You are on page 1of 10

Effective stress, porosity, velocity and abnormal pore pressure

prediction accounting for compaction disequilibrium and unloading


Jincai Zhang
*
Shell Upstream Americas, USA
a r t i c l e i n f o
Article history:
Received 30 July 2012
Accepted 15 April 2013
Available online 25 April 2013
Keywords:
Pore pressure prediction
Effective stress
Porosityedepth relationship
Overpressure
Compaction disequilibrium
Compressional velocity and transit time
Unloading
a b s t r a c t
Abnormal pore pressures, mostly overpressures, exist in many sedimentary formations. The over-
pressures deteriorate drilling safety, causing borehole inux, kicks, and even blowout, if the pressures are
not accurately predicted prior to drilling. Highly anomalous overpressures may also induce instability
and reactivation of faults, causing fault weakness. Formation overpressures are primarily generated by
compaction disequilibrium, which is often recognized by higher than expected porosities at a given
depth and the porosities deviated from the normal porosity trend. Based on this mechanism, the paper
proposes a new generalized theoretical model for porosityedepth relationship for both normally com-
pacted and abnormally compacted formations, i.e., f f
0
e
cZse =sn
. This model leads to a new method
for calculating effective stress and pore pressure in subsurface formations using porosity and
compressional velocity. A new relationship of the transit time and depth has also been derived which
extends the existing model (Chapmans model). It demonstrates that the sonic/seismic travel time and
effective stress have an exponential relationship (i.e., Dt Dt
m
Dt
ml
Dt
m
e
cZse=sn
).
Stress unloading caused by formation uplift has a different path compared to compaction/loading
curve of the stress and velocity, thus a different compaction constant. This causes a smaller effective
stress and lower porosity than those in the loading case; i.e., unloading causes pore pressure increase.
Effective stress and pore pressure calculations accounting for unloading are also proposed. Field data in
several petroleum basins are analyzed and verify the theoretical relationship between effective stress
and sonic transit time. Lab experimental data in sonic velocity and effective stress in both loading and
unloading cases also verify the proposed effective stress and velocity relationship. Case study in an oil
eld is presented to examine the proposed model for pore pressure analysis in subsalt formations.
2013 Elsevier Ltd. All rights reserved.
1. Introduction
1.1. Under-compaction and abnormal pore pressure
Pore pressures in subsurface formations vary from hydrostatic
pressures (normal pore pressures) to severe overpressures (more
than double of the hydrostatic pressures). Overpressures exist in
many geologic basins in the world. If this abnormal overpressure is
not accurately predicted before drilling and while drilling, it can
greatly increase drilling risks and incidents. For examples, in
deepwater of the Gulf of Mexico, incidents associated with pore
pressure and wellbore instability accounted for 5.6% of drilling time
in non-subsalt wells, and 12.6% of drilling time in the subsalt wells
(York et al., 2009). The abnormally high pore pressures also caused
serious drilling incidents, such as the uid kicks and well blowouts
(Skalle and Podio, 1998; Holand and Skalle, 2001). Therefore, pore
pressure prediction is critically important for drilling planning and
operations in oil and gas industry. Abnormally high pressures also
induced geologic hazards and disasters, such as weakness in faults
(e.g., Bird, 1995; Tobin and Saffer, 2009) and mud volcanoes (Davies
et al., 2007; Tingay et al., 2009).
Overpressures can be generated by many mechanisms, such as
compaction disequilibrium (under-compaction), hydrocarbon
generation and gas cracking, aquathermal expansion, tectonic
compression (lateral stresses), mineral transformations (e.g.,
smectiteeillite transition), and hydrocarbon buoyancy (Swarbrick
and Osborne, 1998). The major reason of abnormal pore pressure
is caused by abnormal formation compaction (compaction
disequilibrium). When sediments compact normally, formation
porosity is reduced at the same time as pore uid is expelled.
During burial, increasing overburden stress is the prime cause
of uid expulsion. If the sedimentation rate is slow, normal
* Now with Hess.
E-mail address: zhangjincai@yahoo.com.
Contents lists available at SciVerse ScienceDirect
Marine and Petroleum Geology
j ournal homepage: www. el sevi er. com/ l ocat e/ marpet geo
0264-8172/$ e see front matter 2013 Elsevier Ltd. All rights reserved.
http://dx.doi.org/10.1016/j.marpetgeo.2013.04.007
Marine and Petroleum Geology 45 (2013) 2e11
compaction occurs, i.e. equilibriumbetween increasing overburden
and ability to expel uids is maintained (Mouchet and Mitchell,
1989). This normal compaction generates hydrostatic pore pres-
sure in the formation. When the sediments subside rapidly, or the
formation has extremely low permeability, uids in the sediments
can only be partially expelled, and the remained uid must support
all or part of the weight of overburden sediments. This causes
abnormally high pore pressure. In this case the porosity decreases
less rapidly than it should be with depth, and formations are under-
compacted or in compaction disequilibrium. The compaction
disequilibrium is often recognized by higher than expected po-
rosities at a given depth and the porosities deviated from the
normal porosity trend. Therefore, pore pressure can be calculated
from the formation porosities.
1.2. Hydrostatic pore pressure
Normal pore pressure is the hydrostatic pressure caused by the
column of pore uid from the surface to the interested depth. For
formations with normal uid pressure, the pore pressure follows
the hydrostatic pressure gradient. The magnitude of the pressure is
proportional to the depth below the surface and to the density of
the uid in the pores. That is, the pressure is the same at the same
depth within the uid with a uniform density if the uid is static.
Thus, the hydrostatic pressure can be calculated using the following
equation:
p
n
r
f
gh (1)
where p
n
is the hydrostatic pressure; g is the acceleration due to
gravity; r
f
is the uid density; and h is the vertical height of the uid
column, as shown in Figure 1.
The equation (Eq. (1)) indicates that the hydrostatic pore pres-
sure depends highly on uid/water density in the formation. While
the density of water is a function of water salinity, temperature, and
content of dissolved gases (Chillingar et al., 2002); therefore, there
is a general variation in the hydrostatic pressure gradient (r
f
g) at
different locations due to different water densities. For instance, the
average hydrostatic pressure gradient is usually taken as 0.465 psi/
ft (1.074 kg/cm
3
) in the Gulf of Mexico, and this corresponds to
water with a salinity of 80,000 parts per million (ppm) of sodium
chloride at 77

