You are on page 1of 28

See

discussions, stats, and author profiles for this publication at: https://www.researchgate.net/publication/260986147

Effective stress, porosity, velocity and abnormal


pore pressure prediction accounting for
compaction disequilibrium and unloading
Article in Marine and Petroleum Geology August 2013
DOI: 10.1016/j.marpetgeo.2013.04.007

CITATIONS

READS

463

1 author:
Jon Jincai Zhang
Hess Corporation
64 PUBLICATIONS 689 CITATIONS
SEE PROFILE

All in-text references underlined in blue are linked to publications on ResearchGate,


letting you access and read them immediately.

Available from: Jon Jincai Zhang


Retrieved on: 30 September 2016

Abstract
Abnormal pore pressures, mostly overpressures, exist in many sedimentary formations. The
overpressures deteriorate drilling safety, causing borehole influx, 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 porositydepth relationship for both normally compacted and abnormally compacted formations, i.e.,

0e cZ (

/n )

. 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., t tm 0 (t f tm )ecZ ( e / n ) ).
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
1

effective stress and velocity relationship. Case study in an oil field is presented to examine the
proposed model for pore pressure analysis in subsalt formations.

Keywords
Pore pressure prediction, effective stress, porosity-depth relationship, overpressure, compaction
disequilibrium, compressional velocity and transit time, unloading

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 fluid 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., smectite-illite 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 fluid is
expelled. During burial, increasing overburden stress is the prime cause of fluid expulsion. If the
sedimentation rate is slow, normal compaction occurs, i.e. equilibrium between increasing
2

overburden and ability to expel fluids is maintained (Mouchet and Mitchell, 1989). This normal
compaction generates hydrostatic pore pressure in the formation. When the sediments subside
rapidly, or the formation has extremely low permeability, fluids in the sediments can only be
partially expelled, and the remained fluid 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
porosities 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 fluid from
the surface to the interested depth. For formations with normal fluid 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 fluid in the pores. That is, the pressure is the
same at the same depth within the fluid with a uniform density if the fluid is static. Thus, the
hydrostatic pressure can be calculated using the following equation:
pn f gh

(1)

where pn is the hydrostatic pressure; g is the acceleration due to gravity; f is the fluid density;
and h is the vertical height of the fluid column, as shown in Fig. 1.

Ground surface

Piezometric surface

h
p f f gh
Confined aquifer

Fig. 1. Schematic cross-section showing the hydrostatic pressure caused by water column in a
subsurface formation (aquifer).

The equation (Eq. 1) indicates that the hydrostatic pore pressure depends highly on
fluid/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 ( 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/cm3) in the Gulf of Mexico, and this corresponds to water with a salinity of 80,000
parts per million (ppm) of sodium chloride at 77F (25C) (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):

e V p

(2)

where p is the pore pressure, V is the overburden or vertical stress, e is the vertical effective
stress; is the Biots effective stress coefficient.
In normal pressure case, from the above equation the normal effective stress and normal
pore pressure have the following relationship:

n V pn

(2a)

where n is the normal vertical effective stress; pn is the normal or hydrostatic pressure.
The effective stress can be obtained by correlating to petrophysical 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 (V, H,

h). However, it is commonly 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 formations; hence, it can be
obtained by integrating bulk density logs. Therefore, vertical stress can be calculated by the
following equation:
z

V w gzw g b ( z )dz

(3)

zw

where b (z) is the formation bulk density as a function of depth; w is the density of sea water for
offshore drilling; z is the depth from the sea level; zw is the water depth, for onshore drilling zw =
0.
The bulk density can be obtained from well 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 depth-density curve in density measurements of shales in
northern Oklahoma, Athy (1930) proposed the following equation to interpolate shallow
formation bulk density:

z 0 Am (1 ebZ )

