You are on page 1of 17

Int Journ Earth Sciences (1999) 88 : 392–408 Q Springer-Verlag 1999

ORIGINAL PAPER

H. Dürrast 7 S. Siegesmund

Correlation between rock fabrics and physical properties


of carbonate reservoir rocks

Received: 28 January 1999 / Accepted: 21 June 1999

Abstract Three carbonate core samples from an oil changes with pressure, which can be converted to
and gas reservoir of the NW German basin were different depths. The knowledge gained from such
chosen to study the correlation between rock fabrics correlations may lead to an improved interpretation of
and physical properties of reservoir rocks. Detailed geophysical data for hydrocarbon exploration and
fabric analyses and texture investigations were carried production and therefore to an advanced reservoir
out as well as laboratory measurements of different characterization.
physical properties, e.g. density, porosity, permeability,
electrical conductivity, seismic compressional and shear Key words Rock fabrics 7 Physical properties 7
wave velocities. Although the three core samples come Anisotropy 7 Reservoir rocks 7 Carbonates 7
from a similar depositional facies, they show great Fractures
differences in the occurrence and three-dimensional
distribution of the rock fabric elements. These hetero-
geneities are the result of various diagenetic and Introduction
tectonic processes. For the correlation between the
rock fabrics and the physical properties four main rock Carbonate rocks along with sandstones are the major
fabric types have to be considered: (a) major consti- oil and gas reservoir rocks in the world. The hydro-
tuents, e.g. fossils, ooides, peloides and crystals; (b) carbon resources in these rocks are associated mainly
pore space with different pore types; (c) fractures; and with dolomites (Jardine and Wilshart 1987; Sun 1995).
(d) stylolites. The results of the correlation clearly show In North America, for example, approximately 80% of
that the values and anisotropies of the petrophysical the recoverable oil and gas in carbonate reservoirs are
properties are fairly related to the observed fabric accumulated in dolomite rocks (Zenger et. al 1980).
elements, with their different arrangements, spatial The origin and distribution of pore space within
distributions and preferred orientations. These results dolomites defined by porosity and permeability are the
also provide a fundamental understanding of the petro- most important parameters which control the transport
physical responses, such as seismics, to the different properties of reservoir rocks (see Lucia 1995; Luo and
geological features (e.g. fractures) and their dynamic Machel 1995; Sun 1995). Furthermore, fractures play an
important role in dolomite reservoirs. They can provide
the essential reservoir porosity and permeability and
they can also create significant reservoir anisotropy
H. Dürrast (Y)
Institute of Geology and Dynamics of the Lithosphere, (Nelson 1985; Sun 1995).
Georg-August-University of Göttingen, Goldschmidtstrasse 3, Dolomite reservoir rocks are very heterogeneous, in
D-37077 Göttingen, Germany contrast to clastic reservoirs (Sun 1995). This depends
e-mail: hduerra6gwdg.de mainly on the depositional environment in various
S. Siegesmund facies and on the diagenetic and tectonic processes
Institute of Geology and Dynamics of the Lithosphere, (cementation, dolomitization, recrystallisation, frac-
Georg-August-University of Göttingen, Goldschmidtstrasse 3, turing, etc.) which are associated with carbonate rocks
D-37077 Göttingen, Germany
(Murray and Pray 1965; Wardlaw 1965). These hetero-
Present address: geneities are scale dependent and very important for
H. Dürrast
Department of Geology and Geophysics, University of
reservoir characterization. At a scale of 10 m to 10 km
Wisconsin-Madison, 1215 West Dayton St., Madison, WI 53706, they represent major lithostratigraphic boundaries,
USA large faults and extensive fractures (Weber 1986). At a
393

Fig. 1 The microfabric


elements in carbonate rocks

smaller scale (millimetres to metres) the heterogenei- et al. 1992). Füchtbauer (1959) used the chemical and
ties are caused by the rock fabric. For carbonate reser- mineralogical composition for a classification. Folk
voir rocks four major rock fabric types are distin- (1959, 1962) and Dunham (1962) presented schemes
guished (Fig. 1): (a) major constituents, e.g. fossils, based on the relative amounts of four textural and diag-
ooides, peloides and crystals; (b) pore space; (c) frac- enetic components in the rocks (particles, lime mud
tures; and (d) stylolites. matrix, cement and pores) as well as the characteristics
The distribution and orientation of the various of particles vs matrix. These classifications combine
microfabrics in dolomites influence the different descriptive and genetic aspects (Mazullo et al. 1992).
physical (e.g. the electrical conductivity, seismic veloci- Through the exploration of hydrocarbon-bearing rocks
ties and attenuation of compressional and shear waves) classification schemes of pore space with regard to
and transport properties (porosity, permeability, satu- porosity and permeability were developed (Archie
ration). If there is a spatial distribution (e.g. pores) or a 1952; Choquette and Pray 1970; Lucia 1983, 1995; see
preferred orientation of the fabric elements (e.g. frac- below). But for a correlation between microfabrics and
tures), the physical properties of a reservoir rock show physical properties all important fabric elements have
anisotropic behaviour (see Schön 1996; Siegesmund to be considered. Herein the four main fabric types are
1996). If two or more of these parameters appear distinguished and explained briefly (see Fig. 1).
together, their effect on the physical properties can be
more complex. Figure 2 shows a compilation of the
lithological and physical parameters controlling the Major constituents
physical properties of reservoir rocks under in situ
conditions (for a discussion see Schön 1996). The components in carbonate rocks are fossils, ooides,
peloides, pellets, intraclasts, crystals and lime mud.
Compositionally they generally consist of the various
carbonate minerals, mainly calcite and dolomite.
Microfabrics in carbonate rocks Secondary constituents of importance consist of
sulphate and sulphide minerals as well as organic mate-
Carbonate rocks rial.
The arrangement and distribution of the compo-
Different classification schemes were developed as a nents resulting in various bedding and layering features
result of the various heterogeneities present in is controlled basically by deposition and sedimentation.
carbonate rocks. Limestones and dolomites have been These features are characterized by their texture and
classified on the basis of mineralogy, texture, composi- structure. This includes variations in composition, size,
tion and certain physical parameters such as pore types shape and orientation of the particles and components,
and porosity (see Roehl and Choquette 1985; Mazullo and in their packing (Collinson and Thompson 1989).
394