F (25

C) (Dickinson, 1953).
1.3. Relationship of effective stress, overburden stress, and pore
pressure
Terzaghis or Biots effective stress law (Terzaghi et al., 1996;
Biot, 1941) is the fundamental theory for pore pressure prediction.
The effective stress and pore pressure in vertical direction in one-
dimensional condition can be expressed as following (Biot, 1941):
s
e
s
V
ap (2)
where p is the pore pressure, s
V
is the overburden or vertical stress,
s
e
is the vertical effective stress; a is the Biots effective stress
coefcient.
In normal pressure case, from the above equation the normal
effective stress and normal pore pressure have the following
relationship:
s
n
s
V
ap
n
(2a)
where s
n
is the normal vertical effective stress; p
n
is the normal or
hydrostatic pressure.
The effective stress can be obtained by correlating to petro-
physical and geophysical data of formations (e.g., resistivity logs,
seismic and sonic travel time/velocity). When effective stress and
overburden stress are known, the pore pressure can be calculated
from Eq. (2).
It is commonly assumed that the in-situ stress includes three
mutually orthogonal principal stresses; i.e., vertical, maximum
horizontal and minimum horizontal stresses (s
V
, s
H
, s
h
). However,
it is further assumed that the formation compaction is mainly
caused by the vertical/overburden stress and formation under-
compaction is primarily related to the vertical stress (e.g.,
Chapman, 1983; Osborne and Swarbrick, 1997). Therefore, the pore
pressure caused by compaction and under-compaction can be
calculated from Eq. (2) when one knows vertical and effective
stresses.
Vertical stress is generated by the weight of the overlying for-
mations; hence, it can be obtained by integrating bulk density logs.
Therefore, vertical stress can be calculated by the following
equation:
s
V
r
w
gz
w
g
_
z
zw
r
b
zdz (3)
where r
b
(z) is the formation bulk density as a function of depth; r
w
is the density of sea water for offshore drilling; z is the depth from
the sea level; z
w
is the water depth, for onshore drilling z
w
0.
The bulk density can be obtained fromwell logging. However, in
most cases the shallow density log data are not available. Empirical
equations can be used to estimate the shallow density. Analyzing
the observed depthedensity curve in density measurements of
shales in northern Oklahoma, Athy (1930) proposed the following
equation to interpolate shallow formation bulk density:
r
z
r
0
A
m
_
1 e
bZ
_
(4)
where r
z
is the density at the depth of Z, in g/cm
3
; r
0
is the for-
mation density of the surface; A is the maximum density increase
possible (A
m
r
m
r
0
and A
m
1.3 in Athy, 1930); r
m
is the matrix
density or the grain density of the rock; b is the tting constant.
When the bulk density data (r
z
) are available at certain depths, by
tting the density curve to Eq. (4), the shallow density (r
0
) can be
obtained from Eq. (4).
Another method to calculate shallow formation density is
Millers near surface or mudline density correlation, which can be
expressed as follows (Zhang et al., 2008):
r
s
r
m
1 f
s
r
w
f
s
(5)
where r
m
is the average density of the sediment grains (typically
2.68 g/cm
3
for shales); r
w
is the density of the pore water (typically
Piezometric surface
p h g
f f
=
h
Ground surface
Piezometric surface
p h g
f f
pp =
h
Ground surface
Confined aquifer
Figure 1. Schematic cross-section showing the hydrostatic pressure caused by water
column in a subsurface formation (aquifer).
J. Zhang / Marine and Petroleum Geology 45 (2013) 2e11 3
1.03 g/cm
3
); f
s
is the shallow porosity and can be calculated from
the following empirical equation:
f
s
f
a
f
b
e
kd
1=N
(6)
where f
a
f
b
is the mudline porosity, d is the shallow depth below
mudline, and k and N are empirically determined parameters that
provide a reasonable t to the data; in the Gulf of Mexico, typically
f
s
0:35 0:35e
0:0035d
1=1:09
; d is in ft.
The relationship of pore pressure, overburden stress and effec-
tive stress is illustrated in Figure 2, assuming a 1. Figure 2 shows
that the pore pressure in deep formations is much higher than the
hydrostatic (normal) pressure, hence overpressures exist. Figure 2
also shows when pore pressure and overburden stress are known,
the effective stress can be easily calculated in each post-drill well.
This effective stress can be correlated to petrophysical and
geophysical properties of the formations (e.g., resistivity, velocity,
porosity) and thus be applied for pre-drill pore pressure prediction
in new wells in a similar area.
1.4. Overview of effective stress and pore pressure prediction from
compressional velocity
It should be noted that the pore pressure prediction methods
are based on the rock properties in shales, and the pore pressures
obtained from these methods are the pressures in shales. For the
pressures in sandstones, limestones or other permeable formations,
the pore pressure can be obtained by either assuming that the shale
pressure is equal to the sandstone pressure or using uid ow
model (centroid method) (Dickinson, 1953; Traugott, 1997; Yardley
and Swarbrick, 2000; Zhang, 2011) to do calculation. In addition,
the pore pressure around a wellbore is affected by stress redistri-
bution near wellbore wall due to drilling perturbations (Zhang
et al., 2003; Zhang and Roegiers, 2005). Therefore, the pore pres-
sure in this paper is designated to the far-eld pore pressure, where
the drilling effect is negligible.
1.4.1. Hottmann and Johnsons method
Hottmann and Johnson (1965) made pore pressure prediction
using shale properties derived from well log data. They indicated
that porosity decreases as a function of depth for analyzing acoustic
travel time (or transit time) in Miocene and Oligocene shales in
Upper Texas and Southern Louisiana Gulf Coast. This porositye
depth relationship represents the normal compaction trend as a
function of burial depth, and uid pressures exhibited within this
normal trend are the hydrostatic. If intervals of the abnormal
compacted formations are penetrated, the resulting porosity or
travel time data points diverge from the normal compaction trend.
They concluded that porosity or transit time in shale is abnormally
high relative to its depth if the uid pressure is abnormally high.
Analyzing the data presented by Hottmann and Johnson (1965),
Gardner et al. (1974) proposed the following equation for pore
pressure calculation:
p
f
s
V
a
V
bA
1
B
1
lnDt
3
=Z
2
(7)
where p
f
is the formation uid pressure (psi); s
V
is the overburden
stress (psi); a
V
is the normal overburden stress gradient (psi/ft); b is
the normal uid pressure gradient (psi/ft); Z is the depth (ft); Dt is
the compressional sonic transit time (ms/ft); A
1
and B
1
are the
constants, A
1
82776 and B
1
15695.
The compressional transit time is the inverse of the compres-
sional interval velocity and can be expressed in the following form:
v
p

