Model-Based Pore-Pressure Prediction in Shales: An Example From The Gulf of Mexico, North America

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 7

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

net/publication/314264461

Model-based pore-pressure prediction in shales: An example from the Gulf of


Mexico, North America

Article  in  Geophysics · May 2017


DOI: 10.1190/geo2016-0504.1

CITATIONS READS

4 233

4 authors:

Tongcheng Han Marina Pervukhina


China University of Petroleum Commonwealth Scientific and Industrial Research Organisation, Perth, Australia
58 PUBLICATIONS   314 CITATIONS    149 PUBLICATIONS   1,145 CITATIONS   

SEE PROFILE SEE PROFILE

Michael Clennell David N. Dewhurst


The Commonwealth Scientific and Industrial Research Organisation The Commonwealth Scientific and Industrial Research Organisation
169 PUBLICATIONS   3,240 CITATIONS    218 PUBLICATIONS   4,028 CITATIONS   

SEE PROFILE SEE PROFILE

Some of the authors of this publication are also working on these related projects:

Data-constrained modelling and DCM software development View project

IGCCS-Induced-seismicity Geomechanics for Controlled CO2 Storage in the North Sea View project

All content following this page was uploaded by Tongcheng Han on 21 October 2020.

The user has requested enhancement of the downloaded file.


GEOPHYSICS, VOL. 82, NO. 3 (MAY-JUNE 2017); P. M37–M42, 5 FIGS.
10.1190/GEO2016-0504.1
Downloaded 03/08/17 to 134.7.153.132. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Model-based pore-pressure prediction in shales: An example


from the Gulf of Mexico, North America

Tongcheng Han1, Marina Pervukhina1, Michael Ben Clennell1, and David Neil Dewhurst1

mechanisms (Swarbrick and Osborne, 1998; van Ruth et al., 2004;


ABSTRACT Gutierrez et al., 2006). Accurate overpressure detection and predic-
tion play an important role in understanding the depositional and
Accurate pore-pressure interpretation and prediction play evolutional history of a sedimentary basin, hydrocarbon migration,
important roles in understanding the sedimentary history of and trapping in the reservoir (McPherson and Garven, 1999; Mug-
a basin and in reducing drilling hazards during hydrocarbon geridge and Mahmode, 2012), as well as reducing the risks of drilling
exploitation. Unlike typical reservoir rocks, in which pore hazards during hydrocarbon exploitation (Tingay et al., 2009).
pressures can be directly measured, pore pressures in shale Overpressures often occur in shales due to their extremely low
must be inferred via their state of compaction. We have used permeability that makes the pore pressure less able to equilibrate to
the shale acoustic properties to accomplish this. Shale P- and hydrostatic pressure. Unlike typical reservoir rocks, the pore pres-
S-wave velocities were first calculated using an anisotropic sure in shale cannot be directly measured but must be inferred by
differential effective medium model. This model was built quantifying their state of compaction. Commonly, within the oil and
using elastic properties of wet clay mineralogy, and the total gas industry, the state of shale compaction is interpreted using the
porosity of the shale was obtained from basic open-hole acoustic or electrical properties’ petrophysical measurement. Over-
well-log measurements. These simulated shale velocities pressure results in lower than expected sonic velocity and electrical
were systematically higher than sonic velocities observed on resistivity value at a given depth, relative to what is assumed to
a test data set from a Gulf of Mexico well, penetrating a be normally pressured petrophysical response for the same corre-
1200 m vertical section of overpressured shale. The differ- sponding depth. Methods based on these aforementioned petrophys-
ence between the modeled velocities and the sonic measure- ical anomalies are used to detect and predict the overpressure (Eaton,
ments was then evaluated to estimate the abnormal pore 1975; Bowers, 1995; Dvorkin et al., 1999; Chopra and Huffman,
pressure based on an exponential relationship between the 2006; Schneider et al., 2009). Zhang (2011) comprehensively re-
pore pressure and the ratio of measured to modeled velocity. views the various methods for overpressure prediction, and a detailed
Using a reasonable exponent, the predicted pore pressure review of pore-pressure prediction from seismic data can be found in
was shown to be in good agreement with direct pore-pres- Dutta (2002). These models, however, all depend largely on the de-
sure measurements either made in adjacent sand layers that termination of the normal velocity trend (NVT), which can either be
are thought to be in pressure equilibrium with the shales or
assumed to be linear (Sayers et al., 2002) or be obtained from generic
inferred from drilling mud weights.
or regional compaction trends (van Ruth et al., 2004; Tingay et al.,
2009). In practice, the velocities in normally compacted shales can
vary significantly with depth and from site to site depending on clay
content and clay mineralogy (although according to Pervukhina et al.
INTRODUCTION [2015], the clay mineralogy does not affect the shale velocity for the
model used in this paper).
Overpressure is the case in which the pore pressure is greater than In this study, we propose to determine the apparent NVT using a
the hydrostatic pressure. In subsurface formations, overpressure can theoretical model. We first calculate the P- and S-wave velocities of
be caused by under-compaction, hydrocarbon generation and expan- shales in the depth (depth below sea level with water depth approx-
sion, tectonic compression, and mineral transformations, among other imately 1951 m) range between 5200 and 6400 m of an overpres-