Fig. 2 Factors controlling the


physical properties of
carbonate reservoir rocks

Furthermore, arrangement and distribution of compo- permeability and the saturation, and it is related to the
nents can be modified by secondary processes (e.g. rock fabrics. These considerations and investigations
tectonics) resulting in a crystallographic preferred led to two major pore space groups: (a) the interpar-
orientation (texture) of the rock-forming minerals (see ticle pore space and (b) the vuggy pore space, which is
Schaftenaar and Carlson 1984; O’Brien et al. 1993; divided into separate-vug pores and touching-vug
Ratschbacher et al. 1994). pores. For the first group Lucia (1995) characterized
three rock fabric/petrophysical classes defined by
certain permeability and water saturation fields.
Pore space Together with interparticle porosity and reservoir
height, these classes can be used to relate the petrophy-
In carbonate rocks various pore types can be distin- sical properties to geological observations (Lucia
guished. All of them are the result of the primary distri- 1999).
bution of different components in a matrix influenced
by secondary diagenetic processes. During the early
water expulsion stage, the mechanical compaction Fractures
resulting from the overburden pressure is the dominant
mechanism for porosity reduction (Schlanger and In the pore space classification of Lucia (1995) frac-
Douglas 1974; Kim et al. 1985). tures are one type of touching-vug pores. The fractures
Archie (1952) was the first one who classified in this study are separated from the pore space, because
carbonate rocks by using their pore space. Choquette they can contribute to anisotropy as well as to porosity
and Pray (1970) presented a fabric-selectivity concept, and permeability. The fracture intensity in dolomites is
a classification that found broad use in the oil and gas higher than in limy dolomites and in limestones
industry. Lucia (1983, 1995, 1999) added a more petro- (Stearns 1967; Sinclair 1980), and with decreasing grain
physical view to the classification of the pore space: the size the fracture intensity in dolomites increases
pore-size distribution controls the porosity, the (Sinclair 1980).
395

For the examination of fractures van Golf-Racht


(1996) presented a descriptive classification scheme
defining the following categories: (a) open/closed frac-
tures; (b) macro-/micro-fractures; and (c) natural/
induced fractures. The distinction between natural and
induced fractures is important for the interpretation of
borehole image data (e.g. Özkanli and Standen 1993).
For the genetic interpretation of the fractures two
fundamental assumptions have to be made (Stearns
and Friedman 1972; Nelson 1985): (a) natural fracture
patterns depict the local state of stress at the time of
fracturing; and (b) subsurface rocks fracture in a
manner qualitatively similar to equivalent rocks in
laboratory tests performed under analogous environ-
mental conditions.

Stylolites

In carbonate rocks and sandstones stylolites are


common diagenetic features, and their origin is inde-
pendent of the rock facies and geological age (Nelson
1985; van Golf-Racht 1996). Generally, they are irreg-
ular planes of discontinuity with an orientation from
horizontal to vertical. The stylolite irregularities are
displayed by the stylas, columns and rows of different
width and height. The stylolites themselves are charac-
terized by the concentration of relatively insoluble
constituents of the enclosing rock (Park and Schot
1968). The (continuous) presence of relatively insoluble
material can act as barriers to the fluid dynamic system
of the reservoir rock.
It is generally accepted that stylolites are the result
of a concentration-pressure process or a pressure-disso-
lution process. Both processes are controlled mainly by
the solubility of the components and particles in the
rock, and by the regional stress field (Nelson 1985).
The solution plane is more or less perpendicular to the
maximum stress direction, which is commonly the over-
burden pressure direction in basin regions with less
significant tectonics.
A classification of stylolites in carbonate rocks is
given by Logan and Semeniuk (1976) with respect to
configuration, arrangement, fabric and structure of the
pressure solution phenomena.

Methods of investigation

For the geological and petrophysical investigations of


each sample all the necessary specimens and thin Fig. 3 Sample 1 : Scanning electron microscopy (SEM) images
sections were cut and prepared from a macroscopically made perpendicular to bedding plane (X–Z plane; box shows the
reference system for all samples). The grainstone consists of
homogeneous part of the core segment to ensure that peloids (P) and aggregate grains (A) with grain sizes between 100
analyses were comparable. and 400 mm embedded in a dolomitic and partly anhydritic matrix
The mineral composition was determined by using (grey dolomite, white anhydrite). Open cracks (M), intercrystal-
the X-ray diffraction method. Structural investigations line as well as grain-boundary cracks, with vertical and subvertical
orientation, can be followed over several millimetres, but they are
were done macroscopically on the drilling cores and on not all connected. Black areas are imaging artefacts
three mutually perpendicular standard thin sections
(reference system see Fig. 3) with the optical micro-
396

scope and the scanning electron microscope (SEM). Table 1 Mineral composition (in volume percent), bulk density
For the texture measurements an X-ray goniometer (r) and effective porosity (F) of the three Zechstein carbonate
samples
was used.
Different kinds of petrophysical measurements were Sample 1 Sample 2 Sample 3
carried out, i.e. porosity, permeability, electrical
conductivity and seismic velocities. The total porosity Dolomite 84.0 94.3 94.6
Anhydrite 15.2 5.3 5.1
and the pore radii distribution were determined using Pyrite 0.4 – –
the mercury intrusion method (Carlo Erba Porosi- Fluorite 0.2 – 0.1
meter). The density measurement was performed on a Quartz 0.2 – 0.1
cylindrical specimen using a gas pyknometer. Halite – 0.4 0.1
On dry spherical samples the complete compres- r [g/cm 3] 2.74 2.27 2.60
F [%] 5.1 16.6 9.6
sional wave velocities (VP) were determined up to a
confining pressure of 200 MPa (see Siegesmund et al.
1993). This allows the determination of the maximum
(VPmax) and minimum (VPmin) P-wave velocities and the peloides contain mainly micritic dolomite, whereas
therefore to quantify the real anisotropy the cortices often consists of anhydrite. The effective
[Ap100*(VPmax–VPmin)/VPmax (%)]. porosity of 5.1% is due primarily to small intercrystal-
Cylindrical dry specimens from three orthogonal line pores between the dolomite crystals and second-
directions of the sample were used for the measure- arily due to open cracks (Fig. 3, 4). From the orienta-
ment of the P-wave velocities (VP) and the shear wave tion two groups of open cracks are distinguished: nearly
velocities (VS) in confining pressure steps up to vertical and nearly horizontal ones. Their main occur-
200 MPa. For every specimen VS was carried out in two rence as single cracks and the roughness of the crack
oriented polarization directions perpendicular to each surfaces are indications that most of the open cracks
other (for details see Jahns et al. 1994; Dürrast and were not formed under in situ conditions. Probably
Siegesmund 1996). they are the result of core relaxation processes.
For the investigations of the permeability and the Sample 2 is a grainstone with subhorizontal layering
electrical conductivity cylindrical specimens from two on centimetre-scale (Fig. 5a). In between there are
directions of the sample, parallel and perpendicular to millimetre-scale layers of organic material (biolamina-
the core axis, were prepared. The measurements were tion; see Steinhoff and Strohmenger 1996). The main
performed up to a confining pressure of 80 MPa. The constituents are peloids and fewer oncoides in a crypto-
permeability was carried out using Argon gas on dry crystalline dolomitic and partly anhydritic matrix. Most
samples and the electrical conductivity was determined of the peloids have a diameter between 150 and
on tap water saturated samples. For the measuring 400 mm, but some are larger up to 600 mm diameter.
techniques see Freund and Nover (1995), Nover et al. The effective porosity of 16.6 % is due mainly to inter-
(1995) and Siegesmund et al. (1993). Additionally the particle pores and molds (see Fig. 5b, c). The interpar-
permeabilities were measured under hydrostatic pres-
sure conditions (1.0 MPa) parallel to the X-, Y- and Z-
axes using cubes (0.7-in. edge length).