10
6
Dt
(8)
where v
p
is the compressional interval velocity in ft/s; Dt is the
compressional transit time in ms/ft.
1.4.2. Eatons method
Eaton (1975) presented the following empirical equation for
pore pressure gradient prediction from sonic compressional transit
time:
P
pg
OBG
_
OBG P
ng
_
Dt
n
=Dt
3
(9)
where P
pg
is the formation pressure gradient and equal to the pore
pressure divided by the true vertical depth; OBG is the overburden
stress gradient; P
ng
is the hydrostatic pore pressure gradient; Dt
n
is
the transit time or slowness in shales at the normal pressure; Dt is
the transit time in shales obtained from well logging.
This method is applicable in some petroleum basins (e.g., Sayers
et al., 2002), but it does not consider unloading effects. This limits
its application in geologically complicated area, such as formations
with uplifts. To apply this method, one needs to determine the
normal transit time (Dt
n
).
1.4.3. Modied Eatons method
Zhang (2011) proposed modied Eatons method by using
depth-dependent normal compaction trend line:
P
pg
OBG
_
OBG P
ng
_
_
Dt
m
Dt
ml
Dt
m
e
cZ
Dt
_
n
(10)
where Dt
m
is the compressional transit time in the shale matrix
(normally 65 ms/ft); Dt
ml
is the transit time in the mudline (nor-
mally 200 ms/ft); Z is the depth below the mudline; c is the
compaction constant; n is the exponent, and normally n 3.
The normal compaction trend (Dt
n
) in this modied Eatons
method decreases exponentially with depth as follows:
0
2000
4000
6000
8000
10000
12000
14000
16000
18000
20000
0 2000 4000 6000 8000 1000012000140001600018000
D
e
p
t
h

K
B

T
V
D

(
f
t
)
Stress, pressure (psi)
Overburden stress
MDT
Hydrostatic pressure
pore
pressure
0.465psi/ft
Sea floor
Effective stress
normal
pore pressure
Figure 2. Pore pressure, overburden stress and effective stress versus the true vertical
depth (TVD) in a deepwater well in the Gulf of Mexico. The MDT points are the
measured formation pore pressures from the borehole.
J. Zhang / Marine and Petroleum Geology 45 (2013) 2e11 4
Dt
n
Dt
m
Dt
ml
Dt
m
e
cZ
(10a)
1.4.4. Mann and Mackenzies effective stress model
Mann and Mackenzie (1990) presented the following effective
stress and porosity equation for sedimentary basins:
3 3
0
C
c
log
10
_
s
0
=s
0
0
_
(11)
where 3 is the void ratio, and 3 f/(1 f); f is the porosity; 3
0
is the
void ratio near the surface; s
0
is the effective stress; s
0
0
is the
reference effective stress, and calculations are started at 1 m below
the surface of the sediment where s
0
0
7800 Pa; C
c
is a lithology-
dependent constant, and C
c
0.43 for shale.
1.4.5. Bowers method
Bowers (1995) calculated the effective stresses from measured
pore pressure data and overburden stresses (refer to Eq. (1)) and
analyzed the corresponded sonic velocities from well logging data
in the Gulf of Mexico slope. He proposed that the sonic velocity and
effective stress have a power relationship as follows:
v
p
v
ml
As
B
e
(12)
where v
p
is the compressional velocity at a given depth; v
ml
is the
compressional velocity at the mudline (normally 5000 ft/s); s
e
is
the vertical effective stress; A and B are the parameters calibrated
with offset velocities versus effective stress data.
The effective stress and compressional velocity do not followthe
loading curve if formation uplift or unloading occurs, and a higher
than the velocity in the loading curve appears at the same effective
stress. Bowers (1995) proposed the following empirical relation to
account for unloading effect:
v
p
v
ml
A
_
s
max
s
e
=s
max

1=U
_
B
(13)
where s
e
, v
p
, v
ml
, A and B are as before; U is a parameter;
s
max
v
max
v
ml
=A
1=B
; s
max
and v
max
are estimates of the
effective stress and velocity at the onset unloading.
1.4.6. Millers method
Pore pressure can be also obtained from Millers sonic velocity
method in the following equation (Zhang et al., 2008):
p s
v

1
l
ln
_
v
m
v
ml
v
m
v
p
_
(14)
where p is the pore pressure; v
ml
is the interval velocity of sedi-
ments in the mudline; v
m
is the sonic interval velocity in the ma-
trix; l is an empirical parameter for calibrating the model (normally
0.00025).
1.4.7. Tau model
Dutta (2002) proposed the following relationship that relates
the transit time to effective stress:
s
e