Manuscript received by the Editor 26 September 2016; revised manuscript received 6 December 2016; published online 06 March 2017.
1
CSIRO Energy Flagship, Kensington, Australia. E-mail: tongcheng.han@csiro.au; marina.pervukhina@csiro.au; ben.clennell@csiro.au; david.dewhurst@
csiro.au.
© 2017 Society of Exploration Geophysicists. All rights reserved.

M37
M38 Han et al.

sured well drilled in the Gulf of Mexico, North America (Figure 1). CPS MODEL
The velocities are simulated using the clay-plus-silt (CPS) model
proposed by Pervukhina et al. (2015) from independent wireline The CPS model is based on an experimental observation of a first-
measurements, such as the density and the neutron porosity logs. order linear correlation between the water fraction in the wet clay
The modeled shale velocities are compared with the velocities pack and the elastic coefficients of clays. The wet clay pack is a hypo-
measured by the sonic tool. The abnormal pore pressure is then pre- thetical composite material that includes all clay minerals and water
Downloaded 03/08/17 to 134.7.153.132. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

dicted based on the Eaton’s (1975) equation by using the simulated but without silt inclusions, and the linear correlation is independent
shale velocities as the apparent NVT. of the clay mineralogy and microstructure (Pervukhina et al., 2008,
2015). The anisotropic differential effective
5200 medium (DEM) model (Nishizawa, 1982) is then
used to add a silt (quartz) component into the wet
clay pack to simulate the CPS nature of shales (the
5400 composite of wet clay and silt). It is assumed that
pores in shales exist in the wet clay pack only, and
the silt particles are isolated within the clay ma-
5600
trix. Porosity and clay contents of the shales are
initially estimated from standard well-logging
Depth (m)

5800 interpretation methods, i.e., density and neutron


logs, which are the main information needed for
the execution of the CPS model.
6000 The following procedures illustrate briefly
how the shale velocities are calculated using the
CPS model. A more detailed description and
6200 workflow of the CPS model can be obtained in
Pervukhina et al. (2015).
6400 1) Calculation of shale porosity φ from the bulk
0 100 200 2 2.5 3 0 0.25 0.5 2 3 4 1 2 3
3 density log as
GR (gAPI) RHOB (g/cm ) NPHI (V/V) VP (km/s) VS (km/s)
ρ g − ρb
φ¼ ; (1)
Figure 1. Original well-log data of the example well. The bulk density (RHOB) and ρg − ρw
neutron porosity (NPHI) are used for the calculation of shale velocities in the CPS
model. The sonic P- and S-wave velocities are used for the prediction of abnormal pore
pressure, and the gamma ray (GR) is only for the indication of shales in this sand/shale where ρb is the measured bulk density, ρg is
sequence of the deepwater well, Gulf of Mexico. the grain density and is assumed to be
2.7 g∕cm3 (Skempton, 1970) for the example
well in this study, and ρw ¼ 1.03 g∕cm3 is the
1
saturating fluid density from the seawater salinity.
2) Clay quantification in the solid matrix f 0c using a neutron-den-
sity porosity crossplot (La Vigne et al., 1994; Ellis and Singer,
0.8 2007), as
φn − φ
f 0c ¼ ; (2)
y
Cl y

