You are on page 1of 4

RP 1.

4
Clay mineral elastic properties
Keith W. Katahara, ARC0 Exploration and Production Technology
Summary

Elastic velocities for dry clay minerals have been estimated from published data on well-crystallized phyllosilicates.
The clay velocities are significantly higher than previous values deduced from shale and shaly sand measurements,
possibly because the latter are affected by low-aspect-ratio pores, by surface-adsorbed water, by clay orientation and
anisotropy, and by extrapolation of empirical trends beyond their range of validity. However, because clays in
rocks are usually wet, clay-water-composite velocities may be more appropriate to use in modeling rock properties
than single-crystal velocities. Research should thus focus on the nature and geometry of surface-bound and
interlayer water associated with clays, and on the sizes, shapes and orientations of clay particles.
Introduction

Clay minerals are a major constituent of sediment columns that bear hydrocarbons. As such, their elastic
properties are important to understanding seismic and sonic-log response. However, there is little experimental
data, and while some indirect inferences of clay stiffnesses have been published, they are not definitive. Here some
published data are used to estimate clay-mineral elastic stiffnesses, and the results are compared to other estimates.
The term clay is used here to denote phyllosilicate minerals and not gram size.
Single crystal properties

Alexandrov and Ryzhova


called AR hereafter, report at least partial data for talc, phlogopite, biotite,
muscovite, valuevite (a brittle mica), clinochlore (a magnesian chlorite), and leichtenbergite (a variety of
clinochlore). Following the example of AR, clay symmetry is taken as hexagonal because effects of lower
symmetry (Vaughan and Guggenheim, 1986) may be no larger than the effects of uncertainties in composition and
structure for clays in shales.
As pointed out by Tosaya
elastic properties of illite and muscovite should be similar because they are
structurally and compositionally very similar. Thus the AR muscovite results will be used for illite. These are
given in Table 1 in the form of mode velocities: =
where is an elastic stiffness and p is the density.
and
are the compressional velocities parallel and perpendicular to layering respectively.
is the
normal shear velocity, and
is the shear velocity when propagation and polarization are both layer-parallel.
has no clear physical meaning. Also listed in Table 1 are compressional,
and shear,
velocities using
Reuss-Hill
averaging (e.g., Watt et al.,
and density, p.
Table 1. Clay mineral velocities, in km/s.
and
mi ne ra l
illite
279
8.03 4.44 5.01 2.05 2.28
chlorite
8.22 6.30 4.82 2.06 2.75
kaolinite 2.52
8.25 4.57 5.13 2.42 3.28

are Voigt-Reuss-Hill averages.


3.37 5.82 1.73 2.17
3.35 5.93 1.77 3.06
3.55 6.23 1.75 1.89

Chlorite. AR measured velocities on two magnesian chlorites, but did not measure
Velocity-density trends
Figures 1 and 2 show the AR mode velocities for illite-type micas (biotite, phlogopite,
will be used to estimate
and muscovite) vs. density. AR gave no composition or density for chlorites. A density of 2.71
was
averaged from several literature values for clinochlores. Except for
the measured chlorite velocities fall close
to the trends for the illite micas. Thus
for chlorite was taken from the trend of Figure 1 at the assumed chlorite
density. The velocity-density trends in Figures 1 and 2 might also be used for iron-rich chlorites, which will be
denser and lower in velocity, and may be more common in sediments than magnesian chlorites.
Kaolinite: Elastic mode velocities were estimated from the velocity-density trends in Figures 1 and 2. A kaolinite
was used (La Vigne et al., 1994). The kaolinite estimates are much more uncertain than the
density of 2.52
illite and chlorite values.

Clay mineral elastic properties

Figure Log-log plot of mode velocities vs.


for micas (3 highest density points) and for chlorite at
=
Mode is shown at left of trend line.

Figure 2. Log-log plot of mode velocities vs. density


for micas (3 highest density points) and for chlorite at
=
Mode is shown at
of trend line.