1
k
1
ln
_
f
0
Ds
1=x
_
_
Ds
1=x
1
__
(15)
where Ds Dt/Dt
m
; Dt is the compressional transit time; Dt
m
is the
transit time in the matrix; f
0
is the porosity at Z 0; k
1
is a coef-
cient; and x is an acoustic formation factor dependent on
lithology.
A transit time dependent pore pressure prediction method was
presented through introducing a Tau variable into the effective
stress equation (e.g., Lopez et al., 2004; Gutierrez et al., 2006; Zhang
and Wieseneck, 2011):
s
e
A
s
s
Bs
(16)
where s
e
is the effective stress; A
s
and B
s
are the tting constants; s
is the Tau variable, and s (C Dt)/(Dt D); Dt is the compres-
sional transit time either from sonic log or seismic velocity; C is the
constant related to the mudline transit time; and D is the constant
related to the matrix transit time.
2. Effective stress, porosity and pore pressure relationships
2.1. Effective stress and porosity relationship
It has been veried that porosity decreases exponentially as
depth increases in normally-compacted formations, as described in
the following equation (e.g., Athy, 1930; Mondol et al., 2007). This is
the normal compaction trend in porosity.
f
n
f
0
e
cnZ
(17)
where f
n
is the porosity in the normally-compacted formation; Z is
the depth below the mudline; c
n
is the normal compaction con-
stant; f
0
is the porosity in the mudline; f
0
(r
m
r
0
)/r
m
(Rubey
and Hubbert, 1959); r
m
is the grain density of the rock; r
0
is the
bulk density of the surface or the mudline.
It is commonly accepted that formation porosity and effective
stress have the following relationship (e.g., Rubey and Hubbert,
1959; Dutta, 2002; Flemings et al., 2002; Peng and Zhang, 2007;
Tsuji et al., 2008):
f f
0
e
as
e
(18)
where a is the stress compaction constant.
Eq. (18) indicates that porosity is a functionof the effective stress;
therefore, pore pressure can be estimated from formation porosity.
Figure 3 illustrates the formation under-compaction and over-
pressure fromporosity prole. For a normallycompacted formation,
porosity should decrease exponentially as depth increases (as
described by Eq. (17)), where the formation has normal pore pres-
sure. When the porosity is reversal, the under-compaction occurs
and overpressure generates. The starting point of the porosity
reversal is the topof under-compactionor topof overpressure. Inthe
formation with under-compaction, porosity and pore pressure are
higher than those in the normally compacted section.
The effective stress can be obtained from Eq. (18) in the
following form:
s
e

1
a
ln
f
0
f
(19)
The effective stress at normal pressure condition can also be
obtained fromEq. (18), inwhich the porosity is the normal porosity,
a condition that formations are normally compacted, i.e.:
f
n
f
0
e
an
s
n
(20)
s
n

1
a
n
ln
f
0
f
n
(21)
Combining Eqs. (19) and (21), we have the following equation
for porosity and effective stress:
s
e
s
n

a
n
a
ln f
0
ln f
ln f
0
ln f
n
(22)
J. Zhang / Marine and Petroleum Geology 45 (2013) 2e11 5
Substituting Eq. (17) into Eq. (22), we obtain the following
constitutive relationship between effective stress and porosity:
s
e
s
n
ln f
0
ln f
cZ
(23)
where c is the generalized compaction constant. Therefore, from
Eq. (23) the porosity in both normal compaction and under-
compaction cases can be written in the following generalized form:
f f
0
e
cZs
e=s
n
(24)
In normal compaction condition (s
e
s
n
), the above equation is
simplied to Eq. (17). Therefore, this new equation extends Athys
porosity equation (Athy, 1930) to a generalized form which is
applicable for both normally compacted and under-compacted
formations.
2.2. Pore pressure prediction from porosity
A number of models are proposed for pore pressure prediction
from porosity (e.g., Heppard et al., 1998; Flemings et al., 2002;
Holbrook et al., 2005; Schneider et al., 2009). From Eq. (23) by
noticing s
e
s
V
ap and s
n
s
V
ap
n
, the following relationship
of pore pressure, overburden stress and porosity can be derived:
p
_
s
V
s
V
ap
n

ln f
0
ln f
cZ
__
a (25)
where p is the pore pressure; s
V
is the overburden stress; p
n
is the
normal pore pressure, f
0
is the porosity in the formation of the
mudline; Z is the depth below mudline; c is a constant and can be
obtained from the normal compaction porosity trend line. a is the
Biot effective stress coefcient, and f a 1; it is conventionally
assumed a 1 in the geopressure community.
0
Porosity
Top under-compaction
Under-compaction
Normal compaction
Pressure
V
n
p
p
n
e
Top overpressure
Overpressure
Normal pressure
Figure 3. Schematic porosity (a) and corresponding pore pressure (b) in a sedimentary basin. The dash porosity prole in (a) represents normally compacted formation. In the
overpressured section the porosity reversal occurs (solid line) and the porosity is larger than that in the normally pressured section.
0
2000
4000
6000
8000
10000
12000
14000
16000
18000
20000
0 4000 8000 12000 16000 20000
D
e
p
t
h

(
f
t

B
M
L
)
Pressure (psi)
Hydrostatic-8.65 ppg
Measured pore pressure
0
2000
4000
6000
8000
10000
12000
14000
16000
18000
20000
40 60 80 100 120 140 160 180 200
D
e
p
t
h

b
e
l
o
w

m
u
d
l
i
n
e
(
f
t
)
t (s/ft)
Figure 4. Measured pore pressures and corresponding shale transit time versus depth below the mudline in the studied basins without uplift. (a) measured pore pressures and
hydrostatic pressure (with a gradient of 8.65 ppg); (b) measured sonic transit time in shale and the normal compaction trend line (NCTL, the thicker dash line calculated from Eq.
(10a)).
J. Zhang / Marine and Petroleum Geology 45 (2013) 2e11 6
The advantage of the proposed model in Eq. (25) is that the
pressures calculated from porosity are dependent on depths.
3. Theoretical model of effective stress and transit time/
velocity
3.1. Theoretical relationship of effective stress, transit time and
depth
From Wyllie equation (Wyllie et al., 1956), the porosity in the
interested depth (f) and the porosity in the mudline (f
0
or f
ml
) can
be written as the following equations:
f
Dt Dt
m
Dt
f
Dt
m
(26)
f
0