φcn − φcd
l 1 0% Cla
% Cla
ay
0%
Density porosity

al

0.6 where φn is the neutron porosity and φcn and φcd are the neutron
ica al 5
00
tic
re

i c

and density porosity of a 100% clay, and are assumed to be 0.37


eo

t
re
Th

eo

and 0.07 for the example well, respectively, as shown in


ret
Th
eo

Figure 2.
0.4
Th

3) Determination of the clay and silt volume fraction at each depth,


f c and f s , respectively, as

f c ¼ f 0c ð1 − φÞ and f s ¼ 1 − f c − φ: (3)
0.2
4) Computation of wet clay porosity (WCP) κ as a ratio of the vol-
ume of pores to the total volume of the wet clay constituent,

0 φ
0 0.2 0.4 0.6 0.8 1 κ¼ : (4)
fc þ φ
Neutron porosity

Figure 2. Neutron-density crossplot for the example well. A point 5) Calculation of wet clay elastic moduli ccij based on the exper-
with neutron porosity of 0.37 and density porosity of 0.07 is chosen imentally observed first-order linear dependency of clay elastic
as a 100% clay point, as indicated by the gray-filled circle. moduli on WCP (Pervukhina et al., 2008, 2015),
Shale overpressure prediction M39

ccij ¼ 2c0ij ð0.5 − κÞ; (5) PORE-PRESSURE PREDICTION


One of the most widely used methods for predicting pore pres-
where c0ij are the elastic coefficients of hypothetical clay with
sure from acoustic measurement is proposed by Eaton (1975). The
κ ¼ 0, and are shown to be equal to c011 ¼ 46.4 GPa, c033 ¼
Eaton model evaluates the difference between the measured sonic
29.9 GPa, c013 ¼ 17.9 GPa, c044 ¼ 6.7 GPa, and c066 ¼11.2GPa
transit time (Δt, inverse of compressional velocity) and the normal
(Pervukhina et al., 2008).
Downloaded 03/08/17 to 134.7.153.132. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

compaction trend for shale acoustic transit time (Δtn ), given as


6) Addition of isolated spherical silt particles into the wet clay
 χ
pack using the anisotropic DEM model (Nishizawa, 1982) to Δtn
arrive at the shale elastic moduli csij . In this study, we assume Pp ¼ OB − ðOB − Ph Þ ; (7)
Δt
that the silt particles consist of pure quartz with bulk and shear
moduli of 37 and 45 GPa, respectively (Mavko et al., 2009). where Pp is the pore pressure to be determined; OB and Ph are the
The model predicts the elastic properties of shales at the depth overburden pressure and the hydrostatic pressure, respectively; and
at which the clay volume fraction is greater than the silt fraction, χ is an exponent that can be changed so as to calibrate the method to
a criteria set for the CPS model to run for the simulation of the estimate overpressure generated by different mechanisms (Tingay
shale velocities. et al., 2009). A typical exponent value of 3.0 is found for sediments
7) Calculation of the shale P- and S-wave velocities based on the in which the overpressure has been generated by under-compaction
determined shale elastic moduli and the measured shale bulk (Mouchet and Mitchell, 1989; Hermanrud et al., 1998).
density is We have demonstrated that the CPS modeled velocity does not
sffiffiffiffiffiffiffi sffiffiffiffiffiffiffi take into account pore-pressure effects, and it can therefore be re-
cs33 cs44 garded as the shale apparent normal velocity. We now use the CPS
VP ¼ and VS ¼ : (6) modeled velocity as the apparent normal compaction trend for shale
ρb ρb
velocity to replace Δtn in equation 7, giving
 n
The predictive power of the CPS model is tested by Pervukhina V
et al. (2015), who show a good agreement between the modeled Pp ¼ OB − ðOB − Ph Þ ; (8)
V CPS
shale velocities and the velocities measured by sonic tools in a
well penetrating a 500 m vertical section of shale with normal where V and V CPS are the log-measured and CPS modeled (P- and
pore pressure. S-wave) velocities, respectively, and n is the apparent Eaton ex-
ponent.
To show the application of the CSP model to our overpressured
Figure 4 shows the pore-pressure prediction for the example well
example well, we plot in Figure 3, a comparison between the mod-
using equation 8 from the P- and S-wave velocities, respectively.
eled and the log-measured velocities. The P- and S-wave velocities
By adjusting the apparent Eaton exponent n, a value of 5.0 and
from the model prediction are systematically higher than their log-
2.3 is found to give a satisfactory fit to the measured pore pressure
measured values, which can be interpreted in terms of the depend-
by the modular formation dynamics tester (MDT) tool, for the P-
ency of the CPS model on total porosity as well as the pore-pressure
and S-wave cases, respectively. The MDT pressures were measured
effects on velocity. The CPS model only takes into account the
porosity and gross mineralogy difference of clay and quartz when 4
predicting shale velocity; it does not take into account the pore-pres-
sure effects on velocity. In fact, the CPS model calculates velocity
through the total porosity of a shale. The pore pressure, however,
3.5
affects velocity by enhancing compliant, low-aspect-ratio pores and
through its effect on total porosity, depending on the overpressure
mechanisms. On the one hand, the increase in the compliant low as-
Modeled velocity (km/s)

