Sie sind auf Seite 1von 6

Journal of Geochemical Exploration 6970 (2000) 2328

www.elsevier.nl/locate/jgeoexp

Overpressure and fluid flow in dipping structures of the offshore


Gulf of Mexico (E.I. 330 field)
B. Bishop Stump 1, P.B. Flemings*
Department of Geosciences, Penn State University, University Park, PA 16802, USA

Abstract
Estimated pressures in overpressured Plio-Pleistocene muds of the Eugene Island 330 (E.I. 330) field (offshore Louisiana)
differ from pressure measurements in adjacent sands in a consistent manner. At structural highs, fluid pressure in sand exceeds
pressure in the adjacent mud; at structural lows, the relationship is reversed. Two physical models describe the origin of these
pressure differences: rapid loading and steady-flow. Observation and theory suggest that stratigraphic layering focuses fluid
flow toward structural highs. 2000 Elsevier Science B.V. All rights reserved.
Keywords: overpressure; porosity; hydrofracture; hydrodynamics

1. Introduction
A Permeable sand body encased in mud can have a
different pressure than the adjacent mud (Dickinson,
1953; England et al., 1987; Mann and Mackenzie,
1990). We explore this phenomenon through two
models. In the rapid-loading model, sediments behave
in an undrained manner and sharp pressure differences
occur at the top and bottom of dipping sand beds. In
the steady-flow model, vertical flow occurs and flow is
focused through the sand.
Porosity or some proxy of porosity (e.g. resistivity)
is used to estimate in situ fluid pressure in young
sedimentary basins (Athy, 1930; Rubey and Hubbert,
1959; Eaton, 1975). We examine fluid pressures in
mud predicted by the porosity-effective stress method
* Corresponding author. Tel.: 814-865-2309; fax: 814-8638724.
E-mail addresses: stumpbb@texaco.com (B. Bishop Stump),
flemings@geosc.psu.edu (P.B. Flemings).
1
Present address: Texaco Exploration and Production, New
Orleans, LA, 70130, USA.

with in situ fluid pressures measured in the reservoir


sands of the Eugene Island 330 (E.I. 330) field. We
show that in one highly overpressured horizon, the
pressure regime is most compatible with the steadyflow model.
2. Rapid loading model
A sand initially at hydrostatic pressure is instantaneously buried and rotated (Fig. 1). The pressure
gradients in the mud and sand are as follows:



b
rb rf g dz
1a
DPmud
b bf 1 f
DPsand

1
L




 ZL
b
rb rf g zx dz
b bf 1 f
0
1b

The overpressure at any point in the mud DPmud is a


function of the overlying load at that point (Eq. (1a)),
while the overpressure in the sand DPsand is
controlled by the integrated area over the sand (Eq.

0375-6742/00/$ - see front matter 2000 Elsevier Science B.V. All rights reserved.
PII: S0375-674 2(00)00116-3

24

B. Bishop Stump, P.B. Flemings / Journal of Geochemical Exploration 6970 (2000) 2328

Nomenclature
F
G
L
DP
Pf
Ph
Phc
Pw
Sv
X
Z
bf
b
Dtma
Dt
l
f
fo
rb
rf
sv

acoustic formation factor (dimensionless)


gravitational acceleration (L/T 2)
length of sand (L)
overpressure (M/LT 2)
fluid pressure (M/LT 2)
hydrostatic fluid pressure (M/LT 2)
hydrocarbon phase pressure (M/LT 2)
water phase pressure (M/LT 2)
vertical (lithostatic) stress (M/LT 2)
horizontal position (L)
depth (L)
fluid compressibility (LT 2/M)
bulk compressibility (LT 2/M)
matrix travel time (T/L)
wireline travel time (T/L)
ratio of Pf to Sv (dimensionless)
porosity (L 3/L 3)
reference porosity (L 3/L 3)
bulk density (M/L 3)
fluid density (M/L 3)
vertical effective stress (M/LT 2)

