes, stoichiometry and electrical properties may vary
strongly with the details of growth and annealing treatment. A recent development of
considerable interest is the use of laser annealing which has been shown to increase the
grain size in silicon films from 50 nm up to 2 pm x 25 pm (Gat et al 1978). I n some
cases the individual crystallites show preferential orientation such as the well-known
tendency for CdS grains to grow with their c axes normal to the substrate surface.
This may result in quite marked anisotropy of electrical properties (Kazmerski et al
1972) but we shall chiefly be concerned with measurements in the plane of the film.
T h e two obvious features of these materials which distinguish them from single-
crystal semiconductors are their thickness, typically in the range 10 nm-1 pm, and
the presence of grain boundaries. We now consider in more detail the effect these
have on electrical properties, in particular the Hall constant R and Hall mobility ,U.

3.2. EfJect ofJilm thickness


T h e mobility of carriers in a thin film will be reduced by surface scattering when
the bulk mean free path hc is comparable with the film thickness. This has been dis-
cussed in detail by Many et a1 (1965) who showed that the following elementary
account gives results very close to those derived by solving the Boltzmann equation.
Consider a film of thickness 2d in which the band edges are flat up to the surface.
Suppose also that the surfaces act as diffuse scatterers, i.e. electrons on reflection lose
all memory of their incident velocities. T h e effective collision time is then given by
the sum of surface and bulk scattering rates:

where T~ z d/Zz,E, being the mean velocity towards the surface. I t is then possible to
define an appropriate mean free path for motion normal to the surface such that
A, = T ~ E ,so that T ~ Tcd/h,.
Z Substituting this in equation ( 3 . 1 ) gives the result:

where A, is related to the bulk crystal mobility by an equation analogous to (2.37), i.e.
A, = (mgkT/2.irq2)lizpC. (3.3)
If, as in practice is likely, the semiconductor bands are not flat to the surface but
bend upward the potential barrier will reduce the number of electrons which actually
reach the surface and thus reduce the amount of diffuse scattering. T h e converse
case of surface accumulation implies an increased proportion of surface conduction
and requires more sophisticated treatment. We shall not consider it further.
There appears to be very little clear experimental evidence for surface scattering
in polycrystalline films, which may be related to the requirement for A, to be com-
parable with d. There is evidence (Shallcross 1966, Ling et a1 1972) that grain size
may be roughly proportional to film thickness, which means that A, is likely always to
1286 J W OFton and M J Powell

be limited by grain boundary scattering giving peel rather than showing the variation
predicted by equation (3 -2). Kazmerski et a1 (1972) have reported that in CdS films p
varies according to equation (3.2) but the value of A, derived from fitting theory and
experiment appears to be very much larger than one estimates from equation (3.3),
i.e. A, (expt)= 110 nm, A, (ca1c)z 5 nm, which is difficult to explain. Berry and
Jayadevaiah (1969) similarly derived a value of A, (expt) = 70 nm for PbTe and Berger
et a1 (1969) reported p against thickness curves suggesting A, (expt)z 100 nm for
CdSe films. These latter authors propose an explanation similar to that proposed above
(i.e. based on grain size being proportional to film thickness up to a thickness of about
100nm). We must conclude that there is as yet no evidence for mobility being
limited by surface scattering in polycrystalline thin films.
Amith (1960) has also shown that the Hall constant should vary with thickness
when X,>d but to a much smaller extent than mobility. Again, there appears to be
no clear experimental evidence for this in polycrystalline films.

3.3. Efects due tograin boundaries


The presence of grain boundaries represents the essential difference between
single-crystal and polycrystalline semiconductors. These boundaries are important
in several ways-they generally contain fairly high densities of interface states which
trap free carriers from the bulk of the grains, scatter free carriers by virtue of the
inherent disorder and the presence of trapped charge and may also act as sinks for the
segregation of dopant atoms. We shall be concerned largely with the first two of
these effects.
The interface states referred to may be either intrinsic or extrinsic (for example,
extrinsic states may result from the adsorption of gases such as oxygen) but it seems
reasonable to expect that in either case their density may be influenced by exposing
the film to appropriate ambients. The density Nt (per unit area) is important because
it determines the maximum amount of charge which can be trapped and the trap
energy Et is also important because, as we show below, it may determine the activation
energy for free carrier freeze-out.
Interface charge gives rise to band-bending in the bulk of the grain and is therefore
the origin of the interface barriers discussed in $2, Also, if Nt is high enough (or the
bulk doping level N is low enough) the grains may be almost fully depleted of free
carriers, leading to very high film resistivity with large thermal activation energy. We
can distinguish three situations, depending on the relative magnitude of nt compared
with NI (nt being the density of charge trapped in surface states) and of the Debye
length LD compared to grain size 1.
If nt < NI the resulting depletion layer width W is less than l12 and the band dia-
gram takes the form shown in figure 9(a). The carrier density n is constant at the bulk
value N throughout most of the grain and the mobility is thermally activated. If
n t z N 1 the depletion layer extends right through the grain, giving the band profile
shown in figure 9(b) where n may be much less than N , and p is still barrier-limited.
This profile occurs when LD< 112. If the converse is true, i.e. LD> 112, the conduction
band will be essentially flat throughout each grain and therefore throughout the
whole film. There will be no barriers to current flow and mobility will not be thermally
activated (though n may be). T h e band profile is shown in figure 9(c) and approximate
values of Debye length are given in table 1. The boundaries between the various
regimes are plotted in figure 10 for a range of trap densities.
The Hall eflect in polycrystalline and powdered semiconducton 1287

io I lbl (Cl

Figure 9. Conduction band profiles through single grains of an ti-type film for the three cases
(b) nt % NI, LD< 112 and ( c ) nt z NI, L n > 112.
discussed in the text: ( a ) nt <NI,

Table 1. Approximate values of Debye length LO= (~kT/iVq~)~/2.


~ ~

N (m-3) 1019 1020 1021 1022 1023


L D(nm) 103 300 100 30 10

Bearing in mind the additional criterion concerning the relative size of grain
diameter and mean free path discussed in $2.4, it is clear that considerable care is
needed in interpreting the electrical properties of polycrystalline thin films. It is
also clear that grain size is a very important parameter and should always be measured.
We now proceed to deal with the individual regimes in more detail.