(4)

where z is the density at the depth of Z, in g/cm3; 0 is the formation density of the surface; A is
the maximum density increase possible ( Am m 0 and Am = 1.3 in Athy, 1930); m is the
matrix density or the grain density of the rock; b is the fitting constant. When the bulk density
data (z) are available at certain depths, by fitting the density curve to Eq. 4, the shallow density
(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):

s m (1 s ) ws

(5)

where m is the average density of the sediment grains (typically 2.68 g/cm3 for shales); w is the
density of the pore water (typically 1.03 g/cm3); s is the shallow porosity and can be calculated
from the following empirical equation:

s a be kd

1/ N

(6)

where a + b is the mudline porosity, d is the shallow depth below mudline, and k and N are
empirically determined parameters that provide a reasonable fit to the data; in the Gulf of
Mexico, typically s 0.35 0.35e0.0035d

1 / 1.09

; d is in ft.
5

The relationship of pore pressure, overburden stress and effective stress is illustrated in
Fig. 2, assuming = 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.

Stress, pressure (psi)


0

2000 4000 6000 8000 1000012000140001600018000

0
2000

Overburden stress
MDT
Hydrostatic pressure

Sea floor

4000

Depth KB TVD (ft)

6000

8000

10000
12000
14000
16000
18000

Effective stress
normal
pore pressure
0.465psi/ft

pore
pressure

20000

Fig. 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.

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 fluid flow model (Dickinson, 1953; Traugott, 1997; Yardley and Swarbrick,
2000; Zhang, 2011) to do calculation. In addition, the pore pressure around a wellbore is affected
6

by stress redistribution near wellbore wall due to drilling perturbations (Zhang et al., 2003;
Zhang and Roegiers, 2005). Therefore, the pore pressure in this paper is designated to the farfield 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 porosity-depth relationship represents the normal
compaction trend as a function of burial depth, and fluid 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 fluid
pressure is abnormally high.
Analyzing the data presented by Hottmann and Johnson (1965), Gardner et al. (1974)
presented the following equation for pore pressure calculation:

p f V (V )( A1 B1 ln t )3 / Z 2

(7)

where pf is the formation fluid pressure (psi); V is the overburden stress (psi); V is the normal
overburden stress gradient (psi/ft); is the normal fluid pressure gradient (psi/ft); Z is the depth
(ft); t is the compressional sonic transit time (s/ft); A1 and B1 are the constants, A1 = 82776
and B1 =15695.
The compressional transit time is the inverse of the compressional interval velocity and
can be expressed in the following form:

vp

106
t

(8)

where vp is the compressional interval velocity in ft/s; t is the compressional transit time in
s/ft.
1.4.2. Eatons Method
Eaton (1975) presented the following empirical equation for pore pressure gradient
prediction from sonic compressional transit time:
7

Ppg OBG (OBG Png )tn / t

(9)

where Ppg is the formation pressure gradient and equal to the pore pressure divided by the true
vertical depth; OBG is the overburden stress gradient; Png is the hydrostatic pore pressure
gradient; tn is the transit time or slowness in shales at the normal pressure; t 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
(tn).
1.4.3. Modified Eatons Method
Zhang (2011) proposed modified Eatons method by using depth-dependent normal
compaction trendline:
t ( tml tm )e cZ
Ppg OBG (OBG Png ) m
t

(10)

where tm is the compressional transit time in the shale matrix (normally 65 s/ft); tml is the
transit time in the mudline (normally 200 s/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 (tn) in this modified Eatons method decreases
exponentially with depth as follows:

tn tm ( tml tm )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:

0 Cc log10 ( / 0 )

(11)

where is the void ratio, and = /(1); is the porosity; 0 is the void ratio near the surface;

is the effective stress; 0 is the reference effective stress, and calculations are started at 1
meter below the surface of the sediment where 0 = 7800 Pa; Cc is a lithology-dependent
constant, and Cc = 0.43 for shale.
8

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 vml A eB

(12)

where vp is the compressional velocity at a given depth; vml is the compressional velocity at the
mudline (normally 5000 ft/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 follow the 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 vml A max ( e / max )1/U