Samples

Microfabrics, porosity and density

The correlation between microfabrics and physical


properties are illustrated in detail on three dolomitic
carbonate reservoir rocks (sample 1, 2 and 3; see
Table 1) chosen from a borehole located in the NW
German Basin (Ca2, Zechstein, Permian). The samples
were taken from a 40 m interval at a depth of approxi-
mately 3000 m. The three samples can be assigned to
the shallow subtidal platform facies (W. Dauben, in
preparation). Fig. 4 Sample 1 : Scanning electron microscopy (SEM) image
Sample 1 is a grainstone that consists mainly of made perpendicular to bedding plane (X–Z plane). This image is
peloides and aggregate grains with sizes from 100 to dominated by peloids and aggregate grains. The core of some
peloids consists of dolomite; only the cortex is made of anhydrite
400 mm; only some have a size of a few millimetres (PD). A subhorizontal open intercrystalline crack (M) can be
(Figs. 3, 4). They are embedded in a cryptocrystalline followed over the whole image width. Between the dolomite crys-
matrix of dolomite and partly anhydrite. The cores of tals there are many small (~20 mm) intercrystalline pores (K)
397

Fig. 5 Scanning electron


microscopy (SEM) images of
a–c Sample 2 and d–f Sample
3 made perpendicular to
bedding plane (Y–Z plane). a
Two layers with different
porosity: compare the upper
part of the image with abun-
dant open pores (black) to the
lower part where open pore
space is much less. b Peloides
(P) with grain sizes from 150
to 400 mm are the major
constituents. The porosity is
supported by the interparticle
pores (I) and the moldic
porosity (L). c Higher-resolu-
tion image of a porous layer
with different open pore types
(black, compare with b). d
Horizontal fine-scale layering
of open (black) and anhydrite-
filled (white) pores. e Most of
the open cracks show a
preferred orientation with
nearly vertical dip, some are
horizontal. The pores and
cracks are interconnected. f
Open cracks often run parallel
to the grain boundaries of the
dolomite grains (arrows).
Open intercrystalline (K)
pores are isolated or can be
connected with others by open
cracks

ticle pores are between the peloides, whereas originally result in any preferred lattice orientation of the
the molds were peloid cores, which were dissolved dolomite crystals.
during diagenetic processes (see Huttel 1989; Below Sample 3 originally consisted of calcitic ooides with
1992). Figure 5a shows that the open pores are an average diameter of 0.3 mm. During later stages of
arranged in nearly horizontal layers between the accu- diagenesis a complete dolomitization and recrystallisa-
mulated organic material. But this alignment did not tion resulted in a final dolomite crystal size of 50 mm
398

(average diameter; see Fig. 5d–f). The relict shapes of


the ooides are still visible under the optical microscope.
Neither the deposition nor compaction nor complete
recrystallisation of the primary concentric ooides led to
any lattice-preferred orientation of the dolomite crys-
tals. The inner cores of most ooides went into solution
during the diagenesis which led to a moldic and incom-
plete porosity (Fig. 5e, f). A later anhydritization filled
the pore space, which occurred selectively (only certain
layers) or completely (whole rock). Assuming that the
complete pore space was mineralized, then parts of the
anhydrite must have been dissolved later (see Mausfeld
1987; Huttel 1989). The result of the diagenesis is a
nearly horizontal millimetre fine-scale layering where
layers with open pores alternate with anhydrite-filled
layers (Fig. 5d).
Furthermore, the rock is also intensively fractured
which was probably supported by the small grain size of
the dolomite crystals (see Stearns 1967; Sinclair 1980).
The fractures appear at all scales ranging from macros-
copic (Fig. 6) to submicroscopic (Fig. 5f). They are
open or mineralized with anhydrite or calcite. Most of
the fractures are interconnected and linked with open
pores (Fig. 5e, f) and therefore both provide to the
effective porosity of 9.6%.
Figure 6 shows a macro image of Sample 3 parallel
to the core axis. The nearly horizontal dark structures
are algae laminae (see Steinhoff and Strohmenger
1996) combined with stylolites. Open and mineralized
natural fractures with a nearly vertical orientation cut
and shift these horizontal structures indicating younger
tectonic events.
The fractures show a preferred orientation as seen in
Fig. 5e, f; see Fig. 12e with a nearly vertical dip (only
some are nearly horizontal) and a strike subparallel to
the X-direction. A schematic sketch of the fracture
orientation is shown in Fig. 12d. The preferred orienta- Fig. 6 Image of Sample 3 shows the macroscopic fabric elements:
tion of the fractures is presumable caused by the orien- horizontal stylolites (S) and layers of organic material (O). Open
and mineralized cracks (M) with a nearly vertical orientation shift
tation of the regional stress field in the NW German these algae structures and stylolites indicating younger tectonic
basin. The fracture planes should be oriented vertical events
and parallel to the maximum horizontal stress direction
(e.g.; Crampin 1987). Unfortunately, the drill cores
investigated in this study cannot be reoriented to the the total porosity. Sample 3 shows nearly the same
geographical coordinate system. distribution but with a higher entry point (55 mm)
representing more and large open cracks. But for
Sample 2 the high entry point of 25 mm and the
Effective pore radii increase up to 2 mm effective pore radius expresses
molds with diameters up to 600 mm (5% of the cumula-
The cumulative pore space occupied by mercury injec- tive pore space). Then between 2 and 0.2 mm effective
tion can be determined by continuously raising the pore radius the mainly large interparticle pores account
pressure (here up to 180 MPa), because the capillary for 20% of the total porosity, whereas the last 75% of
pressure is related to the smallest pore aperture (effec- the porosity shows a uniform distribution similar to
tive pore radius) through which the mercury has to Sample 3 due to small intercrystalline pores.
penetrate. The entry point of Sample 1 is at 15 mm and
with only 5% increase in the cumulative pore space up
to 0.2 mm effective pore radius (see Fig. 7). Both are P-wave velocities
indications of the open cracks in this rock. Between 0.2
and 0.02 mm the distribution of pores, here intercrystal- The P-wave velocity measurements on dry spherical
line pores, is uniform and represents the other 95% of samples gave complete VP patterns represented in
399