3.3.1. nt x NI, LD> 112. Because LD> l/2 the bands will be effectively flat and we can
readily calculate n from the condition for charge neutrality. T h e various energy levels
involved are shown schematically in figure 11. We assume the Fermi level to be far

/ I!
Figure 10. Graph showing the demarcation between the three regimes illustrated in figure Y.
1288 J W Orton and M J Powell

I-- - --- -La


Figure 11. Band diagram for the case where nt % NI and L D> Zj2 showing the relative positions
of donor and acceptor levels Ed,Ea, the Fermi level EF and the interface traps Et.

enough removed from the conduction band and shallow donors that we can use
Boltzmann statistics but we use Fermi-Dirac statistics for the traps.
T h e charge neutrality condition may be written:

Zt= l(N- Ed- n) (3.4)


where nd is the number of shallow donors containing electrons and N is the net doping
level, i.e. N = Nd- Na. T h e other relationships we use are as follows:

(3.5)

(-kT-)
n = N , exp Ec - EF

. )(-, E,d EF
na=Nd exp -

Combining equations (3.4)-(3.7) leads to the following quadratic equation for n:

n2 1 +
{: exp ( X d )
I+ n{Nt' - N + Nc exp ( - Xt) + Nd eXp [ - (Xt - xd)]}
-NcNexp(-xt)=O (3.8)
where we have written N( = Nt/Z and xi = (E, - Ei)/kT. Examination of the terms in
curly brackets in (3.8) suggests certain simplifying approximations. If we suppose
Nd exp ( X d ) <Nc and that Nt' dominates the coefficient of n we obtain:

I n a wide-gap semiconductor such as CdS or even Si (E, - Et) = E t may be 0.3 eV or


more which implies n <N at room temperature. Also n is thermally activated with an
activation energy equal to the trap depth E t and, as p is not activated, this implies
that the conductivity will have the same activation energy.
The Hall effect in polycrystalline and powdered semiconductors 1289

3.3.2. nt z NI, LD< 112. When LD< 112 we can no longer assume the conduction band
to be flat and must take account of the fact that n varies continuously across each
grain. T h e band profile is shown in figure 12. There is a real difficulty in this case
to know what the Hall effect would measure, as the approximation of a thin barrier
region used in $2 obviously breaks down. Set0 (1975) implies that the appropriate
quantity is the average carrier concentration though it is not clear that there is any
rigorous justification for this. However, because of the exponential decrease of n with
increasing potential the average value differs by only a small factor from the value at
the centre of the grain, as we now show.
T h e conduction band profile is obtained by solving Poisson's equation. Using the
depletion layer approximation for simplicity we have
E&) = E,O+ az2 (3.10)

where a= q 2 " / 2 ~ . This leads to a carrier density profile:

n ( z )= n(0) exp ( - az2/KT) (3.11)


and an average carrier density obtained by integrating n(z) from x = 0 to z = 1/2 and

(3.12)

If l/2< LD, navZnn(0) which is the case we have already discussed. On the other
hand, if 112 2 3 L D the integral approximates to ( ~ / 2 ) land / ~ navZ (.\/Z%LD/l)n(O)
which implies nav < 0.4 n(0). However, using the relation 2 L ~ / l = ( K T / + b ) 1 / 2 shows
that, if cjb 6 0.5 eV, nav > 0.2 n(0)making it relatively easy to estimate an approximate
value for nav.
T h e activation energy for nav is contained in n(0). We obtain it as we did pre-

-------_

7
------....--.I-
I
I
I
I
I
I 2
-112 0 //2
Figure 12. Band diagram for the case nt N1 and LD < 112. Eco is the bottom of the conduction
band at the centre of the grain.
1290 7 W Orton and M J Powell

viously by using the relationships n(0)= Nc exp [ - (&O- E F ) / ~ Tand


] ( N - nav) =
Nt'[l+ exp (Et-E~)/kT1-1which results again in a quadratic equation for nav:
+
nav2 nav {Nt' - N + KNc exp [ - (xt - x b ) ] )- K", exp [ - (xt - ~ b )=] 0 (3 .13)
where xb = (I$b/kT)and K = (nav/no)x .\/%LD/~. I n the limit where Nt' > N the
solution follows the same lines as in 53.3.1 and we find

no%-----NNc exp [ - (3.14)


(Xt - Xb)l
Nt'-N
showing an activation energy of ( E t - I $ b ) . On the other hand, if Nt'= N we obtain a
different solution :
(3.15)

The activation energy for n,, decreases with increasing N because (from equation
(3.10)) +b = q2NZ2/8~and then drops rather suddenly when N = Nt'. I n fact, it must
drop to zero when N > Nt' because this takes us into the third regime where the
depletion barriers extend over only a small fraction of the grain and n x N. Notice
that if we suppose the mobility to have a thermal activation energy of # J bthe con-
ductivity accna,pxexp (-xt), the term xb cancelling from the exponential in
equation (3.14).

3.3.3. nt <NI. I n this case we are concerned with relatively thin depletion barriers
which result in a thermally activated mobility. T o calculate the size of these barriers
we equate the interface charge to the fixed charge in the depletion layer, i.e. nt=2NW
where W is the depletion layer width. Also, from equation (3.10) #Jb= q2NW2/2~so
we have
+b = q2nt"8eN (3.16)
with nt given by equation (3.5).
Eliminating EF from equation (3 .5) by means of equation (3 . 6 ) with n = N results
in an equation for I$b which we write in the following form:

(3.17)

If Xt$>Xb and E t > 0.3 eV the second term in large square brackets is negligible and
n t z N t , i.e. all the traps are filled and EF lies well above Et at the surface. So long as
this holds we can use the expression #Jb=q2Nt2/8GN to calculate #Jb. Thus, Nt=
1016m-2 and N=1024m-3 gives I$b=O*O23 eV but I$b does not increase without
limit as Nt increases or N decreases. Equation (3.17) implies that once EF falls
below Et, I$b increases only slowly beyond the value E t as illustrated in figure 13.
Comparing the case where I$b varies with N-l in this regime with the case N < Nt'
where it varies linearly with N implies that there is a maximum barrier height for
some value of N . This occurs when N=Nt' as shown schematically in figure 14
(assuming nt= Nt). The maximum barrier is given by
(3.18)

which, for values of Nt= 1016 m-2 and Z= 20 nm, results in #Ib (max) = 0.045 eV.
Larger grains or higher interface state densities may give considerably larger values
but the limit implicit in equation (3.17) cannot be exceeded.
The Hall efect in polyci*ystalline and powdered semiconductors 1291

Figure 13. Plot of the barrier height db against interface trap density Nt calculated from
equation (3,17), assuming a trap depth c t = 0 * 5 eV. This shows how $b tends to be
limited to values only slightly greater than Et.

The importance of the maximum barrier is that it results in a sharp minimum in


mobility when considered as a function of doping level. T h e ratio of maximum to
minimum mobility is just exp (+b (max)/kT) which may easily be as large as several
hundred at room temperature. However, it has to be emphasised that this only
happens for moderate trap densities. Examination of figure 13 shows that if Nt is as
high as 1018 m-2, ,$+c remains close to Et even up to the highest doping levels and p
will therefore be very small-probably too small to measure in most cases.
A brief summary of the results of 53.3 is provided in table 2.

3.4. Experimental results


3.4.1, Polycrystalline silicon. Interest in polycrystalline films of Si has developed
relatively recently but we shall consider results on poly-silicon first because they
provide the most convincing verification of the theory developed in the previous sub-
sections. This stems largely from the existence of a well-developed technology for

N-

Figure 14. Schematic plot of barrier height db against doping level N showing how ab reaches
a maximum when N = Nt.
1292 J W Orton and M J Powell
Table 2. Summary of theoretical results for the three regimes discussed inS3.3.

