You are on page 1of 16

Geophysical Prospecting, 2005, 53, 481–496

Ultrasonic velocities of North Sea chalk samples: influence of porosity,


fluid content and texture
Birte Røgen,1 Ida L. Fabricius,1∗ Peter Japsen,2 Christian Høier,2 † Gary Mavko3
and Jacob M. Pedersen4
1 Environment & Resources DTU, Bygningstorvet Building 115, Technical University of Denmark, 2800 Kongens Lyngby, Denmark,
2 GeologicalSurvey of Denmark and Greenland (GEUS), Øster Voldgade 10, 1350 Copenhagen K, Denmark, 3 Department of Geophysics,
School of Earth Sciences, Stanford University, Stanford, CA 94305-2215, USA, and 4 Ødegaard A/S, Titangade 15, 2200 Copenhagen N,
Denmark

Received November 2002, revision accepted August 2004

ABSTRACT
We have studied 56 unfractured chalk samples of the Upper Cretaceous Tor Forma-
tion of the Dan, South Arne and Gorm Fields, Danish North Sea. The samples have
porosities of between 14% and 45% and calcite content of over 95%. The ultrasonic
compressional- and shear-wave velocities (V P and V S ) for dry and water-saturated
samples were measured at up to 75 bar confining hydrostatic pressure corresponding
to effective stress in the reservoir. The porosity is the main control of the ultrasonic
velocities and therefore of the elastic moduli. The elastic moduli are slightly higher for
samples from the South Arne Field than from the Dan Field for identical porosities.
This difference may be due to textural differences between the chalk at the two loca-
tions because we observe that large grains (i.e. filled microfossils and fossil fragments)
that occur more frequently in samples from the Dan Field have a porosity-reducing
effect and that samples rich in large grains have a relatively low porosity for a given
P-wave modulus. The clay content in the samples is low and is mainly represented
by either kaolinite or smectite; samples with smectite have a lower P-wave modulus
than samples with kaolinite at equal porosity. We find that ultrasonic V P and V S of
dry chalk samples can be satisfactorily estimated with Gassmann’s relationships from
data for water-saturated samples. A pronounced difference between the V P /V S ra-
tios for dry and water-saturated chalk samples indicates promising results for seismic
amplitude-versus-offset analyses.

ified limestone data, Castagna, Batzle and Kan (1993) pro-


INTRODUCTION
vided a second-order polynomial fit to the V P /V S relationship.
Knowledge of the relationships between the petrophysical Lind (1997) found that Wyllie’s time-average model (Wyllie,
properties of reservoir rocks and the parameters that can be Gregory and Gardner 1958) can be modified to give good
observed with geophysical methods has increased during the estimates of the porosity of chalk from velocity data by appli-
last decades, but knowledge of the acoustic properties of chalk cation of the concept of critical porosity. Raiga-Clemenceau,
is still limited, even though several studies have been pub- Martin and Nicoletis (1988) suggested an acoustic formation
lished. Some of the earlier studies are as follows. For unspec- factor as an alternative to Wyllie’s time-average equation and
provided empirical constants for calcite, among other miner-
∗ E-mail: il@er.dtu.dk
als. Walls, Dvorkin and Smith (1998) fitted a modified up-
†Present address: DONG/A/S, Agern Allé 24–26, 2970 Hørsholm, per Hashin–Shtrikman (MUHS, Nur et al. 1998) bound to
Denmark. log data on chalk from the Ekofisk Field, resulting in an


C 2005 European Association of Geoscientists & Engineers 481
482 B. Røgen et al.

empirical porosity–velocity relationship. The effect of clay on present difference in burial depths of about 900 m is counter-
the porosity–velocity relationship in chalk has not been stud- acted by differences in overpressure.
ied extensively, but an effective-medium model for chalk where Based on the work of, for example, Scholle (1977), we ex-
the clay is mainly in suspension was suggested by Fabricius pect compaction to be the primary porosity-controlling factor
(2003). Empirical relationships between sonic-wave velocities, in chalk, and textural variations to be the secondary porosity-
porosity and clay content have been established for sandstone controlling factor. An important textural factor is the grain-
(e.g. Han, Nur and Morgen 1986; Wilkens et al. 1986; Yan size distribution which we have assessed by image analysis
2002). Han et al. (1986) found that as little as 1% of clay of backscatter electron micrographs. Irrespective of miner-
significantly reduces the elastic moduli of sandstones. alogy, particles of clay and fine-silt size are termed ‘mud’
In this study our aim is to establish a systematic data foun- or ‘matrix’ and particles of coarse-silt and sand size (larger
dation for the interpretation of porosity and fluid content of than 20 µm in diameter) are termed ‘large grains’. The calcite
chalk from acoustic data. Using laboratory measurements on ‘mud’ consists mainly of nannofossils whereas the large grains
core material from three Danish hydrocarbon fields, we in- are mainly microfossils and fossil fragments. The term ‘large
vestigate how fluid content, texture and mineralogy influence porosity’ covers intrafossil and moldic porosity. The chalk
the relationship between porosity and ultrasonic velocity in samples are classified according to depositional texture either
unfractured chalk. as mudstones (<10% large grains) or wackestones (>10%
To obtain samples with different burial and hydrocarbon- large grains) (Dunham 1962) (Fig. 2). The proportion of min-
filling histories while maintaining high purity and low strati- erals not soluble in hydrochloric acid is termed the ‘insolu-
graphic variation, we selected 56 chalk samples from the Tor ble residue’ and consists mainly of particles of clay size. Even
Formation of the Chalk Group from the Dan Field (25 oil- though the amount of insoluble residue is less than 5%, it can
zone samples), the South Arne Field (29 oil-zone samples) and have a significant influence on the petrophysical properties of
the Gorm Field (two water-zone samples) in the Danish North the chalk (Røgen and Fabricius 2002).
Sea (Jørgensen 1992; Mackertich and Goulding 1999; Hurst We investigate the effect of pore fluid on the elastic proper-
1983) (Fig. 1). The Tor Formation is of Maastrichtian age and ties of chalk by comparing data from dry core samples and
is the main oil-producing formation in these fields. According from water-saturated core samples. Gassmann (1951) sug-
to information from field operators, present effective stress on gested a low-frequency model for fluid substitution in a porous
reservoir rocks is nearly equal for the three fields, because the medium relating the elastic moduli of rock saturated with dif-
ferent fluids. In line with Borre (1998), Wang (2000), Borre
and Fabricius (2001), Gommesen, Mavko and Mukerji (2002)
and Japsen et al. (2004), we apply Gassmann’s relationships
(e.g. Mavko, Mukerji and Dvorkin 1998) to test whether
dry moduli can be predicted from measurements on water-
saturated chalk samples.
Similarly to Walls et al. (1998), Japsen et al. (2000) used
a modified upper Hashin–Shtrikman bound to model the
velocity–porosity relationship for the Tor Formation, based
on log data from two wells in the Dan and South Arne
Fields. We apply a modified upper Hashin–Shtrikman bound
to ultrasonic laboratory data on chalk taken from the same
two fields and partly from the same wells. To avoid textu-
ral influence only samples classified as mudstone according to
Dunham (1962) are used for this model.
In this study, we show that the influence of fluid (water
versus dry) on the ultrasonic velocities of North Sea chalk
can be estimated with reasonable accuracy using Gassmann’s
relationships. We also show that the presence of large grains
(microfossils and fossil fragments) and smectite influences the
Figure 1 Location map. elastic properties for North Sea chalk.


