You are on page 1of 19

Earth.

Science Reviews
Elsevier Publishing Company, Amsteldam - Printed in The Netherlands

The Microstructure of Clay Sediments


C.F. Moon

ABSTRACT

Moon, C.|;., 1972. The microstructure of clay sediments. Earth-Sci. Rev., 8:303- 321.

The microstructure of clay sediments has been of minor interest to geologists until very recently.
Many of the major advances in the study of clay fabric have been provided not by scdimentologists or
petrologists but by workers in the fields of soil mechanics, pedology, ceramics and colloid chemistry,
and although the findings are not always directly applicable to clay sediments they have proved
extremely useful. A review of the theories of microstructure past and present has been prepared and it
has emerged that, probably due to the almost insuperable difficulties associated with sampling, there
have been certain misconceptions regarding what is meant by a freshly deposited sediment. Ill this
review the structure of a fresh clay is described and an attempt has been made to explain it. Micro-
structural changes produced by compaction are discussed and it appears that single, discrete particles
are of little importance in clay sediments either fresh or compacted, but rather that the particles are
organised into domains of a varying number of oriented clay particles. This concept is incorporated
into the study of the development of fissility in shales in which it appears that a fissile nature in
argillaceous rocks is more a function of the chemistry of the depositional medium than of overburden
pressure. In the light of previous investigations an attempt has been made to produce a revised system
of classification of particle association in fresh and compacted clay sediments.

INTRODUCTION

The study o f the microstructure o f clay sediments is divisible into two main headings:
the microstructure o f (1) freshly deposited clays and (2) c o m p a c t e d clays. Relatively
little w o r k has been devoted to the fabric o f these sediments c o m p a r e d with, for example,
sandstones. Considering that argillaceous deposits account for a m u c h greater p r o p o r t i o n
o f the stratigraphic c o l u m n than do arenaceous sediments this may be thought surprising.
However, a cursory exfimination o f published w o r k shows that the obvious arresting
factor in the study o f clay microstructure has been the difficulty e n c o u n t e r e d in investi-
gating the very fine clay material until the advent o f the scanning electron microscope.
Because o f this, it appears to be true that, as Burnham (1970) states:
" . . . m a n y geological a u t h o r s . . , have been c o n t e n t to describe the larger particles,
and to dismiss the finer materials as a featureless intergranular paste of very finely
crystalline minerals."
Probably the major impetus for recent clay-fabric research has been provided n o t by
geologists but by soil scientists, civil engineers, ceramicists and to a lesser e x t e n t colloid
"chemists. A l t h o u g h m u c h o f this i n f o r m a t i o n is n o t directly applicable to the geological
304 C.F. MOON

study of clays both fresh and compacted, much pertinent information can be glean'ed
from these various approaches. Indeed, the study of clay microstructure has recently
become an interdisciplinary topic. A comprehensive account of clay fabric has been
provided by Brewer (1964) who gives a detailed system of terminology for use in the
description of soil microstructure. This system has been readily adopted by soil scientists
but has not been used widely in the geological or soil-mechanics literature. A summary of
Brewer's terminology is given in Osmond and Bullock (1970, Glossary). The structure of
clay soils seen from a soil-mechanics point of view and the influence of shear and com-
pression on fabric is considered in detail by Ingles (1968).
Prior to the application of electron microscopy to clay fabric research, the methods of
study were at best inadequate. High-power optical microscopes have been used, but due
to the limit of their resolution, one can never be certain whether it is single particles that
are being viewed and not aggregates (see Tchalenko, 1968). Much use has been made of
X-ray diffraction methods for the study of microfabric and this relies on the intensity of
basal reflections from clay particles to measure parallelism or randomness. This has been
found to be very useful in some circumstances but because of its two-dimensional inter-
pretation, is capable of giving some grossly misleading results. Light scattering has
recently been applied to the study of dilute clay suspensions (Schweitzer and Jennings,
1971 ), and this is considered here on p. 309.
The electron microscope has proved to be extremely useful in this field and, especially
with the introduction of the scanning electron microscope, theories are being proved or
disproved by the direct observation of the sedimentary fabric (Gillott, 1969).
The fact that clay minerals have distinct physico-chemical properties that are not
duplicated by other sedimentary minerals has a profound effect on clay fabric. These
properties have received much attention from previous authors in this field (Grim, 1953,
1962; Gillott, 1968) and comprehensive summaries are given by Meade (1964) and
Morgenstern (1969); consequently this will not be dealt with in detail in this review.
The important properties of clay minerals which differentiate them from other clastic
minerals are their surface and internal chemistry which enables them to form charged
particles, the ability to adsorb water layers, their capacity for cation exchange, their platy

Colin F. Moon graduated from London University in 1970


where he obtained an honours degree in geology. In the fol-
lowing year he undertook a Masters degree in engineering
geology at the University of Leeds and began a study of clay
microstructure. After graduating from Leeds he joined the
staff of Sheffield Polytechnic where he is currently under-
taking research into the microstructure of clay soils. His main
interests are in the fields of clay sedimentology and glacial
sediments.
MICROSTRUCTURE OF CLAY SEDIMENTS 305

crystal habit and last but not least, their extremely small size. Lambe (1958)quotes the
following sizes for the most common clay mineral species: typical montmorillonite
particle, 1,000 A by 10 A thick; typical kaolinite particle, 1 /~m by 0.10/am thick (i.e.,
10,000 A by 1,000 )~ thick). Illite crystals are of a similar size order to kaolinite crystals.
It will be appreciated that all the properties of clay minerals listed above are highly
interdependent.
The suggestion that, in general, clay particles are plate-shaped and have negatively
charged crystal faces (and under certain conditions of pH, positively charged edges) has
been incorporated in several theories of clay microstructure. The prime example of this
being the attraction of negative face to positive edge giving rise to the much quoted
"house-of-cards" structure (Lambe, 1958). This theory and many others will be discussed
in the following review.