Regime n tL

po or p1

n(%N
y
,-- LD NNC
2rr -- -~ exp [-(Xt-Xb)] po exp (-Xb)
LD<@ 1 N(-N

depositing Si films and, in particular, for controlled doping. I n the majority of work
on other materials doping levels have been pre-determined by growth conditions and
very little attempt seems to have been made to vary them systematically. Silicon
films have been doped both by ion implantation and by gas phase doping and the
density of donors or acceptors has been varied over a wide range.
Kamins (1971) performed Hall effect measurements on Si films grown by thermal
decomposition of silane onto Si02 surfaces. Both n- and p-type layers were grown
using arsine or diborane as gas phase dopants or by diffusion from doped oxide films,
Doping levels were in the range 1023-1025 m-3 and in all cases there was a marked
fall in ,u with decreasing carrier density consistent with theoretical predictions for the
regime n( < N .
Cowher and Sedgwick (1972) also measured vapour-deposited poly-silicon films
doped n- or p-type during growth. They covered a wider doping range and observed
a dramatic drop in free carrier density when the doping level fell below about 1025 m-3,
Less than one order reduction in N produced four to five orders reduction in n (or p).
They also found a sharp minimum in p for their p-type samples which occurred at
about the same doping level as the reduction in carrier density, i.e. at N z 1024 m-3,
At low carrier levels the mobility showed values close to bulk single crystals, i.e.
p z 0.02 m2 V-1 s-1, it dropped to about 5 x 10-5 m2 V-1 s-1 at the minimum and
finally increased to 10-3 m2 V-1 s-1 at higher doping levels. Rather similar results
were observed on n-type films though the evidence was less clear.
I t is unfortunate that neither set of workers measured the temperature dependence
of n or ,u so we have no direct estimate of barrier heights. However, the mobility
reduction observed by Cowher and Sedgwick suggests a value of +b (max)zO*15eV
and combining this with the reported grain size, lz 0.1 pm, yields the reasonable
value of N ~ 7zx 1015 m-2. T h e very large drop in carrier density below N E 1024 m-3
can be interpreted in terms of an activation energy of about 0.3 eV which suggests a
trap depth of perhaps 0.4 eV but these are only rough estimates.
T h e most complete study of Hall effect and resistivity measurements on poly-
silicon was reported by Set0 (1975). Two important features of this work were the
use of ion implantation to introduce known levels of dopant (p-type) and the measure-
ment of the temperature dependence to provide thermal activation energies.
I n figures 15 and 16 we reproduce Seto's experimental results for the variation of
carrier concentration and Hall mobility with doping concentration. Figure 15 clearly
demonstrates the steep drop in hole density, as the doping level falls below 1024 m-3,
which is predicted by the theory of $93.3.1 and 3.3.2. Similarly, figure 16 shows the
predicted sharp minimum in p which occurs at the same doping level. T h e full
curves in both figures are calculated from theoretical considerations closely similar
to those of $3.3.
The Hall effect in polycrystalline and powdered semiconductors 1293

Figure 15. Variation of carrier concentration with doping level for p-type polycrystalline films
of silicon showing the steep fall in p for doping levels below about 1024 m-8 (from
Set0 1975). x , experimental; -, theoretical.

-E
N

.-c-
n
E
+I
% 10-3-

10-L
102' ioz3 1oZ5
Doping concentration
Figure 16. Variation of hole mobility with doping level for p-type silicon films showing a
sharp minimum at a doping level of about loz4m-8 (from Set0 1975). x , experi-
mental; -, theoretical; ---, single-crystal silicon.
1294 J W Orton and M J Powell

Hall mobilities for the four samples of doping level N = 1 x 1024-5 x 1025 m-3
were shown to be thermally activated with activation energies roughly proportional
to N-1 as expected from equation (3.16). Assuming these activation energies corres-
pond to the band-bending q$, it is possible to calculate interface state densities from
equation (3.16) (assuming nt=Nt). Values of 41,and Nt are given in table 3. Also
shown in table 3 are values of po determined from the relation p = po exp ( - +b/JZT)
(equation (2.19)) which are to be compared with poz 1.5 x 10-2 m2 V-1 s-1 calculated
from equation (2,20), using the experimentally measured grain diameter 1z 20 nm.
The sudden jump in po as iV decreases from 5 x 1024 to 1 x lO24m-3 is unexpected
and, at first sight, difficult to understand (we discuss this point briefly in $3.4.2).
Having obtained a value for Art it is now possible to explain the results in figure 1.5.
For the lowest doped sample N = 1 x 1022 m-3 the Debye length is 30 nm, roughly
three times the grain radius so there can be very little band-bending and the theory of
$3.3.1 should apply. Using equation (3.9) we derive a value for the trap depth 1 of
0.32 eV. Set0 quotes a somewhat larger value, 0-37eV, which may mean he has
allowed for a small amount of band-bending but it is not clear from his paper. There
is a difficulty in explaining the rather large experimental values of activation energy

Table 3. Values of barrier height, interface trap density and po for four samples of poly-
crystalline silicon (from Set0 1975).

N (m-3) 1 x 1024 5 x 10L 1 x 1025 5 x 1025


$b (eV) 0.15 0.0335 0.022 0.005
N t (in-2) 2.98~1016 3.41 x10lfi 3.64~1010
p (m3 V 1 s-1) 1.23 x 10-4 I 2 x 10-4 1s x 10-4 27 x 10-4
po (ni2 V-1 s -1) 4.7x10-2 4.5~10-3 4 . 3 ~ 1 0 - ~ 3.3~10-3

for the resistivity at low doping levels. According to the argument at the end of
$3.3.2 this energy should be E t whereas the value measured at N = 1 x 1022 m-3 is
slightly over 0-5 eV.
There are, therefore, one or two difficulties when theory and experiment are com-
pared in detail but the work of Set0 has considerably improved our understanding of
transport properties in polycrystalline films. The importance of measuring grain size,
controlling and measuring the doping level, together with measuring the temperature
dependence of both resistivity and Hall coefficient, is clearly seen.
Before leaving the subject of poly-silicon we make brief reference to more recent
work by Yaron (1979) and Baccarani et aZ(1978) who studied phosphorus-implanted
films. Both sets of workers found the resistivity to be thermally activated with an
activation energy dependent on doping level. Baccarani et a1 found E,,zO..55 eV for
N = 2 x 1023 m-3 which they interpreted as representing the interface trap energy et.
I n both papers E,,decreased approximately as N-1 for AT>1024 m-3 and Nt was esti-
mated to be 8 x 1016 and 4 x 1016 m-2, respectively. Though Baccarani et al measured
Hall effect on some of their samples this was only at one temperature, thus precluding
a more detailed analysis. Finally, Ghosh et al (1980) have summarised the experi-
mental work on silicon (including previously unpublished results) and show there
exists a clear correlation between the doping level for minimum mobility and grain
size, i.e. 106/Z ( N in m-3, 1 in m) as predicted by equation (3.18).
The Hall epect in polycrystalline and powdered semiconductors 1295