C 2005 European Association of Geoscientists & Engineers, Geophysical Prospecting, 53, 481–496
Ultrasonic velocities of North Sea chalk samples 483

DEPOSITIONAL TEXTURE RECOGNIZABLE DEPOSITIONAL


Original components not bound together during deposition Original components were TEXTURE NOT
Contains mud Lacks mud bound together during RECOGNIZABLE
(particles of clay and fine-silt size) and is deposition
(Subdivide according to
Mud-supported Grain-supported grain- classifications designed to
Less than More than supported bear on physical texture
10% grains 10% grains or diagenesis.)
Crystalline
Mudstone Wackestone Packstone Grainstone Boundstone
Carbonate
Figure 2 Classification of carbonate rocks according to depositional texture (Dunham 1962). Grains are defined as particles with a diameter
larger than 20 µm, and the term ‘mud’ covers smaller (mainly carbonate) particles of clay and fine-silt size. All samples in this study are classified
as either mudstone or wackestone.

METHODS and a porosity-dependent factor c. The liquid permeability can


be approximated by the Klinkenberg correction. An empirical
The sample-selection criteria were to cover both the main trend
conversion from measured gas permeability kg to Klinkenberg
and the scatter in a cross-plot of porosity and permeability
permeability kk of chalk was given by Mortensen et al. (1998).
for the available chalk samples. In order to avoid complicat-
Thus, the Kozeny equation is given by
ing factors, samples with fractures or other visible inhomo-
geneities were avoided. We used 1.5-inch vertical plugs.
ϕ3
Plugs were Soxhlet-cleaned with methanol and toluene for kl = c(ϕ) ≈ kk = 0.52kg1.083 , (1)
S2
salt and hydrocarbons and then dried at 110◦ C. Porosity was
measured as helium-porosity with an uncertainty at 0.1% where c(ϕ) can be calculated from ϕ (Mortensen et al. 1998).
point. Gas permeability was measured at a differential pres- In the porosity interval, 20–50%, c(ϕ) varies from 0.21 to
sure between 0 and 1 bar, with a confining sleeve pressure of 0.25. The Kozeny equation implies that permeability is lower
400 psi. Uncertainty of permeability data is 4%. for chalk rich in clay and silica compared with clean chalk,
The calcite content was measured by adding a surplus of due not only to sorting (poorer sorting reduces porosity) but
0.5 m hydrochloric acid to the powdered sample, and subse- also to the higher surface area (i.e. higher resistance to flow)
quent titration by 0.5 m Na-hydroxide solution. The miner- (Røgen and Fabricius 2002).
alogical composition of the insoluble residue was identified For quality control using the Kozeny equation, the specific
by X-ray diffraction (Cu-Kα) of orientated, subsequently gly- surface area was obtained by the BET method using nitrogen
colated (3 days, 60◦ C) and heated (500◦ C, 2 hours) samples. (99.999% pure) as adsorbate (Brunauer, Emmet and Teller
A Philips diffractometer with graphite monochromator and 1938). Prior to the BET measurements, the crushed samples
automatic divergence slit was used. Thin sections were used were outgassed at 70◦ C using nitrogen as a carrier gas.
for a visual classification of depositional texture according to In four samples, a high measured gas permeability compared
Dunham (1962) (Fig. 2). with the estimated matrix permeability may indicate fractures
in the samples (A782, B216, N169 and P177, Fig. 3). The
four samples that appear fractured in the permeability mea-
Quality control
surements do not appear to have lower P-wave moduli than
As a test of sample homogeneity, the gas permeability of a sam- unfractured samples when measured at 75 bar hydrostatic
ple can be compared with the estimated matrix permeability confining pressure (Fig. 14a).
predicted from the Kozeny equation. Due to the homogene- The velocity of all samples was measured at 25 bar, 50 bar
ity and connectivity of the pore space, the Kozeny equation and 75 bar hydrostatic confining pressures to test for pressure-
is valid for chalk (Mortensen, Engstrøm and Lind 1998). We dependent behaviour. Borre and Fabricius (2001) found that
can use the Kozeny equation to test whether the measured velocities stabilize as the effective stress increases to above
permeability is true matrix permeability or is dominated by half the in situ vertical effective stress. Our data confirm this
fracture permeability. The Kozeny equation expresses a rela- by showing a velocity increase as low as 4% associated with
tionship between liquid permeability kl [m2 ], porosity ϕ [frac- an increase in hydrostatic confining pressure from 25 to the
tion], specific surface area relative to bulk volume S [m2 /m3 ] approximate in situ effective stress of 75 bar (Fig. 4). This


C 2005 European Association of Geoscientists & Engineers, Geophysical Prospecting, 53, 481–496
484 B. Røgen et al.

10 (a)
Estimated matrix permeability [mD]

Dry
6 Saturated
1

vP [km/s]
4
0.1 South Arne 23% porosity
Dan
Gorm 41% porosity
2
1:1
0.01
0.1 1 10 100
0
Measured gas permeability [mD] 0 25 50 75 100
Figure 3 Estimated matrix permeability versus measured gas per- Hydrostatic confining pressure [bar]
meability. The estimated matrix permeability was calculated using
Kozeny’s equation and an empirical relationship between gas and (b)
Klinkenberg permeability (Mortensen et al. 1998). In four samples
Dry
(A782, B216, N169 and P177), a high measured gas permeability 6 Saturated
compared with the estimated matrix permeability may indicate frac-
tures in the samples.

vS [km/s]
4

shows that the measurements at 75 bar hydrostatic confining


pressure are comparable with in situ conditions. 2 23% porosity
The fluid-substitution relationships of Gassmann apply for 41% porosity

slightly moist rock samples (Mavko et al. 1998). To test if the


0
dry ultrasonic measurements were made on samples that were 0 25 50 75 100
too dry, three samples were re-measured at room moisture Hydrostatic confining pressure [bar]
conditions. This resulted in a slight reduction in the velocities,
Figure 4 Example of pressure dependence of velocity measurements
but the reduction was smaller than the estimated standard
for samples with low and high porosities. (a) Compressional-wave
deviations. Therefore it was concluded that the dry samples velocity (V P ); (b) shear-wave velocity (V S ). The increase in velocity is
were not too dry and thus the measurements are valid input approximately 4% between 25 bar and 75 bar hydrostatic confining
to Gassmann’s relationships. pressure. Eighty-seven per cent of the samples have an increase in
velocity lower than 6%.