symmetry disappears and changes into a nearly


isotropic VP distribution. Therefore, the
Cumulative percent pore space

DVPpVP(100 MPa) – VP(0.1 MPa) pattern in Fig. 9a shows


also a nearly orthorhombic symmetry. In the DVP
diagram the max. DVP values represent the highest
increase in the P-wave velocities from the 0.1 to the
100 MPa pressure stage. Perpendicular to the orienta-
tion of these values most of the open pores and cracks
are closed.
Based on this correlation the DVP diagram can be
directly compared with a calculated density pattern of
open cracks. The maximum in the crack density pattern
Effective pore radii (µm) represents an orientation where most cracks have been
penetrated along a unit distance (Fig. 9d; see Voll-
Fig. 7 Mercury porosimetry data of samples 1, 2 and 3 with the brecht et al. 1994). For Sample 1 three crack popula-
cumulative pore space as the function of the effective pore radii
tions are distinguished, a horizontal one and two nearly
vertical ones (Fig. 9b, c), which are all verified by the
microfabrics investigations (Figs. 3, 4). This quantified
Schmidt net diagrams as a function of confining pres- crack density pattern of Sample 1 has a good correla-
sures up to 200 MPa. The projection plane for all the tion with the DVP diagram, so that the orthorhombic
experimental derived VP plots is horizontal (X–Y symmetry can be explained only by open cracks.
plane) with the Z-direction parallel to the drilling For Sample 2 the complete VP pattern shows a trans-
axis. verse symmetry at all pressures with the lowest P-wave
The complete VP pattern of Sample 1 shown in velocities in the Z-direction and the highest in the X-Y
Fig. 8 has a nearly orthorhombic symmetry at 0.1 MPa plane (Fig. 10). VPmin increases slightly from 3.4 km/s
confining pressure with the maximum and minimum P- (0.1 MPa) to 4.75 km/s (200 MPa), as well as VPmax
wave velocities in the horizontal plane resulting in a
real anisotropy of about 20%. With increasing pressure
the open cracks and pores get progressively closed Fig. 8 Experimentally determined directional dependence of
which leads to an increase in the velocities and to a compressional wave velocities (VP in kilometres/second) of
Sample 1 at confining pressures up to 200 MPa. The directions of
decrease in the anisotropy: VPminp6.0 km/s and the highest VP are coloured in dark grey, the ones of the lowest
VPmaxp6.3 km/s (200 MPa) and the anisotropy is less VP in light grey. Projection plane is in the X–Y plane (Schmidt
than 5%. Between 20 and 50 MPa the orthorhombic net, lower hemisphere)
400

Fig. 9a–d. Correlation between


the pressure dependent VP
measurements and the rock
fabric of Sample 1. a The
difference between the
complete VP distribution at
100 and 0.1 MPa gives the
DVP diagram (VP in kilo-
metres/second). b Schematic
sketch and c pole figure of the
open crack populations used
for d calculation of the
diagram showing the crack
density distribution. All
Schmid net, lower hemisphere,
except b

from 3.95 km/s up to 5.4 km/s. This is caused by the The pressure dependency of the VP exhibits a
large spherical molds and interparticel pores between symmetry change with increasing confining pressure.
the peloides, but less by open cracks. Both poretypes The most pronounced velocity increase is in a hori-
cannot really be closed with increasing confining pres- zontal direction parallel to the VPmin at low confining
sure because of their shape (molds) and position pressures. This increase is also reflected in the DVP
between spherical peloides (interparticle pores). The diagram [VP(100 MPa)–VP(0.1 MPa)] with the highest value
DVP diagram between VP at 100 MPa and VP at of 2.6 km/s (Fig. 12c). Since the open fractures and
0.1 MPa therefore shows an increase in the P-wave pores are mainly closed at 100 MPa, the DVP diagram
velocities of only approximately 1.3 km/s and with represents the bulk pattern of open fractures and pores
nearly random distribution. The anisotropy decreases which effectively influences the VP anisotropy below
very slightly but on a high level, from 13.7% (0.1 MPa) the crack closure pressure (ca. 50 MPa).
down to 11.2% (200 MPa) due to the arrangement of From the DVP diagram the orientation of the frac-
open pores in horizontal layers. tures and pores can be deduced (Fig. 12). In corre-
Sample 3 shows the highest velocity anisotropy with spondence with the microstructural observations there
nearly 25% at 0.1 MPa confining pressure (Fig. 11). is a prominent population of open fractures with a
The lowest VP at this pressure level are horizontal and preferred orientation. The fracture planes show an
subparallel to the Y-direction. Nearly subparallel to the almost vertical dip and a strike subparallel to the X-
Z-direction the highest values are found. With direction (Fig. 12d, e). The closure of these fractures
increasing pressure at around 10–20 MPa, the direction causes the highest DVP values (2.6 km/s). The minor
of VPmax rotates from the nearly vertical position to a increase of VP in the nearly vertical direction (1.3 km/s)
horizontal orientation, subparallel to the X-direction. is induced by the small population of horizontal open
The orientation of VPmin is stable up to a confining fractures (Fig. 5e, f) and open pores in discrete small
pressure ranging from 20 to 50 MPa. At this pressure horizontal layers (Fig. 5d). With their nearly unique
range the VPmin direction rotates from a horizontal to a aspect ratio (ratio width to length) the pores cannot be
nearly vertical orientation. Above 50 MPa the orienta- closed completely at confining pressures of approxi-
tions are fixed while the VPmax and VPmin values mately 200 MPa, similar to Sample 2.
increase. At 200 MPa the VPmax value has reached The P-wave velocity distributions at pressure levels
6.7 km/s and the VPmin 6.1 km/s with an anisotropy of higher than 50–100 MPa shows transversal isotropic
9%. behaviour. The symmetry axis is vertical and parallel to
401