3.4.2. Other materials. T h e electrical properties of a wide range of other semiconduct-


ing thin films have been measured and in this subsection we shall comment on
aspects of these results. T h e fact that electrical properties are sensitive to preparative
conditions and to ambient conditions after film growth makes it very difficult to draw
general conclusions. However, we shall try to examine the experimental data against
the background of the theory outlined in 402 and 3.3.
T h e first conclusion to be drawn from a study of the appropriate literature is that
the Hall mobility is almost always found to be thermally activated. This provides
strong evidence for the existence (and importance) of inter-grain barriers which con-
trol electrical conductivity. I n quite a number of cases (Ma and Bube 1977, Chan
and Hill 1976, Shigeta et a1 1976, Ling et aZl972, Kubovy and Janda 1977, Okuyama

1IT -
Figure 17. Schematic plot of lg (mobility) against reciprocal temperature observed on many
polycrystalline films. In the high-temperature region current flow is assumed to
be limited by thermionic emission over inter-grain barriers, and the low-temperature
region by thermally assisted tunnelling.

1976, Capers and White 1973) a plot of lgp against T-I shows two different slopes as
illustrated in figure 17, the high-temperature region corresponding to thermionic
emission over the barrier, the low-temperature region to thermally assisted tunnelling.
T h e knee occurs typically at about 200 K. I n what follows we shall refer always to
barrier heights $b derived from the slope of the high-temperature portion of the
Arrhenius plot.
I n table 4 we have collected together some representative data on Hall mobilities
for a number of different materials where they have been measured over a range of
temperatures. With only one or two exceptions the barrier heights lie between zero
and about 0.2 eV which is consistent with equation (3.16) for reasonable values of the
parameters I and Nt. However, it should be noted that significantly larger barriers
would probably lead to unmeasurably small values of p. For example, if $bb= 0.4 eV,
then exp (- xb) z 10-7 at room temperature and p may be 10-7 m2 V-1 s-1 or less
83
1296 J W Orton and M J Powell

N
?
0
3 2 vi0
0 a CQ
X X
N m

N N N N N d N

b b b b oI ol ol
33-3113
x x x x x x x
m 3 <? cy vi 3 b
1 r J Q

vi
m cl in
"hlON
"????
0 0 0 0 0 0 0 0 0 0 0 0

a m m m e ~ m m m ~ m m
l l l l l l l l l l l l
0 0 0 0 0 0 0 0 0 0 0 0
3-333 31333333

x x x x x
3CV3viN
II /I I/ /I I/ x x x x x x x x x x x x x x x x
C4JaQiQc: "mc-4vimCQvia3"aylm~
22 3 3 3

n
d w Gb
The Hall effect in polycrystalline and powdered semiconductors 1297

n
m
z
n

a\
b
0
s
s Y
U

Gm U

3
a E
8 B
E k2

N
In
m
c? 3 2* N0 0 ??.e* ..??
Ln

0 0 xi, 0 0 3 O O N 0 0 0
1z '? V
b

m m
N
3

0
3
0
5..
0 0
?
0
Ln
0
0
c??"
ooow
Ln
m-? . .
0 0 0
A\ V

n N N N N N m "" m a N N ~
I I 1 / 1 1 l l
OObbo0o
3 3 3 3 3 3 3 22220 22
/
0 0 0
3 3 3
I /

x x x x x x x x x x x: x x x x x
m ~ m m m y mb m m - ?'? " 3
m 11 3 m

*3
w m $ w 3 mhl
* b
b3w
N
0000.. ?? ?YO
0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0
?1 12 21 11 ??

10 (D m b m d il n n m b d
I I I I I I I l l l l
0 0 0 0 0 0 0 0 0 0 0 0
3 3 3 3 3 3 3 - 3 3 4 3
x x x x x x x x x x x x
wwo\ew?mm N m 3 L n 3
3 3 N

i
N N

-
H I 1 0 0
ZEE"0"O 2 1 x x x x x x s
3 - 3
x x"0
0 0 0 3 3
11 11 22 N m d- m Ln hl 'I, m 3 -
x x x x x x x x x I1 I/ 11
LnN Lnco ymco w m L'LQ,
11 3
1298 J W Orton and M J Powell

whereas the experimental limit usually lies at about 10-5 m2 V-1 s-1. (The one or
two cases in the table where p is reported as being smaller than this were derived
from a combination of resistivity and thermopower measurements.)
Having said that the measured barrier heights are reasonable, one can say rather
little more. There is very little clear evidence for variation of $b with doping level,
for example. Saitoh and Matsubara (1977) report on a range of n-type InP films doped
between and 1025m-3 where $b shows a clear tendency to decrease as N increases,
even though Nt must range between 4 x 1015 and 3 x 1016 m-2 in order to satisfy
equation (3.16). There is a similar qualitative dependence of $b on AT for the n-type
InSb samples measured by Ling et a1 (1972) but that appears to be about the extent
of the evidence.
T h e other major point of interest concerns po and its relation to grain size. Accord-
ing to the argument of $2.3, po should be determined by grain size and should be
given numerically by equation (2.20). This is the origin of the figures in table 4
under the heading po(ca1c). Comparing them with the experimental values shows
reasonable agreement in some cases but also a considerable number of discrepancies.
A striking feature of results from CdS (Wu and Bube 1974) and InP (Saitoh and
Matsubara 1977) is the very large range of po values obtained for samples with nomin-
ally the same grain sizes. Wherever there is a large increase in $b this seems to be
associated with a sympathetic increase in po. We have already noted a similar effect
for poly-silicon in the previous subsection (see table 3). A possible explanation for
this behaviour can be given in terms of a temperature-dependent barrier height such
as has been observed by Seager and Pike (1979) for silicon bi-crystals. Thus, if we
write $b = $o( 1 + a T ) , it is easy to show (Orton et ~ 1 1 9 8 0that
) the measured activation
energy is equal to $0, the barrier height extrapolated to absolute zero. Also, the pre-
exponential factor is proportional to exp ( - a$o/KT) which accounts for the observed
dependence of po on $b. Orton et a1 present experimental evidence on a range of
CdS samples which supports this interpretation.
Note that in some cases po is larger than the single-crystal mobility pc. T h e
measured mobility is still thermally activated provided p c > po exp ( - $b/kT) as we
discussed in the context of equation (2.25).
I n spite of these unexpected features of po there is, nevertheless, evidence that in
some cases the mobility does vary with grain size as expected from equation (2.20).
Capers and White (1973) report on a range of T e samples for which po varies linearly
with grain size (1=0.04-0.2 pm) though the slope of the po against 1 plot is only a
quarter of that predicted by equation (2.20). Other evidence that p varies linearly
with Z for T e films was reported by Okuyama (1976) and for InSb by Ling et aZ(1972)
and Greene and Wikersham (1976). Because of varying barrier heights this does not
translate directly into a linear variation of po--the comparison cannot be made
because the authors do not provide the necessary information on barrier height. A
qualitatively similar dependence of mobility on grain size is also found for films of
CdS (Kazmerski et al 1972), CdSe (Dhere et aZ 1975, 1977) and PbS (Espevik et al
1971).
There is no doubt, therefore, of the importance of grain size in influencing mobil-
ity, though further work is necessary to elucidate details.
Another important aspect of the theoretical model, which is clearly demonstrated
by experiments on poly-silicon, is the strong carrier freeze-out characteristic of samples
with low doping levels. Once again the evidence from other materials is less con-
clusive due to poor control over and lack of knowledge of the doping level.
The Hall effect i n polycrystalline and powdered semiconductors 1299