Ultrasonic measurements on chalk samples


automated analysis, using the arrival picker software devel-
Ultrasonic compressional- and shear-wave velocities were
oped by Ødegaard A/S. Readings of traveltime for both wave
measured on 56 samples. All samples were measured in a dry
types were taken as first peak. The P- and S-wave transducers
state but, due to time constraints, only 34 samples were mea-
have a centre frequency of 700 kHz. The confining pressure
sured when saturated with calcite-equilibrated tap-water. The
was increased gradually to the next pressure step over a time
ultrasonic velocity V of the samples was calculated from the
period of 30 minutes. When unloading the core holder after
transit time t it takes an ultrasonic P- or S-wave to travel a
testing, the confining pressure was decreased gradually from
certain distance d (i.e. the sample length) through a sample,
75 bar to 0 bar over a time period of 90 minutes. For water-
i.e.
saturated plugs, the saturation was calculated from helium
d porosity and the mass of the dry and the saturated sample
V = , (2)
t − tdelay with an uncertainty of 1% point. The water saturation was
where tdelay is the system delay time. The system delay time was higher than 98%.
determined by measuring the transit time without any plugs For fluid substitution, we applied Gassmann’s relationships
and on three aluminium plugs of different lengths. Transit in the form given by Mavko et al. (1998):
times for P- and S-waves were measured on a digital oscil- Ksat Kdry Kfluid
= + , Gsat = Gdry , (3)
loscope. The P- and S-wave data were saved digitally for later K0 − Ksat K0 − Kdry ϕ(K0 − Kfluid )


C 2005 European Association of Geoscientists & Engineers, Geophysical Prospecting, 53, 481–496
Ultrasonic velocities of North Sea chalk samples 485

where K0 and Kfluid are the mineral and fluid bulk moduli, Modified upper Hashin–Shtrikman model
Ksat and Kdry are the effective bulk moduli of saturated and
The empirical modified upper Hashin–Shtrikman model
dry rock, respectively, ϕ is the porosity, and Gsat and Gdry
(MUHS) model describes how the bulk and shear moduli
are the effective shear moduli of saturated and dry rock, re-
(KMUHS and GMUHS ) increase as porosity ϕ is reduced from the
spectively. For isotropic elastic materials, the elastic moduli
high-porosity end member ϕ max to zero porosity. This model
are expressed by the compressional- and shear-wave veloci-
is given by the equations,
ties, V P and V S , respectively, and the bulk density ρ. The P-
wave modulus M is given by M = ρV 2P , the shear modulus 1 1 − ϕmax ϕmax
= + , (4)
G is given by G = ρV 2S and the bulk modulus K is given by KRmax Ks Kf
K = M − 4/3G.
KMUHS
Standard deviations of sample length, transit time and den- ϕ/ϕmax
sity were estimated as 0.1 mm, 0.09 µs and 0.02 g/cm2 , re- = Ks + ,
(KRmax − Ks )−1 + (1 − ϕ/ϕmax ) (Ks + 4/3Gs )−1
spectively. Standard deviations of ultrasonic velocities and of
(5)
elastic moduli were estimated by error propagation. The esti-
mated standard deviations s are as follows: s(V P ) < 0.08 km/s, ϕ/ϕmax
GMUHS = Gs +  , (6)
−1 2(1−ϕ/ϕmax )(Ks +2Gs )
s(V S ) < 0.03 km/s, s(M) < 1.76 GPa, s(G) < 0.41 GPa and (−Gs ) + 5Gs (Ks +4/3Gs )
s(K) < 1.84 GPa, where M is the P-wave modulus, G is the
shear modulus and K is the bulk modulus. MMUHS = KMUHS + 4/3GMUHS . (7)

At the high-porosity end member, the bulk modulus is given


Microtextural analysis of chalk samples as the modulus at the Reuss suspension curve (KRmax ) and the
by image analysis shear modulus is zero. Ks and Gs denote the estimated bulk
and shear moduli of the solid, respectively, and Kf denotes
Image analyses for this study were performed on backscat-
the bulk modulus of the fluid. The modified upper Hashin–
ter micrographs, following the method for determination of
Shtrikman model is fitted to bulk, shear and P-wave moduli
grain-size distribution described by Røgen, Gommesen and
at the same time, resulting in a set of estimated parameters:
Fabricius (2001). In this method, grain-size distributions are
Ks , Gs and ϕ max .
defined by cross-sectional area in a two-dimensional plane.
Vertically orientated samples with a height and width of ap-
proximately 1 cm were impregnated with epoxy and polished R E S U LT S A N D D I S C U S S I O N
before imaging in a scanning electron microscope. A standard
Overall porosity–depth trend
imaging procedure was followed for each sample. This in-
cludes an overview image of the sample to check for inho- A general loss of porosity with depth may be observed in
mogeneities in the sample or in preparation. Then four image chalk (Scholle 1977; Borre and Fabricius 1998; Japsen 1998).
fields are chosen along the centre-line in the vertical direc- This trend reflects compaction of the chalk due to an increase
tion with spacing of 2 mm. Each image field is represented by in effective stress. However, textural variations of the chalk
two images. One image at large magnification (representing overshadow this general trend in the depth intervals studied
60 µm × 80 µm) shows only the chalk matrix, the insoluble (Fig. 6).
residue and the matrix porosity. The second image is a con-
centric image at 10 times smaller magnification (representing
Influence of porosity on ultrasonic velocity
600 µm × 800 µm) showing microfossils, chalk clasts and
shell debris as well as large porosity (intraparticular and We found negative correlations between porosity and P- and
moldic porosity), but not resolving the matrix. By combin- S-wave velocities for dry as well as saturated samples (Fig. 7;
ing the information from each of the magnifications and as- Table 1). In the porosity range 14–45%, we found V P in
suming a homogeneous matrix, we obtain information cor- the range 4.3–1.9 km/s and V S in the range 2.6–1.3 km/s;
responding to 100 images at the large magnification for these results are similar to those of Borre (1998). We found
each image field of the small magnification. Examples of that the V P /V S ratios are distinctly different for dry and satu-
grain-size distributions from image analysis are shown in rated samples (Fig. 8). For dry samples, we observe a negative
Fig. 5. correlation with porosity, whereas for saturated samples the


C 2005 European Association of Geoscientists & Engineers, Geophysical Prospecting, 53, 481–496
486 B. Røgen et al.

Mudstone
10

1
0.1

10
0.01

100

1000
0.001

10000

100000
Wackestone
10

10
0.1

100

1000
0.01

10000
0.001

100000
Wackestone
10

0
1

10
0.1
0.01

100
0.001

1000

10000

100000

Grain area [µm2]

200 µm

Figure 5 Backscatter micrographs representing 600 µm × 800 µm and final grain-size distributions of a mudstone (< 10% grains) and two
wackestones (> 10% grains) (micrographs at large magnification showing matrix particles are not shown). The isolated peak to the left in the
histograms represents insoluble residues that are not resolved in the images. The mudstone contains hollow microfossils. The two wackestones
contain filled microfossils.

correlation between V P /V S and porosity is positive. For dry compared at, for example, 30% porosity, the V P /V S ratios in
samples, the V P /V S ratio ranges from 1.70 to 1.46 in the this study are similar to those for saturated samples found in
porosity range 14–45%. For saturated samples, the V P /V S ra- other studies, but they are slightly smaller than those found
tio ranges from 1.80 to 2.10 in the same porosity range. When for dry samples in other studies (Table 2).