(13)

where e, vp, vml, A and B are as before; U is a parameter; max (vmax vml ) / A

max and vmax 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 v

v vml

ln m
vm v p
1

(14)

where p is the pore pressure; vml is the interval velocity of sediments in the mudline; vm is the
sonic interval velocity in the matrix; 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:
9

1
ln[0 1 / x /( 1 / x 1)]
k1

(15)

where =t/tm; t is the compressional transit time; tm is the transit time in the matrix; 0 is
the porosity at Z = 0; k1 is a coefficient; 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):

e As B

(16)

where e is the effective stress; As and Bs are the fitting constants; is the Tau variable, and

(C t ) (t D) ; t is the compressional 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.
2.1.

Effective stress, porosity and pore pressure relationships


Effective stress and porosity relationship
It has been verified 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.

n 0e cZ

(17)

where n is the porosity in the normally-compacted formation; Z is the depth below the mudline;
c is the compaction constant; 0 is the porosity in the mudline; 0 ( m 0 ) / m (Rubey and
Hubbert, 1959); m is the grain density of the rock; 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):

0 e a

(18)

where a is the stress compaction constant.


10

Eq. 18 indicates that porosity is a function of the effective stress; therefore, pore pressure
can be estimated from formation porosity. Fig. 3 illustrates the formation under-compaction,
smectite-illite transformation and overpressure from porosity profile. For a normally compacted
formation, porosity should decrease exponentially as depth increases (as described by Eq. 17),
where the formation has normal pore pressure. When the porosity is reversal, the undercompaction occurs and overpressure generates. The starting point of the porosity reversal is the
top of under-compaction or top of overpressure. In the formation with under-compaction,
porosity and pore pressure are higher than those in the normally compacted section.

Porosity
0

Pressure

Normal compaction

Normal pressure
Top overpressure

Top under-compaction

p
pn

Under-compaction

(a)

Overpressure

(b)

Fig. 3. Schematic porosity (a) and corresponding pore pressure (b) in a sedimentary basin. The
dash porosity profile 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.

The effective stress can be obtained from Eq. 18 in the following form:

1
a

e ln

(19)

The effective stress at normal pressure condition can also be obtained from Eq. 18, in
which the porosity is the normal porosity, a condition that formations are normally compacted,
i.e.:

n 0e a

(20)

11

1
a

n ln

0
n

(21)

Combining Eqs. 19 and 21, we have the following equation for porosity and effective
stress:

e ln 0 ln

n ln 0 ln n

(22)

Substituting Eq. 17 into Eq. 22, we obtain the following constitutive relationship between
effective stress and porosity:

e n

ln 0 ln
cZ

(23)

Therefore, from Eq. 23 the porosity in both normal compaction and undercompaction
cases can be written in the following generalized form:

0e cZ (

/n )

(24)

In normal compaction condition ( e n ), the above equation is simplified 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 e V p and n V pn , the following relationship of pore pressure, overburden
stress and porosity can be derived:

ln ln

p V (V pn ) 0
/
cZ

(25)

where p is the pore pressure; V is the overburden stress; pn is the normal pore pressure, 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 trendline. is the Biot effective stress coefficient,
and 1; it is conventionally assumed = 1 in the geopressure community.
The advantage of the proposed model in Eq. 25 is that the pressures calculated from
porosity are dependent on depths.

12

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 () and
the porosity in the mudline (0 or ml) can be written as the following equations:

t tm
t f tm

(26)

where t is the compressional transit time; tm is the transit time in the matrix; tml is the transit
time in the formation of the mudline; tf is the transit time in the fluid of the pores.
Substituting Eq. (26) into Eq. (23), we can obtain the relationship of effective stress and
transit time in the following form:

n
cZ

ln

0 (t f tm )

(27)

t tm

Or

t tm 0 (t f tm )ecZ ( e / n )

(28)

In normal compaction case ( e n ) Eq. 29 becomes the following form:

t tm 0 (t f tm )ecZ

(29)