I n a few cases activation energies of n, En, in the range 0.2-0.5 eV have been obser-
ved for carrier densities below 1022 m-3 whereas at the more commonly occurring
levels of 1023-1025 m-3 n shows very little temperature dependence. T h e majority of
the evidence refers to CdS where Shallcross (1966) measured En~0.5-0.1eV for
n~ 101*-1020 m-3, Mankarious (1964) found En=0*33eV and 0.2 eV for samples
with n = 8 x 1020 and 1 x 1022 m-3, respectively. Muller and Watkins (1964) found
En=0*35eV for n = 3 x 1020 m-3, Wilson and Woods (1973) reported conductivity
activation energies E,,% 0.5-0.29 eV for n~ 1018-1022 m-3 and Wu and Bube (1974)
measured E,= 0.24 eV after compensating highly doped films by Cu diffusion. Other
examples are provided by CuInS2 where Kazmerski et aE (1975) measured E n = 0.29 eV
at n = 2 x 1022 m-3 and E n = 0.23 eV at n = 4 x 1021 m-3 and p-type PbS (Espevik et a1
1971) which showed E , = 0.35 eV after vacuum baking. While some of these results
probably represent the effect of grain boundary trapping it should be borne in mind
that bulk trapping cannot be ruled out as an alternative explanation.

3.4.3. Barrier modulation effects. We have emphasised the value of Hall efiect measure-
ments in interpreting the transport properties of polycrystalline semiconductors and
the study of barrier modulation effects provides a good example of this. I n this
subsection we shall briefly discuss the effects of light, applied electric fields and ambient
gases on these transport properties.
As we have seen, the conductivity of thin films and powders is generally dominated
by the presence of excess charge trapped at the interface between grains. It follows
therefore that if the interface charge density can be changed by an external agency this
may result in quite large variations of film conductivity. Two possibilities exist: in
the regime where nt < N , changing nt will result in a change in barrier height and
therefore modulation of the effective mobility. Jn the regime Nt > N where ne N
changes in nt can be expected to modulate the carrier density n.
Consideration of numerical values shows that these modulation effects can be
rather large. If we suppose p 0 = 3 x 10-2 m2 V-1 s-1 and +b=0*05 eV we find
p=4*2x 10-3 m2 V-1 s-1 at room temperature. A change of nt by a factor of 1.5 in
either sense changes p from 3.3 x 10-4 to 1.3 x 10-2 m2 V-1 s-1, a factor of almost 40.
T h e modulation of n is less easy to calculate precisely but if N t z N it requires only
a small change in nt to move from a non-activated to a strongly activated carrier
density and n may change by several orders of magnitude. However, it must be borne
in mind that the inevitable spread in grain sizes will smear the transition region and
percolation effects will become important. Thus we can envisage a situation where a
particular change in nt results in almost complete depletion of some grains while
having only minimal effects on others. Some grains will be switched from conducting
to non-conducting states while others remain unchanged. As a result, not only will
the overall resistance be increased but the distribution of current paths will be
appreciably modified and this can only be treated in terms of percolation theory.
This makes detailed calculation much more complex but the general conclusion that
n may be very sensitive to changes in nt remains valid.
One of the major applications of thin films is as photoconductors so there has been
interest in photoconductivity since the 1950s and in the early days a considerable
controversy grew up as to whether the measured changes in cr arose from modulation
of mobility or carrier density (see, for example, Espevik et a1 1971). Hall measure-
ments are obviously essential but it is equally important that they can be correctly
interpreted. T h e discussion in $2 provides evidence that (except in the unusual case
1300 J W Orton and M J Powell
of surface accumulation) the measured Hall coefficient relates to the carrier concen-
tration in the grains and the mobility is given by equation (2.19). This allows an
unambiguous interpretation of experimental results, which show that modulation of
either or both mobility and carrier concentration can occur.
T h e mechanism presumably involves minority carriers generated by photon
absorption being swept to grain boundaries where they neutralise charge of opposite
sign trapped there. This reduces inter-grain barriers or increases n as discussed
above. The question of whether n or p is likely to be more sensitive to such effects
clearly depends on the relative magnitudes of N and nt in the initial (dark) situation.
I n general, one may expect much larger changes in n when the dark value of n is small
and larger changes in p when the barrier energy is large. (For the latter case the
exponential dependence on (+b/kT) implies much larger changes in p at low
temperatures.)
Experimental evidence on a number of materials which show behaviour consistent
with the above ideas has been published in the following references: PbS, Espevik
et a1 (1971); CdS, Mitcheletti and Mark (1967), Wu and Bube (1974), Ma and Bube
(1977); CdSe, Fowler (1961); Cd(S0.45Se0.55); Robinson and Bube (1965). A word of
caution must be added, however, to emphasise that the above interpretation is not
unique. Photoconductivity may be sensitised by minority carrier trapping in bulk
states as is well established for single-crystal photoconductors and it is far from easy
to distinguish the two interpretations. Nevertheless an interesting example of photo-
sensitivity which is directly attributable to the presence of inter-grain barriers is
provided by PbTe films (Green and Miles 1973).
There is considerable evidence that polycrystalline films are sensitive not only to
light but also to their environment. I n particular the effect of atmospheric oxygen
has been shown to be important in modifying the conductivity of CdS, CdSe and
PbTe films (see the references above) and Ziebert (1972) and McLane and Zemel
(1971) have investigated the effect of various gases on the Hall mobility in CdSe,
PbSe and PbTe. Ma and Bube (1977) have shown that heating CdS films, prepared
by spray pyrolysis, in hydrogen can increase both n and p by a factor of ten, presum-
ably by removing oxygen from grain boundaries and possibly from the grain material
too. Our own work (B J Goldsmith, M J Powell, J W Orton and J A Chapman,
unpublished), has shown quite dramatic increases in n ( b y z 105 times) resulting from
Hz treatment of CdS and that in certain circumstances both n and p are significantly
reduced by exposing films to the atmosphere at room temperature. The use of a
hydrogen plasma for passivating grain boundaries in poly-silicon has also been des-
cribed by Seager and Ginley (1979) though they did not study the Hall effect.
The third method of modulating thin-film conductivity we wish to describe con-
cerns the use of high surface electric fields. Measurements of Hall mobility as a
function of gate voltage in a specially modified thin-film transistor were reported by
Waxman et a1 (1965) on CdS and by van Heek (1968) on CdSe. By applying a voltage
to the TFT gate, charge is induced at the gate insulator-semiconductor interface and
the resulting field reduces the inter-grain barriers and increases the Hall mobility.
At sufficiently high gate voltages (e.g. 15 V across an insulator of 100 nm thickness)
the inter-grain barriers are reduced to zero and p saturates at a value corresponding
to po (see equation (2.19)). Thus Waxman et a1 found p o ~ x6 10-3 m2 V-1 s-1 for
CdS and van Heek found po z 10-2 m2 V-1 s-1 for CdSe (this latter to be compared
with a value of 5 x 10-2 m2 V-1 s-1 derived from the quoted grain size). The meas-
ured carrier density increased with increasing gate voltage, due to the formation of a
The Hall effect in polycrystalline and powdered semiconductors 1301

