Sie sind auf Seite 1von 13

Marine and Petroleum Geology 19 (2002) 527±539

www.elsevier.com/locate/marpetgeo

A compaction trend for non-reservoir North Sea Chalk


A.J. Mallon*, R.E. Swarbrick
Science Laboratories, Department of Geological Sciences, University of Durham, South Road, Durham DH1 3LE, UK
Received 30 October 2001; received in revised form 13 March 2002; accepted 21 March 2002

Abstract
A compaction trend for the non-reservoir Chalk of the Central North Sea has been constructed using sonic interval transit time and density
log data from 59 wells. The dataset used has avoided high porosity reservoir units and hydrocarbon-bearing chalks in order to emphasise the
lower porosity, potentially pressure sealing units that make up the bulk of the Chalk Group. The compaction trend exhibits two stages: ®rstly
an approximately linear trend from 60 to 70% at the surface to 18% at 1300±1500 m, followed by a second linear trend from 18% at 1500 m
to 5% at 4500 m. Porosity loss within the Chalk is predominantly due to mechanisms such as stylolitisation and grain-to-grain pressure
solution. The decrease in the rate of porosity loss at about 1500 m depth may be due to the reduction of the effectiveness of these processes by
overpressure development. q 2002 Elsevier Science Ltd. All rights reserved.
Keywords: Chalk; Porosity; Compaction

1. Introduction from core and wireline data in wells across the North
Central Graben. The data allow construction of representa-
The Upper Cretaceous±Lower Palaeocene Chalk Group tive compaction curve behaviour for the bulk of the North
is a signi®cant stratigraphic unit in the Central North Sea. It Sea Chalk. Data from hydrocarbon-bearing provinces such
is an important hydrocarbon reservoir in the UK, Norwegian as the Greater Eko®sk area have been excluded from the
and Danish sectors of the North Sea Central Graben. dataset.
Previous research has largely concentrated on reservoir
quality chalks (Childs & Reed, 1975; D'Heur, 1984;
2. Previously published porosity trends
Jensenius, 1987; Scholle, 1977). However, reservoir quality
chalk, typically associated with allochtonous units accounts
Published porosity trends and pro®les for the Chalk have
for only a small part of the succession (Fig. 1). The Chalk
in many cases used data from restricted geographical areas
Group acts as a seal to hydrocarbons in a number of Central
(Scholle, 1977; Scholle & Halley, 1985) or single hydrocar-
North Sea ®elds. Examples include the Judy ®eld (Block 30/
bon bearing chalk reservoirs (D'Heur, 1984; Norbury, 1987;
7a) where the Jurassic Fulmar sandstone reservoirs are over-
Scholle, Albrechtssen, & Tirsgaard, 1998). Some generic
lain by chalk (Swarbrick et al., 2000) and the Fulmar Field.
porosity±depth trends have also been generated to describe
The Chalk also can act as an effective pressure barrier to
the `normal' behaviour of the Chalk. These trends are based
formation waters in underlying high-pressure reservoirs
on datasets considered to have the characteristics of chalk
(examples include wells 22/12a-1, 22/19-1 and 22/23a-2).
with normal compaction behaviour (Scholle, 1977; Scholle
These wells contain highly overpressured Triassic reser-
& Halley, 1985) or on the theoretical behaviour of chalk
voirs (15±30 MPa overpressure) separated from normally
compaction (Sorensen, Jones, Hardman, Leutz, & Schwartz,
pressured Palaeocene sandstone succession by the Chalk.
1986). These trends show a wide variation of porosity with
Units that could have contributed to pressure sealing, such
depth (Fig. 2). Some of the trends can be ascribed to chalks
as Jurassic or Lower Cretaceous shales, are thin or absent
that have been in¯uenced by meteoric ¯uids which show
(0±4 m).
some of the most rapid porosity decline (e.g. `Berkshire' in
This paper presents results from widespread porosity data
Fig. 2). Chalk successions that have been inverted show
anomalously low porosity values for their shallow depth
* Corresponding author. Tel.: 144-191-374-4785; fax: 144-191-374-
(e.g. Berkshire, `Norfolk' or `Celtic Sea' in Fig. 2).
2522. Hydrocarbon reservoir chalks typically show porosity
E-mail address: a.j.mallon@durham.ac.uk (A.J. Mallon). preservation at depth due to the retardation of cementation
0264-8172/02/$ - see front matter q 2002 Elsevier Science Ltd. All rights reserved.
PII: S 0264-817 2(02)00027-2
528 A.J. Mallon, R.E. Swarbrick / Marine and Petroleum Geology 19 (2002) 527±539

Fig. 1. Cross-section of the Chalk Group through part of the North Sea. Allochtonous chalks (solid black), typically associated with hydrocarbon reservoirs
represent only a small proportion of the succession. Non-reservoir chalks make up the bulk of the Group and contain hydrocarbon and pressure sealing units
(redrawn from Gatliff et al. (1994)).

Fig. 2. Compilation of data showing the wide variety of chalk porosity±depth trends. Some of the variation is due to the in¯uence of hydrocarbons (e.g. Hod,
Tomimeliten, Valhall), meteoric ¯uids and/or uplift (e.g. Norfolk, Berkshire). Based on Scholle (1977) with added data from Childs and Reed (1975), D'Heur
(1984), Jensenius (1987), Norbury (1987), Scholle and Halley (1985), Scholle et al. (1998), and Sorensen et al. (1986).
A.J. Mallon, R.E. Swarbrick / Marine and Petroleum Geology 19 (2002) 527±539 529