Fig. 10 Sample 2 : Experimentally determined VP up to 200 MPa At low confining pressures up to 20 MPa VS is
confining pressure (see also Fig. 8). Additionally, the DVP higher in the Z-direction than in the Y-direction. The
diagram between VP at 100 MPa and VP at 0.1 MPa is presented
highest values in the Z-direction are parallel to the
X–Z plane (3.16 km/s at 5 MPa) and therefore subpa-
rallel to the prominent open fracture population with
the VPmin direction. Up to a confining pressure of their nearly vertical preferred orientation (Fig. 12e).
200 MPa there is no significant change in this This leads to a shear wave splitting of 2.9% at 5 MPa.
symmetry. At this high-pressure level most of the open With increasing pressure the fractures are progressively
fractures and pores are closed, and therefore the VP closed and consequently the shear wave splitting
anisotropy is caused by intrinsic properties (Fig. 12). decreases to 1.4% (200 MPa). This remaining shear
One reason is the fine-scale layering of the open and wave splitting is probably due to mineralized calcite
anhydrite filled pores (Fig. 5d) and the horizontal accu- cracks, which have the same orientation as the open
mulation of organic material by algae growth and cracks (Fig. 12e). From approximately 50 MPa both
stylolites (see Fig. 6). This caused a less porous hori- shear waves in the Z-direction are lower than those in
zontal layer in the more porous rock leading to higher the Y-axis.
VP parallel to this plane rather than perpendicular to For the Y-direction the shear wave splitting
it. decreases from 2.8% at 5 MPa to 1.2% at 200 MPa with
the faster shear wave polarized in the X–Y plane. This
orientation of the faster and slower shear wave is also
S-wave velocities controlled by the horizontal microfabrics (organic
material, stylolites and fine-scale layering). The inter-
The results of the shear wave measurements are shown section of the VS-pressure curves between 20 and
for Sample 3. The data of the direction parallel (Z) and 50 MPa (Fig. 13) is therefore comparable to the
perpendicular to the core axis (Y) are presented in symmetry change in the P-wave velocity distribution at
Fig. 13, whereas the orientations of the different shear the same pressure range (Fig. 11).
wave polarisation planes are shown in Fig. 12d. In The complete P-wave velocity distribution and addi-
correspondence with VP the shear wave velocities (VS) tional measurements of two shear waves in three ortho-
in both directions show a distinct pressure dependency. gonal directions (X, Y and Z; see Table 2) provide the
With increasing confining pressure the VS values opportunity to calculate the elastic tensor without any
increase up to approximately 50 MPa, the crack closure assumption (see Jahns 1995; Weiss 1998). Based on this
pressure for this rock. elastic tensor the P-wave velocities can be recalculated
402

Fig. 11 Sample 3 : Experimentally determined P-wave velocities directly correlated with the calculated crack density
up to 200 MPa confining pressure (see also Fig. 8) pattern (Fig. 9d): in the Z-direction the polarisation
direction is parallel to the strike of the nearly vertical
cracks, whereas in the X–Y plane the faster shear
as well as the faster and slower shear wave and the waves are oriented parallel to the horizontal ones
orientation of their polarization directions. (Fig. 14c).
The results for Sample 1 at 10 MPa and for Sample 3 For Sample 3 the P-wave velocity distribution at
at 50 MPa confining pressure are shown in Fig. 14b and 50 MPa is due mainly to the horizontal microfabric
f. For both samples there is a good agreement between elements (see Figs. 5d, 6) so that the faster shear waves
the laboratory measurements and the recalculated VP are polarized in the X–Y plane. In the Z-direction the
distribution in the absolute values as well as in the situation is much more complex. The laboratory mea-
symmetry of the distribution. surements have revealed that above the crack closure
At low confining pressure only three sets of open pressure the shear waves are oriented parallel to the
cracks are responsible for the VP distribution of Sample mineralized cracks (Fig. 12e). Therefore, the orienta-
1 (see Figs. 3, 4, 9). Therefore, the orientation of the tion of the polarization direction of the faster shear
polarization direction for the faster shear waves can be waves in the Z-direction is not only affected by the
horizontal microfabrics but also by the vertical mineral-
ized cracks (Fig. 14g).
Table 2 Results of the ultrasonic velocity measurements (in kilo-
metres/second) for Sample 1 (at 10 MPa) and Sample 3 (at
50 MPa): Vp and two shear wave velocities (VSfast, VSslow) in three Permeability and electrical conductivity
orthogonal directions (X, Y and Z) used for the calculations of
the elastic tensors The permeability measured on cubes in three ortho-
gonal directions (X, Y and Z) is generally below 1 milli-
VP VS_fast VS_slow
darcy (mD), except the Y- and Z-direction of Sample 3
Sample 1 (Table 3). Here a single open crack leads to distinct
X 4.441 2.974 2.958 higher values. Similar low values ranging between 1
Y 4.556 2.855 2.851 and 5 mD were found in dolomitized oolitic and pelle-
Z 4.521 2.878 2.873 toidal grainstones (Ca2; e.g. Wijhe 1981).
Sample 3 The permeability is also a directional-dependent
X 6.337 3.438 3.326 property (see Table 3). Sample 1 shows in all directions
Y 6.272 3.440 3.389
Z 5.911 3.370 3.314 nearly the same values representing the matrix permea-
bility. The open relaxation cracks did not contribute to
403

Fig. 12 Experimentally deter-


mined VP of Sample 3 at a
100 MPa and b 0.1 MPa
confining pressures. c Calcula-
tion of the DVP diagram from
the VP distributions at 0.1 and
100 MPa confining pressure. d
From the DVP diagram the
orientation of the major rock
fabric elements can be
deduced: e the horizontal
alignment of the open pores
and the preferred orientation
of the open fractures, which is
in agreement with the macros-
copical observations of the
open and mineralized cracks
(X-Y plane). The intrinsic
rock fabrics with the hori-
zontal structures (pores, stylol-
ites, organic material) can be
concluded from the VP distri-
bution at 100 MPa (a)

the permeability because they occurred rarely and they


are not interconnected (see Figs. 3, 4). Sample 2 has
higher values in the X–Y plane in accordance with the
horizontal fine-scale layering of permeable and
impermeable layers, whereas Sample 3 exhibits higher
permeabilities in the X–Z plane caused by vertically
oriented open cracks. A large single crack in a second
cube of Sample 3 increases the permeability up to
5.8 mD.
The pressure-dependent permeability and electrical
conductivity measurements of samples 2 and 3 up to a
confining pressure of 80 MPa are shown in Fig. 15 (here
Fig. 13 Sample 3 : Shear wave velocity measurements parallel to Y- and Z-direction). For Sample 2 the permeability in
the core axis (Z-direction) and in the horizontal Y-direction up to
200 MPa confining pressure. For the orientation of the polariza-
the Y-direction is significantly higher than in the Z-
tion directions see Fig. 12d. Arrowheads indicate the faster shear direction, comparable to the values from the cube
wave of each direction measurements. With increasing pressure up to 80 MPa
404