3
pect ratio pores caused by the increasing pore pressure (decreasing
effective stress) can reduce the elastic moduli of a rock significantly,
resulting in a profound decrease in velocity (Eberhart-Phillips et al.,
2.5
1989), and therefore the modeled shale velocity should be higher be-
cause it does not take into account this low aspect ratio pore effect.
On the other hand, the total porosity may either be enhanced or un-
changed by overpressure, and the modeled shale velocity based on 2
this total porosity is therefore expected to decrease slightly or stay the
same, respectively, on top of the increased modeled velocity due to
the low-aspect-ratio pores. This not only explains the higher than 1.5
VP
measured CPS velocity in overpressured wells, but it also implies
VS
that the modeled velocity tends to be what the shale velocity should
be if the abnormal pore pressure has not changed the total porosity. 1
1 1.5 2 2.5 3 3.5 4
The above analysis paves the way for the employment of the CPS
Well-log determined velocity (km/s)
modeled velocity as the apparent NVT (see the “Discussion” section
for the difference between the apparent NVT and the NVT) for the Figure 3. A comparison between the CPS modeled and the well-log
pore-pressure prediction. determined P- and S-wave velocities in the example well.
M40 Han et al.

in the interbedded sands and were assumed to be in pressure equi- with the conventional Eaton (1975) model. The Eaton model de-
librium with the shales. pends on knowing the normal compaction trend for shale velocity.
To correctly determine the normal compaction trend, a large number
of wells with normally compacted sediments in the nearby area need
DISCUSSION
to be analyzed (Athy, 1930; van Ruth et al., 2004; Tingay et al.,
The proposed pore-pressure prediction method based on the 2009). For new exploration regions without many wells drilled, it
Downloaded 03/08/17 to 134.7.153.132. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

CPS-modeled shale velocity offers several advantages compared is difficult to predict such a normal compaction trend, and in this case
the proposed approach will be more appropriate to
a) 5200 b) predict pore pressure once the exponent n is cali-
brated against a well with known pore pressure.
Another advantage of the new method is that
5400 the CPS model on which the method is based cal-
culates the shale velocity using porosity and the
gross mineralogy difference of clay and quartz of
5600 a specific well. This makes the modeled shale
velocity more specific to the well of interest than
Depth (m)