on hydrocarbon entry. Even when these anomalous is almost completely composed of calcite the matrix density
examples are removed, considerable uncertainty exists on of the Chalk (i.e. zero porosity) is taken as 2.71 g/cm 3. It
the overall trend of the porosity decline, especially at low was also assumed that the pore ¯uid has a density of 1.01 g/
(0±10%) porosities. cm 3 (i.e. a low salinity brine). Drilling mud ®ltrate invasion
Japsen (1998) displayed three normal compaction trends into pore space is regarded as minimal due to the low perme-
for the Chalk of the North Sea in addition to his own. Those ability of non-reservoir chalk. With these assumptions
of Bulat and Stoker (1987) and Hillis (1995) are interval porosity has been calculated from density as follows:
transit time-depth, Sclater and Christie (1980) porosity±
depth and Japsen (1998) velocity±depth. To allow direct f ˆ ……rma 2 rlog †=…rma 2 rfl ††
comparisons between the four trends Japsen (1998)
converted the four trends to velocity±depth and porosity± where f is the fractional porosity; r ma, the matrix density
depth pro®les. The conversions between porosity and (2.71 g/cm 3); r log, the density log reading; and r ¯ is the ¯uid
velocity are based on the following equations: density (1.01 g/cm 3).
To calculate porosity from interval transit time, a density
f ˆ 1:15 2 2:83 £ 1024 V; 1600 , V , 2700 m=s; to sonic transform was developed. The transform developed
here as a speci®c relationship for non-reservoir chalk is a
f ˆ 1 2 …V=6403†0:57 ; V , 2700 m=s more appropriate way to derive porosity from interval transit
The porosity±depth trends are shown in Fig. 3. The data time than conventional relationships (Raiga-Clemenceau,
show high variability, especially below 2000 m. Martin, & Nicoletis, 1988; Wyllie, Gregory, & Gardiner,
1956) that were developed for sandstones and Eko®sk
oil®eld Chalk, respectively. The two wells located on the
3. Data sources and porosity determination Norwegian Platform area immediately east of the Central
Graben (Fig. 4; Table 1) were used since both sonic and
Data was taken from 59 wells covering a wide geo- density logs were available through the chalk interval at
graphical area to construct the compaction curve for the relatively shallow depths. The porosity was calculated
Chalk of the Central North Sea (Fig. 4; Table 1). In all from the density log at 1 m intervals in these wells.
cases the data was screened so that wells in which hydro- It has been assumed that the chalks in these wells are not
carbon bearing chalks occurred were excluded. It is overpressured and hence the sonic data are not subject to
assumed that none of the wells has been uplifted in the anomalously low effective stress. The density-derived
area as it has been undergoing continuous and rapid subsi- porosity is plotted against interval transit time and a best-
dence and sedimentation since the deposition of the Chalk ®t linear relationship obtained (Fig. 5). The line is anchored
Group (Swarbrick et al., 2000). Sonic log data was used for to a limestone matrix interval transit time value of 48 ms/ft
all wells and the density log from three wells. As the Chalk at zero porosity (Schlumberger, 1972). This relationship

Fig. 3. Normal chalk compaction curves converted to porosity±depth trends by Japsen (1998). Note the variability below 2000 m.
530 A.J. Mallon, R.E. Swarbrick / Marine and Petroleum Geology 19 (2002) 527±539

Fig. 4. Map of the Central North Sea showing the location of the wells used in this study.

was used to convert the sonic interval transit time data to for the non-reservoir section of the Chalk. The background
porosity for wells where only sonic data are available. sonic interval transit time was estimated visually from each
Eight wells from the Mid North Sea High were used to well log (Fig. 6). If the majority of the formation was of a
give additional coverage to the compaction curve between constant sonic interval transit time, except for a few
350 and 1900 m (Fig. 4; Table 1). Only sonic log data were horizons of slower velocity (higher sonic values) then the
available for these wells, so the density-sonic transform modal value for the formation was used (Fig. 6). This repre-
from wells Nor 8/3-2 and 9/2-2 (Fig. 5) was used to calcu- sents the typical interval transit time of the low porosity
late the porosity from the sonic data. Values for sonic were intervals. Where high and low porosity intervals are in
taken at 10-foot interval averages. The wells were subdi- even proportions, the average for the entire section was
vided into the Eko®sk, Tor and Hod Formations. Most of taken. A single value of interval transit time was taken for
these wells have no Danian (Eko®sk) chalk as this formation each of the three Chalk Formations (Tor, Eko®sk and Hod)
has been largely removed by Tertiary erosion in these areas within each well and converted to porosity using the
(Gatliff et al., 1994). Log biostratigraphy in many cases was density-interval transit time transform. In each case porosity
not detailed enough to separate the Tor and the Hod, and in is plotted against midpoint depth for the formation. The
addition Gatliff et al. (1994) state that the two formations calculation of background interval transit time was also
cannot be distinguished lithologically in the southern repeated on the other sample subsets (Norwegian Platform,
Central North Sea (especially in Quadrants 37 and 38). Mid North Sea High and Deep Sea Drilling Project (DSDP))
Consequently the Tor±Hod boundary in these cases was so that the resulting values could be compared directly.
extrapolated from available cross-sections and other data.
The bulk of the data for the compaction curve below 3.1. Non-North Sea data sources
2000 m has come from 49 wells in Quadrants 21, 22, 23,
29, 30 and 31 (Fig. 4; Table 1). In each case the `background As there are no data available for the Chalk buried
sonic interval transit time' of the Eko®sk, Tor and Hod between 0 and 400 m from the Central North Sea, published
Formations has been calculated. The background interval porosity information from the DSDP was used to ®ll the
transit time is used in order to calculate the porosity value shallow depth data gap. The nannofossil oozes sampled by
A.J. Mallon, R.E. Swarbrick / Marine and Petroleum Geology 19 (2002) 527±539 531

Table 1 Table 1 (continued)


North Sea wells used to construct the Chalk compaction trend (depths
below sea ¯oor) Well Eko®sk (m) Tor (m) Hod (m)