Adsorbed water, smectites and mixed-layer clays:


space precludes detailed discussion of smectites, for
which it is only possible to make speculative velocity estimates. Smectites are composed of aluminosilicate sheets
nm thick separated by one or more molecular monolayers of water. In contrast, illite particles are typically 7 nm
thick. Thus one adsorbed monolayer of water (0.3 nm) per particle implies 20% adsorbed water by volume for
smectite, compared to 4% for illite, and vanishing amounts for silt-sized or larger grains. Smectite properties will
thus be dominated by effects of interlayer water. Water in smectite is less compressible than bulk water (Sun et al.,
and its properties may depend on the type of exchange cations present (e.g., Guven, 1992).
Discussion

Several published empirical relations give velocities in terms of porosity and clay content for shaly sands and
shales. Figure 3 compares the present results with velocities extrapolated to pure clay using some correlations. The
clay contents of Tosaya (1982) and Han et al. (1986) are based largely on thin-section point-counting. This may
overestimate clay mineral content because small non-clay particles are included.
The correlation-derived velocities in Figure 3 are lower than the present velocities. Because a wide variety of rocks
was used in these studies, it is unlikely that expanding clays could be causing all of the difference. The present
values are unlikely to be far wrong for illite, which is probably the most abundant clay in the rocks used to develop
the correlations. Crystal imperfections and compositional variations may explain some of the difference.
Correcting for the clay-size/clay-mineral distinction would make things worse. A possible explanation of the
disagreement is that extrapolating the rock data to 100% clay is inaccurate. Since clay mineral content seldom
exceeds 70% in shales, the correlations must be extrapolated well beyond the data range. Furthermore, there
appears to be significant nonlinearity in the dependence of elastic velocities on clay content at the transition from
grain-supported to clay-matrix-supported rocks (e.g.,Yin et al., 1993).
In shales and laminated shaly-sands, clay particles tend to align parallel to bedding during compaction. In a
perfectly aligned nonporous claystone, the vertical compressional and shear velocities would be
and
which
are lower than the
averages (Table 1). Alignment may account for some of the discrepancy between the VRH
velocities and the rock-derived estimates, because log and lab velocities are often roughly normal to bedding. Of
course, when plate-like clay particles are aligned, interparticle pores will tend to be crack-like. Such
ratio pores will reduce velocities much more than water in equant pores, and will tend to accentuate the anisotropy
Thus an abundance of crack-like pores may also help explain the
if they are strongly aligned.
discrepancy.

1692

Clay mineral elastic properties

3. Clay mineral and


3 pairs are from this study. The next 3 pairs are
published
shaly-sand correlations extrapolated to pure clay.
and Castagna (1986) values are
logs in
claystones. The present values are relatively high. Some possible reasons are that (1) the rock data reflect
significant alignment of anisotropic clay particles so that
and
are more appropriate for this comparison;
(2) the effective
of clay-bearing rocks is decreased by tightly-bound surface-adsorbed water on clay
particles, which have huge surface areas; (3) the combination of clay-particle alignment and surface-bound water
implies the presence of water-filled crack-like pores, which will lower velocities much more than equant pores;
and (4) it is not valid to extrapolate the correlations to pure clay, i.e., beyond the range of the experimental data.
and Castagna (1986) used well-logs to identify nearly-pure claystones without porosity. Their
velocities (Fig.
although higher than the extrapolated rock values, are still low compared to the present values.
Again, the discrepancy may be due to alignment of anisotropic clay particles, and to small undetected amounts of
water in aligned low-aspect-ratio pores.
Homby et al. (1994) and Sayers (1994) have proposed that clay-water composites be used as basic elements in
modeling of shales. The properties of such composites depend not only on dry clay properties, but on the shape
and size of particles and pores, and the extent of particle and pore alignment. Obviously such textural variables can
differ from one shale to another, and in a given shale they will evolve with compaction and diagenesis.
of the rock framework (Tosaya, 1982; Vemik
In another context, clay in shaly sands generally reduces the
and Nur, 1992). If average bulk properties are considered, this will not occur with dry clays: the quartz bulk
modulus (38
is lower than the present
values
and the VRH compressional-wave stiffnesses
of quartz and dry-clay are comparable. If anisotropic clay layers are aligned with grain contacts, then more
and
of clay
contrast is seen between the VRH compressional stiffness of quartz (97
Moreover, if adsorbed water monolayers remain on the clay surfaces at intergranular contacts, then an even lower
clay-water composite stiffness should be compared to the quartz stiffness.
The present results may be valid for dry clays, but clays are wet in most rocks of interest. Because water is tightly
bound to clay
and because clay particles are plate-like, pores associated with clays will tend to be
shaped. The high specific surface area of clays then implies a relatively high volume of associated low-aspect-ratio
pores. Such pores may dominate elastic behavior to the extent that dry-clay properties are almost irrelevant. The