the universally determined normal compaction


5800 1.1 1.2 1.3 1.4 1.5 1.6 trend. Although the simulated shale velocities
are higher, they will follow the trend of the varia-
6000
tion of the measured velocity with depth (as
shown in Figures 3 and 5a). Therefore, the ratio
Hydrostatic Hydrostatic between the measured and the modeled velocity
Overburden Overburden will tend to have less scatter compared with the
6200
Mud weight Mud weight
ratio between the log-measured velocity and the
MDT MDT
From VP From VS
normal compaction trend for velocity (Figure 5b).
6400 This makes the predicted pore-pressure trend not
40 60 80 100 120 40 60 80 100 120
Pressure (MPa) Pressure (MPa)
only less scattering, i.e., more precise, but also
more accurate compared with the conventional
Figure 4. Pore-pressure prediction from CPS modeled P- and S-wave velocities, respec- Eaton (1975) model.
tively. The light-gray lines show the pore pressure with the overpressure coefficients (ra- Apart from the above-mentioned advantages,
tios between the actual pore pressure and the hydrostatic pressure) indicated by the texts. there are also uncertainties regarding the param-
The overburden pressure is derived from the integration of the density log plus the pressure
from the seawater with a density of 1.03 g∕cm3 . The Eaton exponents are 5.0 and 2.3 for eters (i.e., grain density and 100% clay neutron
the P- and S-waves, respectively. and density porosity) used in the CPS model for
the calculation of shale velocity. Grain density,
which is required to calculate the total porosity,
a) 5200 b)
Log from the bulk density well log, was assumed to
NVT be 2.7. This is a typical grain density value for
Model shale formation in the Gulf of Mexico (Meissl
5400
et al., 2010), although slightly higher values are
suggested by different authors (Krushin, 2005,
5600 2014). The grain density value can be improved
from the elemental capture spectroscopy (ECS)
Depth (m)

tool and known mineral grain densities if such


5800 information is available. The 100% clay neutron
and density porosity determine the clay content
calculated for the formation, and it is empirically
6000 chosen in this work to make the silt-clay curves
envelop most of the measured data points. The
Hydrostatic
Overburden
values 0.37 and 0.07 are used for the clay point
6200 neutron and density porosity, respectively. The
Mud weight
MDT neutron porosity of this empirically chosen
Eaton model 100% clay point is close to the average neutron
6400
2 2.5 3 3.5 4 4.5 5 40 60 80 100 120 porosity of illite, namely, 0.373 as reported by La
VP (km/s) Pressure (MPa) Vigne et al. (1994). Although smectite is the
dominant clay mineral at shallow depths in the
Figure 5. (a) Comparison between the log measured and CPS modeled P-wave velocity Gulf of Mexico, it becomes progressively illi-
and the NVT and (b) pore-pressure prediction using the Eaton equation based on the NVT
and an exponent of 3.0. The NVT is calculated as NVT ðin km∕sÞ ¼ 304.8∕Δtn, where tized with depth, driven by temperature and the
Δtn ðin μs∕ftÞ ¼ 70þ130e −0.00080381d , where d is the depth below the seafloor in meters presence of potassium (Burst, 1969; Hower et al.,
(Zhang, 2011). 1976; Elliott and Matisoff, 1996; Bauluz, 2007).
Shale overpressure prediction M41

In the formation investigated, depths are approximately 3300– almost constant (i.e., approximately 1.45). This justifies the single
4500 m below the seafloor with temperatures of 75°C–105°C, con- apparent Eaton exponent used for the prediction of pore pressure for
ditions, which usually result in a 50%–80% illitization of smectite the entire depth range.
(Hower et al., 1976). This generally validates the empirically se- In this work, the modeled shale velocity was integrated into Ea-
lected 100% clay point used in the neutron-density porosity cross- ton’s (1975) equation to predict pore pressure. There may be other
plot. For wells with ECS information measured or X-ray diffraction alternative models (Bowers, 1995; Dutta, 2002) that can be used for
Downloaded 03/08/17 to 134.7.153.132. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