Well Eko®sk (m) Tor (m) Hod (m) 22/13a-1 2903.8±2970.6 2970.6±3215.3 3215.3±3443.3
22/15-1 2980.0±3050.0 3050.0±3326.0 3326.0±3604.0
Norwegian Shelf 22/21-4 2824.3±2944.3 2944.3±3186.4 3186.4±3661.2
8/3-2 1015±1050 1050±1274 1274±1503 22/23a-3 2734.3±2784.6 2784.6±3016.0 3016.0±3025.1
9/2-2 682±751 751±1040 1040±1304 22/24a-2 2952.2±3015.6 3015.6±3279.6 3279.6±3402.5
Mid North Sea High 22/28a-1 3098.9±3204.0 3204.0±3582.6 3582.6±4354.3
37-23-1 836±842 842±1040 1040±1354 22/30c-10 3295.0±3393.0 3393.0±3921.0 3921.0±4933.0
37/10-1 Absent 1192±1310 1310±1497 23/11-383z 2625.2±2693.5 2693.5±2789.2 2789.2±2987.6
38/1-1 Absent 1396±1450 1450±1566 23/16a-2 2929.7±3040.6 3040.6±3326.6 3326.6±3519.5
38/25-1 Absent 1466±1680 1680±1820 23/16d-6 2983.4±3048.3 3048.3±3314.1 3314.1±3561.6
38/16-1 Absent 1132±1290 1290±1561 23/22b-4 3123.0±3201.5 3201.5±3606.0 3606.0±3876.5
27/3-1 Absent 265±415 415±631 23/26b-8 3243.0±3342.0 3342.0±3773.0 3773.0±4168.0
27/10-1 Absent 458±657 657±858 23/26a-9 3305.2±3399.7 3399.7±3808.1 3808.1±4412.9
28/10-1 Absent 526±669 669±892 30/1c-3 3218.3±3315.0 3315.0±3736.4 3736.4±4113.9
30/7a-4A 2841.9±2912.3 2912.3±3095.5 3095.5±3182.6
Central Graben 30/11b-3 3190.3±3290.0 3290.0±3597.2 3597.2±3979.7
21/1-6 1983.4±2043.1 2043.1±2125.4 2125.4±2447.2 30/12b-4 3030.3±3159.8 3159.8±3489.0 3489.0±3784.1
21/1-7st 1975.4±2033.2 2033.2±2123.6 2123.6±2444.2 30/13-3 3072.1±3163.2 3163.2±3355.5 3355.5±3711.2
21/2-5 2394.8±2456.7 2456.7±2675.5 2675.5±3184.5 30/17b-6 3008.0±3080.3 3080.3±3410.4 3410.4±3547.5
21/4b-5 2497.8±2552.1 2552.1±2819.6 2819.6±3480.2 30/22b-1 2624.3±2690.4 2690.4±2759.0 2759.0±2847.1
21/6-1 2165.5±2194.5 2194.5±2315.0 2315.0±2574.0 31/26a-10 2290.2±2335.4 2335.4±2389.0 2389.0±2415.2
21/11-3 1792.2±1850.1 1850.1±2029.9 2029.9±2182.3 29/2a-4 2784.6±2912.6 2912.6±3141.2 3141.2±3485.7
21/13b-1 2299.1±2393.6 2393.6±2551.8 2551.8±2755.4 29/2b-5 3007.4±3058.9 3058.9±3568.9 3568.9±3932.5
21/15b-4a 2625.8±2759.0 2759.0±2932.7 2932.7±3475.9 29/2c-9 2734.6±2795.0 2795.0±2941.0 2941.0±3346.4
21/18-5 1920.5±1957.4 1957.4±1980.3 1980.3±2079.3 29/3A-2 2780.4±2849.2 2849.2±3199.5 3199.5±3574.4
21/19-2 2354.2±2431.7 2431.7±2556.0 2556.0±2806.3 29/4a-2 2942.2±3026.3 3026.3±3480.5 3480.5±3961.5
21/25-1 2143.9±2215.6 2215.6±2246.0 2246.0±base not seen 29/5a-5 2996.8±3092.5 3092.5±3542.7 3542.7±4072.7
21/28-1 1146.0±1159.7 1159.7±1305.7 1305.7±1456.3 29/5b- 3261.6±3377.5 3377.5±3883.7 3883.7±4805.8
22/1a-4 2816.3±2879.4 2879.4±3067.8 3067.8±3674.3 4 1 st
22/3a-1 2940.9±3001.3 3001.3±3266.8 3266.8±3978.8 29/12-2 2487.4±2524.9 2524.9±2565.2 2565.2±2705.1
22/4b-5 2871.8±2966.0 2966.0±3223.8 3223.8±3846.9 29/10-3sti 2957.1±3048.6 3048.6±3496.9 3496.9±4027.0
22/5b-12 2623.4±2638.3 2638.3±2781.9 2781.9±2978.5 29/27-1 1032.3±1043.0 1043.0±1216.7 1216.7±1399.6
22/10b-6 2610.3±2679.5 2679.5±2737.4 2737.4±2873.0
22/12a-1 2695.6±2819.0 2819.0±2996.0 2996.0±3058.6