THE MICROSTRUCTUREOF FRESH CLAY SEDIMENTS

A fresh clay deposit is taken here to mean that which has not been compacted by the
pressure of superincumbent sediments; in a sedimenting system this will be that deposit
lying at the depositional interface. Because of the difficulty involved in sampling these
types of sediment the majority of microstructural studies have been purely theoretical
and thus, as will be seen in the following section, nearly all experimental study has been
devoted to compacted (or in the engineering sense, consolidated), clays. Some authors
have suggested that misconceptions may have arisen from sampling and the observational
difficulties that are involved. Gipson (1966) writes:
"Because of the difficulties involved in making detailed textural and petrographical
studies of fine grained sediments, many of the present theories regarding such sediments
are based on gross generalizations."
Even in the most recent studies using model clays of pure kaolinite or illite, the
samples are usually compacted in order to expel excess pore water and thus make them
easier to handle in subsequent study. As Greene-Kelly and Mackney (1970) state, even
small stresses have a large effect on particle orientation, and it will be appreciated that
any clay once buried or subjected to loading will have suffered some degree of structural
change, and therefore it is not feasible to regard the fabric of these clays as "original".
Meade (1966) suggested this very fact when he wrote:
"Some recent experiments on clay-mineral slurries and pastes indicate that, given
enough water in the initial sediments, most of the preferred orientation produced during
the initial stages of compaction might be expected to develop under the first few meters
of overburden."
Because of the complex electrochemistry of the clay minerals, the spatial arrangement
of the sedimented particles (i.e., the fabric) is vastly influenced by the chemical composi-
tion of the solution from which they are deposited. In this way it would be expected that
marine clays (deposited in electrolyte-rich conditions) would have a different fabric from
those deposited in fresh water.
306 C.F. MOON

Probably the first theoretical attempt at describing clay fabric was that of Terzaghi
(1925) who listed a number of hypothetical soil fabrics ("Strukturarten"). Among these
was the "honeycomb" structure formed in a marine clay in which clay particles form
chain-like arrangements of flocs into which silt particles are enmeshed. This was taken up
by Casagrande (1932) who suggested how this fabric may be modified during compac-
tion.
Urbain (1937) attempted to reconstruct the environment of deposition of several
argillaceous rocks using optical microscopy to study microstructure. As the sediment was
lithified and thus subjected to a certain degree of compaction his conclusions are not
wholly tenable in a discussion of fresh sediments.
Since the work of Terzaghi and Casagrande the subject of particle association was
largely neglected until the extensive studies of Lambe (1958) who envisaged three main
modes of clay structure. These were: (1) the "salt-flocculated" structure formed in an
electrolyte-rich solution; (2) the "non salt-flocculated" structure produced in electrolyte-
free conditions; and (3) the "dispersed" structure produced during the presence of
dispersing agents (Fig. 1).

c
Eig. 1. Sediment structures, after Lambe (1958). a. Salt-flocculated. b. Non salt-flocculated, c.
Dispersed.

Lambe developed this theory to account for marine-deposited and fresh-water clays
and introduced the idea that undisturbed marine clay had a very random, open structure
dominated by edge-to-face particle contacts. This structure has subsequently been termed
the "house-of-cards" or "cardhouse" structure and has been adopted by numerous
workers in this field. Among them is Tan (1959) who gave a schematic three-dimensional
impression of cardhouse fabric. Rosenqvist (1962) produced transmission electron micro-
graphs (although somewhat unclear) as proof of the existence of the cardhouse structure
having earlier (1953) incorporated this structure into his theory of quickclay formation.
The researches of Rosenqvist are of great importance as most of the opinions on the
fabric of recently deposited clays were formed on the basis of his transmission electron
MICROSTRUCTURE OF CLAY SEDIMENTS 307

microscopy of Scandinavian quickclays. However, none of the clays he studied were


"freshly sedimented" in the fullest meaning of the term, and although they were either
Recent or Quaternary in age were nevertheless sampled from depth and thus could have
undergone some structural reorganization due to this overburden pressure. His results
were in accordance with those of Lambe in that he identified the cardhouse structure in
marine clays and found fresh-water clays to have a greater degree of parallel orientation
of particles, the more open, random arrangement of the marine clays being due to the
flocculating effect of sea water.
Thus the cardhouse structure was, and still is, strictly adhered to, and many recent
authorities using transmission electron microscopy have produced evidence to prove its
existence. These studies, however, have been largely confined to the quickclays of Eastern
Canada and of Scandinavia which may not be regarded as typical clay sediments (see
p.311) and neither are they freshly deposited (O'Brien and Harrison, 1967). Another
factor worthy of consideration is that the transmission electron microscope really only
gives a two-dimensional impression of structure and, as Smart (1967) has shown, even
replication techniques can be grossly misleading. The only truly three-dimensional view is
provided by the scanning electron microscope.

o b

c d

Ill ,,

e f g
Fig. 2. Particle association after Van Olphen (1963). a. "Dispersed" and "deflocculated". b. "'Aggre-
gated" but "deflocculated". c. EF flocculated but "dispersed". d. EE flocculated and "aggregated". e.
EF flocculated and "aggregated". f. EE flocculated and "aggregated" g. EF and EE flocculated and
"aggregated".
308 C.F. MOON

Many authors have studied the fabric of "undisturbed" clays and noted the changes" of
fabric after remoulding and compaction. Such a study is that of Mitchell (1956) who
observed, by optical microscopy, the structure of several natural clays. Although the
structure may have been non-remoulded it was certainly not undisturbed as the samples
were not freshly deposited.
Van Olphen (1963) approached particle association in clay suspensions from a colloid
chemical standpoint and envisaged seven major particle arrangements (reproduced in
Fig. 2). It is to be noted that this scheme reveals that the structure is not as simple as that
offered by Lambe. In addition to the flocculated and dispersed structures, Van Olphen
introduces the "aggregated" mode. In Fig. 2b dispersed units form aggregates having a
parallel oriented structure. The various structures produced in a clay sediment will be
largely governed by electrolyte concentration which will determine which of the three
modes of association will occur: face-to-face (FF), edge-to-face (EF) and edge-to-edge
(EE). Van Olphen states that the results of these three types of interaction will be quite
different; FF associations will lead to thicker and probably larger flakes (or flake aggre-
gates) whilst EE and EF associations will cause voluminous cardhouse structures. Un-
fortunately, although Van Olphen's ideas have been widely accepted, there has been little
experimental evidence to support or disprove it as the practical difficulties involved in
studying particles suspended in solution have proved almost insurmountable. Recently,
however, a new technique has been utilized by Schweitzer and Jennings (1971). Tiffs
involves measuring the change in scattered light intensity from a colloidal suspension
upon the application of an alternating current. The authors conclude that certain of Van
Olphen's structures appear very likely to exist, especially that shown in Fig. 2c, and they
believe that their study supports the existence of the cardhouse model.
An aggregate structure was proposed by Von Engelhardt and Gaida (1963) for clays
deposited in electrolyte-rich (i.e., marine) conditions. However, their aggregates were
composed of units of large numbers of particles having edge-to-face contacts said to