This is the normal compaction trend for normal pressure case. That is, in normal compaction
case with 0 = 1 Eq. 28 simplifies to Chapmans model (Chapman, 1983, P.50, Eq.3.8).
Therefore, Eq. 28 extends the Chapmans transit time-depth 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 fields 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 Fig. 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 Fig. 4b. The normal compaction trend line in the transit
13

time was calculated from Eq. 10a using tml = 203 s/ft, tm = 60 s/ft, and c = 0.00021ft-1.
Figure 4 shows that the overpressure corresponds to a higher transit time (or slower velocity)
compared to the normal transit time-depth trend. Figure 4b also shows that the transit time does
not always decrease monotonically with depth. In the shallow depth with a normal pore pressure,
the transit time follows the normal compaction trend line (NCTL from Eq. 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 proposed relationship in Eq. 29 can
better describe the transit time-depth behavior in both normally compacted and under-compacted
cases.

t (s/ft)

Pressure (psi)
0

4000

8000

40

12000 16000 20000


Hydrostatic-8.65 ppg

4000

Depth (ft BML)

80 100 120 140 160 180 200

2000

Measured pore pressure

Depth below mudline (ft)

2000

60

6000
8000
10000
12000
14000

4000
6000
8000
10000
12000
NCTL

14000

16000

16000

18000

18000

20000

20000

(a)

(b)

Fig. 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).

In order to correlate the transit time to effective stress, the vertical effective stresses are
firstly calculated from Eq. 1 by assuming = 1 using the pore pressure data shown in Fig. 4,
while overburden stresses are obtained from integrating bulk density log data using Eq. 3. Then,
14

the effective stresses and corresponding sonic transit time are plotted versus depth, shown in Fig.
5. The measured data in those petroleum basins plotted in Fig. 5 indicate an exponential
relationship between the vertical effective stress (e) and compressional transit time (t):
t tm 175.49e0.000267Z ( e / n )

(30)

where Z is the depth below the mudline in ft; e and n are in psi; t and tm are in s/ft. Eq. 30
can be expressed as the following general form:

t tm Me kZ ( e / n )

(31)

where M and k are the fitting constants. This relationship verifies the proposed theoretical
solution (i.e., Eq. 28).
Plotting the data with formation uplift (unloading case) to Fig. 5, it shows that the
unloading curve is different from the original compaction/loading curve, as shown in Fig. 6. The
unloading occurs along a flatter effective stress-transit time path than the initial
compaction/loading curve. This unloading curve still defines an exponential relationship between
the vertical effective stress and compressional transit time, but it is flatter 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 Fig. 6. Figure 7 shows the loading and
unloading curves of the velocities having very different trends.

100
DT-DTm
Expon. (DT-DTm)

t t m (s/ft)

80
y = 175.491175e-0.000267x
R = 0.912911

60
40

20
0
0

2000

4000

6000

(e/n)*Zbml (ft)

15

8000

10000

Fig. 5. Vertical effective stresses versus sonic transit time (DT or t) in the studied basins
without uplift (with normal pore pressure gradient of 8.65 ppg and tm = 60 s/ft); Zbml is the
depth below the mudline in ft.

100
DT-DTm
Expon. (DT-DTm)

t t m (s/ft)

80
60
40
20
unloading

0
0

2000

4000

6000

8000

(e/n)*Zbml (ft)

10000

12000

Fig. 6. Relationship between the vertical effective stresses and sonic transit time with unloading
effect in the studied basins.
15000
Vp in shale
Expon. (Vp in shale)

14000

Vp in shale (ft/s)

13000

unloading

12000
11000
10000
9000
8000
7000
6000
5000
0

2000

4000

6000

(e/n)*Zbml (ft)

8000

10000

Fig. 7. Relationship between the vertical effective stresses and compressional velocities with
unloading effect plotted from Fig. 6.