strong accumulation layer at the semiconductor surface, and eventually reached


degeneracy.
Different combinations of semiconductor and insulator result in a variety of
different behaviours as observed by Waxman et a1 (1965) and Tyagi (1975) for CdS,
Snejdar et a1 (1973) for CdSe and Ling et a1 (1972) for InSb.

4. The Hall effect in zinc oxide powders

Powdered semiconductors show some obvious similarities to thin films insofar as


they may have similar grain sizes and the surfaces of the grains can accommodate
surface states which trap charge just as do the grain boundaries in films. T h e import-
ant difference between the two lies in the constrictions to current flow imposed by the
rather small grain to grain contacts as we discussed in $2.3.
I n terms of Hall measurements we should expect results rather similar to those
on films, the Hall coefficient measuring the carrier density in the bulk of the grains
while the mobility is limited by inter-grain barriers and the geometrical effects just
mentioned. There is evidence for this in some powder systems but we shall not
discuss it here. Instead we wish to concentrate on one particular powder, ZnO,
because it shows interesting behaviour of a slightly different kind.
Experiments on single crystals of ZnO (Many 1974, Eger et al 1975) have shown
that oxygen adsorption can result in quite strong surface depletion layers, oxygen
atoms capturing electrons from the bulk n-type crystal and being chemisorbed as
0- or 0 2 - ions. Absorption of uv photons generates holes which are swept to the
surface by the depletion layer field, neutralise oxygen ions and can swing the surface
through the flat-band condition into strong accumulation. Positive surface charge
densities as large as 1017 m-2 can be produced in this way giving rise to electron
densities of 1026 m-3 close to the surface, i.e. the surface is strongly degenerate.
T h e corresponding behaviour of ZnO powder-binder layers under uv illumination
shows the importance of these band-bending effects by the large changes of resistivity
which occur (Orton and Powell 1978). A typical dark value of 106 !2 m can be reduced
to about 10-2 C2 m by irradiating with uv in vacuum. Hall measurements (Clarke
et a1 1978) on a large number of irradiated powder layers gave repeatedly the same
values for carrier density and mobility, i.e. n= (3.0 5 0.5) x 1023 m-3, p= (1.1 i-0.2) x
10-3m2V-1s-1 even though we expect the bulk doping level to vary between
different samples.
We can explain this result by reference to the discussion of the Hall coefficient in
$2. As the grain surfaces are expected to be strongly accumulated under the experi-
mental conditions employed, the surface region is more conducting than the bulk and,
according to $2.3, when p2 the Hall coefficient is determined by the properties of
the inter-granular material or, in this case, by the accumulation layer. This is borne
out by the way in which n and p vary with changes in ambient conditions following
illumination and subsequent exposure to air.
I n figure 18 we plot experimental points showing the variation of n with p as the
grain surfaces recover from strong accumulation towards strong depletion. n varies
approximately inversely with p over two orders of magnitude and clearly does not
correspond to the bulk doping level (which remains constant during these surface
changes).
We can account rather well for the observed results in terms of the model outlined
1302 J W Ovton and M J Powell

!- \\

I L.--
10-2 1 102
p i R ml

Figure 18. Variation of effective carrier density n with layer resistivity p for ZnO powder-
binder layers as the grain surfaces recover from strong accumulation to strong
depletion. The full curves are calculated on a composite sphere model for a range
of bulk doping levels N .

at the end of $2.3. Each grain is represented by a composite sphere of core diameter
21, resistivity p l , etc, with a thin outer shell thickness 12/2, resistivity p2, etc, the
relation between the real and the model carrier densities being illustrated in figure 19.
We can then compare the theoretical expressions of equations ( 2 . 3 2 ) , (2.35) and
(2.36) with the measured values of p, n and p,
For grains of diameter 1 ~ 0 . 6p.m we estimate $zO.1 and therefore A z ~ Many .
(1974) has calculated values of surface carrier concentration n2 as a function of surface
charge An (m-2) and also reports Hall mobilities p2 against An as measured on single
crystals of ZnO. We can therefore use these results to test the viability of our powder
model. Table 5 provides a comparison between calculated and measured parameters
following prolonged uv irradiation of ZnO powder and shows satisfactory agreement
considering the simplicity of the model.

Table 5. Comparison of measured and calculated values of n, p and p for ZnO powder-binder
layers after prolonged uv irradiation.

Accumulation layer Calculated parameters Measured parameters


- -_
An = 1017 ni-2 x 1023 m-3
n+ = 5 n = 3 x 1023 m-3
n2 = 1026 m-3 p*=7~10-4m2V-ls-l p=11 ~ 1 0 - 4 m 2 V - l s - l
pz = 20 x 10-4 m2 11-1 s-1 p*=0-025 O m p=0*02 Q m
12 = 2 nm
The Hall effect in polycrystalline and powdered semiconductors 1303

1 I

la I (bl
Figure 19. Relationship between the parameters of the composite sphere and the real carrier
distribution in a powder grain for (a) depleted surface and (b) accumulated surface.
(a) (ni- n2) 1 2 / 2 = 1 (ni- n) dr.
(b) (nz- ni) 12/2 = (n- nl) dr.

These ideas can be extended to calculate the way in which n may be expected to
vary with p as the irradiated layer reabsorbs oxygen and slowly reverts to a state of
surface depletion. The full curves in figure 18 represent the result of the calculation
for a range of bulk doping levels N ranging from 1020-1024 m-3. The different types
of behaviour relate to the ratio of depletion layer width W to grain radius, 1/2. At
high doping levels W<1/2 and in surface depletion the Hall effect measures N even
though p is increasing due to the build-up of inter-grain barriers. At low doping
levels W>1/2, the whole grain is depleted and the free carrier density n < N . There
is very little band-bending so ncc p-1 (i.e. p remains roughly constant). A comparison
between theory and experiment shown in figure 18 suggests that the ZnO grains have
a bulk doping N < 1022 m-3. It is interesting to note that Hall measurements are not
able to measure N under these circumstances but only allow us to estimate an upper
limit.