Fig. 14 Experimentally deter-


mined VP for Sample 1 a at
10 MPa and e Sample 3 at
50 MPa. P-wave velocities for
the two samples at the certain
pressure levels from the calcu-
lated elastic tensor which was
determined from the complete
VP distribution and shear
wave velocity measurements
(b,f). c,g Orientation of the
faster shear wave based on
the calculated elastic tensor
for each sample. d Crack-
density distribution of Sample
1 (see Fig. 9d). h Distribution
of the open and mineralized
cracks in the X-Y plane of
Sample 3 (see Fig. 12e)
405

Table 3 Permeabilities (in millidarcy) of samples 1, 2 and 3


measured on cubes in three orthogonal directions (X, Y and Z)

Direction Sample 1 Sample 2 Sample 3

Large single
crack

X 0.0093 0.3735 0.0840 0.1810


Y 0.0072 0.3755 0.1950 4.0200
Z 0.0085 0.1675 0.2590 5.8600

Fig. 16 Formation factors calculated from the electrical conduc-


tivity values (see Fig. 15) for samples 2 and 3 as a function of
confining pressure

except until at high pressures the open pores are


collapsed (see Christensen 1965).
Vertical open cracks, however, provide the major
pathways in Sample 3. Accordingly, the main values of
the permeability and electrical conductivity are
measured in the Z-direction, whereas in the Y-direction
the values are lower, especially for the permeability.
The horizontal stylolite planes in this sample have no
influence on the anisotropy of the permeability and
electrical conductivity, because their impermeable
planes were cut by the vertical cracks (see Fig. 6). With
increasing pressure the open cracks get progressively
closed, but the roughness of the crack surfaces prevents
a complete closure. So even at 80 MPa confining pres-
sure there are still open pathways through the rock
which lead to a permeability of 0.13 mD and an elec-
trical conductivity of 0.003 S/m (both in Z-direction).
From the resistivity of the samples (R0) and the
resistivity of the tap water (Rwp26 Vm) used for the
electrical measurements the formation factor (FpR0/
Fig. 15 Determination of the permeability and the electrical Rw) can be calculated. The results vs confining pressure
conductivity parallel to the core axis (Z-direction) and in the for sample 2 and 3 are plotted in a semi-logarithmic
horizontal Y-direction up to 80 MPa confining pressure for a plot (Fig. 16).
Sample 2 and b Sample 3
In correspondence to the conductivity data the
formation factors of Sample 3 in the Y-direction are
higher than in the Z-direction. The absolute difference
the values in both directions declined. The electrical between the values of both directions is relatively small
conductivity shows nearly the same behaviour with in comparison with Sample 2. Here the Y-direction
increasing pressure and the values in the Z-direction increases slightly with increasing confining pressure,
are also lower than in the Y-direction. Even at 80 MPa whereas the formation factors in the Z-direction jump
confining pressure a distinct permeability and electrical from approximately 2 at 0.1 MPa up to over 450 at
conductivity in the Y-direction is found. 100 MPa. Sample 2 generally consists of good and of
Both physical properties, the permeability and the low permeable horizontal layers. At 0.1 MPa confining
electrical conductivity, are linked to open pathways in pressure probably small open pores and cracks connect
the rock sample: (a) open pore throats as connections the nearly impermeable layers. But with increasing
between pores; (b) open cracks; and (c) networks pressure these pathways were closed. Then the
between open pores and cracks. For Sample 2 mainly conducting water is trapped and nearly isolated in the
the open pore throats get progressively closed with permeable layers. Therefore, the horizontal layers do
increasing pressure so that the values of both proper- not provide a continuous path to the electrical current,
ties decrease. But for a complex network of pores and which probably leads to the high-resistivity values in
pore throats it is not possible to close all pathways, the Z-direction (see Sen 1997).
406