Fig. 3. The aggregate structure of Von Engelhardt and Gaida (1963).


M1CROSTRUCTURE OF CLAY SEDIMENTS 309

completely break down under pressure (Fig. 3).


In 1960 Aylmore and Quirk introduced the concept that lightly compacted clays were
composed of "domains" or "turbostratic groups" (sometimes referred to as "tactoids")
of oriented packets of clay flakes in turbulent array. This received great support in
subsequent investigations and will be discussed in more detail in the following section,
but little attention has been paid to the possible existence of domains in freshly deposited
clays. It is of interest to note that Van Olphen's "aggregated" but "deflocculated" struc-
ture (Fig. 2b) strongly suggests domain-like units. Meade (1964) suggests that for pre-
ferred orientation to occur in naturally compacted clays, a partly oriented fabric may" be
required at the onset of compaction. That domains do form with increasing pressure now
appears indisputable and it seems that with increasing pressure more particles are incor-
porated into each domain. In this case it seems feasible to assume that packets of a few
flakes are present in fresh clays. Burnham (1970) recognized ill-defined domains in a
Recent marine alluvium from Somerset. However, they cannot really be considered fresh
by the present criterion as they were sampled from a depth of 1.6 m.
There have been few studies of the fabric of freshly deposited clays as defined in this
review. O'Brien (1970)prepared flocculated clays in the laboratory and freeze-dried the
samples to preserve the structure. They were then examined by scanning electron micro-
scope. O'Brien concludes:
"It (the fabric) consists mainly of an open framework of randomly oriented groups of
face-attracted flakes in turn joined in an edge-to-face or edge-to-edge arrangement" (i.e.,
randomly oriented domains).
An extremely interesting study is that of Bowles et al. (1969)in which clay sediment
samples from the Gulf of Mexico were subjected to various loads in an oedometer and the
fabric examined by transmission electron microscopy. It was observed that packets of
oriented particles were present in a sample that had been subjected to a very low pressure
(0.5 kg/cm 2). As a result of this, the authors suggest that the packets may be, in fact, an
original feature of the clay fabric. In other words, they were present before the applica-
tion of pressure. This may be regarded as pure speculation but this conclusion strongly
supports the results of experiments carried out by the present author. In these experi-
ments, kaolinite was sedimented in fresh water directly onto specimen mounts as used in
the scanning electron microscope. The microstructure of the clay deposit so produced
was studied and distinct domains observed which were composed of two to three particles
per packet (Fig.4). This lends support to the observations of both O'Brien and Bowles et
al. that packets are present in unconsolidated clay sediments.
It has been considered that, although the various structures postulated by Van Olphen
have not been observed in clay deposits, the work of Schweitzer and Jennings has
provided some support for at least one of them in clay suspensions. A recent paper by
Barden (1972) suggests that single-plate structures such as those of Van Olphen are only
relevant in dilute colloidal suspensions and in dense clay systems (i.e., clay >> water),
only multiple-plate structures occur. Considering the evidence before us, this seems
perfectly feasible. The overriding question is why? It can only be hoped that further
310 C.F. MOON

research will resolve this intriguing problem.


It could be suggested that the formation of domains is initially a purely random
process; a few localized patches of parallel particles causing the orientation and formation
of other similar patches. The author has attempted to demonstrate this by plotting an

Fig. 4. Scanning electron micrograph of kaolinite deposited in fresh water showing small packets of
oriented particles (i.e., domains).

initial random arrangement of particles and simulating compaction by gradually reducing


the vertical dimension of the random plot. From this it was found that small oriented
patches do occur which, in turn, cause the formation of other similar patches. This
method however, takes no account of interparticle forces and as there is no evidence as
such to support this approach, it must be regarded as pure speculation.
To summarize, it can be inferred that the fabric of freshly deposited clay (i.e., that at
the depositional interface) has a structure composed of a random arrangement of small
packets (domains) of clay particles in parallel orientation and these probably contain a
low number (t~vo or three) of clay flakes. There is thus decreasing evidence for the
M1CROSTRUCTURE OF CLAY SEDIMENTS 31 1

existence of the true cardhouse structure as described by Lambe.