C 2005 European Association of Geoscientists & Engineers, Geophysical Prospecting, 53, 481–496
Ultrasonic velocities of North Sea chalk samples 487

Porosity [%] Porosity [%] ships between V P and V S for dry and saturated samples (Figs 8
0 25 50 0 25 50 and 10). The bulk and P-wave moduli are higher for saturated
x00 y00 samples than for dry samples, whereas the shear moduli for
saturated samples are slightly lower than for the dry samples;
this is probably a result of fluid–solid interactions weakening
20
x20 y20
20 grain contacts (Fig. 11; compare Japsen et al. 2004). This ef-
fect was also observed by Assefa, McCann and Sothcott (2003)
for oolitic limestone. When we apply Gassmann’s relationship
Relative depth [m]

Relative depth [m]


x40
40 y40
40 (3) to data for saturated samples and predict dry moduli, we
obtain good agreement with the measured dry data (Fig. 12).
The Gassmann-estimated P- and S-wave velocities of dry sam-
60
x60 60
y60 ples compare well with the measured velocities of dry samples
(Fig. 13). In general, velocities are slightly underestimated by
Gassmann’s relationship, although V P values of high-porosity
80
x80 80
y80 samples are an exception with higher estimated than measured
A
B
values. This is in agreement with the local flow model of a
M
C N rock with medium-to-low aspect ratio pores (Mavko et al.
x100
100 y100
100 1998; Wang 2000). When defining the Gassmann deviation
Figure 6 Porosity versus vertical depth below uppermost sample for as ((V Gassmann − V measured )/V measured ) (Wang 2000), an average
the South Arne Field (wells A, B and C) and the Dan Field (wells M and underestimation of dry V P of 2% and of dry V S of 4% is
N). Normal compaction resulting in porosity reduction for increasing observed (Fig. 13). This observation is consistent with results
depth can be observed within wells A, C and N. The variations are
obtained by Borre (1998) and Borre and Fabricius (2001) for
interpreted as variations in texture (here focused on the proportion of
large grains, clay mineralogy and cementation). chalk samples from the Gorm and Tyra Fields.
In line with the results of Borre (1998), we conclude that
5
Gassmann’s relationships give good estimates of velocities of
4 dry samples when applied to ultrasonic measurements on
water-saturated chalk. In contrast, Wang (1997) found that
3
v [km/s]

V P of carbonates is underestimated by Gassmann’s relation-


2 ships by up to 25%. Carbonate rocks can be heterogeneous
vp(saturated)
vp(dry) and have high permeability so the results of Wang (1997) may
1
vs(dry) not be comparable with the pelagic chalk we deal with in this
vs(saturated)
0 study. The reason that Gassmann’s relationships give good re-
0 10 20 30 40 50
sults for chalk may be the homogeneity and low permeability
Porosity [%]
of the chalk, where individual particles and pores are so much
Figure 7 P- and S-wave velocities versus porosity of 33 saturated and smaller than the applied wavelength that the ultrasonic wave
55 dry chalk samples.
experiences the chalk as a homogeneous medium. Matrix par-
ticles are about 3 µm in diameter, the largest observed grain
Influence of fluid content on velocity
is approximately 100 µm in diameter, and the applied fre-
We observe that V P values for saturated samples are equal to quency results in wavelengths between 700 µm and 6000 µm.
V P values for dry samples with the highest velocities, but are Gassmann’s relationships represent a low-frequency theory, so
slightly higher for samples with the lowest velocities (i.e. high when we observe that it applies to ultrasonic data, it probably
porosities) (Fig. 9; Table 1). This increase in saturated V P val- also applies to the lower-frequency logging data and seismic
ues can be seen for samples with porosities above 35% (V P data.
below 3 km/s) and is due to the increasing influence of fluid
properties on acoustic properties at high porosities. For shear-
Influence of microtexture on velocity
wave velocities we observe lower velocities for saturated sam-
ples than for dry samples, in accordance with the difference in Most Dan Field samples have a lower P-wave modulus for
densities (Fig. 9). Consequently, we observe distinct relation- a given porosity than the corresponding South Arne Field


C 2005 European Association of Geoscientists & Engineers, Geophysical Prospecting, 53, 481–496
488 B. Røgen et al.
Table 1 Data for 56 chalk samples of the Tor Formation in the North Sea

Carbonate Specific
Gas content surface Matrix <3.16 µm2 >316 µm2 X-ray data VP VS VP VS Water
Sample Porosity permeability (% of area porosity % (rel. to % (rel. to (dry) (dry) (sat) (sat) saturation
no. (%) (mD) solids) (m2 /g) Texture (%) solid) solid) Quartz Kaolinite Smectite Other (km/s) (km/s) (km/s) (km/s) (%)

A755 24.5 0.52 97.9 1.75 M 25 16 5 2 3 b, i 4.05 2.42 4.08 2.24 99


A759 27.2 1.42 97.2 1.75 M 28 25 8 2 2 f, i- s, p 3.59 2.15 – – –
A760 29.3 1.15 98.2 1.90 W 30 13 14 1 3 f, i 3.48 2.16 3.60 1.98 99

C

A761 30.4 1.48 98.6 1.70 W 32 14 12 2 2 1 b, f, i, p – – 3.47 1.86 99


2005 European Association of Geoscientists & Engineers, Geophysical Prospecting, 53, 481–496

A763 28.0 1.47 98.7 1.53 W 30 12 15 2 3 f, i 3.52 2.18 3.60 1.98 100
A767 30.9 2.15 98.9 1.57 M 33 15 11 2 i-s 3.47 2.12 3.53 1.96 100
A771 32.1 2.41 98.7 1.68 W 34 16 13 1 b, f, i, i-s 3.33 2.03 3.40 1.82 100
A774 28.0 1.34 99.0 1.62 M 30 20 8 3 f, i 3.66 2.23 – – –
A779 22.6 0.40 99.3 1.17 M 23 15 5 3 f, i-s, p 3.97 2.37 – – –
A780 30.9 1.99 99.3 1.49 M 32 16 6 3 1 f, i-s, p 3.38 2.08 – – –
A782 24.8 3.28 99.6 1.23 M 25 15 2 3 1 f, i-s, p 3.68 2.25 – – –
A783 20.1 0.96 98.7 1.29 M 21 16 3 3 f, i-s, p 3.98 2.42 – – –
B216 13.5 0.54 93.1 2.04 W 15 20 15 3 2 f, i-s, p 4.30 2.60 – – –
B220 30.1 1.25 92.1 2.13 M 30 26 2 2 2 f, i 3.33 2.07 3.33 1.82 100
B236 35.5 2.52 95.6 1.63 M 37 23 9 3 2 f, i-s, p 3.15 2.05 – – –
B240 37.6 4.07 96.7 1.54 M 39 23 9 3 2 f, i-s 2.57 1.68 – – –
B244 37.4 3.50 97.1 1.60 M 39 21 8 3 2 f, i-s, p 2.86 1.81 – – –
B260 41.1 4.10 96.0 1.92 M 41 25 5 2 2 f, i, p 2.45 1.62 2.63 1.37 100
B264 40.2 4.51 97.4 1.94 M 40 26 3 2 2 b, f, i 2.43 1.58 2.58 1.34 100
B274 38.0 3.62 97.5 1.40 M 38 20 4 3 2 i-s 2.68 1.76 – – –
C121 42.5 8.36 96.6 1.77 M 42 23 7 2 2 i, p 2.09 1.42 2.44 1.25 100
C128 42.7 7.17 96.3 1.73 M 43 21 7 2 2 i, p 2.06 1.37 2.37 1.13 100
C136 41.4 5.76 96.5 1.71 M 42 22 6 2 2 i, p 2.07 1.39 2.43 1.22 100
C151 43.7 7.40 95.9 1.42 M 43 34 1 2 2 i, p 2.20 1.44 2.48 1.26 99
C162 43.2 8.87 96.9 1.62 M 44 23 5 3 2 f, i, p 2.32 1.51 – – –
C176 39.5 4.78 96.1 1.65 M 40 25 4 2 2 f, i, p 2.70 1.73 – – –
C200 45.0 8.60 97.0 1.37 M 44 43 1 2 2 i, p 1.98 1.3 2.31 1.12 99
C204 41.3 5.93 97.4 1.78 M 43 17 9 2 2 f, i, p 1.96 1.34 2.36 1.19 100
C212 42.7 6.29 96.9 1.90 M 43 25 4 2 2 b, f, i, p 2.06 1.39 2.43 1.23 100