Dt
ml
Dt
m
Dt
f
Dt
m
(27)
where Dt is the compressional transit time; t
m
is the transit time in
the matrix; t
ml
is the transit time in the formation of the mudline; t
f
is the transit time in the uid of the pores.
Substituting Eqs. (26) and (27) into Eq. (23), we can obtain the
relationship of effective stress and transit time in the following
form:
s
e

s
n
cZ
ln
Dt
ml
Dt
m
Dt Dt
m
(28)
Or
Dt Dt
m
Dt
ml
Dt
m
e
cZs
e=s
n
(29)
In normal compaction case (s
e
s
n
) Eq. (29) becomes the
following form:
Dt Dt
m
Dt
ml
Dt
m
e
cZ
(10a)
This is the normal compaction trend for normal pressure case.
That is, in normal compaction case Eq. (29) simplies to Chapmans
model (Chapman, 1983, P.50, Eq. (3.8)). Therefore, Eq. (29) extends
the Chapmans transit timeedepth relationship to both normally
compacted and under-compacted formations.
3.2. Effective stress and velocity relationship from well logging data
Well logging data in several Tertiary and Jurassic petroleum
basins of offshore Gulf of Mexico and U.S.A. onshore elds were
analyzed to determine sonic velocity/transit time and effective
stress relationship (published data can be found in Jones, 1969;
Bowers, 1995; Issler, 1992; Flemings et al., 2002; Nelson and Bird,
2005; etc.). The downhole measured pore pressure data ranging
from normal pressures to overpressures were analyzed, as shown
in Figure 4a. The sonic transit time (slowness) at each data point of
the measured pore pressure was carefully picked from the nearest
shale and plotted in Figure 4b. The normal compaction trend line in
the transit time was calculated from Eq. (10a) using Dt
ml
203 ms/
ft, Dt
m
60 ms/ft, and c 0.00021 ft
1
. Figure 4 shows that the
overpressure corresponds to a higher transit time (or slower ve-
locity) compared to the normal transit time-depth trend. Figure 4b
also shows that the transit time does not always decrease mono-
tonically with depth. In the shallow depth with a normal pore
pressure, the transit time follows the normal compaction trend line
(NCTL fromEq. (10a)). When the formation is overpressures (under-
compacted), the transit time reversal occurs (Fig. 4b); i.e., the
transit time increases as the depth increases. Therefore, the pro-
posed relationship in Eq. (29) can better describe the transit timee
depth behavior in both normally compacted and under-compacted
cases.
In order to correlate the transit time to effective stress, the
vertical effective stresses are rstly calculated from Eq. (1) by
assuming a 1 using the pore pressure data shown in Figure 4,
while overburden stresses are obtained from integrating bulk
density log data using Eq. (3). Then, the effective stresses and cor-
responding sonic transit time are plotted versus depth, shown in
Figure 5. The measured data in those petroleum basins plotted in
Figure 5 indicate an exponential relationship between the vertical
effective stress (s
e
) and compressional transit time (Dt):
Dt Dt
m
175:49e
0:000267Zs
e=s
n
(30)
where Z is the depth belowthe mudline in ft; s
e
and s
n
are in psi; Dt
and Dt
m
are in ms/ft. Eq. (30) can be expressed as the following
general form:
Dt Dt
m
Me
kZs
e=s
n
(31)
where M and k are the tting constants. This relationship veries
the proposed theoretical solution (i.e., Eq. (29)).
Plotting the data with formation uplift (unloading case) to
Figure 5, it shows that the unloading curve is different from the
original compaction/loading curve, as shown in Figure 6. The
unloading occurs along a atter effective stress-transit time path
than the initial compaction/loading curve. This unloading curve
still denes an exponential relationship between the vertical
effective stress and compressional transit time, but it is atter and
with a different compaction constant (Fig. 6). Figure 7 plots the
effective stress and the compressional velocity converted from the
transit time of the same data shown in Figure 6. Figure 7 shows the
loading and unloading curves of the velocities having very different
trends.
3.3. Experimental results of effective stress and velocity in loading
and unloading cases
An experimental study of compaction effects on the acoustic
velocity in soils was conducted with a conventional triaxial cell
apparatus at the University of Mississippi (Lu et al., 2004). In the
y = 175.491175e
-0.000267x
R = 0.912911
0
20
40
60
80
100
0 2000 4000 6000 8000 10000
t
t

m
(
s
/
f
t
)
( e/ n)*Zbml (ft)
DT-DTm
Expon. (DT-DTm)
Figure 5. Vertical effective stresses versus sonic transit time (DT or Dt) in the studied
basins without uplift (with normal pore pressure gradient of 8.65 ppg and Dt
m
60 ms/
ft); Z
bml
is the depth below the mudline in ft.
J. Zhang / Marine and Petroleum Geology 45 (2013) 2e11 7
experiments, the device was modied to measure the velocity of a
compressional wave propagating through a soil sample during
triaxial compressive tests. Three soil samples taken from sites in
Sharkey, Neshoba, and Marshall Counties, Mississippi were com-
pacted vertically to simulate compaction processes. The compres-
sional wave velocity in the axial direction was measured along with
the measurement of the stress-strain response with different
conning stresses during both loading and unloading cases (Lu
et al., 2004). We plot the vertical effective stresses and acoustic
P-wave velocity responses in the soil compaction (loading) and
unloading processes based on the data provided by Dr. Lu (2011).
The compressional wave velocities and travel time in the
compaction test for Neshoba soil/clay are shown in Figures 8 and 9.
The acoustic velocities increase or travel time decreases as the
effective stress increases in the compaction stage. The acoustic
velocity during the unloading test does not recover to its original
loading path and decreases sharply with the change in the effective
stress. The reloading path follows approximately an opposite di-
rection of the unloading path. The unloadingeloading cycle forms a
clockwise loop representing a hysteresis for the acoustic velocity.
The acoustic behavior resumes its normally consolidated line after
passing the point where the unloadingereloading cycle starts. The
same results are in other tests, such as Sharkey clay (Lu et al., 2004).
Due to unloading, the relationship of the effective stress and the
velocity does not follow the loading curve, and a higher velocity
exists than the velocity in the loading curve at the same effective
stress. The experimental results in loading and unloading cases are
consistent to the eld data (Figs. 6 and 7) and the derived effective
stress and transit time relationship (Eq. (29)).
4. Theoretical model of pore pressure calculation from
transit time or velocity and its application
4.1. Pore pressure model without unloading
From Eq. (28) and noticing s
e
s
V
ap and s
n
s
V
ap
n
, we
obtain the following equation to calculate pore pressure (p):
p
_
s
V

s
V
ap
n

cZ
ln
Dt
ml
Dt
m
Dt Dt
m
__
a (32)
where Dt is the measured compressional transit time; Dt
m
is the
transit time in the matrix; Dt
ml
is the transit time in the mudline; c
is the compaction constant and can be determined from Eq. (10a),
0
20
40
60
80
100
0 2000 4000 6000 8000 10000 12000
t