data available on core samples, the clay point neutron and density the pore-pressure prediction. However, most of those models will
porosity can be better calibrated and more fully constrained. need more unknown parameters than does the Eaton equation,
As mentioned previously, the CPS modeled velocity is consid- which requires only one, i.e., the apparent Eaton exponent. Indeed,
ered to be what the velocity should be if the total porosity has not the Eaton equation is empirical and does not reveal the physical
been changed by overpressure, and it is used as the apparent NVT. relationship between the acoustic velocity and pressure (Eberhart-
This apparent NVT differs from the true NVT depending on the over- Phillips et al., 1989; Jones, 1995; Khaksar et al., 1999). Thus, the
pressure mechanisms that have distinct effects on the total porosity best way to accurately estimate pore pressure from the difference of
(Tingay et al., 2009; Zhang, 2011; Suwannasri et al., 2014). Because velocity in shales might be to conduct an experimental measure-
the CPS model calculates shale velocity through its total porosity, for ment of velocity at varying pore pressures. However, this is a time-
overpressured shales with no abnormal porosity (e.g., postdeposi- consuming experiment due to the extremely low permeability of
tional fluid expansion), the modeled velocity is greater than the mea- shales (Domnesteanu et al., 2002). The obtained relationships may
sured velocity and in this case, the apparent NVT is expected to be the not be universal, especially because shales can show frequency-de-
true NVT. On the other hand, for overpressures generated by mech- pendent behavior (Duranti et al., 2005; Fjaer et al., 2013) meaning
anisms with higher porosity (e.g., under-compaction), the higher that laboratory, wireline log, and seismic velocities may be different
porosity will make the difference between the modeled velocity for the same shale because they are typically obtained at ultrasonic
and the log measurement smaller, and therefore the apparent NVT (approximately 0.5–1.0 MHz), sonic (approximately 1–10 kHz),
is slower than the true NVT. This is confirmed by the results given and seismic (approximately 1–50 Hz) frequencies, respectively.
in Figure 5a for the example well with overpressure generated by Nevertheless, given the reasonable agreement between the predicted
and the measured pore pressure presented in this context, the Eaton
under-compaction, which is indicated by fitting the measured pore
equation based on the CPS modeled velocity can be used as a prac-
pressure using the Eaton equation with an exponent of 3.0 (Fig-
tical method for accurate pore-pressure prediction.
ure 5b), as demonstrated by Ebrom et al. (2003) for most under-com-
The proposed method predicts pore pressure through sonic logs
paction wells in the Gulf of Mexico. The difference between the
and the modeled shale velocity from density and neutron porosity
modeled shale velocity (i.e., the apparent NVT) and the true NVT
logs. All of these measurements are available from logging-while-
can be used to distinguish under-compaction from other overpressure
drilling (LWD) tools. Therefore, it is possible that the pore pressure
mechanisms that do not cause abnormal porosity.
can be predicted in real-time based on LWD data, provided that the
The apparent Eaton exponent n was determined to be 5.0 and 2.3
grain density and 100% clay neutron and density porosity for the
for P- and S-wave velocities, respectively, whereas the original Eaton
calculation of the shale velocity and the Eaton exponent for the es-
exponent χ used 3.0 for the P-wave velocity to fit the measured pore
timation of pore pressure are fully calibrated. Application of the
pressure. The difference between the two exponents arises due to the
current approach to real-time pore-pressure prediction based on
different NVTs, which in turn is a result of the under-compaction LWD data will be a topic for future study.
mechanism that enhances the shale porosity. As discussed above,
the enhanced porosity caused by under-compaction reconciles part
of the difference between the expected (i.e., the true NVT) and CONCLUSION
the measured velocity, making the apparent NVT slower than the true We have shown that the CPS-modeled velocities are systemati-
NVT. Hence, a higher exponent is needed to amplify the reduced cally higher than the log-measured velocities for 1200 m of over-
velocity difference. This confirms that a larger value of the exponent pressured shales in a deepwater well, Gulf of Mexico. The CPS
indicates the insensitivity of the velocities to changes in pore pressure model works through total porosity and gross mineralogy variables
(Ebrom et al., 2003). In this sense, the S-wave velocity is more sen- of clay- and silt-sized quartz of a shale, and it does not take into
sitive to the variation in pore pressure because the apparent Eaton account the pore-pressure effects; the modeled velocities are there-
exponent obtained for the S-wave is smaller than that for the P-wave. fore considered to be what the velocities should be if no abnormal
For overpressures with no abnormal porosity, the apparent NVT porosity is associated with the overpressure. Using the CPS-mod-
should coincide with the true NVT, and therefore the apparent Eaton eled velocities as the apparent normal compaction trend for shale
exponent n is expected to be equal to the original Eaton exponent χ. velocity, the abnormal pore pressure is ultimately predicted based
In addition to the different overpressure mechanisms, the degree on an exponential relationship between the pore pressure and the
of overpressure may also affects the apparent Eaton exponent. It is ratio of measured to modeled velocity. The pore pressure predicted
understandable that varying degrees of overpressure will lead to a using this new approach shows good agreement with direct pore-
difference between the measured and modeled velocities, resulting pressure measurements in the example well.
in changing apparent Eaton exponents. Therefore, a single value of
the exponent is insufficient to predict the pore pressure for overpres-
ACKNOWLEDGMENT
sures generated by the same mechanisms or for the entire section of
a single well (Sarker and Batzle, 2008). Fortunately, as shown in The authors would like to thank J. Krushin for kindly providing
Figure 4, the overpressure coefficient (defined as the ratio between the log data presented in this work and for his helpful comments on
the actual and hydrostatic pore pressure) for our example well is the manuscript.
M42 Han et al.