oozes make a reasonable analogue for the early compaction


the DSDP are considered to be the closest possible analogue of the Chalk.
to the Chalk of the North Sea. In both the North Sea Chalk Twelve DSDP sites from a variety of locations and water
and the nannofossil oozes coccolithic material makes up the depths were selected (Table 3) to give a representative
majority of the sediment and the mineralogy is almost porosity±depth trend. The sites selected have been the
completely low magnesium calcite. Ninety-®ve percent of subject of previously published work on carbonate physical
the forams in the sediments are planktonic in origin rather properties, diagenesis and burial trends (see Table 3 for
than benthic (Jeans & Rawson, 1980). There are, however, references to speci®c DSDP sites). Porosity values were
some differences, principally concerning the composition of derived from cores (e.g. using GRAPE analysis) or calcu-
the sediment and the effect of the deeper depositional depth lated using log parameters. The porosity±depth information
of the nannofossil oozes (Table 2). The differences outlined, for each site obtained was checked against data for CaCO3
however, are minimised by selecting sites in which the sedi- content in the DSDP database and any interval containing
ment has a similar coccolithic component to Central North marls or clays was removed from the dataset. This ®ltering
Sea Chalk and not examining oozes from extreme water is important for two reasons: (1) to ensure that different
depths where calcite dissolution in the water column is sections being compared are as alike as possible and (2)
common. Molluscan debris and bryozoan fragments are the grain density in pure carbonate sequences can be consid-
common in Cretaceous Chalks at basin margins that were ered a constant so that porosity is inversely proportional to
not part of this study, elsewhere they are relatively uncom- density and seismic velocity (Schlanger & Douglas, 1974).
mon. The statement that nannofossil oozes may contain a At these sites pelagic sedimentation is the main process of
higher clay content refers only to comparisons with `white sedimentation. Other processes such as the formation of
Chalk' facies. Chalks in the North Sea do contain this facies turbidites, are restricted or absent.
but also contain Chalks with a range of clay contents, To provide a check on the porosity values obtained from
particularly in the Hod Formation. Therefore, with these the sonic and density logs direct porosity measurements
minor differences taken into consideration nannofossil were also made. Twenty-®ve core samples and six sets of
532 A.J. Mallon, R.E. Swarbrick / Marine and Petroleum Geology 19 (2002) 527±539

Fig. 5. Sonic-density transform constructed from the Chalk of wells Nor 8/3-2 and 9/2-2.

cuttings samples were measured using mercury poros- 30% per km). Below 1300±1500 m there is a change in the
imetry. The 25 core samples were selected from an original rate of porosity loss to approximately 4% per km resulting
dataset of 128 samples taken from Central Graben cores and in a porosity of about 5% at 4500 m. The relatively sharp
represent the wide variety of chalk lithotypes from the Tor, change in the rate of porosity loss in the Chalk, as opposed
Hod and Eko®sk Formations. The six cuttings samples were to a gradual curve normally associated with sediment
selected from Norwegian Shelf wells 8/3-2 and 9/2-2. compaction can be seen in areas where the Chalk is currently

4. The Chalk compaction curve

The combined porosity±depth information describes a


new compaction curve for the non-reservoir Chalk of the
Central North Sea (Fig. 7). Although there is some scatter,
an overall trend is clearly observed. Both the wireline-
derived data and direct measurements of North Sea Chalk
show good agreement.
Fig. 8 removes any bias in the data due to differing
sampling intervals by displaying a single background poros-
ity/midpoint depth for each of the formations in each well
(or in the case of the DSDP data a single value for each site).
A best-®t trendline was calculated for the DSDP, shallow
chalk (500±1500 m) and Central Graben subsets (Fig. 8).
The most noticeable feature is the differing porosity±depth
trend for the shallow and Central Graben chalks. The DSDP
data, however, appears to describe adequately the shallow
portion of the compaction trend as the trend derived for this
portion of the data is very similar to that of the shallow
(400±1300 m) North Sea Chalks.
Fig. 6. The principle of `background interval transit time'. In this schematic
The dataset suggests therefore that the Chalk has a two-
sonic wireline log of the Chalk mean values have been assigned to the
stage compaction history. The ®rst stage shows a steady ¯uctuating Eko®sk and Hod, while the rare high interval transit time
decrease from approximately 65% at the surface to about zones of the Tor have been omitted to give a value appropriate to the
18% at between 1300 and 1500 m (a rate of approximately bulk of the formation.
A.J. Mallon, R.E. Swarbrick / Marine and Petroleum Geology 19 (2002) 527±539 533

Table 2 reduction of Central North Sea Chalk porosity can be reached


A comparison between Cretaceous Chalk and recent oozes (data from Jeans by examining the DSDP nannofossil sediments as an analo-
and Rawson (1980))
gue. In particular the onset of diagenetic processes, such as
Cretaceous white chalk facies Recent deep-sea nannofossil cementation, dissolution and stylolitisation are of interest.
oozes Mechanical compaction begins immediately after
deposition of the calcareous ooze. Sea¯oor cementation
High coccolithic component Lower coccolithic
(75±90%) componentÐ20±70% (but this and hardground formation are not recognised as important
includes samples of within the deep-water basinal regions typically drilled by
foraminiferal ooze) DSDP which experience continuous sedimentation
Little or no etching of Extensive etching of laths in (Garrison, 1981). This sedimentation pattern is in contrast
coccolith laths deeper water
to the Cretaceous Shelf chalk of the North Sea and elsewhere
Molluscan debris Molluscan debris rare
(inoceramids) and bryozoan in which hiatuses are clearly recognised in pelagic
debris prominent in coarser sequences in the form of hardgrounds (Hancock, 1975;
fractions Matthews, Sellwood, & Farmer, 1996).
Possibly a lower clay content Possibly a higher clay content Up to 10% porosity can be lost by mechanical processes
in the ®rst 50 m burial of the ooze (Matter, Douglas, &
buried to a depth of approximately 1300±1500 m. Well Nor Perch-Nielsen, 1975). However, mechanical compaction is
8/3-2 shows the change in rate of porosity loss in both well not deemed to be important once chemical compaction
log data and directly measured samples (Fig. 9). begins. Estimates for changes from dominantly mechanical
If the dataset is broken out into the constituent formations of to chemical processes vary from 50 to 200 m (Fig. 10)
the Chalk (Fig. 8), below 2000 m the Tor Formation generally (Garrison, 1981; Matter, 1974; Matter et al., 1975; Wilkens,
has the lowest porosity values while the Eko®sk and Hod are 1987). Dissolution and reprecipitation become the dominant
typically more scattered. These observations are in agree- processes and several studies (summarised in Garrison
ment with the directly measured chalk samples (Fig. 7). (1981)) show that certain bioclasts, usually planktonic
foraminfera and the very small elements of coccoliths are
5. Discussion preferentially dissolved. The dissolved calcium carbonate
reprecipitates on other bioclasts such as larger coccoliths
5.1. The nature of chalk compaction and discoasters (the overgrowths appear to be highly site-
and species-selective) and chamber ®lls. Signi®cant
An understanding of the processes involved in the early precipitation of calcite on other fossils commences at
Table 3
DSDP sites used in this study