3.3. Experimental results of effective stress and velocity in loading and unloading cases
16

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 experiments, the device was modified 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 compacted
vertically to simulate compaction processes. The compressional wave velocity in the axial
direction was measured along with the measurement of the stress-strain response with different
confining 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 (Lu, 2011).
Acoustic travel time (s/m)

3500
Pc=103.4 kPa

3000
2500

2000
1500
1000
Neshoba soils

500
0

100

200

300

400

500

600

700

Vertical effective stress (kPa)

Fig. 8. Vertical effective stress versus the acoustic travel time in the unconsolidated undrained
test for air-dry remolded Neshoba soils with a confining pressure of 103.4 kPa (Data provided by
Lu et al., 2011).

17

Acoustic velocity (m/s)

1200
Pc=103.4 kPa

1100
1000

900
800
700
600
500

400
300

Neshoba soils

200
0

100

200

300

400

500

600

700

Vertical effective stress (kPa)

Fig. 9. Vertical effective stress versus the acoustic velocity in the unconsolidated undrained test
for air-dry remolded Neshoba soils with a confining pressure of 103.4 kPa (Data provided by Lu
et al., 2011).

The compressional wave velocities and travel time in the compaction test for Neshoba
soil/clay are shown in Figs. 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 direction of the unloading
path. The unloading-loading 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 unloading-reloading 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 field 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

18

4.1. Pore pressure model without unloading


From Eq. 27 and noticing e V p and n V pn , we obtain the following
equation to calculate pore pressure (p):

( pn ) 0 (t f tm )
p V V
ln
/
cZ
t tm

(32)

where t is the measured compressional transit time; tm is the transit time in the matrix; tml is
the transit time in the mudline; c is the compaction constant and can be determined from Eq. 10a,
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 Figs. 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):

t tm 0 (t f tm )e(bc ) ymax bZ ( e / n )

(33)

where b is the compaction constant in the unloading case; b = c if no unloading occurs, and b > c
in unloading case; ymax is defined in Fig. 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):

n b c
bZ c

ln

0 (t f tm )
tu 0 tm

ln

0 (t f tm )
t tm

(34)

where tu0 is the transit time at the starting point of the unloading, as defined in Fig. 10.
From Eq. 34 and noticing e V p and n V pn , we obtain the following
equation to calculate pore pressure in unloading case:

(t tm )
( pn ) b c 0 (t f tm )

/
p V V
ln
ln 0 f
bZ
c

t
u
0
m
m

19

(35)

loading

tu0

unloading

ymax

y Z ( e / n )

Fig. 10. Simplified plot from Figs. 6 and 8 showing the relationship between the vertical
effective stress and transit time in loading and unloading cases.

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 tml = 131 s/ft, tm = 73 s/ft, Png = 8.75 ppg, = 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 influx (fluid 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 tml = 131 s/ft (instead of 200
s/ft in conventional cases) to make a better pore pressure estimation in subsalt formations. Fig.
11 also demonstrates that the pore pressure calculation from the proposed method can excellently
catch the pore pressure regression, which is a common phenomenon in some areas of the Gulf of
Mexico.
Figure 12 shows another case application in pore pressure estimates in the clastic (nonsalt) formations. In this case the sonic transit time from well logs are used for pore pressure
calculations from modified Eatons method (Eq. 10) and the proposed method (Eq. 32) using the
20

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 fit to the normal
compaction line by using Eq. 10a with parameters of tml = 205 s/ft, tm = 70 s/ft, and c =
0.00026 ft-1. The calculated pore pressure from the proposed method (with Png = 8.5 ppg, = 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.

Fig. 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 figure, the gamma ray and shale base
lines are shown in the left track; the resistivity (Res) and filtered shale points of resistivity
(SHPT Res) are plotted in the second track; the sonic transit time (DT) and filtered shale points
of the transit time (SHPT DT) are shown in the third track; and the calculated pore pressures
from the filtered 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).