(1b)). A pressure discontinuity is predicted at the


sandmud interface (Fig. 1c).
2.1. Steady-flow model
We assume steady flow and that sand permeability
is much greater than mud permeability (Fig. 3). Phillips (1991) presented this problem. We extend his
model and describe its implications for geopressured
sands. Details are given in Stump (1998).
The presence of a dipping and relatively more
permeable sand lens results in a focusing of flow
toward the sand at the structural low and away from
the sand at the structural high (Fig. 2a). The vertical
overpressure gradient far from the sand is constant
and described by Eq. (1a). Close to the structural
low, pressure contours are depressed, and at the structural high, they are elevated (Fig. 2b).
Overpressure in the sand is constant (Fig. 2c). At the
structural high, overpressure in the sand is greater than
the overpressure in the overlying mud, which records
flow upward out of the crest. At the structural low, the

sand has a lower overpressure than the bounding mud


and flow goes from the shale into the sand (Fig. 2a).
3. Fluid pressures in sands and muds of the E.I. 330
field
We compare the predictions of the two models with
observations of the Lentic 1 sand in the Eugene Island
330 (E.I. 330) field.
3.1. Fluid pressure analysis in mud
We combine the definition of effective stress (Eq.
(2)) and an exponential porosity-effective stress relationship (Eq. (3)) to calculate fluid pressure from
porosity in muds (Eq. (4)) after Hart et al. (1995).

s v Sv P f

f fo e bs v

P f Sv



1
f
ln o
b
f

The coefficients b and f o are determined in the


hydrostatic zone where fluid pressure, and therefore
effective stress, are known. Porosity is calculated
from the wireline sonic (travel time, Dt) log through
an empirical relationship (Eq. (5)), developed by
Raymer et al. (1980) and enhanced by Raiga-Clemenceau et al. (1986).


Dtma 1=F
f1
5
Dt
Dtma is assumed to be 220 ms/m and the acoustic
formation factor (F) is 2.19.
3.2. Fluid pressure analysis in sands
We used only pre-production pressures in our
analysis. In hydrocarbon-bearing sands the pressure
of the hydrocarbon phase (Phc) exceeds the water
phase pressure (Pw), due to fluid density differences.
We calculate Pw for comparison with our predicted
fluid pressure in the mud.
3.3. Comparison of sand and mud pressures
Pressures in the Lentic 1 sand exceed the porositybased estimation of pressures in the bounding mud

B. Bishop Stump, P.B. Flemings / Journal of Geochemical Exploration 6970 (2000) 2328

25

Fig. 1. (a, b) The sand is rapidly buried and rotated. (c) Shale pressures (triangles) parallel the lithostatic gradient (Eq. (1a)); sand pressures
(circles) parallel the hydrostatic gradient but are shifted by DP (Eq. (1b)). Dashed lines represent hydrostatic pressure (10.5 MPa/km) and
lithostatic stress (21 MPa/km).

(Fig. 4, top). On a depthpressure plot, the vertical


pressure gradient appears approximately lithostatic.
At the structural high, the pressure in the sand is
4.2 MPa greater than that in the overlying mud. The
sand and mud pressures converge with depth and are
equal at 2316 m.
The pressure gradient in the sand is hydrostatic and
the overpressure is constant and equal to 16.3 MPa
(2363 psi) (Fig. 4, bottom). The overpressure in the
sand is everywhere greater than the overpressure in its
bounding mud. The overpressure contours in the mud

are sub-parallel to the sand and converge slightly


toward the top of structure.

3.4. Assessment of observed pressure differences


In the mud overlying the Lentic 1 sand, overpressure contours are sub-parallel to the dip of the sand
(Fig. 4, bottom). This observation is most similar to
the overpressure profile predicted by the steady-flow
model (Fig. 2b). The steep pressure gradient in the

26

B. Bishop Stump, P.B. Flemings / Journal of Geochemical Exploration 6970 (2000) 2328

Fig. 2. Steady-flow model. (a) Streamlines indicate primarily vertical flow, with enhanced flow into the sand lens at the structural low and out of
the sand at the structural high. (b) Overpressure contours are elevated near the structural high and are depressed near the structural low. (c)
Overpressure at the structural high (solid line), midpoint (dashed line), and low (dotted line). (d) Pressure profiles. Circles denote the
intersection of the profile with the sand lens.

Fig. 3. The E.I. 330 field is in the Gulf of Mexico, 272 km southwest of New Orleans, LA, USA at a water depth of 77 m (Holland et al., 1990).
Plus symbols indicate straight holes; solid lines represent deviated well paths; filled circles show the bottom hole locations of deviated wells.
Dashed line BB locates overpressure cross-section (Fig. 4b).