REFERENCES Krushin, J. T., 2014, Quantifying shale pore pressure by modeling the con-
trols on compaction and porosity: Interpretation, 2, no. 1, SB79–SB88,
doi: 10.1190/INT-2013-0110.1.
Athy, L. F., 1930, Density, porosity, and compaction of sedimentary rocks: La Vigne, J., M. Herron, and R. Hertzog, 1994, Density-neutron interpre-
AAPG Bulletin, 14, 1–24. tation in shaly sands: Presented at the SPWLA 35th Annual Logging
Bauluz, B., 2007, Illitization processes: Series of dioctahedral clays and Symposium, paper EEE.
mechanisms of formation, in F. Nieto, and J. Millan, eds., Diagenesis and Mavko, G., T. Mukerji, and J. Dvorkin, 2009, The rock physics handbook:
low-temperature metamorphism. Theory, methods and regional aspects: Cambridge University Press.
Downloaded 03/08/17 to 134.7.153.132. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Seminarios de la Sociedad Española de Mineralogía, Sociedad Española McPherson, B., and G. Garven, 1999, Hydrodynamics and overpressure
de Mineralogía, 31–39. mechanisms in the Sacramento basin, California: American Journal of
Bowers, G. L., 1995, Pore pressure estimation from velocity data: Account- Science, 299, 429–466, doi: 10.2475/ajs.299.6.429.
ing for overpressure mechanisms besides undercompaction: SPE Drilling Meissl, S., J. Behrmann, and J. H. Behrmann, 2010, Data report: Preliminary
& Completion, 10, 89–95, doi: 10.2118/27488-PA. assessment of Pleistocene sediment strength in the Ursa Basin (Gulf of
Burst, J. F., 1969, Diagenesis of Gulf Coast clayey sediments and its possible Mexico continental slope) from triaxial and ring shear test data, in P.
relation to petroleum migration: AAPG Bulletin, 53, 73–93. B. Flemings, J. H. Behrmann, and C. M. John, and , the Expedition 308
Chopra, S., and A. Huffman, 2006, Velocity determination for pore-pressure Scientists, eds., Proceedings of IODP, 308: Integrated Ocean Drilling Pro-
prediction: The Leading Edge, 25, 1502–1515, doi: 10.1190/1.2405336. gram Management International Inc., 1–18.
Domnesteanu, P., C. McCann, and J. Sothcott, 2002, Velocity anisotropy and Mouchet, J. P., and A. Mitchell, 1989, Abnormal pressures while drilling:
attenuation of shale in under- and overpressured conditions: Geophysical Editions Technip.
Prospecting, 50, 487–503. Muggeridge, A., and H. Mahmode, 2012, Hydrodynamic aquifer or reser-
Duranti, L., R. Ewy, and R. Hofmann, 2005, Dispersive and attenuative voir compartmentalization?: AAPG Bulletin, 96, 315–336, doi: 10.1306/
nature of shales: Multiscale and multifrequency observations: 75th An- 06141110169.
nual International Meeting, SEG, Expanded Abstracts, 1577–1580. Nishizawa, O., 1982, Seismic velocity anisotropy in a medium containing
Dutta, N. C., 2002, Geopressure prediction using seismic data: Current status oriented cracks: Transversely isotropic case: Journal of Physics of the
and the road ahead: Geophysics, 67, 2012–2041, doi: 10.1190/1.1527101. Earth, 30, 331–347, doi: 10.4294/jpe1952.30.331.
Dvorkin, J., G. Mavko, and A. Nur, 1999, Overpressure detection from com- Pervukhina, M., D. N. Dewhurst, B. Gurevich, U. Kuila, A. F. Siggins, M.
pressional- and shear-wave data: Geophysical Research Letters, 26, 3417– Raven, and H. M. Hordgård-Bolås, 2008, Stress-dependent elastic proper-
3420, doi: 10.1029/1999GL008382. ties of shales: Measurement and modeling: The Leading Edge, 27, 772–