It was stated above that the concept of the cardhouse structure has been utilized to
explain the anomalous behaviour of the quickclays of Eastern Canada and Scandinavia
that are prone to become spontaneously liquefied. The Rosenqvist theory (1953), which
provides the basis for most of the current theories, is that these clays were Post- or
Late-Glacial marine clays having an initially stable cardhouse structure. Due to the later
isostatic uplift of Scandinavia, the deposits were raised above sea-level and subsequently
the stabilizing ions were leached out by the action of fresh water causing the cardhouse
structure to remain but to become extremely unstable. This theory accounts for the
behaviour of many Scandinavian clays but becomes inadequate when considering the
Leda Clay of the St. Lawrence River area in Canada (Penner, 1965).
Swedish workers, notably Pusch (1966, 1970) and Soderblom (i966), have suggested
that the unstable nature of quickclays may be due to the presence of organic substances
which would act as dispersing agents: the surface charges on the clay particles would be
effectively neutralized, thus eliminating particle attraction. Pusch (1970) has advanced a
model for the fabric of Swedish quickclay that is strongly reminiscent of the honeycomb
structure of Casagrande. From a series o f electron micrographs (transmission), he
observed a network of small aggregates with particle links. On shearing, these broke down
and formed groups resembling domains.
The factor of highest interest concerns the fact that the actual clay-mineral content of
these clays is, in many cases, very low. In fact, many mineralogical analyses of quickclays
show that the quartz content is surprisingly high (Bjerrum and Rosenqvist, 1957; Berry
and JOrgensen, 1971).
Smalley (1971) has advanced a theory which takes a completely new look at these
enigmatic clays. He claims that the clay minerals play a very minor role in these deposits
and that they are in fact composed of predominately clay-size (i.e., less than 2/am) quartz
particles derived from the glacial abrasion of crystalline Shield rocks. These very fine
quartz particles would have none of the long-range bonding prevalent in real clays but
would have short-range, inactive bonds giving rise to spontaneous failure, the quickclay
slide being propagated in much the same way as are flow-slides (Bishop et al., 1969).
Thus according to Smalley, quickclays are "clays" by virtue of their having particles
predominately in the clay size range (less than 2/am) but they are not clays in the sense
that they contain few clay minerals.

THE MICROSTRUCTUREOF COMPACTEDCLAY

It is now generally accepted that the effect of a superincumbent load on a clay


sediment produces an increase in the parallelism of the particles. A considerable amount
of research has been carried out in this field from a soil mechanics point of view by both
loading and compacting (monomineralic) clays and interpreting the results by various
methods. It should be noted here that the soil mechanics definition of compaction is
somewhat different from that used in the geological sense. Scott (1965)defines compac-
tion as:
312 C.F. MOON

"Artificial increase in the dry density of a granular soil by mechanical means such as
rolling the surface layers... " This implies that in most compactive methods used in soil
mechanics, there is generally a certain amount of remoulding of the clay. Probably the
term "consolidation" as used by soil mechanics experts approximates better to the
geological meaning of compaction although it must be borne in mind that in the geologi-
cal context "consolidation" means something quite different. In this review the term
compaction will be used in the geological sense, i.e.:
"Decrease in volume of sediments, as a result of compressive s t r e s s . . . " (American
Geological Institute, 1962). In one or two instances the soil mechanics definition applies
to the text and where this is so, it is stated.
In an extensive study of clay sediment compaction, Weller (1959) raises a highly
pertinent point with respect to the use of soil mechanics data in geological theory. He
states:
"The results of compression tests on mud are presented mostly in engineering publica-
tions . . . . . . there is reason to doubt that such laboratory tests show results exactly
comparable with the natural compaction of mud because the destruction of a sediment's
original structure may permit compaction to proceed farther under comparable pressure
in the laboratory than in nature."
The first notable study of clay sediment compaction due to increasing burial is that of
Sorby (1908). He compacted kaolinite by applying a pressure of 30 tons per square inch
equivalent to 75,000 ft. of overburden and noted the resulting voids ratio. In addition he
measured the "percentage cavities" and water content of settled samples with increasing
mechanical compaction. On leaving a kaolinite suspension to settle for over a year he
found it still to contain 75% water.
A later study by Hedberg (1936) provided a comprehensive account of the compaction
of clays and shales and considered compaction as a process that results in a sediment
passing from a watery mud to a dense shale or slate. He envisaged four major stages in the
progressive compaction of clays:
(1) Rearrangement of particles producing a new packing mode of greater stability and
expulsion of the majority of free water.
(2) Expulsion of adsorbed water, i.e., that electrochemically bound to the particles.
(3) Particles come into contact due to stage (2) and further reduction in volume is
only possible by mechanical deformation of the particles.
(4) Reduction of voids below about 10% is an extremely slow process and needs great
pressure; under this pressure molecular transformations occur and voids reduction is
accomplished by recrystallization.
Hedberg does not consider fabric changes but alludes to them in the early stages of his
compaction system. He states that the initial sediment of high porosity (according to
Boswell, 1961, this is about 70-90% in fresh sediments) is "repacked" to form a more
stable structure. We can only surmise that he looked upon fresh sediment as having a very
open, random (cardhouse?) structure which progressively becomes more preferentially
oriented and hence more stable. Obviously, the stage will be attained when the ultimate
MICROSTRUCTURE OF CLAY SEDIMENTS 3 13

state of parallelism will be reached and further reduction of voids can only be achieved by
crystal deformation and later by recrystallization.
The subject of compacted clay fabric was somewhat neglected after Hedberg's classic
work and the subject was studied in more detail by researchers in the rapidly growing
study of soil mechanics. Lambe (1958) stated that the rearrangement of clay particles is
dependent upon compactive effort and on water content. As previously considered,
"compaction" in the engineering sense has not strictly the same meaning as in the
geological sense; in the former definition some remoulding is implied (except in static
compaction). With regard to the second of Lambe's variables, water content, this would
most certainly be expected to influence structure by a mechanism of "lubrication"
(Lambe, 1958) of the excess water which permits particles to slide past each other. The
main argument against this approach is that in soil mechanics compaction the water
content may easily be varied; under geological conditions of compaction, it may not: the
clay has a fixed initial water content which will only decrease with increasing pressure.
Again, the concept of soil mechanics compaction is not directly analogous to that in the
geological sense. Meade (1966) stresses the importance of water content in the early
stages of compaction and suggests that a fast rate of deposition of the sediment is the
cause of a high initial water content. He states:
"The initial water content, then, seems to be a critical factor in determining the degree
of preferred orientation that develops during the early stages of compaction - a certain
amount of water apparently is necessary to allow the particles to move easily past one
another."
If, as will be considered, the domain approach is employed, the sliding of particles may
not be such a simple matter.
Lambe's conclusions that compaction (or more properly, compression) causes an
increase in particle parallelism, produced considerable impetus for further research in this
field and two causes of particle orientation were postulated:
(1) Compression causes an increase in preferred orientation of clay particles.
(2) Clay particles are aligned parallel to planes of shear in a shear zone.
In this study we are not concerned with the latter of these changes since clay sedi-
ments are seldom subjected to shear stresses during compaction; it is the production of a
parallel fabric due to compressive stresses that is pertinent.
Several workers have studied clay structure from an engineering viewpoint and have
attempted, often successfully, to correlate engineering properties of cgmpacted clays with
their microstructure (Mitchell, 1956; Seed and Chart, 1959; Trollope and Chart, 1960).
Trollope and Chart suggested that packets of oriented clay particles could build up on
compression. This was also taken up by Aylmore and Quirk (1960) whose studies were
expanded in a later paper (1962) describing an extensive transmission electron micro-
scope study (using carbon replicas) of compacted illite. The existence of turbostratic
groups was supported by their results.
Previous to the investigation above, studies of clay-particle orientation had been
carried out largely by indirect methods and usually by utilizing the intensity of basal
314 C,F. MOON