three examined samples have at least two different


Conclusion pore types but vary from sample to sample. This
The combined study of rock fabrics and physical prop- makes a correlation between the rocks’ pore types
erties of reservoir rocks is an important aspect of reser- and the physical properties much more difficult (see
voir characterization, because it directly links the also Wang 1997).
geophysical exploration tools to the reservoir proper- The anisotropies of the physical properties also
ties. Results of such a study provide a fundamental depend on the pressure, an important (dynamic)
understanding of the petrophysical responses to parameter of oil and gas reservoirs. The compressi-
geological features and their dynamic changes. The aim bilities of the different pores, cracks and network
is to directly convert the attributes of the physical prop- among pores and between cracks and pores are
erties to the geological reservoir parameters (see also directly related to the pressure. Changes in the
Wang 1997). permeability or seismic wave velocities, for example,
Carbonate reservoir rocks are very heterogeneous in are the results of these pressure dependencies.
contrast to clastic reservoirs. The foremost reasons are 3. Open crack and fracture systems with a vertical and
the occurrence and the spatial distribution of the preferred orientation are of natural occurrence
different rock fabric elements due to the various depo- (Sample 3). They can be open as well as mineralized
sitional facies and the diagenetic and tectonic pro- and range from a macroscopic to a submicroscopic
cesses. For carbonate rocks four major rock fabric scale. Their preferred orientation is probably due to
types are distinguished: (a) major constituents; (b) pore the regional stress field.
space; (c) fractures; and (d) stylolites. Detailed investi- Sample 3 also shows the common situation of reser-
gations of these different microstructural elements and voir rocks from a sedimentary basin: horizontal
their influence on the physical properties from three layering/bedding structures with vertical- and
dolomitic carbonate reservoir rocks lead to the preferred-oriented cracks and fractures. Therefore,
following main conclusions: the complete VP pattern of Sample 3 at low
1. Although the depositional facies of the three confining pressures shows a complex (monoclinic or
samples is similar (shallow subtidal platform facies) triclinic) symmetry. With increasing pressure
and the vertical range of the corresponding core combined with the closure of open cracks the
segments is approximately 40 m, the three samples symmetry of the P-wave velocities changes into
show very different rock fabrics. Therefore, not only transversal isotropic with a vertical rotation axis
the sedimentary processes and depositional factors, (TIV). A similar complexity is shown by Sample 1,
but also the diagenetic (e.g. mineralization and where three sets of relaxation cracks (vertical and
dissolution) and tectonic processes, lead to varied horizontal) create an orthorhombic symmetry of the
rock fabrics. These processes together control the complete VP pattern.
arrangement, distribution and orientation of the The classical and often used transverse isotropic
major constituents, the open pore space and the symmetry is common, but to the same extent other
open pathways, the fractures and the stylolites in the and more complex VP symmetries are found in
rock. Consequently, these main fabric elements have rocks, especially in fractured reservoirs. This has to
to be considered when working with physical prop- be taken into consideration by the interpretation of
erties of carbonate reservoir rocks. All these fabric borehole and seismic measurements.
elements together control the present values of the 4. The results of the detailed investigations have also
physical properties such as transport and seismic shown that shear waves are more sensitive to the
parameters. distribution and orientation of open and mineralized
2. The detailed fabric investigations have shown that cracks and fractures than the P-wave velocities.
all three samples have a heterogeneous arrange- Even above the crack closure pressure (ca. 50 MPa
ment, a spatial distribution and/or a preferred orien- for the three samples) a shear wave splitting parallel
tation of one or more of the main fabric elements. to the borehole axis is a good indication for nearly
This leads to anisotropies of the physical properties vertical cracks with a preferred orientation (see
which are directional dependent (e.g. seismic wave samples 1 and 3). This becomes important for the
velocities, permeability, electrical conductivity). For detection of vertical open cracks using seismic field
example, the spatial distribution of open molds and and borehole measurements (e.g. vertical seismic
interparticle pores in Sample 2 cause an intrinsic VP profiling; Dey-Barsukov et al., accepted), because
anisotropy of 12% ( 1 50 MPa confining pressure). open cracks can provide a major contribution to the
The pore types in a rock are important for the permeability. Identifying cracks are a key point for
physical properties, because different pore types the characterization of fractured reservoirs.
have different compressibilities so that they affect
the seismic wave velocities in a different manner. Acknowledgements The authors gratefully acknowledge
Preussag Energie GmbH, Lingen, Mobil Erdgas Erdöl GmbH,
The electrical conductivity and permeability are also Celle, Wintershall AG, Kassel, BEB Erdgas and Erdöl GmbH,
closely related to the type of pores in the rock. Addi- Hannover, and RWE-DEA, Hamburg, for providing the samples
tionally, investigations have shown that each of the and supporting this investigation. The data publication was done
407