C
2005 European Association of Geoscientists & Engineers, Geophysical Prospecting, 53, 481–496

Table 1 (Continued)

Carbonate Specific
Gas content surface Matrix <3.16 µm2 >316 µm2 X-ray data VP VS VP VS Water
Sample Porosity permeability (% of area porosity % (rel. to % (rel. to (dry) (dry) (sat) (sat) saturation
no. (%) (mD) solids) (m2 /g) Texture (%) solid) solid) Quartz Kaolinite Smectite Other (km/s) (km/s) (km/s) (km/s) (%)

M001 24.0 0.96 95.0 1.19 W 26 21 15 2 2 f, i, p 3.28 2.11 3.36 1.82 99


M002 27.1 1.27 96.9 1.28 W 29 14 13 3 1 f, p 3.51 2.13 3.57 – 99
M003 23.9 0.52 95.6 1.41 M 25 16 10 2 2 f, i 3.39 2.12 3.52 1.89 100
M004 28.0 0.55 95.8 1.63 W 29 16 13 2 2 i, p 2.97 1.89 3.11 1.62 100
M005 18.7 0.44 93.9 1.87 M 19 30 8 2 2 f, i, p 3.67 2.32 3.71 2.01 95
M006 23.7 0.63 97.6 1.22 M 25 12 9 2 2 p 3.83 2.33 3.91 2.14 100
M008 29.9 3.79 96.6 1.19 M 32 19 12 3 1 f, i, p 3.02 1.91 3.14 1.70 100
M014 25.8 1.28 96.1 1.49 M 26 27 6 2 2 f, i, p 3.18 2.01 3.20 1.74 100
M015 29.9 2.69 97.8 1.08 M 32 12 12 2 2 p 3.11 1.97 3.25 1.75 100
M017 31.3 1.83 97.1 1.27 W 39 12 32 3 1 p 3.10 1.93 3.22 1.72 100
M020 29.4 1.60 98.7 1.16 W 32 11 16 2 2 i, p 3.09 1.97 3.26 1.75 100
M021 30.3 2.58 97.4 1.14 M 31 23 8 2 2 f, i, p 3.18 1.95 3.22 1.70 99
M022 29.0 1.37 96.4 1.45 W 32 13 15 2 2 i, p 2.93 1.89 3.09 1.67 99
M030 29.7 1.28 98.4 1.31 M 30 13 6 2 2 i, p 3.14 1.96 3.23 1.74 100
M031 31.5 3.28 98.9 1.05 M 34 9 12 3 1 i 2.84 1.84 2.97 1.60 99
M032 28.8 2.29 97.1 1.41 M 30 21 9 3 1 f, i, p 3.21 2.00 3.27 1.75 100

Ultrasonic velocities of North Sea chalk samples 489


M035 32.3 3.23 97.9 1.24 M 33 25 5 2 2 f, i, p 2.84 1.81 2.95 1.57 100
M037 31.7 2.32 98.0 1.17 M 32 12 3 3 1 i, p 2.96 1.85 3.11 1.66 100
N166 26.7 1.91 98.7 1.28 W 33 14 28 2 2 fl, i, p 2.99 1.91 – – –
N167 26.5 2.45 98.8 1.00 W 31 14 21 1 3 f, fl, i 3.32 2.07 – – –
N169 25.6 5.31 98.1 1.26 W 31 16 23 2 2 f, i 3.19 2.00 – – –
N170 26.9 2.24 98.3 1.18 W 30 16 19 2 2 fl, p 3.10 1.95 – – –
N171 25.8 1.66 97.4 1.34 M 27 24 7 2 2 i, p 3.26 2.06 – – –
N172 23.5 1.30 98.5 1.11 W 28 13 21 2 2 i, p 3.46 2.17 – – –
N175 22.2 0.94 97.9 1.33 W 26 17 17 2 3 i, p 3.67 2.23 – – –
P176 23.1 0.83 97.9 1.29 W 28 18 21 2 2 b, f, i 4.02 2.37 – – –
P177 20.8 34.03 98.5 1.06 W 27 14 30 2 1 3 i, p 4.18 2.46 – – –

The letter in sample numbers refers to the well. Wells A, B and C are in South Arne Field; wells M and N are in Dan Field; well P is in Gorm Field. Depositional texture (Dunham 1962): M denotes
mudstone, W denotes wackestone (see Fig. 2). Grain-size distribution from image analysis is given as proportion of solids with a cross-sectional area smaller than 3.16 µm2 and proportion of solids with
a cross-sectional area larger than 316 µm2 . Key for X-ray data: 1, present; 2, contributor; 3, dominating; 4, sole detected. For other minerals: b, barite; f, feldspar; fl, fluorite; i, illite; i-s, illite–smectite
interstratified; p, pyrite. Velocities are measured at a hydrostatic confining pressure of 75 bar. Water saturation of dry samples is 0–1%.
490 B. Røgen et al.