(001) X-ray reflections as a measure of parallelism. This may indeed be feasible in a


system composed of single, discrete particles but its validity in identifying domains or
packets of particles is surely questionable.
It was thus generally accepted that compressed clays were composed of packets and
Sloane and Kell (1966) in studying the fabric of kaolinite which had been mechanically
compacted by various methods, including static loading, reinforced this view by pro-
ducing transmission-electron micrographs. They added that the term "bookhouse" should
be used for the structure in which clay plates were stacked one above the other (Fig. 5a).

| I
| t
I I
ra I |

a b c

Fig. 5. Domain structures, a. "Bookhouse" according to Sloane and KeU (1966). b. and c. "Stepped
face-to-face" according to Smalley and Cabrera (1969).

This arrangement was later challengeci by Smalley and Cabrera (1969) who, whilst in
agreement with Sloane and Kell with regard to the existence of packets, suggested that in
real systems they should be thought of as having a stepped stacking - the "stepped
face-to-face" structure (Fig. 5a and b). This appears to be a more stable and thus more
likely arrangement since in the bookhouse structure, like (negative) charges are juxta-
posed with edges. Their idea was supported by scanning electron micrographs of statically
compacted kaolinite. O'Brien (1968a), studying fissile Palaeozoic shales of varying ages
from Cambrian to Carboniferous found a fabric strongly suggestive of the stepped face-to-
face:
" . . . the reader will note that the clay flakes overlap like pages in a book." (My
italics).
The idea of the existence of turbostratic groups or domains in compacted clays has
continually received support from many researchers. Smart (i967) produced a series of
transmission electron micrographs of kaolinite and was able to trace the domain pattern
in one of his compacted samples. They appear to have a completely random orientation.
He also noted the existence of small domains in remoulded but uncompacted specimens.
Recent work, notably by Barden and Sides (1970, 1971) and Barden etal. (1969) has
shown that the presence of domains in compacted clays is virtually indisputable.
An interesting observation by Blackmore and Miller (1961) suggests that as pressure is
increased on a clay, the number of particles in each domain also increases. They found
that for montmorillonite at 100 kg/m 2 there were about eight 10 A sheets per domain.
If we consider that an argillaceous sedimentary rock is equivalent to a clay that has
been subjected to a high degree of compaction, we should expect to observe preferred
orientation and perhaps turbostratic groups. At this point we have to consider the micro-
structural fabric of argillaceous rocks. It must be specified that we are not concerned with
MICROSTRUCTUREOF C L A Y SEDIMENTS 3 15

those sediments that have undergone intense deformation (i.e., metasediments) but only
those that have been indurated by compaction. Clark (1970) indicates that distortion, not
compaction, seems to be a much more important cause of preferred orientation in
deformed argillites. Extensive studies by ceramicists (Weymouth and Williamson, 1953;
Williamson, 1960) on the effects of extrusion on plastic clays have indicated that this is
very likely to be so. This may in fact, be analogous to the relatively easy development of
preferred orientation by the remoulding of clays in soil mechanics.
Very little detailed study of argillaceous rocks has been carried out at high magnifica-
tions and hence reliable information is scanty. In many sediments that have been studied
by X-ray diffraction methods, preferred orientation is not apparent, but this could quite
easily be due to the fact that domains are formed which would not be discernible by
X-ray diffraction techniques. Many workers have noticed from the study of deep-sea
cores that preferred orientation does not increase with depth, but this does not mean that
the number of particles per domain does not increase with depth. For a complete con-
sideration of this topic the electrolyte concentration of the depositional medium and
pore solutions of the sediment must be discussed. If we assume that a flocculated fabric is
one of randomly arranged packets of few individual particles, then perhaps a dispersed
fabric is composed of parallel or sub-parallel domains of a few particles. This may only be
surmised, but it would seem likely on the basis of the foregoing evidence.
Many authors have equated fissility with a high degree of preferred orientation and the
present writer considers this to be a feasible assumption. As preferred orientation has
previously been thought of as being applicable to single, discrete particles and not to
domains, it has been considered by many that preferred orientation and hence fissility
must increase with compression (i.e., burial depth). In fact Baldwin (1971) gives 350 m
depth as the beginning of the formation of fissility in a clay to finally produce a shale.
Von Engelhardt and Gaida (1963) are less specific and hold that clay sediments form
shales at depths between 100 and 3,000 m. If the domain structure is invoked this need
not necessarily be so; increasing burial depth would probably increase the number of
particles in each domain but these domains may still be randomly oriented in a floc-
culated sediment and parallel or sub-parallel in a dispersed clay deposit. Hence it can be
envisaged that a dispersed clay sediment would be likely to produce a fissile rock due to
the overall initial preferred orientation of packets whilst a flocculated sediment would
produce a non-fissile rock (i.e., mudstone).
The study of fissility in shale has for long been an enigma of sedimentary geology.
White (1961) stated that fissility is not due solely to the degree of compaction, as some
marine cores show that argillaceous rocks from a depth of 1 km are less fissile than those
found nearer to the surface. Odom (1967) similarly found no relation between fissility
and depth of burial, nor did he observe any correlation between clay fabric and burial
depth. He did find, however, that shales with a high degree of particle orientation were
characterized by a fissile structure and summarized his observations in tabular form
(Fig. 6). The absence of any obvious relation between fissile structure and overburden
depth may be explained by a consideration of the chemistry of the depositional medium.
316 C.F. MOON

Degr£e of
Rock Fabric prefe,rr,e
d Structure
type index' onentot~on
of cloy minerals

0.10 very good


shale 0.20 good .e
fair
, 0.30 "= I
claysItoneor 0.40 poor I i.i
muds-tone 0.50 very poor

Fig. 6. Fabric indices as related to rock type and macrostructure (after Odom, 1967). 1 Fabric index is
a measure of particle parallelism as determined by the intensity of X-ray diffraction lines. A fabric
index of 0.50 signifies complete randomness of particles; a value of 0 indicates perfect parallelism.