5. Conclusions

In this review we have shown that the Hall effect provides a valuable method for
understanding the transport properties of powdered and polycrystalline thin-film
semiconductors. Its importance lies in the ability to separate carrier density and
1304 J W Orton and M J Powell

mobility and thereby to distinguish, for example, a thermally activated mobility from
an activated carrier density. I n this respect it plays the same role as in studies of
crystalline semiconductors but with one important proviso. I n the polycrystalline
case the interpretation of Hall effect measurements depends on a model and the only
ones which have so far proved tractable are highly idealised geometrical models which
we discussed in $2. T h e application of results derived from these models to real
materials involves an extrapolation which cannot be rigorously justified. Nevertheless,
an encouraging feature is the fact that predictions based on widely differing geometries
are closely similar, suggesting that precise geometry is unimportant.
These models consider materials containing two phases, a bulk grain having
conductivity al, carrier density nl and mobility p1 which is surrounded by a thin
skin of inter-grain material characterised by 02, n2 and p ~ We . have concentrated
on the two extremes, (i) 01 B a2 and (ii) a1 < u2 and have demonstrated that the pre-
dictions for these two cases apply, at least semi-quantitatively, to the behaviour of
real polycrystalline materials. T h e commonly occurring case corresponds to a1 9 uz
where the high-resistance inter-grain material takes the form of a depletion barrier
resulting from charge trapped in interface states. However, there exists at least one
example of the opposite type where strong surface accumulation on ZnO powder
grains provides a highly conducting skin.
I n considering the behaviour of polycrystalline thin films in more detail it becomes
apparent that the presence of inter-grain depletion barriers is of overriding importance,
giving rise to a thermally activated Hall mobility which is the characteristic feature of
such materials. Linked with this and of almost equal significance is the grain size
and, more specifically, the ratio of grain radius to depletion width. When these are
equal the inter-grain barrier height reaches its maximum value, resulting in a minimum
in the Hall mobility as has been clearly demonstrated by experiments on polycrystalline
silicon films. At low doping levels where NI< Nt the carrier density (as derived from
the Hall coefficient) becomes thermally activated and the ratio n/N decreases rapidly
with decreasing N . This feature has also been demonstrated on polycrystalline
silicon.
Experimental results on other materials such as CdS, CdSe, InP and InSb show
much greater spread though there is a limited amount of evidence that Hall mobility
varies with doping level in the way predicted by the theoretical model, i.e. 4boCN-l
for the case NI> Nt. Some evidence also exists for thermally activated carrier densities
when Nk < Nt but there is clearly a need for further experiments. The central diffi-
culty lies in achieving the same degree of control over doping level in these materials
as has been achieved with silicon films.
An important detail in our understanding of Hall mobility in polycrystalline
materials is the interpretation of the parameter po in the relation p= po exp ( - $b/kT).
According to Petritzs (1956) theoretical prediction, po should be determined by grain
size and should be independent of doping level and y$,, whereas experimentally
observed values show a marked tendency for po to increase as 4 b increases. This can
be explained on the basis of q$, being temperature-dependent.
An interesting question in the interpretation of photoconductivity in powders and
polycrystalline films concerns the relative importance of carrier density or mobility
modulation. Hall measurements provide an essential tool for separating these effects
and have been applied extensively to materials such as CdS, CdSe and PbS. NO
complete theory of their photoconducting properties exists but the Hall effect demon-
strates the importance of inter-grain barriers. Minority carrier trapping at grain
The Hall efJect in polycrystalline and powdered semiconductors 1305

boundaries reduces f#b and thus enhances mobility. This mechanism is more signifi-
cant at moderate to high doping levels while carrier density modulation is likely to
dominate for low doping when N1 c Nt.
By their nature, the geometrical models are highly ordered whereas any real
material must contain inhomogeneities of a random nature. One can cite random
variations in barrier height in thin films or of contact area in powders. I n these
situations electric current will follow the path of least resistance and we are therefore
concerned with percolation. Percolation theories of conductivity have been studied
in some detail and we have given a brief outline in $2. For a random mixture of
conducting and non-conducting material the net conductivity drops to zero at the
percolation threshold, i.e. when the volume fraction of conducting material falls
below a critical value of about 15%. Above this threshold the conductivity is given
by a power law of the form U= ao(p-pc)t with t = 1.65 in three dimensions. Recently
there have been attempts to derive a percolation theory for the Hall coefficient, giving
R= Ro(p-pc)-g which suggests that R should diverge at the percolation threshold
in a three-dimensional system (g= 0 in two dimensions).
Experimental evidence for this divergence is so far lacking as measurements have
only been made at values of p remote from pc. Though there is evidence for a small
increase in R a s p decreases from unity, measurements closer top, become very difficult
because the effective Hall mobility decreases (theory suggests ,UH-+0 a s p + p,). It is
fair to say that the development of a percolation theory for the Hall effect is still in
its infancy.
Our main conclusion is that the use of idealised geometrical models to represent
the Hall effect in inhomogeneous materials has been justified. T h e application to the
study of powdered and polycrystalline semiconductors has played an important role
in improving our understanding of their electrical properties.