2.5
Saturated
6
Dry

vP (saturated) [km/s]
2
4
vP/vS [-]

1.5 2
Calcite (Mavko et al . 1998)
1:1
0
1 0 2 4 6
0 10 20 30 40 50 vP (dry) [km/s]
Porosity [%]

Figure 8 V P /V S ratio versus porosity for dry and saturated samples. 6


In the porosity range, 14–45%, dry samples have V P /V S ratios from

vS (saturated) [km/s]
1.70 to 1.46 and saturated samples have V P /V S ratios from 1.80 to
2.10. Compare with V P /V S of 1.94 for calcite (Mavko et al. 1998). 4

Table 2 V P /V S ratio at 30% porosity from different studies, with


origin of chalk samples 2

This study Borre (1998) Hvid (1998) 1:1


0
V P /V S @ Dan and South Gorm Copenhagen
0 2 4 6
30% porosity Arne Fields Field limestone
vS (dry) [km/s]

Dry 1.5–1.6 1.7–1.9 1.6–2.1 Figure 9 Velocity for saturated versus dry samples. Mineral endpoint
Saturated 1.8–1.9 1.8–2.1 – indicated with asterisk at highest velocity and fluid endpoint indicated
with asterisk at lowest velocity. See text for explanation.
4
Dry
samples (Fig. 14a; Table 1). This may be due to textural differ- Saturated
ences where we observe that the samples from the Dan Field 3
contain smectite and many are classified as wackestones. In
vS [km/s]

contrast, the samples from the South Arne Field are domi- Castagna
2
nated by mudstone texture and contain kaolinite. Due to the
limited amount of large porosity (less than 2%) present in the 1
studied samples, we cannot reach a conclusion on the effect of
large porosity on the elastic properties (Fig. 14d). 0
In backscatter micrographs we observe that the microfossils 0 2 4 6
appear hollow in some samples and filled with cement in other vP [km/s]
samples (Fig. 15). The samples with filled microfossils do not
Figure 10 Relationship between V P and V S for both dry and satu-
generally have pore-filling cement in the matrix. The presence rated samples. Mineral and fluid endpoints are indicated with aster-
of large grains in the form of filled microfossils in a mud- isks (Mavko et al. 1998). Note the distinct relationships between V P
supported chalk reduces porosity and indicates softer rather and V S for dry and saturated samples. The measured V S values of
than stiffer chalk, compared with a pure mudstone with the saturated samples are slightly higher than V S values estimated from
V P values by application of the empirical relationship for limestone
same porosity (Fig. 14b). The porosity-reducing effect of filled
of Castagna et al. (1993). Observe that the empirical relationship of
microfossils is thus more significant than a possible sediment- Castagna et al. (1993) does not fit the endpoints.
stiffening effect. This may also be due to the differences in clay
mineralogy associated with this texture. stiffness of the samples. Smectite is present in the softest sam-
The samples have high calcite content, mostly over 95%, ples which also tend to have wackestone texture (from the
and all samples contain quartz. Despite the low clay content Dan Field), whereas kaolinite is present in the stiffest sam-
we observe that the mineralogy of the clay is related to the ples, which are dominated by mudstone texture (from the


C 2005 European Association of Geoscientists & Engineers, Geophysical Prospecting, 53, 481–496
Ultrasonic velocities of North Sea chalk samples 491

25 25
1:1 1:1

K(Gassmann dry) [GPa]


20 20
K(saturated) [GPa]

15 15

10 10

5 5

0 0
0 5 10 15 20 25 0 5 10 15 20 25
K(dry) [GPa] K(dry) [GPa]
15 15
1:1 1:1

G(Gassmann dry) [GPa]


G(saturated) [GPa]

10 10

5 5

0 0
0 5 10 15 0 5 10 15
G(dry) [GPa] G(dry) [GPa]
40
40
1:1
1:1
M(Gassmann dry) [GPa]
M(saturated) [GPa]

30
30

20
20

10
10

0
0
0 10 20 30 40
0 10 20 30 40
M(dry) [GPa]
M(dry) [GPa]
Figure 11 Elastic moduli of saturated samples versus moduli of dry
Figure 12 Estimated moduli of dry samples calculated from saturated
samples. Bulk (K), shear (G) and P-wave (M) moduli are plotted in
samples using Gassmann’s relationships versus measured moduli of
the upper, middle and lower figures, respectively. The shear moduli
dry samples. Bulk (K), shear (G) and P-wave (M) moduli are plot-
(G) of the saturated samples are slightly lower than those of the dry
ted in the upper, middle and lower figures, respectively. In general,
samples, whereas the bulk (K) and P-wave moduli of water-saturated
Gassmann’s relationships give good estimates of the elastic moduli of
samples are higher than those of dry samples due to the influence of
dry chalk samples calculated from saturated samples. Note, however,
the moduli of the fluid present. The estimated standard deviations of
the increasing error for the shear modulus with increasing velocity
measurements of K, G and M are 1.84 GPa, 0.41 GPa and 1.76 GPa,
(see Fig. 13). The correlation coefficients between measured dry mod-
respectively.
uli and Gassmann-predicted dry moduli are 0.99, 0.96 and 0.98 for
K, G and M, respectively.

South Arne Field) (Fig. 14c). Variations in clay mineralogy


can be due either to geographical depositional differences or in the ϕ–M domain is 5 percentage points. The differences in
to temperature-related diagenesis. The volume of clay minerals moduli for samples with different dominant clay mineralogies
is less than 1% (Røgen and Fabricius 2002) whereas the dif- can have several explanations. Kaolinite has higher mineral
ference in porosity between the kaolinite and smectite trends moduli than smectite, but with the small volumes present, this


C 2005 European Association of Geoscientists & Engineers, Geophysical Prospecting, 53, 481–496
492 B. Røgen et al.

(a) (c)
5 5
Gassmann vP(dry) [km/s]

Gassmann vS(dry) [km/s]


4 4

3 3

2 2

1 1
1 2 3 4 5 1 2 3 4 5
Measured vP(dry) [km/s] Measured vS(dry) [km/s]

(b) (d)
6 6
Gassmann deviaton of vP [%]

Gassmann deviaton of vS [%]


4 4
2 2
0 0
-2 -2
-4 -4
-6 -6
-8 -8
-10 -10
10 20 30 40 50 10 20 30 40 50
Porosity [%] Porosity [%]

Figure 13 (a) Gassmann-estimated V P (dry) versus measured V P (dry). (b) Gassmann deviation of V P versus porosity. The Gassmann deviation
is defined as ((V Gassmann − V measured )/V measured ) (Wang 2000). Gassmann’s relationship underestimates V P on average by 2%. (c) Gassmann-
estimated V S (dry) versus measured V S (dry). (d) Gassmann deviation of V S versus porosity. Gassmann’s relationship consequently underestimates
V S (on average by 4%).

is hardly the main reason. Smectite may have a softening effect and 0.95 for bulk, shear and P-wave moduli, respectively. The
on the calcite frame that kaolinite does not have and may thus empirical value for elastic properties at 0% porosity is lower
be an indicator of a softer calcite frame. than the mineral value for calcite, and is not well determined
when low-porosity samples are not present in the data set. It
can be seen that the same correlation coefficients are obtained
Empirical velocity–porosity relationship with the modified upper Hashin–Shtrikman model fitted to
In order to establish a texturally independent moduli–porosity log data (Japsen et al. 2000) or even with a simple linear re-
relationship for chalk, a modified upper Hashin–Shtrikman gression (Fig. 16). The linear regression model is: Kdry = 27 −
bound was fitted to the elastic moduli (bulk, shear and P-wave) 0.6ϕ; Gdry = 19 − 0.4ϕ; Mdry = 53 − ϕ. We conclude that
and porosity data for 37 dry mudstone samples. Mudstones in the present case, there is no justification for preferring the
are a better choice than wackestones because mudstones con- modified upper Hashin–Shtrikman model to a linear model.
stitute a more homogeneous group with respect to grain-size
distribution. The bulk and shear moduli of the fluid (air) were
set at 150 kPa and zero, respectively. The estimated parameters
CONCLUSIONS
defining the model are bulk and shear moduli at zero poros-
ity of 40 GPa and 28 GPa, respectively, and a high-porosity On the basis of the experimental results from 56 chalk samples
end member at 52%. Correlation coefficients are 0.93, 0.96 from the Tor Formation in the Dan, South Arne and Gorm