under the permission of the DGMK-Project ’LITASEIS’. We Logan BW, Semeniuk V (1976) Dynamic metamorphism, proc-
thank P. Krajewski (Lingen), G. Nover, St. Heikamp (both esses and products in Devonian carbonate rocks, Canning
Bonn), D. Freund (Potsdam), G. Braun (Kiel) and T. Chlupac Basin, Western Australia. Geol Soc Austral Spec Publ
(Prague) for their assistance and the many helpful discussions. 6 : 1–183
Thanks also to C. Gross for proofreading the manuscript. H.D. Lucia FJ (1983) Petrophysical parameters estimated from visual
thanks the State of Lower Saxony (Germany) for a scholarship of description of carbonate rocks: a field classification of
the Graduiertenförderung, S.S. the DFG for a Heisenberg fellow- carbonate pore space. J Petrol Tech 35 : 626–637
ship. Appreciation is also extended to the reviewers, F. Jordan Lucia FJ (1995) Rock fabric/petrophysical classification of
and an anonymous reviewer, for their many helpful corrections carbonate pore space for reservoir characterization. Am
and suggestions which improved the quality of the manuscript. Assoc Petrol Geol Bull 79 : 1275–1300
Thanks to N.I. Christensen (Madison, Wisconsin) for final Lucia FJ (1999) Carbonate reservoir characterization. Springer,
comments and suggestions. Berlin Heidelberg New York, pp 1–226
Luo P, Machel HG (1995) Pore size and pore throat types in a
heterogeneous dolostone reservoir, Devonian Grosmont
Formation, Western Canada sedimentary basin. Am Assoc
Petrol Geol Bull 79 : 1698–1720
References Mausfeld S (1987) Der Plattformrand des Stassfurtkarbonats
(Ca2) südlich von Oldenburg: Sedimentologie, Fazies und
Diagenese. PhD thesis, Univ Marburg, pp 1–98
Archie GE (1952) Classification of carbonate reservoir rocks and Mazullo SJ, Chilingarian GV, Bissell HJ (1992) Carbonate rock
petrophysical considerations. Am Assoc Petrol Geol Bull classification. In: Chilingarian GV, Mazullo SJ, Rieke HH
36 : 278–298 (eds) Carbonate reservoir characterization: a geologic-engi-
Below A (1992) Fazies- und geochemische Diagenesestudie im neering analysis, part I. Elsevier, Amsterdam, pp 59–108
Zechstein 2-Karbonat (Ca2) Nordwestdeutschlands. PhD Murray RC, Pray LC (1965) Dolomitization and limestone dia-
thesis, Univ Kiel, pp 1–147 genesis – an introduction. Soc Econ Paleont Min Spec Publ
Choquette PW, Pray LC (1970) Geologic nomenclature and clas- 13 : 1–2
sification of porosity in sedimentary carbonates. Am Assoc Nelson RA (1985) Geologic analysis of naturally fractured reser-
Petrol Geol Bull 54 : 207–250 voirs. Gulf Publishing, Houston, pp 1–320
Christensen NI (1965) Compressional wave velocities in meta- Nover G, Heikamp S, Kontny A, Duba A (1995) The effect of
morphic rocks at pressures to 10 kilobars. J Geophys Res pressure on the electrical conductivity of KTB rocks. Surv
70 : 6147–6164 Geophys 16 : 63–81
Collinson JD, Thompson DB (1989) Sedimentary structures. O’Brien DK, Manghnani MH, Tribble JS, Wenk H-R (1993)
Chapman and Hall, London, pp 1–207 Preferred orientation and velocity anisotropy in marine clay-
Crampin S (1987) The geological and industrial implications of bearing calcareous sediments. In: Rezak R, Lavoie DL (eds)
extensive-dilatancy anisotropy. Nature 3 : 16–20 Carbonate microfabrics. Springer, Berlin Heidelberg New
Dauben W (in preparation) Faziesverteilung, Sequenzstratigra- York, pp 1–313
phie und Speicherqualität des Zechstein-2 Karbonats im Özkanli M, Standen EJW (1993) Fracture morphology in Creta-
West-Emsland, NW-Deutschland. PhD thesis, Tech Univ ceous carbonate reservoirs from south-east Turkey. First
Berlin Break 11 : 323–333
Dey-Barsukov S, Dürrast H, Rabbel W, Siegesmund S, Wende S Park WC, Schot EH (1968) Stylolite: their nature and origin. J
(accepted) Aligned fractures in carbonate rocks: Evidence Sediment Petrol 38 : 175–191
from laboratory and in situ measurements of seismic aniso- Ratschbacher L, Wetzel A, Brockmeier H-G (1994) A neutron
tropy. Geol. Rundschau texture goniometer study of the preferred orientation of
Dunham RJ (1962) Classification of carbonate rocks according to calcite in fine-grained deep-sea carbonate. Sediment Geol
depositional texture. Am Assoc Petrol Geol Mem 1 : 108–121 89 : 315–324
Dürrast H, Siegesmund S (1996) Gesteinsphysikalische Untersu- Roehl PO, Choquette PW (1985) Introduction. In: Roehl PO,
chungen an Kernsegmenten karbonatischer Reservoirges- Choquette PW (eds) Carbonate petroleum reservoirs.
teine. In: Tessmer G (ed) Lithologieerkundung für den Springer, Berlin Heidelberg New York, pp 1–15
Tiefenaufschluss mit seismischen Methoden (LITASEIS). Schaftenaar CH, Carlson RL (1984) Calcite fabric and acoustic
DGMK Forschungsbericht 397-2/2, Hamburg, pp 217–237 anisotropy in deep-sea carbonates. J Geophys Res
Folk RL (1959) Practical petrographic classification of limestones. 89 : 503–510
Am Assoc Petrol Geol Bull 43 : 1–38 Schlanger SO, Douglas RG (1974) The pelagic ooze-chalk-limes-
Folk RL (1962) Spectral subdivision of limestone types. Am tone transition and its implication for marine stratigraphy. Int
Assoc Petrol Geol Mem 1 : 62–84 Assoc Sediment Spec Publ 1 : 117–148
Freund D, Nover G (1995) Hydrostatic pressure tests for the Schön JH (1996) Physical properties of rocks: fundamentals and
permeability–formation factor relation on crystalline rocks principles of petrophysics. Pergamon Elsevier, Oxford, pp
from the KTB drilling project. Surv Geophys 16 : 47–62 1–583
Füchtbauer H (1959) Zur Nomenklatur von Sedimentgesteinen. Sen NP (1997) Resistivity of partially saturated carbonate rocks
Erdöl Kohle 12 : 605–613 with microporosity. Geophysics 62 : 415–425
Huttel P (1989) Das Stassfurt-Karbonat (Ca2) in Süd-Oldenburg Siegesmund S (1996) The significance of rock fabrics for the
– Fazies und Diagenese eines Sediments am Nordhang der geological interpretation of geophysical anisotropies. Geotekt
Hunte-Schwelle. Gött Arbeit Geol Paläont 39 : 1–94 Forsch 85 : 1–123
Jahns E (1995) Quantifizierte seismische Anisotropie: Ein skalen- Siegesmund S, Vollbrecht A, Chlupac T, Nover G, Dürrast H,
übergreifender Vergleich am Beispiel der KTB. Cuvillier, Müller J, Weber K (1993) Fabric-controlled anisotropy of
Göttingen, pp 1–110 petrophysical properties observed in KTB core samples. Sci
Jahns E, Siegesmund S, Chlupac T (1994) In situ seismic veloci- Drilling 4 : 31–54
ties versus laboratory-measurements: an example from the Sinclair SW (1980) Analysis of macroscopic fractures on Teton
KTB. Sci Drilling 4 : 215–226 anticline, northwestern Montana. MS thesis, Texas A&M
Jardine D, Wilshart JW (1987) Carbonate reservoir description. University, pp 1–93
Soc Econ Paleont Mineral Spec Publ 40 : 129–152
Kim D-C, Manghnani MH, Schlanger SO (1985) The role of dia-
genesis in the development of physical properties of deep-sea
carbonate sediments. Mar Geol 69 : 69–91
408

Stearns DW (1967) Certain aspects of fracture in naturally Wang Z (1997) Seismic properties of carbonate rocks. In: Palaz I,
deformed rocks. In: Ricker RE (ed) NSF advanced sciences Marfurt KJ (eds) Carbonate seismology. Geophys Develop
seminar in rock mechanics. Natural Science Foundation, Series 6, Soc Explor Geophys, Tulsa, pp 29–52
Bedford, Mass., pp 97–118 Wardlaw NC (1965) Pore geometry of carbonate rocks as
Stearns DW, Friedman M (1972) Reservoirs in fractured rocks. revealed by pore casts and capillary pressure. Am Assoc
Am Assoc Petrol Geol Mem 16 : 82–106 Petrol Geol Bull 60 : 245–257
Steinhoff I, Strohmenger C (1996) Zechstein 2 carbonate plat- Weber KJ (1986) How heterogeneity affects oil recovery. In:
form subfacies and grain-type distribution (Upper Permian, Lake LW, Carroll HB Jr (eds) Reservoir characterization.
Northwest Germany). Facies 35 : 105–132 Academic Press, Orlando, pp 487–544
Sun SQ (1995) Dolomite reservoirs: porosity evolution and reser- Weiss T (1998) Gefügeanisotropie und ihre Auswirkung auf das
voir characterization. Am Assoc Petrol Geol Bull seismische Erscheinungsbild: Fallbeispiel aus der Lithosphäre
79 : 186–204 Süddeutschlands. Geotekt Forsch 91 : 1–156
van Golf-Racht TD (1996) Naturally fractured carbonate reser- Wijhe DH van (1981) The Zechstein 2 carbonate exploration in
voirs. In: Chilingarian GV, Mazullo SJ, Rieke HH (eds) the eastern Netherlands. Proc Int Symp Central European
Carbonate reservoir characterization: a geologic-engineering Permian, pp 574–586
analysis, part II. Elsevier, Amsterdam, pp 683–771 Zenger DH, Dunham JB, Ethington RL (1980) Concepts and
Vollbrecht A, Dürrast H, Weber K (1994) Eine einfache models of dolomitization. Soc Econ Paleont Min Spec Publ
Methode zur dreidimensionalen Quantifizierung von Riss- 28 : 1–320
schardichten auf der Basis von U-Tisch-Messungen. Rep
German Continental Deep Drilling Program (KTB) 94 : B89

You might also like