t

m
(
s
/
f
t
)
( e/ n)*Zbml (ft)
DT-DTm
Expon. (DT-DTm)
unloading
Figure 6. Relationship between the vertical effective stresses and sonic transit time
with unloading effect in the studied basins.
5000
6000
7000
8000
9000
10000
11000
12000
13000
14000
15000
0 2000 4000 6000 8000 10000
V
p

i
n

s
h
a
l
e

(
f
t
/
s
)
( e/ n)*Zbml (ft)
Vp in shale
Expon. (Vp in shale)
unloading
Figure 7. Relationship between the vertical effective stresses and compressional ve-
locities with unloading effect plotted from Figure 6.
500
1000
1500
2000
2500
3000
3500
0 100 200 300 400 500 600 700
A
c
o
u
s
t
i
c

t
r
a
v
e
l

t
i
m
e

(

s
/
m
)
Vertical effective stress (kPa)
Pc=103.4 kPa
Figure 8. Vertical effective stress versus the acoustic travel time in the unconsolidated
undrained test for air-dry remolded Neshoba soils with a conning pressure of
103.4 kPa. Data provided by Lu, 2011.
200
300
400
500
600
700
800
900
1000
1100
1200
0 100 200 300 400 500 600 700
A
c
o
u
s
t
i
c

v
e
l
o
c
i
t
y

(
m
/
s
)
Vertical effective stress (kPa)
Pc=103.4 kPa
Figure 9. Vertical effective stress versus the acoustic velocity in the unconsolidated
undrained test for air-dry remolded Neshoba soils with a conning pressure of
103.4 kPa. Data provided by Lu, 2011.
J. Zhang / Marine and Petroleum Geology 45 (2013) 2e11 8
once the normal compaction trend is known. The advantage of this
model is that the calculated pore pressures are dependent on
depth, and both the effects of the matrix and mudline transit time
are considered.
4.2. Pore pressure model accounting for unloading
Unloading case has a different compaction path, thus, a different
compaction constant (b), as shown in Figures 6, 8 and 10. The
compressional transit time and vertical effective stress in unloading
case have the following relationship (refer to Appendix A for the
derivation):
Dt Dt
m
Dt
ml
Dt
m
e
bcymaxbZs
e=s
n
(33)
where b is the compaction constant in the unloading case; b c if
no unloading occurs, and b > c in unloading case; y
max
is dened in
Figure 10.
The relationship between the effective stress and transit time in
unloading case can be expressed in the following form (refer to
Appendix A for the derivation):
s
e

s
n
bZ
_
b c
c
ln
Dt
ml
Dt
m
Dt
u0
Dt
m
ln
Dt
ml
Dt
m
Dt Dt
m
_
(34)
where Dt
u0
is the transit time at the starting point of the unloading,
as dened in Figure 10.
From Eq. (34) and noticing s
e
s
V
ap and s
n
s
V
ap
n
, we
obtain the following equation to calculate pore pressure in
unloading case:
p
_
s
V

s
V
ap
n

bZ
_
bc
c
ln
Dt
ml
Dt
m
Dt
u0
Dt
m
ln
Dt
ml
Dt
m
Dt Dt
m
___
a
(35)
4.3. Case applications
A Tertiary petroleum basin of subsalt formations in deepwater
Gulf of Mexico, as described in Zhang et al. (2008), is examined to
verify the proposed model for pore pressure calculation. This case
study presents the pore pressure analysis in a post-drill well with
water depth of 3560 ft. The formations are primarily Tertiary shales
and sandstones, and the target zone is located in the Middle
Miocene sandstones. Figure 11 shows a post-drill pore pressure
analysis to examine the proposed pore pressure model. Pore
pressure gradient is calculated from the proposed equation (Eq.
(32)) using Dt
ml
131 ms/ft, Dt
m
73 ms/ft, P
ng
8.75 ppg, a 1,
and c 0.00009 ft
1
. The pore pressure gradient is also estimated
using Eatons resistivity method (Eaton, 1975). Compared to the
measured pore pressure results (MDT) and well inux (uid gain),
the proposed method (Eq. (32)) gives an excellent result in pore
pressure calculation. Also, the pore pressure calculation from the
proposed method gives a better result than the resistivity method.
It should be noted that the mudline transit time needs to be
adjusted Dt
ml
131 ms/ft (instead of 200 ms/ft in conventional cases)
to make a better pore pressure estimation in subsalt formations.
Figure 11 also demonstrates that the pore pressure calculation from
the proposed method can excellently catch the pore pressure
Figure 10. Simplied plot from Figures 6 and 8 showing the relationship between the
vertical effective stress and transit time in loading and unloading cases.
Figure 11. Pore pressure calculation from the sonic transit time using the proposed method (Eq. (32)) in subsalt formations of deepwater Gulf of Mexico. In this gure, the gamma
ray and shale base lines are shown in the left track; the resistivity (Res) and ltered shale points of resistivity (SHPT Res) are plotted in the second track; the sonic transit time (DT)
and ltered shale points of the transit time (SHPT DT) are shown in the third track; and the calculated pore pressures from the ltered shale transit time (Pp DT) and resistivity (PP
res e1.2) are shown in the right track with comparison to the measured formation pressures (MDT) and mud weights (MWIN).
J. Zhang / Marine and Petroleum Geology 45 (2013) 2e11 9
regression, which is a common phenomenon in some areas of the
Gulf of Mexico.
Figure 12 shows another case application in pore pressure es-
timates in the clastic (non-salt) formations. In this case the sonic
transit time from well logs are used for pore pressure calculations
from modied Eatons method (Eq. (10)) and the proposed method
(Eq. (32)) using the exponential normal compaction trend line (Eq.
(10a)). To determine the normal compaction trend line, the sonic
transit time in shallow section (with normal pore pressure) is t to
the normal compaction line by using Eq. (10a) with parameters of
Dt
ml
205 ms/ft, Dt
m
70 ms/ft, and c 0.00026 ft
1
. The calculated
pore pressure fromthe proposed method (with P
ng
8.5 ppg, a 1,
and c 0.00026 ft
1
) matches the measured pore pressures from
the drill stem tests (DST). To use the Eatons method (Eq. (10)) in
this case, the exponent needs to be adjusted (n 2.4) to match the
measured pore pressure result.
5. Conclusions
Porosity is not only dependent on depth, but also a function of
the effective stress. Porosity does not always decrease with depth;
however, it increases when the increase of the effective stress
with depth is smaller than the effective stress in the normal
compaction condition. Effective stress from porosity and
compressional velocity is derived from compaction disequilib-
rium theory. This theoretical relationship of effective stress and
velocity (or transit time) is veried by eld data and lab experi-
mental results. Theoretical pore pressure-porosity model is pro-
posed for pore pressure prediction in shales based on the
compaction disequilibrium. Using this theoretical model, pore
pressure prediction from compressional velocity (transit time) is
obtained, and unloading case is also considered for pore pressure
calculation. Case study indicates that pore pressure can be accu-
rately obtained from velocity and well logging data using pro-
posed method with necessary calibrations.
Appendix A. Derivation of effective stress in unloading case
The effective stress and transit time in the loading case has the
following relationship (i.e., Eq. (29)):
Dt Dt
m
Dt
ml
Dt
m
e
cZs
e=s
n
(A1)
Based on Eq. (A1) the following relationship between effective
stress and transit time is assumed in the unloading case (Dt
ul
), as
shown in Figure 10:
Dt
ul
Dt
m
Be
bZs
e=s
n
(A2)
where B and b are constants.
Assuming y Z(s
e
/s
n
) as shown in Figure 10, Eqs. (A1) and (A2)
become:
Dt Dt
m
Dt
ml
Dt
m
e
cy
(A3)
Dt
ul
Dt
m
Be
by
(A4)
At the starting point of the unloading where the transit time
reversal starts and the maximum velocity (or the minimum transit
time, Dt
u0
) occurs in the loading curve, we have the following re-
lationships, because the both unloading and loading curves (Eqs.
(A3) and (A4) intercept at the point (Dt
u0
, y
max
). Hence, the
following equations exist:
Dt
u0
Dt
m
Dt
ml
Dt
m
e
cymax
(A5)
Dt
u0
Dt
m
Be
bymax
(A6)
y
max