Leg and site Location Water depth (m) Maximum depth (m) References

Leg 17 site 167 Magellan rise, Central Paci®c 3176 560.86 Schlanger and Douglas
(1974) and Garrison (1981)
Leg 21 site 206 Southwest Paci®c 3196 384.05 Packham and van der
Lingen (1973) and Garrison
(1981)
Leg 21 site 207 Southwest Paci®c 1389 44.4 Packham and van der
Lingen (1973) and Garrison
(1981)
Leg 21 site 208 Southwest Paci®c 1545 261.3 Packham and van der
Lingen (1973) and Garrison
(1981)
Leg 23 site 220 Arabian Sea 4036 265.3 Matter (1974) and Garrison
(1981)
Leg 30 site 288 Ontong Java Plateau 3000 230.41 Van der Lingen and
Packham (1975) and
Garrison (1981)
Leg 30 site 289 Ontong Java Plateau 2206 1234.39 Van der Lingen and
Packham (1975) and
Garrison (1981)
Leg 32 site 305 Shatsky rise, Northwest Paci®c 2903 240 Matter et al. (1975) and
Garrison (1981)
Leg 90 site 588c Southwest Paci®c 1533 462.69 Morin (1985)
Leg 90 site 590b Southwest Paci®c 1299 493.26 Morin (1985)
Leg 90 site 592 Southwest Paci®c 1088 382.42 Morin (1985)
Leg 90 site 593 Southwest Paci®c 1068 562.61 Morin (1985)
534 A.J. Mallon, R.E. Swarbrick / Marine and Petroleum Geology 19 (2002) 527±539

Fig. 7. Results of all wireline derived and measured porosity values. The trend of porosity decline is clearly visible. Note also the good agreement between the
log and lab measured data.

about 200 m, after a grain-supporting framework has been occurs (Hamilton, 1959; Fig. 2) are at variance with data
established which triggers pressure solution on the solution generated by direct sediment measurements, where both
prone forams and coccoliths. These processes continue to mechanical and chemical compaction processes are active)
porosity values as low as 25%. (Audet, 1995).
The early onset of chemical processes within the oceanic The formation of pressure dissolution phenomena in the
carbonate sediment explains why the experimental compac- form of solution seams and stylolites is an important
tion of deep sea oozes in which only mechanical compaction chemical process within calcareous sediments. Stylolite

Fig. 8. Data converted to one `background interval transit time' datapoint per formation per well (DSDP one datapoint per site) to enable direct comparisons
between subsets of data. Note the close agreement in the best-®t line of the DSDP and shallow chalk datasets. Note also the difference in the trendline of the
more deeply buried chalks. The change in the rate of porosity decline appears to occur at approximately 1500 m.
A.J. Mallon, R.E. Swarbrick / Marine and Petroleum Geology 19 (2002) 527±539 535

Fig. 9. Porosity±depth trend of the Chalk in well Nor 8/3-2. Note the change of slope in this well at 1200 m.

development in pelagic limestones is thought to begin at formity within the sediment (especially the shallower occur-
about 40% porosity (Tada & Siever, 1989; Fig. 10). A rences). However, at and below 750 m they also appear
review of solution structures in DSDP holes is given by within apparently continuous sequences (Hill, 1987). In
Hill (1987) who records such features between 335 and addition to dissolution seams, well-developed stylolites
1270 m (Fig. 10). These solution structures are macrosco- have been recorded in chalk and limestone below 830 m
pically identical to wispy dissolution seams seen in the from the Ontong Java Plateau (Lind, 1993; Fig. 10).
English Chalk of the Isle of Wight. The solution structures In conclusion, the study of calcareous sediments drilled at
seen in the DSDP cores are often associated with a discon- DSDP sites shows that they are a highly reactive sediment

Fig. 10. Timing and nature of the processes responsible for the porosity reduction of carbonate oozes and chalk. Note the predominance of chemical processes.
536 A.J. Mallon, R.E. Swarbrick / Marine and Petroleum Geology 19 (2002) 527±539

with chemical processes dominating its compaction in all carbonates (Moore, 1989). However, calcite solubility
but the ®rst 50±200 m. As these sediments are the closest decreases with increasing temperature (solubility is impor-
analogues to the Chalk of the North Sea it can be expected tant as the relatively high solubility of calcium carbonate
that the Chalk underwent similar chemical compaction promotes grain-to-grain pressure solution and stylolitisation
during its early diagenesis. Indeed the Chalk may have (Bathurst, 1976)). Together these two factors favour `the
experienced a greater degree of porosity loss than the precipitation of calcite cements in the subsurface' (Moore,
DSDP sites due to the effects of sea¯oor cementation, 1989). A comparison of data from two otherwise similar
especially in the shallower shelf and platform areas (e.g. DSDP sites shows that carbonate oozes contain more
Mid North Sea High). cement in the area with a higher geothermal gradient
The overall trend developed from the Norwegian and Mid (Wetzel, 1989). Increasing temperature by progressive
North Sea High data links well with the DSDP data. The rate burial favours porosity occlusion by cementation. It is there-
of porosity decline appears to be broadly similar. The simi- fore unlikely that temperature can account for the decrease
larity in data trends with DSDP suggest that the processes in the rate of porosity loss at 1500 m in the Chalk.
controlling porosity decline in the shallow North Sea wells
are similar (i.e. dissolution and reprecipitation of calcite). 5.2.2. Lithostatic pressure
Stylolites have been recorded during the course of this study Calcite solubility increases with increasing pressure
from well cuttings from well 9/2-2 at depths of 996 m and (Moore, 1989; Morse & McKenzie, 1990). However, as
deeper and are a common feature within the Central Graben noted earlier calcite solubility decreases with increasing
chalk cores. In addition authigenic cements are abundant in temperature. Therefore, as a carbonate rock such as the
Central North Sea Chalk samples as recorded using catho- Chalk is buried the increasing temperature and pressure
doluminescence and scanning electron microscope. The have opposing in¯uences on calcite solubility. Morse and
percentage of this cement also increases with depth (Mallon McKenzie (1990) have shown (Fig. 11) that for typical
& Swarbrick, in preparation). burial conditions calcite solubility is little affected by
It is therefore apparent that dominantly chemical combined changes in pressure and temperature. Therefore,
processes such as dissolution and reprecipitation are respon- it is unlikely that the temperature and lithostatic pressure
sible for the rapid porosity loss in the ®rst section of the conditions at 1300±1500 m are responsible for the in¯ection
compaction trend. The more gradual porosity decline in the seen within the Chalk data.
Chalk from the deep North Sea wells suggests that chemical
compaction has been retarded. A number of processes may
have such an effect and should be considered. 5.2.3. Salinity
Pore ¯uid salinity has a major effect on calcite solubility.
5.2. Controls on porosity loss in Chalk