C 2005 European Association of Geoscientists & Engineers, Geophysical Prospecting, 53, 481–496
Ultrasonic velocities of North Sea chalk samples 493

(a) (c)
50 50
South Arne Kaolinite
Dan Smectite
40 Gorm 40
Other
High permeability
M(dry) [MPa]

M(dry) [MPa]
30 30

20 20

10 10

0 0
10 20 30 40 50 10 20 30 40 50
Helium porosity [%] Helium porosity [%]

(b) (d)
50 50
Mudstones Up to 1% large porosity
Wackestones > 1% large porosity
40 40
M(dry) [MPa]

M(dry) [MPa]
30 30

20 20

10 10

0 0
10 20 30 40 50 10 20 30 40 50
Helium porosity [%] Helium porosity [%]

Figure 14 P-wave moduli for dry samples versus helium porosity. (a) Fields are marked; (b) depositional texture; (c) dominant clay mineral;
(d) amount of large porosity. The South Arne and Dan Field samples form almost separate trends that may be explained by differences in texture.
The increased content of filled microfossils in wackestones compared with mudstones reduces porosity and indicates softer rather than stiffer
chalk, even in the limited amounts present in the studied samples. Clay minerals are present in such small amounts that the porosity-reducing
effect can be discounted. The differences in moduli for samples with different dominant clay mineralogies can have two explanations: firstly,
that kaolinite has higher mineral moduli than smectite; secondly, either that smectite has a softening effect on the calcite frame that kaolinite
does not have, or the opposite, that kaolinite has a stiffening effect on the calcite frame that smectite does not have. The limited amount of large
porosity in hollow microfossils present in these samples prevents conclusions about a possible effect. Four samples that appear fractured in the
permeability measurements are marked in (a), and do not appear to have lower P-wave moduli than the unfractured samples.

Figure 15 Backscatter micrographs of samples with hollow microfossils to the left and with filled microfossils to the right. Each image represents
800 µm × 600 µm. Together with fossil fragments, the microfossils are termed large grains (visible in thin section and larger than 316 µm2 in
image grain-size analysis). Large grains are seen as white areas, large porosity as black areas and the matrix particles are not resolved. Matrix
particles are nannofossils, small fossil fragments and insoluble residue of silt and clay size (smaller than 3.16 µm2 in image grain-size analysis).


C 2005 European Association of Geoscientists & Engineers, Geophysical Prospecting, 53, 481–496
494 B. Røgen et al.

hydrocarbon fields in the North Sea, we conclude that porosity


is the main factor controlling the ultrasonic velocity of North
Sea chalk, and that a decrease in ultrasonic velocity due to
large grains and smectite can be detected as a secondary factor.
We also conclude that Gassmann’s relationships are applicable
to ultrasonic measurements on North Sea chalk.
Gassmann’s relationships give good estimates of the elastic
moduli of dry samples when predicted from ultrasonic V P and
V S data from water-saturated samples. For 32 samples within
the porosity range 14–45%, dry V P and V S are underestimated
by 2% and 4% on average, respectively, with 10% maximal
error.
Within the porosity range 14–45%, samples from the South
Arne Field are found to have slightly higher elastic moduli
than samples from the Dan Field, 90 km away. This observa-
tion is surprising, considering the high carbonate content of
all samples, and we suggest that the difference is related to
textural differences between the chalk at the two locations, as
indicated by differences in the large-grain content (microfos-
sils and fossil fragments) and differences in clay mineralogy.
For a given velocity, the presence of filled microfossils indi-
cates lower porosity. For a given velocity, samples with smec-
tite tend to have lower porosity than samples with kaolinite.
Indurated water-zone samples from the Gorm Field with pore-
filling cement have relatively high P-wave velocities for a given
porosity.
Comparison of ultrasonic V P and V S data shows negative
correlation between porosity (18–45%) and ultrasonic veloc-
ity for dry as well as water-saturated chalk samples. A modi-
fied upper Hashin–Shtrikman model was fitted to the data for
37 dry samples classified as mudstone, but linear regression of
moduli and porosity data gave an equally good fit. Within the
porosity range 14–45%, the V P /V S ratio is found to show neg-
ative correlation with porosity for dry samples, and positive
correlation with porosity for water-saturated samples. This
difference between the V P /V S ratio of dry and saturated sam-
ples indicates that fluid differences may be detected through
amplitude-versus-offset analyses of seismic data. The results
Figure 16 Empirical models and measured data of moduli versus he- of this study indicate that separation of porosity and fluid ef-
lium porosity for dry mudstone samples (less than 10% grains). In
fects may be possible on seismic data obtained from chalk.
black: plug data, modified upper Hashin–Shtrikman model (MUHS)
This could have important implications for exploration and
for plug data (full line) and linear regression (dotted line). In grey: Dan
Field log data and corresponding modified upper Hashin–Shtrikman reservoir monitoring of hydrocarbon reservoirs in chalk.
model (Japsen et al. 2000). Correlation coefficients between any of
the three models (MUHSplug , MUHSlog , linear regression) and mea- ACKNOWLEDGEMENTS
sured values of elastic moduli of core plugs are 0.93, 0.96 and 0.95
for bulk (K), shear (G) and P-wave modulus (M), respectively. Error This paper presents results from the ‘Rock Physics of Chalk’
bars indicate the estimated standard deviations of measurements of K, research project. The project was funded by the Danish Energy
G and M; these are 1.84 GPa, 0.41 GPa and 1.76 GPa, respectively. Research Programme, EFP-98. Core material, as well as log


C 2005 European Association of Geoscientists & Engineers, Geophysical Prospecting, 53, 481–496
Ultrasonic velocities of North Sea chalk samples 495