A Pennsylvanian black shale was examined by transmission electron microscope (O'Brien,


1968b) and a well-developed preferred orientation (of domains? ) was observed. O'Brien
attributed the black colour of the rock to organic material resulting from the high:organic
content of the depositional medium. This could quite easily be the dispersing agent
necessary for the production of a fabric of parallel or sub-parallel domains. He notes that
the addition of a few parts per thousand of certain organic ions can cause the dispersion
of clays by neutralization of surface charges and thus the black shale may have been
deposited in the dispersed state. If the depositional environment of black muds is con-
sidered as anaerobic with high organic content (Strgm, 1955), then surely the clay par-
ticles or domains in these muds will be dispersed and give rise, on compaction, to fissile
black shales. The importance of organic content in the development of a fissile material
has been emphasized by Odom (1967) who states:
" . . . samples with a high organic content generally show the best degree of clay
particle orientation and fissile s t r u c t u r e . . . " This was earlier recognized by Gipson
(1965) who found that planes of fissility tended to occur along zones of organic material.
Some writers, however, have associated a dispersed system, and hence fissility, in
argillaceous rocks simply with a low electrolyte content, which would be insufficient to
flocculate the clay and would allegedly produce single particles that would take up a
parallel orientation (White, 1961). As we have seen, the status of the single particle in
clay fabric is to be questioned.
A factor which may have importance but which has in the main been overlooked is
outlined by Gipson (1966). He suggests that the action of benthonic organisms may
disorganize some sediment layers and not others, thus resulting in juxtaposed strata of
preferred and random clay mineral orientation.
Meade (1964, 1966) has placed great importance on electrolyte concentration and
cations present in the formation of clay fabrics, and as stated, these factors cannot be
ignored.
A recent paper by Burnham (1970), in which the marine Oxford Clay was studied,
found that the majority of the deposit was of clay in small domains (up to 50/am across)
MICROSTRUCTUREOF CLAY SEDIMENTS 317

showing some alignment to the bedding. A low preferred orientation of domains would
be expected in a marine clay which was deposited in flocculating conditions.
The importance of clay particle size has often been neglected in studies of clay fabric.
Bolt (1956) found that his experiments suggested that the larger particles of illite and
kaolinite developed a preferred orientation much more easily than small particles. This
has similarly been reported by Von Engelhardt and Gaida (1963). Meade (1964) found
that the larger particles appear to form domains more rapidly than the smaller clay
minerals. In connection with this', it must be borne in mind that domains of very small
particles such as montmorillonite are extremely difficult to identify even using sophistic-
ated electron microscope techniques.
Probably the major argument against the domain hypothesis in argillaceous rocks is
that so few electron microscope studies of these rocks have been made to either confirm
or dispute the existence of domains in fissile or non-fissile argillites. Taylor (1971) has
observed edge-to-face structures in Coal Measures marine clays but as the study was
carried out by optical microscopy it is doubtful if domains would be observed if present.
In a paper on deep sea sediments and their liquefaction by earthquakes, Francis (1971)
states:
"One can conclude that even at depths.of 1 km below the ocean floor, the porosity of
unconsolidated silt-clay sediments is probably in the region of 50%. The fact that high
porosity extends to depth implies that a house-of-cards type of structure does also, with
only gradual orientation of the sediment particles towards the structure of laminated
siltstone or shale." (My italics). Despite the mounting body of evidence against it, it is
seen that the cardhouse structure is still favoured by many investigators.

CONCLUSIONS

In the light of the evidence which has been presented above, the author has attempted
to produce a system describing clay-particle re-arrangement with increasing pressure (i.e.,
increasing burial).
From the foregoing, two major observations on the nature of clay fabric are suggested:
(1) The cardhouse structure of random arrangement of individual clay particles having
predominately edge-to-face contacts appears to be increasingly unlikely and the concept
of the domain or turbostratic group has importance even in freshly deposited clays. In
short, units of more than one plate appear to be present in uncompacted clays. However,
in dilute clay systems (suspensions) where particle groups are widely spaced, there is some
evidence for the existence of cardhouse-like single plate clusters. It can be inferred there-
fore, that at a certain clay concentration at or very near the depositional interface,
particle clusters take up the domain arrangement. Further research is needed to determine
at which point this transformation occurs and more important, why it occurs.
(2) With compaction there is an increase in particle orientation or, more properly, an
increase in domain orientation. It also seems likely that under normal conditions, more
particles are incorporated into each domain with compaction. The orientation of the
318 C.F. MOON

domains will undoubtedly be greatly influenced by electrolyte concentration; a dispersed


solution will be conducive to a structure with a high degree o f preferred orientation of
domains; in a flocculating medium this may not necessarily be so. Experimental evidence
suggests that particle size is also of high importance. High pressure compaction studies are
needed incorporated with electron microscopy in order to determine the fabric changes
that occur with increasing pressure, under the influence of different pore fluids and with
different clay mineral species. It may then become apparent which of these parameters is
favourable for the development or absence of fissility in argillaceous rocks.
With this in mind, the author has proposed a tentative scheme for the arrangement and
re-arrangement of particles with progressive compaction of clay sediments (Fig. 7).

FLOCCULATED DISPERSED
Fresh
clay

a b

Compacte~
clay
C d
Fig. 7. A suggested scheme of particle arrangement in clay sediments, a. Open, random arrangement of
domains of 2-3 particles per packet, b. Parallel or sub-parallel arrangement of domains of 2 3
particles per packet, c. Increased parallelism and more particles incorporated into each domain than in
(a), i.e., mudstone, d. Complete parallelism and more particles per packet than in (b), i.e., shale.