1693

Clay mineral elastic properties


high clay velocities estimated here tend to imply that lower rock-derived clay velocities may really be velocities of
clay-water composites in which water occupies low-aspect ratio pores. It only takes a little water in such crack-like
pores to cause a large decrease in velocities relative to the dry minerals.
That illite, chlorite, and kaolinite have about the same dry velocities does not mean that the corresponding wet
clays will have similar velocities. To the contrary, these clays can vary widely in surface area and volume of
clay properties,
associated water, as well as in their surface
far water. Many other factors will also
including sorting, bioturbation, and water composition. Perhaps research should now focus on characterizing the
geometry and orientation of clays and associated pores, and on measuring elastic properties, especially rigidity, of
interlayer and surface water. Given the properties and microgeometry of the constituents, an effective-medium
model (e.g., Homby et al., 1995) could yield appropriate clay-water composite properties, which could then be used
in building rock models. However, such composites will differ for different clay types and degrees of compaction.
It seems unlikely that there will be a single clay-water composite that can be used to -model all rocks.
Acknowledgements

I received valuable help from T. K. Kan, John Castagna, Chuan-Sheng Yin, Mike Batzle,
Franks, Steve Bergman, and Mike
I thank
for permission to publish.

Brown, Steve

References

Alexandrov, K. S., and Ryzhova, T. V., 1961, Elastic properties of rock-forming minerals. II. Layered silicates:
Sci., Geophys. Ser., no.
Bull. U.S.S.R.
Castagna, J. P., Batzle, M. L., and Eastwood, R. L., 1985, Relationships between compressional-wave and
wave velocities in
silicate rocks: Geophysics, 50,571.581.
ratios from sonic logs, in Danbom, S. H., and
Eastwood, R. L., and Castagna, J. P., 1986, Interpretation of
Expl. Geophys., 139-153.
Domenico, S. N., Eds., Shear-wave exploration:

Clay-water
Guven, N., 1992, Molecular aspects of clay-water interactions, in Guven, N., and Pollastro, R. M.,
Interface and its
Implications: CMS Workshop Lectures Vol. 4, The Clay Minerals Society, 2-79.
D., Nur, A., and Morgan, D., 1986, Effects of porosity and clay content on wave velocities in sandstones:
Geophysics,
Homby, B. E., Schwartz, L. M., and Hudson, J. A., 1994, Anisotropic effective-medium modeling of the elastic
properties of shales: Geophysics,
La Vigne, J., M.
and R. Hertzog, 1994, Density-neutron interpretation in shaly sands: SPWLA 35th
Annual Logging Symposium Transactions, June 19-22, paper EEE.

Sayers, C. M., 1994, The elastic anisotropy of shale: J. Geophys. Res., 99,767.774.
Sun, Y., Lin, H., and Low, P. F., 1986, The nonspecific interaction of water with the surface of clay minerals:
Colloid Interf. Sci., 112,556.564.
Tosaya, C., 1982, Acoustical properties of clay-bearing rocks, Ph.D. dissertation, Stanford University.
Vaughan, M. T., and Guggenheim, S., 1986, Elasticity of
Geophys. Res.,

and its relationship to crystal structure:

Watt, J. P., Davies, G. F., and OConnell, R. J., 1976, The elastic properties of composite media: Rev. Geophys.
Space Phys., 14,541.563.
Yin, H., Nur, A., and
G., 1993, Critical porosity a physical boundary in poroelasticity:
Min. Sci.

Geomech. Abstr., 30,805.808.

Vemik, L., and Nur, A., 1992, Petrophysical classification of siliciclastics for lithology and porosity prediction

seismic velocities: AAPG Bull.,

1694

You might also like