data and seismic data from recent wells, were kindly provided Adjacent Onshore Areas (eds J.P. Kaasschieter and T.J.A. Reijers),
by the operators of the Dan and the South Arne Fields, Mærsk pp. 157–168. Geologie en Mijnbouw 62. Royal Geological and
Mining Society of the Netherlands, The Hague.
Olie og Gas AS and Amerada Hess A/S, respectively. For tech-
Hvid J.M. 1998. Influence of depositional texture and porosity on
nical assistance, we thank the staff at GEUS Core Laboratory,
ultrasonic wave propagation in Danian Limestone from Eastern
at the Department of Physics, DTU, and at Environment & Denmark. In: Nordic Petroleum Technology Series: IV, Research in
Resources, DTU. We also thank Mark Sams, Kelvin Gylden- Petroleum Technology (ed. M.F. Middleton), pp. 125–154. Nordisk
holm and an anonymous reviewer for thorough reviewing and Energiforskningsprogram, Ås, Norway.
suggestions for improvements. Japsen P. 1998. Regional velocity-depth anomalies, North Sea chalk.
A record of overpressure and Neogene uplift and erosion. AAPG
Bulletin 82, 2031–2074.
REFERENCES Japsen P., Bruun A., Fabricius I.L., Rasmussen R., Vejbcek O.V.
and Pedersen J.M. et al. 2004. Influence of porosity and pore
Assefa S., McCann C. and Sothcott J. 2003. Velocities of compres- fluid on acoustic properties of chalk: AVO response from oil,
sional and shear waves in limestones. Geophysical Prospecting 51, South Arne Field, North Sea. Petroleum Geoscience 10, 319–
1–13. 330.
Borre M. 1998. Ultrasonic velocity of North Sea Chalk –predicting sat- Japsen P., Wagner H., Gommesen L. and Mavko G. 2000. Rock
urated data from dry. In: Nordic Petroleum Technology Series: IV, physics of chalk: modelling the sonic velocity of the Tor Formation,
Research in Petroleum Technology (ed. M.F. Middleton), pp. 71– Danish North Sea. 62nd EAGE conference, Glasgow, Scotland, Ex-
98. Nordisk Energiforskningsprogram, Ås, Norway. tended Abstracts, P0170.
Borre M. and Fabricius I.L. 1998. Chemical and mechanical processes Jørgensen L.N. 1992. Dan Field; Denmark, Central Graben, Danish
during burial diagenesis of chalk: an interpretation based on spe- North Sea. In: AAPG Treatise of Petroleum Geology, Atlas of Oil
cific surface data of deep-sea sediments. Sedimentology 45, 755– and Gas Fields (eds N.H. Foster and E.A. Beaumont), pp. 199–218.
769. American Association of Petroleum Geologists.
Borre M. and Fabricius I.L. 2001. Ultrasonic velocities of water satu- Lind I. 1997. A modified Wyllie equation for the relationship between
rated chalk from Gorm field, Danish North Sea: sensitivity to stress porosity and sonic velocity of mixed sediments and carbonates from
and applicability of Gassmann’s equation. In: Nordic Petroleum the Caribbean Sea. In: Nordic Petroleum Technology Series: III,
Technology Series: V, Research in Petroleum Technology (ed. Research in Petroleum Technology (ed. M.F. Middleton), pp. 123–
I.L. Fabricius), pp. 1–18. Nordisk Energiforskningsprogram, Ås, 137. Nordisk Energiforskningsprogram, Ås, Norway.
Norway. Mackertich D.S. and Goulding D.R.G. 1999. Exploration and ap-
Brunauer A., Emmet P.H. and Teller E. 1938. Adsorption of gases in praisal of the South Arne field, Danish North Sea. In: Petroleum
multimolecular layers. Journal of the American Chemical Society Geology of Northwest Europe: Proceedings of the 5th Conference
60, 309–319. (eds A.J. Fleet and S.A.R. Boldy), pp. 959–974. Geological Society,
Castagna J.P., Batzle M.L. and Kan T.K. 1993. Rock physics – London.
the link between rock properties and AVO response. In: Offset- Mavko G., Mukerji T. and Dvorkin J. 1998. The Rock Physics Hand-
Dependent Reflectivity – Theory and Practice of AVO Analysis (eds book. Cambridge University Press.
J.P. Castagna and M.M. Backus), pp. 135–171. Investigations in Mortensen J., Engstrøm E. and Lind I. 1998. The relation among
Geophysics 8. Society of Exploration Geophysicists. porosity, permeability, and specific surface of chalk from the Gorm
Dunham R.J. 1962. Classification of carbonate rocks according to de- Field, Danish North Sea. SPS Reservoir Evaluation and Engineer-
positional texture. In: Classification of Carbonate Rocks (ed. W.E. ing, 245–251.
Ham), pp. 108–123. Memoir 1. American Association of Petroleum Nur A., Mavko G., Dvorkin J. and Galmudi D. 1998. Critical poros-
Geologists. ity: a key to relating physical properties to porosity in rocks. The
Fabricius I.L. 2003. How burial diagenesis of chalk sediments Leading Edge 17, 357–362.
controls sonic velocity and porosity. AAPG Bulletin 87, 1755– Raiga-Clemenceau J., Martin J.P. and Nicoletis S. 1988. The con-
1778. cept of acoustic formation factor for more accurate porosity deter-
Gassmann F. 1951. Über die Elastizität poröser Medien. Viertel- mination from sonic transit time data. The Log Analyst 29, 54–
jahrsschrift der Naturforschenden Gesellschaft in Zürich 96, 1– 60.
23. Røgen B. and Fabricius I.L. 2002. Influence of clay and silica on per-
Gommesen L., Mavko G. and Mukerji T. 2002. Fluid substitution meability and capillary entry pressure of chalk reservoirs in the
studies for North Sea chalk logging data. 72nd SEG meeting, Salt North Sea. Petroleum Geoscience 8, 287–293.
Lake City, USA, Expanded Abstracts, 340–343. Røgen B., Gommesen L. and Fabricius I.L. 2001. Grain size distribu-
Han D., Nur A. and Morgen D. 1986. Effects of porosity and clay tions of chalk from image analysis of electron micrographs. Com-
content on wave velocities in sandstones. Geophysics 51, 2093– puter and Geosciences 27, 1071–1080.
2107. Scholle P.A. 1977. Chalk diagenesis and its relation to petroleum ex-
Hurst C. 1983. Petroleum geology of the Gorm field, Danish North ploration; oil from chalks, a modern miracle? AAPG Bulletin 61,
Sea. In: Petroleum Geology of the Southeastern North Sea and the 982–1009.


C 2005 European Association of Geoscientists & Engineers, Geophysical Prospecting, 53, 481–496
496 B. Røgen et al.

Walls J.D., Dvorkin J. and Smith B.A. 1998. Modeling seismic velocity Wilkens R.H., Simmons G., Wissler T.M. and Caruso L. 1986. The
in Ekofisk Chalk. 68th SEG meeting, New Orleans, USA, Expanded physical properties of a set of sandstones – Part III. The effects of
Abstracts, 1016–1019. fine-grained pore-filling material on compressional wave velocity.
Wang Z. 1997. Seismic properties of carbonate rocks. In: Carbonate International Journal of Rock Mechanical Mining Science and Ge-
Seismology (eds I. Palaz and K.J. Marfurt), pp. 29–52. Geophysical omechanics Abstracts 23, 313–325.
Developments Series 6. Society of Exploration Geophysicists. Wyllie M.R.J., Gregory A.R. and Gardner G.H.F. 1958. An experi-
Wang Z. 2000. The Gassmann equation revisited: comparing labo- mental investigation of factors affecting elastic wave velocities in
ratory data with Gassmann’s predictions. In: Seismic and Acous- porous media. Geophysics 23, 459–493.
tic Velocities in Reservoir Rocks: 3, Recent Developments (eds Z. Yan J. 2002. Reservoir parameter estimation from well log and core
Wang and A. Nur), pp. 8–23. Geophysics Reprint Series 19. Society data: a case study from the North Sea. Petroleum Geoscience 8,
of Exploration Geophysicists. 63–69.


C 2005 European Association of Geoscientists & Engineers, Geophysical Prospecting, 53, 481–496

You might also like