References

Adkins C J 1979 J. Phys. C : Solid St. Phys. 12 3389


Ambegoakar V, Halperin B I and Langer J S 1971 Phys. Rev. B 4 2612
Amith A 1960 J. Phys. Chem. Solids 14 271
Anderson J C 1973 Thin Solid Films 18 239
Baccarani G, Ricco B and Spadini G 1978 J. Appl. Phys. 49 5565
Barna A, Barna P B, Bodo Z , Pocza J F, Pozsgai I and Radnoczi G 1974 Thin Solid Films 23 49
Bednarczyk D and Bednarczyk J 1979 Thin Solid Films 59 279
Beekmans N M 1978 J . Chem. Soc. Faraday Trans. I 74 31
Beer A C 1963 Galvanomagnetic EfSects in Semiconductors. Solid St. Phys. Suppl. 4 ed F Seitz
and D Turnbull (New York: Academic)
Berger H, Janiche G and Grachovskaya N 1969 Phys. Stat. Solidi 33 417
Berry W B and Jayadevaiah T S 1969 Thin Solid Film 3 77
Blood P and Orton J W 1978 Rep. Prog. Phys. 41 157
Blount GH, Bube R H and Robinson A L 1970J. Appl. Phys. 41 2190
Bruggeman D A G 1935 Ann. Phys., Lpz. 24 636
Bube R H 1968 AppZ. Phys. Lett. 13 136
Capers M J and White M 1973 Thin Solid Films 15 5
Chan D S H and Hill A E 1976 Thin Solid Films 35 337
Clarke P S, Orton J W and Guest A J 1978 Phys. Rev. B 18 1813
Cohen M H and Jortner J 1973 Phys. Rev. Lett. 30 696
Cowher M E and Sedgwick T 0 1972 J . Electrochem. Soc. 119 1565
Davidson A and Tinkham M 1976 Phys. Rev. B 13 3261
Dawar A L, Krishna K V , Taneja 0 P and Mathur P C 1978 Phys. Stat. Solidi a 47 375
1306 J W Orton and M J Powell
Dhere N G , Parikh N R and Ferreira A 1975 Tlzin Solid Films 36 133
-1977 Thin Solid Films 44 83
Eger D, Goldstein Y and Many A 1975 R C A Reo. 36 508
Eggarter T P and Cohen M H 1970 Phys. Rev. Lett. 25 807
Espevik S, Wu C and Bube R H 1971 J. Appl. Phys. 42 3513
Finney J L 1977 Nature 266 309
Fowler A B 1961 J. Phys. Chem. Solids 22 181
Gat A, Gerzberg L, Gibbons J F, Magee T J, Peng J and Hong J D 1978 Appl. Phys. Lett. 33
775
de Gennes P G 1976 J . Physque Lett. 37 L1
Ghosh .4K, Fishman C and Feng T 1980J. Appl. Phys. 51 446
Giroux G 1963 Can.J. Phys. 41 1840
Goldin E and Juretschke H J 1958 Trans. Metall. Soc. A I M E 212 357
Granquist C G and Hunderi 0 1978 Phys. Reo. B 18 1554
Green M and Miles R E 1973 J. Phys. D : Appl. Phys. 6 L45
Greene J E and Wickersham C E 1976 J. Appl. Phys. 47 3630
van Heek H F 1968 Solid St. Electron. 11 459
Heleskivi J and Sal0 T 1972 J. Appl. Phys. 43 740
Jerhot J and Snejdar V 1978 Thin Solid Films 52 379
Kamins T I 1971 J. Appl. Phys. 42 4357
Kazmerski L L, Ayyagari M S and Sanborn G A 1975 J . Appl. Phys. 46 4865
Kazmerski L L, Berry W B and Allen C W 1972 J. Appl. Phys. 43 351 5
Kirkpatrick S 1973 Rev. Mod. Phys. 45 574
-1978 A I P Conf. Proc. 40 99
Kubovy A and Janda M 1976 Plays. Stat. Solidi a 36 101
-1977 Phys. Stat. Solidi a 40 225
Landauer R 1952 J. Appl. Phys 23 779
Le Contellec M , Richard J and Henaff J 1976 Thin Solid Films 36 151
Levinshtein M E , Shklovskii B I, Shur M S and Efros A L 1976a Son. Phys.-JETP 42 197
Levinshtein M E , Shur M S and Efros A L 1976b Sou. Phys.-JETP 42 1120
Ling C H, Fisher J H and Anderson J C 1972 Thin Solid Films 14 267
Lipskis K, Sakalas A and Viscakas J 1971 Phys. Stat. Solidi a 4 K217
Ma Y Y and Bube R H 1977 J . Electrochem. Soc. 124 1430
McLane G and Zemel J N 1971 Thin Solid Films 7 229
hlankarious R G 1964 Solid St. Electron 7 702
Many A 1974 Crit. Reo. Solid St. Sci. 4 515
Many A, Goldstein Y and Grover N B 1965 Semiconductor Surfaces (Amsterdam: North-
Holland) chap 8
Mathew M G and Mendelson K S 1974J. Appl. Phys. 45 4370
Meskauskas A and Viscakas J 1976 Thin Solid Films 36 281
Mitcheletti F B and Mark P 1967 Appl. Phys. Lett. 10 136
Muller R S and Watkins B G 1964 Proc. IEEE 52 425
Nasby R D and Thompson J C 1970 J. Chem. Phys. 53 109
Okuyama K 1976 Thin Solid Films 33 165
Okuyama I( and Kumagai Y 1973 Japan. J. Appl. Phys 12 1884
Orton J W, Goldsmith B J, Powell M J and Chapman J A 1980 Appl. Phys. Lett. 37 557
Orton J W and Powell M J 1978 Phil. Mag. B 38 491
Petritz R L 1956 Phys. Rec. 104 1508
Pike G E and Seager C H 1974 Phys. Reo B 10 142 1
Powell M J 1980a Powder Technol. 25 45
-1980b Phys. Rev. B 21 3725
Putley E H 1960 The Hall EfJect and Related Phenomena (London: Butterworth)
Robinson A L and Bube R H 1965 J. Electrochem. Soc. 112 1002
Romeo N, Sberveglieri G and Tarricone L 1978 Thin Solid Films 55 413
Saitoh T and Matsubara S 1977 J . Electrochem. Soc. 124 1065
Scher H and Zallen R 1970 J . CJiem. Phys. 53 3759
Seager C H and Ginley D S 1979 AppZ. Phys. Lett. 34 337
Seager C H and Pike G E 1974 Phys. Rev. B 10 1435
-1979 Appl. Phys. Lett. 35 709
The Hall eflect in polycrystalline and powdered semiconductors 1307

Set0 J Y W 1975 J . Appl. Phys. 46 5247


Shallcross F V 1966 Trans. Metall. Soc. A I M E 236 309
Shigeta J, Kotera N and Tetsu 0 19763. Appl. Phys. 47 621
Shikalgar A G and Pawar S H 1979 Solid S t . Corninun. 32 361
Shklovskii B I 1977 Sov. Phys.-JETP 45 152
Shklovskii B I and Efros A L 1976 Sov. Phys.-Usp. 18 845
Skal A S and Shklovskii B I 1975 Sou. Phys.-Semicond. 8 1029
Snejdar V and Jerhot J 1976 Thin Solid Films 37 303
Snejdar V, Jerhot J, Berkova D and Zboncak M 1973 Phys. Stat. Solidi a 15 691
Straley J P 1976J. Phys. C: Solid St. Phys. 9 783
-1977 Phys. Rev. B 15 5733
-1978 A I P Conf. Proc. 40 118
-- 1980J. Phys. C: Solid St. Phys. 13 819
Thomas PA, Sebenne C and Balkanski M 1970 Reo. Phys. Appl. 5 683
Tyagi M S 1975 Phys. Stat. Solidi a 30 609
Volger J 1950 Phys. Rev. 79 1023
Waxman A 1966 Solid St. Electvon. 9 303
Waxman A, Henrich VE, Shallcross F V , Borkan I-I and Weimer P K 1965 J . Appl. Phys.
36 168
Webman I, Jortner J and Cohen M H 1975 Phys. Rev. B 11 2885
-1976 Phys. Rev. B 14 4737
-1977a Phys. Rev B 15 1936
-- 1977b Phys. Rev. B 16 2593
Wilson J I B and Woods J 1973 J . Phys. Chem. Solids 34 171
Wu C and Bube R H 1974 J . Appl. Phys. 45 648
Yaron G 1979 Solid St. Electron. 22 1017
Yoshikawa A Kondo R and Sakai Y 1973 Japan. J . Appl. Phys. 12 1096
Zallen R and Scher H 1971 Phys. Rev. B 4 4471
Ziebert V 1972 Thin Solid Films 11 53

You might also like