A number of processes have the potential to control


calcite dissolution and precipitation in the Chalk, these
include:

1. temperature;
2. lithostatic pressure (overburden stress);
3. salinity;
4. overpressure (high pore pressure).

Mineralogy is often a factor in controlling dissolution and


precipitation in carbonates (Bathurst, 1976; Moore, 1989).
However, as the carbonate fraction of Chalk is composed
almost completely of low magnesium calcite the in¯uence
of more unstable phases such as aragonite do not have to be
considered.
The extent to which these parameters change in a typical
basin setting needs to be considered in order to assess
whether they result in the observed change at the 1300±
1500 m interval.
Fig. 11. The in¯uence of dissolved NaCl on pK p (logarithm of stoichio-
metric solubility constant for calcite) under typical burial conditions. Note
5.2.1. Temperature the variation resulting from differing salinities but little variation with
Increasing temperature increases the rate of chemical increasing burial (increasing temperature and pressure). From Morse and
reactions such as dissolution and ionic diffusion within McKenzie (1990).
A.J. Mallon, R.E. Swarbrick / Marine and Petroleum Geology 19 (2002) 527±539 537

Fig. 12. Modelled overpressure through time for a mid Chalk unit from block 30/11 using the basin modelling software Temispack (top). Note a rapid increase
in overpressure beginning at 24 Ma. The corresponding burial curve for the same mid chalk unit (bottom) shows that at 24 Ma the Chalk is at 1500 m.

Increasing the ionic strength of a solution from that of icant variations in calcite solubility, greater than solubility
meteoric water (0 M NaCl) to a 0.725 M NaCl solution changes owing to increasing pressure and temperature
will increase calcite solubility by a factor of 30 (Fig. 11) associated with sediment burial. Within the Chalk of the
(Morse & McKenzie, 1990). If the ionic strength is further North Sea differences can be seen in the rate of porosity
increased up to 6 M NaCl there is relatively little change in decline between chalks now onshore which have experienced
solubility (Fig. 11) and are `approximately equal to the meteoric ¯uids and those offshore in which the ¯uids are domi-
combined effects of pressure and temperature for diagenetic nantly or exclusively marine (Fig. 2). However, examination
conditions typical for a sedimentary basin' (Morse & of log data (e.g. the SP log in well Nor 8/3-2) indicates that
McKenzie, 1990). there is not a major change in ¯uid composition at the depth
Morse and McKenzie (1990) conclude that even small at which the rate of porosity loss changes. Therefore,
variations in subsurface ¯uid composition can cause signif- salinity changes are not the cause of the rate of compaction.
538 A.J. Mallon, R.E. Swarbrick / Marine and Petroleum Geology 19 (2002) 527±539