1
c
ln
Dt
ml
Dt
m
Dt
u0
Dt
m
(A7)
Figure 12. Pore pressure calculation from the sonic transit time using the proposed method (Eq. (32)) and modied Eaton method (Eq. (10)). In this gure, the exponential normal
compaction trend line (from Eq. (10a)) and ltered shale points of sonic transit time (SHPT DT) are shown in the left track, and the calculated pore pressures from the ltered shale
transit time from the proposed method (Pp DT) and modied Eaton method (Pp Eaton) are shown in the right track with comparison to the measured formation pressures (DST).
J. Zhang / Marine and Petroleum Geology 45 (2013) 2e11 10
Substituting Eq. (A5) into Eq. (A6), we obtain:
B Dt
ml
Dt
m
e
bcymax
(A8)
Substituting Eq. (A8) into Eq. (A2), we obtain the transit time as
a function of the effective stress in unloading case:
Dt Dt
m
Dt
ml
Dt
m
e
bcymaxbZs
e=s
n
(A9)
The effective stress can be solved fromEq. (A9) with substituting
Eq. (A7) as following:
s
e

s
n
bZ
_
b c
c
ln
Dt
ml
Dt
m
Dt
u0
Dt
m
ln
Dt
ml
Dt
m
Dt Dt
m
_
(A10)
References
Athy, L.F., 1930. Density, porosity, and compaction of sedimentary rocks. AAPG
Bulletin 14 (1), 1e24.
Biot, M.A., 1941. General theory of three-dimensional consolidation. Journal of
Applied Physics 12 (1), 155e164.
Bird, P., 1995. Lithosphere dynamics and continental deformation. Reviews of
Geophysics (Suppl.), 379e383.
Bowers, G.L., June, 1995. Pore pressure estimation fromvelocity data; accounting for
overpressure mechanisms besides undercompaction. SPE Drilling and Com-
pletions, 89e95.
Chapman, R.E., 1983. Petroleum Geology. Elsevier.
Chillingar, G.V., Serebryakov, V.A., Robertson, J.O., 2002. Origin and Prediction of
Abnormal Formation Pressures. Elsevier.
Davies, R.J., Swarbrick, R.E., Evans, R.J., Huuse, M., 2007. Birth of a mud volcano: East
Java, 29 May 2006. GSA Today 17 (2). http://dx.doi.org/10.1130/GSAT01702A.1.
Dickinson, G., 1953. Geological aspects of abnormal reservoir pressures in Gulf Coast
Louisiana. AAPG Bulletin 37 (2), 410e432.
Dutta, N.C., 2002. Geopressure prediction using seismic data: current status and the
road ahead. Geophysics 67 (6), 2012e2041.
Eaton, B.A., 1975. The Equation for Geopressure Prediction from Well Logs. Society
of Petroleum Engineers of AIME. Paper SPE 5544.
Flemings, P.B., Stump, B.B., Finkbeiner, T., Zoback, M., 2002. Flow focusing in over-
pressured sandstones: theory, observations, and applications. American Journal
of Science 302, 827e855.
Gardner, G.H.F., Gardner, L.W., Gregory, A.R., 1974. Formation velocity and density e
the diagnostic basis for stratigraphic traps. Geophysics 39 (6), 2085e2095.
Gutierrez, M.A., Braunsdorf, N.R., Couzens, B.A., Dec., 2006. Calibration and ranking
of pore-pressure prediction models. The Leading Edge, 1516e1523.
Heppard, P.D., Cander, H.S., Eggertson, E.B., 1998. Abnormal pressure and the
occurrence of hydrocarbons in offshore eastern Trinidad, West Indies. In:
Law, B.E., Ulmishek, G.F., Slavin, V.I. (Eds.), Abnormal Pressures in Hydrocarbon
Environments. AAPG Memoir, vol. 70, pp. 215e246.
Holbrook, P.W., Maggiori, D.A., Hensley, R., 2005. Real-time pore pressure and
fracture gradient evaluation in all sedimentary lithologies. SPE Formation
Evaluation 10 (4), 215e222.
Holand, P., Skalle, P., 2001. Deepwater Kicks and BOP Performance. SINTEF Report
for U.S. Minerals Management Service.
Hottmann, C.E., Johnson, R.K., 1965. Estimation of formation pressures from log-
derived shale properties. Paper SPE1110. JPT 17, 717e722.
Issler, D.R., 1992. A newapproach to shale compaction and stratigraphic restoration,
Beaufort-Mackenzie Basin and Mackenzie Corridor, Northern Canada. AAPG
Bulletin 76, 1170e1189.
Jones, P.H., 1969. Hydrodynamics of geopressure in the Northern Gulf of Mexico
basin. SPE2207. JPT.
Lopez, J.L., Rappold, P.M., Ugueto, G.A., Wieseneck, J.B., Vu, K., Jan., 2004. Integrated
shared earth model: 3D pore-pressure prediction and uncertainty analysis. The
Leading Edge, 52e59.
Lu, Z., Hickey, C.J., Sabatier, J.M., 2004. Effects of compaction on the acoustic velocity
in soils. Soil Science Society of America Journal 68, 7e16.
Lu, Z., 2011. Personal communication.
Mann, D.M., Mackenzie, A.S., 1990. Prediction of pore uid pressures in sedimentary