In conclusion, it has emerged that clays are extremely complex deposits when one
considers their microstructure, and it appears that Sorby at the turn of the century was
fully aware o f this fact when he wrote:
"Possibly many may think that the deposition and consolidation of fine-grained mud
must be a very simple matter and the results o f little interest. However, when carefully
studied experimentally, it is soon found to be so complex a question and the results
dependent on so many variables that one might feel inclined to abandon the enquiry were
it not that so much of the history of our rocks appears to be written in this language."
(Sorby, 1908).
MICROSTRUCTURE OF CLAY SEDIMENTS 3 19

ACKNOWLEDGEMENTS

The a u t h o r w o u l d like to t h a n k Dr. lan Smalley o f the University o f Leeds for reading
the m a n u s c r i p t and for o f f e r i n g s o m e e x t r e m e l y helpful c o m m e n t s .

REFERENCES

American Geological hrstitute, 1962. Dk'tionary ~ff Geological Terms. Dolphin Books, New York,
N.Y., 545 pp.
Ayhnore, L.A.G. and Quirk, J.P., 1960. Domain or turbostratic structure of clays. Nature,
187:1046 1048.
Aylmore, L.A.G. and Quirk, J.P., 1962. Tile structural status of clay systems. Clays Clay Minerals,
Proc. Natl. ( b n f Clays Clay Minerals, Lafayette, Ind., 9:104 - 130.
Baldwin, B., 1971. Ways of deciphering compacted sediments. J. Sediment. Petrol., 41:293 301.
Barden, L., 1972. The influence of structure oil deformation and failure in clay soil. Geotechnique,
22:159-163.
Bardcn, L. and Sides, G.R., 1970. Engineering behaviour and the structure of compacted clay. Proc.
Am. Soc. Cir. Eng., 96. SM4:1171 1200.
Barden, L. and Sides, G.R., 1971. The microstructure of dispersed and flocculated samples of kaoli-
nite, illite and montmorillonite. Can. Geotech. J., 8:391--399.
Barden, .L., Sides, G.R. and Karunaratne, J.P., 1969. A microscopic examination of aspects of clay
structure. Proc. S.L Asian Con l; Soil Eng., 2nd., Bangkok, 1969:67-72.
Berry, R.W. and J~rgensen, P., 1971. Grain size, mineralogy and chemistry of a quickclay sample from
the Ullensaker slide, Norway. Eng. Geol., 5:73-84.
Bishop, A.W., Hutchinson, J.N., Penman, A.D., Evans, H.E., Nash, J.K.T.L., Williams, G.M.J., Wood-
land, A.W., Moore, L.R., Wardell, K., Piggott, R.J. and Bleasdale, A., 1969. Inquiry into the
Aberfan disaster. A selection o f technical reports submitted to the Aher]an Tribunal. Welsh Office.
H.M.S.O, London, ltem 3:114 115.
Bjerrum, L. and Rosenqvist, 1.'1"11., 1957. Some experiments with artificially sedimcnted clays.
Geoteehnique, 6:124-136.
Blackmore, A.V. and Miller, R.D., 1961. Tactoid size and osmotic swelling in calcium mont-
morillonite. Soil Sci. Soc. Am. Proc., 25:169-173.
Bolt, G.H., 1956. Physico-chemical analysis of the compressibility of pure clays. Geotechnique,
6:86-93.
Boswell, P.G.H., 1961. Muddy Sediments. Heffer, Cambridge, 140 pp.
Bowles, [:.A., Bryant, W.R. and Wallin, C., 1969. Microstructure of unconsolidated and consolidated
marine sediments. J. Sediment. Petrol., 39:1546 1551.
Brewer, R., 1964. Fabric and Mineral Analysis of Soils. Wiley, New York, N.Y., 470 pp.
Burnham, C.P., 1970. The micromorphology of argillaceous sediments: particularly calcareous clays
and siltstones. In : D.A. Osmond and P. Bullock (Editors), Mieromorphological Techniques attd
Applications. Soil Sun,. Tech. Monogr., 2:83 96.
Casagrande, A., 1932. The structure of clay and its importance in foundation,engineering. J. Boston
Soc. Cir. Eng., 19:168 221.
Clark, B.R., 1970. Mechanical formation of preferred orientation in clays. Am. J. Sci., 269:250 266.
Francis, T.J.G., 1971. Effect of earthquakes on deep sea sediments. Nature, 233: 98-102.
Gillott, J.E., 1968. Clay in Engineering Geology. Elsevier, New York, N.Y., 296 pp.
Gillott, J.E., 1969. Study of the fabric of fine grained sediment with the scanning electron micro-
scope. J. Sediment. Petrol., 39:90- 105.
Gipson, M., 1965. Application of the electron microscope to the study of fissility in shale. J.
Sediment. Petrol., 35:408 414.
Gipson, M., 1966. A study of the relations of depth, porosity and clay mineral orientation in
Pennsylvanian shales. J. Sediment. Petrol., 36:888-903.
320 C.F. MOON