5.2.4. Overpressure in¯ection on the compaction curve both occur in the region
Much of the chemical compaction occurring within the of 1500 m. This coincidence is consistent with a change in
Chalk is in the form of stylolite and dissolution seam forma- the rate of effective stress increase during burial, which
tion, as well as grain to grain pressure dissolution. Such occurs near the Oligocene±Miocene boundary, and
processes are dependent on the stress experienced at grain provides the most plausible explanation for the change in
boundaries (Carrio-Schaffhauser, Raynaud, LatieÁre, & rate of porosity loss in the Central North Sea Chalks.
Mazerolle, 1990; Tada & Siever, 1989). Overpressures
would have the effect of reducing the effective stress of
the system and thus would reduce the effectiveness of 6. Conclusions
pressure solution. Mechanical compaction would also be
retarded. ² A new compaction curve for non-reservoir chalk has
The Chalk is known to be overpressured in the Central been created using a combination of DSDP porosity
North Sea. Overpressures up to 32.0 MPa (Drill stem test data for nannofossil oozes and log-derived porosity
from well 30/1c-3 at 3827 m) are recorded. Overpressures data for shallow to deeply buried chalk from the North
are thought to be mainly due to rapid burial in the late Sea.
Tertiary to Quaternary (Japsen, 1998; Swarbrick et al., ² The rate of porosity decline is in two stages: ®rstly an
2000). Overpressure in Chalk reservoirs such as Eko®sk approximately linear trend from 60 to 70% at the surface
has also been held as a mechanism for porosity preservation to 18% at 1300±1500 m, followed by a second linear
(Brasher & Vagle, 1996; Scholle, 1977). However, in the trend from 18% at 1500 m to 5% at 4500 m.
non-reservoir sections studied here direct pressure measure- ² Porosity loss is principally chemically controlled in all
ments are rare due to low permeability. but the ®rst 50±200 m of burial and is due to dissolution±
In assessing the likelihood of overpressure as a control on reprecipitation processes.
the rate of porosity decline in the Chalk a means of deter- ² The change from a rapid to slower rate of porosity loss at
mining the nature and timing of overpressure generation is approximately 1300±1500 m is most likely due to the
required. This was accomplished by running a number of generation of overpressure within the Chalk, reducing
basin models for part of the Central North Sea (UK Quad the effectiveness of stress driven chemical compaction.
30; Fig. 4) using the basin modelling package, Temispack.
The models were originally developed as part of a case
study along a 74 km line of the Central North Sea used to
model results from the GeoPOP project (Seldon et al., Acknowledgements
unpublished data). In order to accurately model the chemi-
cal compaction processes such as pressure solution that This study was carried as part of the GeoPOP 2 project
occur in the Chalk the Temispack chemical calculator and the authors wish to thank the sponsors (Amerada Hess,
`tpack2.9.3BIG' was used. The chemical calculator is BP, BG plc, Conoco, Enterprise Oil, ExxonMobil, JNOC,
based on the model of Schneider, Potdevin, Wolf, & Faille, Norsk Hydro, Phillips Petroleum, Statoil, Texaco and Total-
(1996). The chemical compaction calculator incorporates a FinaElf) for permission to publish this paper. In particular
viscoplastic term into the compaction equation and `simu- Sven Hansen at Norsk Hydro and Lars Wensaas at Statoil
lates viscous compaction phenomena such as pressure are thanked for providing data. Ben Seldon at Durham is
solution' (Schneider et al., 1996). thanked for running the Temispack model. The paper also
The porosity±depth relationship from this study was used bene®ted from the valuable comments from Richard Hillis
to characterise the non-reservoir section of the Chalk in the and an anonymous reviewer.
model. The reservoir section was described using the trends
of Brasher and Vagle (1996). Porosity±permeability References
relationships were derived from measurements on non-
reservoir chalk (Mallon & Swarbrick, in preparation) and Audet, D. M. (1995). Modelling of porosity evolution and mechanical
reservoir chalk data. compaction of calcareous sediments. Sedimentology, 42, 355±373.
The Chalk succession in UK Block 30/11 was examined. Bathurst, R. G. C. (1976). Carbonate sediments and their diagenesis. Devel-
In the model a mid Chalk unit (Age: 71 Ma in the lower Tor opments in Sedimentology, 12.
Brasher, J. E., & Vagle, K. R. (1996). In¯uence of lithofacies and diagen-
Formation) was selected, which indicated that signi®cant esis on Norwegian North Sea Chalk reservoirs. American Association of
overpressure began to be generated at a depth of 1500 m Petroleum Geologists, Bulletin, 80, 746±769.
(Fig. 12(a)). In the model this depth was reached at 24 Ma Bulat, J., & Stoker, S. J. (1987). Uplift determination from interval velocity
(Fig. 12(b)) and is coincident with an increase in the rate of studies, UK, southern North Sea. In J. Brooks & K. W. Glennie (Eds.),
sedimentation at the top Oligocene±base Miocene. Other Petroleum geology of north west Europe (pp. 239±305). London:
Graham & Trotman.
models run for UK blocks 30/7, 30/8, 30/12 and 30/17 Carrio-Schaffhauser, E., Raynaud, S., LatieÁre, H. J., & Mazerolle, F.
areas show the onset of overpressure at similar depths. (1990). Propagation and localization of stylolites in limestone.
The onset of signi®cant overpressure in the model and In R. J. Knipe & E. H. Rutter (Eds.), Deformation mechanisms,
A.J. Mallon, R.E. Swarbrick / Marine and Petroleum Geology 19 (2002) 527±539 539