B. Bishop Stump, P.B. Flemings / Journal of Geochemical Exploration 6970 (2000) 2328

27

Fig. 4. (a) Fluid pressures in the sand (circles) are greater than pressures in the mud (triangles) at most depths; pressures converge near the
structural low. (b) Overpressure cross-section BB (located in Fig. 3) shows overpressure contours dipping parallel to structure.

mud overlying the sand implies fluid flow from the


highly overpressured sand into the shale.
Our observations rely upon the assumption that
porosity is a proxy for pressure. Bowers (1994)
proposed that late-stage decreases in effective stress
are not recorded in the porosity signature and thus
porosity-predicted pressure can underestimate in situ
fluid pressure. Thus, an alternative explanation for the
abrupt pressure jump between sand and shale at the

peak of the Lentic 1 sand (Fig. 4) is that pressuring by


lateral flow is occurring at the structural highs but is
not being recorded in the porosity.

4. Implications
The models presented illustrate the interactions
between layered strata with strong permeability

28

B. Bishop Stump, P.B. Flemings / Journal of Geochemical Exploration 6970 (2000) 2328

contrasts, structural relief, and overpressure. Overpressures are elevated at structural highs and
depressed at structural lows. Given enough structural
relief, pressures in the sand will converge on the minimum principal stress in the overlying mud and fluids
will migrate by fracture permeability. Observations of
pressure and porosity suggest that we can image
dynamic flow processes at the level of individual
beds.
References
Athy, L.F., 1930. Density, porosity, and compaction of sedimentary
rocks. Am. Assoc. Petrol. Geol. Bull. 14, 122.
Bowers, G.L., 1994. Pore pressure estimation from velocity data:
accounting for overpressure mechanisms besides undercompaction. Soc. Petrol. Engng SPE 27488, 515529.
Dickinson, G., 1953. Geological aspects of abnormal reservoir pressures in Gulf Coast Louisiana. Am. Assoc. Petrol. Geol. Bull.
37, 410432.
Eaton, B.A., 1975. The equation for geopressure prediction from
well logs. Soc. Petrol. Engng SPE 5544.
England, W.A., MacKenzie, A.S., Mann, D.M., Quigley, T.M.,
1987. The movement and entrapment of petroleum fluids in
the subsurface. J. Geol. Soc., London 144, 327347.
Hart, B.S., Flemings, P.B., Deshpande, A., 1995. Porosity and pres-

sure: Role of compaction disequilibrium in the development of


geopressures in a Gulf Coast Pleistocene basin. Geology 23,
4548.
Holland, D.S., W.E. Nunan, D.R. Lammlein, 1990. Eugene Island
Block 330 fieldU.S.A., offshore Louisiana. In: Beaumont,
E.A., Foster N.H. (Eds.), Structural Traps III, Tectonic Fold
and Fault Traps. American Association of Petroleum Geologists
Treatise of Petroleum Geology, Atlas of Oil and Gas Fields, pp.
103143.
Mann, D.M., Mackenzie, A.S., 1990. Prediction of pore fluid pressures in sedimentary basins. Mar. Petrol. Geol. 7, 5565.
Phillips, O.M., 1991. Flow and Reactions in Permeable Rocks,
Cambridge University Press, New York (285pp).
Raiga-Clemenceau, J., J.P. Martin, S. Nicoletis, 1986. The concept
of acoustic formation factor for more accurate porosity determination from sonic transit time data. SPWLA 27th Annual
Logging Symposium Transactions, Paper G.
Raymer, L.L., E.R. Hunt, J.S. Gardner, 1980. An improved sonic
transit time-to-porosity transform. SPWLA 21th Annual
Logging Symposium Transactions, Paper P.
Rubey, W.W., Hubbert, M.K., 1959. Overthrust belt in geosynclinal
area of western Wyoming in light of fluid pressure hypothesis. 2.
Role of fluid pressure in mechanics of overthrust faulting. Geol.
Soc. Am. Bull. 70, 167205.
Stump, B.B., 1998. Illuminating Basinal Fluid Flow in Eugene
Island 330 (Gulf of Mexico) Through in situ Observations,
Deformation Experiments, and Hydrodynamic Modeling, The
Pennsylvania State University, MS thesis.

Das könnte Ihnen auch gefallen