basins. Marine and Petroleum Geology 7, 55e65.
Mondol, N.H., Bjrlykke, K., Jahren, J., Heg, K., 2007. Experimental mechanical
compaction of clay mineral aggregates e changes in physical properties of
mudstones during burial. Marine and Petroleum Geology 24 (5), 289e311.
Mouchet, J.-C., Mitchell, A., 1989. Abnormal Pressures While Drilling. Editions
TECHNIP, Paris.
Nelson, H.N., Bird, K.J., 2005. Porosity-depth Trends and Regional Uplift Calculated
from Sonic Logs. National Reserve in Alaska. Scientic Investigation Report
2005-5051, U.S. Dept. of the Interior and USGS.
Osborne, M.J., Swarbrick, R.E., 1997. Mechanisms for generating overpressure in
sedimentary basins: a reevaluation. AAPG Bulletin 81, 1023e1041.
Peng, S., Zhang, J., 2007. Engineering Geology for Underground Rocks. Springer.
Rubey, W.W., Hubbert, M.K., 1959. Role of uid pressure in mechanics of overthrust
faulting, II. Overthrust belt in geosynclinal area of western Wyoming in light of
uid-pressure hypothesis. Geological Society of America Bulletin 70, 167e205.
Sayers, C.M., Johnson, G.M., Denyer, G., 2002. Predrill pore pressure prediction using
seismic data. Geophysics 67, 1286e1292.
Schneider, J., Flemings, P.B., Dugan, B., Long, H., Germaine, J.T., 2009. Overpressure
and consolidation near the seaoor of Brazos-Trinity Basin IV, northwest
deepwater Gulf of Mexico. Journal of Geophysical Research 114, B05102.
Skalle, P., Podio, A.L., June, 1998. Trends extracted from 1,200 Gulf Coast blowouts
during 1960e1996. World Oil, 67e72.
Swarbrick, R.E., Osborne, M.J., 1998. Mechanisms that generate abnormal pressures:
an overview. In: Law, B.E., Ulmishek, G.F., Slavin, V.I. (Eds.), Abnormal Pressures
in Hydrocarbon Environments. AAPG Memoir, 70, pp. 13e34.
Tobin, H.J., Saffer, D.M., 2009. Elevated uid pressure and extreme mechanical
weakness of a plate boundary thrust, Nankai Trough subduction zone. Geology
37 (8), 679e682.
Tsuji, T., Tokuyama, H., Costa Pisani, P., Moore, G., 2008. Effective stress and pore
pressure in the Nankai accretionary prism off the Muroto Peninsula, south-
western Japan. Journal of Geophysical Research 113, B11401.
Terzaghi, K., Peck, R.B., Mesri, G., 1996. Soil Mechanics in Engineering Practice, third
ed. John Wiley & Sons.
Tingay, M.R.P., Hillis, R.R., Swarbrick, R.E., Morley, C.K., Damit, A.R., 2009. Origin of
overpressure and pore-pressure prediction in the Baram province, Brunei.
AAPG Bulletin 93 (1), 51e74.
Traugott, M., 1997. Pore and fracture pressure determinations indeepwater. World
Oil 218 (8), 68e70.
Wyllie, M.R.J., Gregory, A.R., Gardner, L.W., 1956. Elastic wave velocities in hetero-
geneous and porous media. Geophysics 21, 41e70.
Yardley, G.S., Swarbrick, R.E., 2000. Lateral transfer: a source of additional over-
pressure? Marine and Petroleum Geology 17, 523e537.
York, P., Prithard, D., Dodson, J.K., Dodson, T., Rosenberg, S., Gala, D., Utama, B., 2009.
Eliminating Non-productive Time Associated Drilling Trouble Zone. OTC20220
Presented at the 2009 Offshore Tech. Conf. held in Houston.
Zhang, J., Bai, M., Roegiers, J.-C., 2003. Dual-porosity poroelastic analyses of well-
bore stability. International Journal of Rock Mechanics and Mining Sciences 40,
473e483.
Zhang, J., Roegiers, J.-C., 2005. Double porosity nite element method for borehole
modeling. Rock Mechanics and Rock Engineering 38, 217e242.
Zhang, J., Standird, W.B., Lenamond, C., 2008. Casing Ultradeep, Ultralong Salt
Sections in Deep Water: a Case Study for Failure Diagnosis and Risk Mitigation
in Record-depth Well. Paper SPE 114273 presented at SPE Annual Technical
Conference and Exhibition, 21e24 September 2008, Denver, Colorado, USA.
Zhang, J., 2011. Pore pressure prediction from well logs: methods, modications,
and new approaches. Earth-Science Reviews 108, 50e63. http://dx.doi.org/
10.1016/j.earscirev.2011.06.001.
Zhang, J., Wieseneck, J., 2011. Challenges and Surprises of Abnormal Pore Pressures in
the Shale Gas Formations. Paper SPE 145964 presented at SPE Annual Technical
Conference and Exhibition, 30 Oct.-2 Nov. 2011, Denver, Colorado, USA.
J. Zhang / Marine and Petroleum Geology 45 (2013) 2e11 11

You might also like