21

Fig. 12. Pore pressure calculation from the sonic transit time using the proposed method (Eq. 32)
and modified Eaton method (Eq. 10). In this figure, the exponential normal compaction trend
line (from Eq. 10a) and filtered shale points of sonic transit time (SHPT DT) are shown in the
left track, and the calculated pore pressures from the filtered shale transit time from the proposed
method (Pp DT) and modified Eaton method (Pp Eaton) are shown in the right track with
comparison to the measured formation pressures (DST).

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 disequilibrium
theory. This theoretical relationship of effective stress and velocity (or transit time) is verified by
field data and lab experimental results. Theoretical pore pressure-porosity model is proposed 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

22

pressure can be accurately obtained from velocity and well logging data using proposed 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. 28):

t tm 0 (t f tm )ecZ ( e / n )

(A1)

Based on Eq. A1 the following relationship between effective stress and transit time is
assumed in the unloading case (tul), as shown in Fig. 10:
tul tm Be bZ ( e / n )

(A2)

where B and b are constants.


Assuming y Z ( e / n ) as shown in Fig. 10, Eqs. A1 and A2 become:

t tm 0 (t f tm )ecy

(A3)

tul tm 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, tu0) occurs in the loading curve, we have the
following relationships, because the both unloading and loading curves (Eqs. A3 and A4)
intercept at the point (tu0, ymax). Hence, the following equations exist:

tu 0 tm 0 (t f tm )ecymax

(A5)

tu 0 tm Be bymax

(A6)

1 (t tm )
ymax ln 0 f
c
tu 0 tm

(A7)

Substituting Eq. A5 to Eq. A6, we obtain:

B 0 (t f tm )e(bc ) ymax

(A8)

Substituting Eq. A8 to Eq. A2, we obtain the transit time as a function of the effective
stress in unloading case:

t tm 0 (t f tm )e(bc ) ymax bZ ( e / n )

(A9)

23

The effective stress can be solved from Eq. A9 with substituting Eq. A7 as following:

n b c
bZ c

ln

0 (t f tm )
tu 0 tm

ln

0 (t f tm )
t tm

(A10)

References
Athy, L. F., 1930. Density, porosity, and compaction of sedimentary rocks: AAPG Bulletin, 14
(1):1-24.
Biot, M.A., 1941. General theory of three-dimensional consolidation. J. Appl. Phys. 12(1):155164.
Bird, P., 1995. Lithosphere dynamics and continental deformation. Reviews of Geophysics,
Suppl. 379-383.
Bowers, G. L., 1995. Pore pressure estimation from velocity data; accounting for overpressure
mechanisms besides undercompaction. SPE Drilling and Completions, June, 1995:89-95.
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), doi: 10.1130/GSAT01702A.1.
Dickinson, G., 1953. Geological aspects of abnormal reservoir pressures in Gulf Coast
Louisiana: AAPG Bulletin, 37(2):410-432.
Dodson, J.K., 2004. Gulf of Mexico trouble time creates major drilling expenses. Offshore,
64(1).
Dutta, N.C., 2002. Geopressure prediction using seismic data: Current status and the road ahead.
Geophysics, 67(6):20122041.
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 overpressured
sandstones: theory, observations, and applications. American J. of Science, 302:827855.
Gardner, G.H.F., Gardner, L.W. and Gregory, A.R. (1974), Formation Velocity and Density
the Diagnostic Basis for Stratigraphic Traps, Geophysics, Vol 39, No 6, pp.2085-2095.