rheology and tectonics. Geological Society Special Publication, 54, Morse, J. W., & McKenzie, F. T. (1990). Geochemistry of sedimentary
193±199. carbonates. Developments in Sedimentology, 49, 707.
Childs, F. B., & Reed, P. E. C. (1975). Geology of the Dan Field and the Norbury, I. (1987). In A. M. Spencer, et al. (Eds.), Geology of the Norwe-
Danish North Sea. In A. W. Woodland (Eds.), Petroleum and the conti- gian oil and gas ®elds (pp. 107±116). London: Graham & Trotman.
nental shelf of North-West Europe, Vol. 1. (pp. 429±438). Great Britain: Packham, G. H., & van der Lingen, G. J. (1973). Progressive carbonate
The Institute of Petroleum. diagenesis at Deep Sea Drilling Sites 206, 207, 208, and 210 in the
D'Heur, M. (1984). Porosity and hydrocarbon distribution in the North Sea southwest Paci®c and its relationship to sediment physical properties
Chalk reservoirs. Marine and Petroleum Geology, 1, 211±238. and seismic re¯ectors. Initial Reports DSDP, 21, 495±521.
Garrison, R. E. (1981). Diagenesis of oceanic carbonate sediments: a Raiga-Clemenceau, J., Martin, J. P., & Nicoletis, S. (1988). The concept of
review of the DSDP perspective. In J. E. Warme, R. G. Douglas & acoustic formation factor for more accurate porosity determination
E. L. Winterer (Eds.), The Deep Sea Drilling Project; a decade of from sonic transit time data. The Log Analyst, 29, 54±60.
progress. SEPM Special Publication, 32, 181±207. Schlanger, S. O., & Douglas, R. G. (1974). The pelagic ooze±chalk±lime-
Gatliff, R. W., Richards, P. C., Smith, K., Graham, C. C., McCormac, M., stone transition and its implications for marine stratigraphy. In K. J. Hsu
Smith, N. J. P., Long, D., Cameron, T. D. J., Evans, D., Stevenson, A. G., & H. C. Jenkyns (Eds.), Pelagic sediments: on land and under the sea.
Bulat, J., & Richie, J. D. (1994). United Kingdom offshore regional International Association of Sedimentologists Special Publication, 1,
report: The geology of the Central North Sea, London: HMSO for 117±148.
the British Geological Survey, 118 pp. Schlumberger. (1972). Log interpretation. Volume 1 ± principles, USA:
Hamilton, E. L. (1959). Thickness and consolidation of deep-sea sediments. Schlumberger Ltd.
Bulletin of the Geological Society of America, 70, 1399±1424. Schneider, F., Potdevin, J. L., Wolf, S., & Faille, I. (1996). Mechanical and
Hancock, J. M. (1975). The petrology of the Chalk. Proceedings of the chemical compaction model for sedimentary basin simulators. Tecto-
Geological Association, 86, 499±535. nophysics, 263, 307±317.
Hill, P. R. (1987). Chalk solution structures in cores from Deep Sea Drilling Scholle, P. A. (1977). Chalk diagenesis and its relation to petroleum
Project Leg 94. In W. F. Ruddiman & R. B. Kidd (Eds.), Initial reports exploration: oil from chalks, a modern miracle? American Association
DSDP, leg 94, Norfolk Virginia to St John's Newfoundland (pp. 1129± of Petroleum Geologists Bulletin, 61, 982±1009.
1143). Washington: US Printing Of®ce. Scholle, P. A., & Halley, R. B. (1985). Burial diagenesis: out of sight, out of
Hillis, R. R. (1995). Quanti®cation of Tertiary exhumation in the United mind. In N. Schneidermann & P. M. Harris (Eds.), Carbonate cements.
Kingdom southern, North Sea using sonic velocity data. American Society of Economic Palaeontologists and Mineralogists Special Publi-
Association of Petroleum Geologists Bulletin, 79, 130±152. cation, 36, 309±335.
Japsen, P. (1998). Regional velocity±depth anomalies, North Sea Chalk: a Scholle, P. A., Albrechtssen, T., & Tirsgaard, H. (1998). Formation and
record of overpressure and Neogene uplift and erosion. American Asso- diagenesis of bedding cycles in uppermost Cretaceous chalks of the Dan
ciation of Petroleum Geologists Bulletin, 82, 2031±2074. Field, Danish North Sea. Sedimentology, 45, 223±243.
Jeans, C. V., & Rawson, P. F. (1980). Andros Island, chalk and oceanic Sclater, J. G., & Christie, P. A. F. (1980). Continental stretching; an expla-
oozes. Unpublished work of Maurice Black. Yorkshire Geological nation of the post-Mid-Cretaceous subsidence of the Central North Sea
Society Occasional Publication No. 5, 100 pp. basin. Journal of Geophysical Research, 85, 3711±3739.
Jensenius, J. (1987). High-temperature diagenesis in shallow Chalk reser- Sorensen, S., Jones, M., Hardman, R. F. P., Leutz, W. K., & Schwartz, P. H.
voir, Skjold oil ®eld, Danish North Sea: evidence from ¯uid inclusions (1986). Reservoir characteristics of high- and low-productivity chalks
and oxygen isotopes. American Association of Petroleum Geologists from the Central North Sea, Habitat of hydrocarbons on the Norwegian
Bulletin, 71, 1378±1386. Continental Shelf. Norwegian Petroleum Society, pp. 91±110.
Lind, I. L. (1993). Stylolites in chalk from Leg 130, Ontong Java Plateau. Swarbrick, R. E., Osborne, M. J., Grunberger, D., Yardley, G. S., Macleod,
Proceedings of Scienti®c Results ODP, Leg 130, Ontong Java Plateau, G., Aplin, A. C., Larter, S. R., Knight, I., & Auld, H. A. (2000). Inte-
445±451. grated study of the Judy Field (Block 30/7a)Ðan overpressured Central
Mallon, A. J., & Swarbrick, R. E. (in preparation). Diagenesis and perme- North Sea oil/gas ®eld. Marine and Petroleum Geology, 17, 993±1010.
ability characteristics of non-reservoir Chalks from the Central North Tada, R., & Siever, R. (1989). Pressure solution during diagenesis. Annual
Sea. Review of Earth and Planetary Sciences, 17, 89±118.
Matter, A. (1974). Burial diagenesis of pelitic and carbonate deepsea sedi- Van der Lingen, G. J., & Packham, G. H. (1975). Relationships between
ments from the Arabian Sea. Initial Reports DSDP, 23, 421±470. diagenesis and physical properties of biogeneic sediments of the
Matter, A., Douglas, R. G., & Perch-Nielsen, K. (1975). Fossil preservation, Ontong-Java Plateau (Sites 288 and 289, Deep Sea Drilling Project).
geochemistry and diagenesis of pelagic carbonates from Shatsky rise, Initial Reports DSDP, 30, 569±615.
northwest Paci®c. Initial Reports DSDP, 32, 891±922. Wetzel, A. (1989). In¯uence of heat ¯ow on ooze/chalk cementation: Quan-
Matthews, A., Sellwood, B., & Farmer, C. (1996). Submarine cementation ti®cation from consolidation parameters in DSDP sites 504 and 505
in Chalks of the Valhall and Hod Fields: Origin of dense zones as sediments. Journal of Sedimentary Petrology, 59, 539±547.
submarine hardgrounds. Fifth North Sea Chalk Symposium, 19. Wilkens, R. (1987). The effects of diagenesis on the microstructure of
Moore, C. H. (1989). Carbonate diagenesis and porosity. Developments in Eocene sediments bordering the Baltimore Canyon Trough. Initial
Sedimentology, 46, 338. Reports DSDP, 95, 527±547.
Morin, R. H. (1985). Physical properties of calcareous sediments from the Wyllie, M. R. J., Gregory, A. R., & Gardiner, L. W. (1956). Elastic wave
southwest Paci®c. Initial Reports DSDP, 90, 1239±1246. velocities in heterogeneous and porous media. Geophysics, 21, 41±70.

Das könnte Ihnen auch gefallen