Greene-Kelly, R. and Mackney, D., 1970. Preferred orientation of clay in soils: the effect of drying
and wetting. In: D.A. Osmond and P. Bullock (Editors), Micromorphological Techniques and
Applications. Soil Surv. Tech. Monogr., 2 : 4 3 - 5 2 .
Grim, R.E., 1953. Clay Mineralogy. McGraw-Hill, New York, N.Y., 384 pp.
Grim, R.E., 1962. Applied Clay Mineralogy. McGraw-Hill, New York, N.Y., 422 pp.
Hcdberg, H.D., 1936. Gravitational conapaction o f clays and shales. Am. J. ScL, 31:241 287.
hrgles, O.G., 1968. Soil chemistry relevant to the engineering behaviour of soils. In: I.K. Lee (Editor),
Soil Mechanies Selected Topics. Butterworths, London, pp. 1-34.
Lambc, T.W., 1958. The structure of compacted clay. Proc. Am. Soc. Cir. Eng., 91. SM4:85-106.
Meade, R.H., 1964. Removal of water and rcarrangement of particles during the compaction of clayey
sediments review. U.S. Geol. Surv. Pro]. Pap., 497-B:BI-B23.
Meade, R.tt., 1966. Factors influencing the early stages of compaction of clays and sands review. J.
Sediment. Petrol., 36:1085 - 1101.
Mitchell, J.K., 1956. The fabric of natural clays and its relation to engineering properties. Proc.
ttighway Res. Board, 35 : 6 9 3 - 713.
Morgenstern, N.R., 1969. Structural and physico-ehemical effects on the properties of clays. Proc. Int.
Conf Soil Mech. Found. Eng., 7th, Mexico, 3:455 471.
O'Brien, N., 1968a. Clay fabric of very fissile Paleozoic gray shales. Marit. Sedfnents, 4:104 105.
O'Brien, N., 1968b. Electron microscope study of a black shale fabric. Naturwiss., 5 5 : 4 9 0 491.
O'Brien, N., 1970. l:abric of flocculated clay sediments. Geol. Soc. Am., Abstr. Programs, 2:637.
O'Brien. N. and Harrison, W., 1967. Clay flake orientation in a sensitive Pleistocene clay. Marit.
Sediments. 3:1 -4.
Odom, I.E., 1967. Clay fabric and its relation to structural properties in mid-Continent Pennsylvanian
scdiments. J. Sediment. Petrol., 37 : 6 1 0 - 623.
Osmond, D.A. and Bullock, P., 1970. Micromorphologieal Techniques and Applications. Soil Sun'.
Tech. Monogr., 2, 110 pp.
Penner, E.. 1965. A study of sensitivity in Leda clay. Can. J. Earth Sci., 2:425-441.
Pusch, R., 1966. Quickclay nricrostructure. Eng. Geol., 1:433-443.
Pusch, R., 1970. Microstructural changes in soft quickclay at failure. Can. Geotech. J., 7:1 7.
Rosenqvist, 1. Th., 1953. Considerations on the sensitivity of Norwegian quickclays. Geoteehnique,
3:195 200.
Rosenqvist, l.Th., 1962. The influence of physico-chemical factors upon the mechanical properties of
clays. Clays Clay Minerals, Proc. Natl. Conf Clays Clay Minerals, Lafayette, Ind., 9: 12-27.
Schweitzer, J. and Jennings, B.R., 1971. The association of montmorillonite studied by light scattering
in electric fields. J. Colloid lnterJ?lee Sei., 37:443 -457.
Scott, J.S., 1965. A Dictionary of Civil Engineering. Penguin, London, 348 pp.
Seed, H.B. and Chan, C.K., 1959. Structure and strength characteristics of compacted clays. Proc. Am.
Soc. Civ. Eng., 85, SM5:87-128.
Sloane, R.L. and Kell, T.F., 1966. The fabric of nrechanically compacted kaolin. Clays Clay Minerals,
Proc. Natl. Conf. Clays Clay Minerals, Berkeley, Carl[, 14:289-296.
Smalley, l.J.: 1971. Naturc of quickclays. Nature, 231 : 310.
Smalley, l.J. and Cabrera, J.G., 1969. Partielc association in compacted kaolinite. Nature, 2 2 2 : 8 0 - 8 1 .
Smart, P., 1967. Particle arrangements in kaolin. C/avs Clay Minerals, Proc. Natl. Conf. Clays Clay,
Minerals, Pittsburgh, Pa., 15:241 254.
Sodcrbtom, R., 1966. Chemical aspects of quickclay formation. Eng. Geol., 1 : 4 1 5 - 4 3 1 .
Sorby, 1t.C., 1908. On the application of quantitative methods to the study of the structure and
history of rocks. Q. J. Geol. Sot'. Lond., 6 4 : 1 7 1 - 2 3 2 .
StrCm, K.H., 1955. Land-locked waters and the deposition of black muds. In: P.D. Trask (Editor),
Recent Marine Sediments. Soc. Econ. Paleontologists Mineralogists, Spee. PubL, 4:356 372.
"l'an, T.K., 1959. Structure mechanics o f clays. Sc~ Sinica, 8 : 8 3 - 9 6 .
Taylor, R.K., 1971. The petrography of the Mansfield Marine Band cyclothenr at Tinsley Park,
Sheffield. Proc. Yorks. Geol. Sot'., 38:299 328.
Tchalenko, J.S., 1968. The microstructure of London Clay. Q. J. Eng. Geol., 1:155-168.
Terzaghi, K., 1925. Erdbaumechanik aujbodenphysikalischer Grundlage. Liepzig, Vienna, 399 pp.
MICROSTRUCTURE OF CLAY SEDIMENTS 321

Trollope, D.H. and Chan, C.K., 1960. Soil structure and the stepstrain phenomenon. Proc. Am. Soc.
Cir. Eng., 86, SM2:1-39.
Urbain, P., 1937. Texture microscopique des roches argileuses. Bull. Soc. G~ol. Fr., Ser. 5,
7:341-346.
Van Olphen, H., 1963. An Introduction to Clay Colloid Chemistry. Interscience, New York, N.Y..
301 pp.
Von Engelhardt, W. and Gaida, K.H., 1963. Concentration changes of pore solutions during the
compaction of clay sediments. J. Sediment. Petrol., 33:919-930.
Weller, J.M., 1959. Compaction of sediments. Bull. Am. Assoc. Pet. Geologists, 43:273-310.
Weymouth, J.H. and Williamson, W.O., 1953. The effects of extrusion and some other processes on
the microstructure of clay. Am. J. ScL, 251:89 108.
White, W.A., 1961. Colloid phenomena in the sedimentation of argillaceous rocks. J. Sedhnent.
Petrol., 31:560 570.
Williamson, W.O., 1960. Some effects of deformation on the structure and properties of clay. Miner.
Ind., Pa. State Univ., 29: 1-8.

(Accepted for publication June 7, 1972)

You might also like