24

Gutierrez, M.A., Braunsdorf, N.R., Couzens, B.A., 2006. Calibration and ranking of porepressure prediction models. The leading Edge, Dec., 2006:1516-1523.
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., G.F. Ulmishek, and
V.I. Slavin eds., Abnormal pressures in hydrocarbon environments: AAPG Memoir 70, p.
215246.
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):215-222.
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, 17:717-722.
Issler, D.R., 1992. A new approach to shale compaction and stratigraphic restoration, BeaufortMackenzie Basin and Mackenzie Corridor, Northern Canada. AAPG Bulletin, 76:1170-1189.
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., 2004. Integrated shared
earth model: 3D pore-pressure prediction and uncertainty analysis. The leading Edge, Jan.,
2004:52-59.
Lu, Z., Hickey, C.J., Sabatier, J.M., 2004. Effects of compaction on the acoustic velocity in soils.
Soil Sci. Soc. Am. J. 68:716.
Lu, Z., 2011. Personal communication.
Mann, D.M., Mackenzie, A.S., 1990. Prediction of pore fluid pressures in sedimentary basins.
Marine and Petroleum Geology 7, 5565.
Mondol, N.H., Bjrlykke, K., Jahren, J., Heg, K., 2007. Experimental mechanical compaction
of clay mineral aggregates-Changes in physical properties of mudstones during burial.
Marine and Petroleum Geology, 24(5), p. 289-311.
Mouchet, J.-C., Mitchell, A., 1989. Abnormal pressures while drilling. Editions TECHNIP,
Paris.

25

Nelson, H.N., Bird, K.J., 2005. Porosity-depth trends and regional uplift calculated from sonic
logs, National Reserve in Alaska. Scientific 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, 1023-1041.
Peng, S., Zhang, J., 2007. Engineering geology for underground rocks. Springer.
Rubey, W. W., and M. K. Hubbert, 1959. Role of fluid pressure in mechanics of overthrust
faulting, II. Overthrust belt in geosynclinal area of western Wyoming in light of fluidpressure hypothesis: Geological Society of America Bulletin, v.70, p.167205.
Sayers, C.M., Johnson, G.M. and Denyer, G., 2002. Predrill pore pressure prediction using
seismic data. Geophysics, 67: 1286-1292.
Schneider, J., Flemings, P.B., Dugan, B., Long, H., Germaine, J.T., 2009. Overpressure and
consolidation near the seafloor of Brazos-Trinity Basin IV, northwest deepwater Gulf of
Mexico. J. Geophys. Res., 114, B05102,
Skalle, P., Podio, A.L., 1998. Trends extracted from 1,200 Gulf Coast blowouts during 19601996. World Oil, June, 1998:67-72.
Swarbrick, R.E. and M.J. Osborne, 1998, Mechanisms that generate abnormal pressures: an
overview, in Law, B.E., G.F. Ulmishek, and V.I. Slavin eds., Abnormal pressures in
hydrocarbon environments: AAPG Memoir 70, p. 1334.
Tobin, H.J., Saffer, D.M., 2009. Elevated fluid pressure and extreme mechanical weakness of a
plate boundary thrust, Nankai Trough subduction zone. Geology, 37 (8), 679-682.
Tsuji, T., H. Tokuyama, P. Costa Pisani, and G. Moore (2008), Effective stress and pore pressure
in the Nankai accretionary prism off the Muroto Peninsula, southwestern Japan. J. Geophys.
Res., 113, B11401.
Terzaghi, K., Peck, R.B., Mesri, G., 1996. Soil Mechanics in Engineering Practice (3rd Edition).
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):5174.
Traugott, M., 1997. Pore pressure and fracture gradient determinations in deepwater. World Oil,
August, 1997.
26

Wyllie, M.R.J., Gregory, A.R., Gardner, L.W., 1956. Elastic wave velocities in heterogeneous
and porous media. Geophysics, 21:41-70.
Yardley, G.S., Swarbrick, R.E., 2000. Lateral transfer: a source of additional overpressure?
Marine and Petroleum Geology 17, 523-537.
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 wellbore
stability. Int. J. Rock Mech. Min. Sci. 40, 473483.
Zhang, J., Roegiers, J.-C., 2005. Double porosity finite element method for borehole modeling.
Rock Mech. Rock Eng. 38, 217242.
Zhang, J., Standifird, 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, 21-24
September 2008, Denver, Colorado, USA.
Zhang, J., 2011. Pore pressure prediction from well logs: Methods, modifications, and new
approaches, Earth-Sci. Rev., 108, 50-63. doi: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.

27

You might also like