Eaton, B. A., 1975, The equation for geopressure prediction from well logs: 779, doi: 10.1190/1.2944162.
SPE of AIME, Paper 5544. Pervukhina, M., P. Golodoniuc, B. Gurevich, M. B. Clennell, D. N.
Eberhart-Phillips, D., D. H. Han, and M. D. Zoback, 1989, Empirical rela- Dewhurst, and H. M. Hordgård-Bolås, 2015, Prediction of sonic velocities
tionships among seismic velocity, effective pressure, porosity, and clay in shale from porosity and clay fraction obtained from logs: A North Sea
content in sandstone: Geophysics, 54, 82–89, doi: 10.1190/1.1442580. well case study: Geophysics, 80, no. 1, D1–D10, doi: 10.1190/
Ebrom, D., P. Heppard, M. Mueller, and L. Thomsen, 2003, Pore pressure geo2014-0044.1.
prediction from S-wave, C-wave, and P-wave velocities: 73rd Annual Inter- Sarker, R., and M. Batzle, 2008, Effective stress coefficient in shales and its
national Meeting, SEG, Expanded Abstracts, 1370–1373. applicability to Eaton’s equation: The Leading Edge, 27, 798–804, doi: 10
Elliott, W. C., and G. Matisoff, 1996, Evaluation of kinetic models for the .1190/1.2944165.
smectite to illite transformation: Clay and Clay Minerals, 44, 77–87, doi: Sayers, C. M., G. M. Johnson, and G. Denyer, 2002, Predrill pore pressure
10.1346/CCMN. prediction using seismic data: Geophysics, 67, 1286–1292, doi: 10.1190/
Ellis, D. V., and J. M. Singer, 2007, Well logging for earth scientists: 1.1500391.
Springer. Schneider, J., P. B. Flemings, B. Dugan, H. Long, and J. T. Germaine, 2009,
Fjaer, E., A. M. Stroisz, and R. M. Holt, 2013, Elastic dispersion derived from Overpressure and consolidation near the seafloor of Brazos-Trinity Basin
a combination of static and dynamic measurements: Rock Mechanics and IV, northwest deepwater Gulf of Mexico: Journal of Geophysical Re-
Rock Engineering, 46, 611–618, doi: 10.1007/s00603-013-0385-8. search, 114, B05102.
Gutierrez, M. A., N. R. Braunsdore, and B. A. Couzens, 2006, Calibration Skempton, A. W., 1970, The consolidation of clays by gravitational com-
and ranking of porepressure prediction models: The Leading Edge, 25, paction: Quarterly Journal of the Geological Society of London, 125,
1516–1523, doi: 10.1190/1.2405337. 373–411, doi: 10.1144/gsjgs.125.1.0373.
Hermanrud, C., L. Wensas, G. M. G. Teige, H. M. Nordgård Bolås, S. Han- Suwannasri, K., W. Promrak, S. Utitsan, V. Chaisomboonpan, R. J. Groot,
sen, and E. Vik, 1998, Shale porosities from well logs on Haltenbanken H. I. Sognnes, and C. K. Morley, 2014, Reducing the variation of Eaton’s
(offshore mid-Norway) show no influence of overpressuring, in B. E. exponent for overpressure prediction in a basin affected by multiple over-
Law, G. F. Ulmishek, and V. I. Slavin, eds., Abnormal pressures in hydro- pressure mechanisms: Interpretation, 2, no. 1, SB57–SB68, doi: 10.1190/
carbon environments: AAPG Memoirs, 65–85. INT-2013-0100.1.
Hower, J., E. V. Eslinger, M. Hower, and E. A. Perry, 1976, Mechanism of Swarbrick, R. E., and M. J. Osborne, 1998, Mechanisms that generate ab-
burial metamorphism of argillaceous sediment. 1: Mineralogical and chemi- normal pressures: An overview, in B. E. Law, G. F. Ulmishek, and V. I.
cal evidence: Geological Society of America Bulletin, 87, 725–737, doi: 10 Slavin, eds., Abnormal pressures in hydrocarbon environments: AAPG,
.1130/0016-7606(1976)87<725:MOBMOA>2.0.CO;2. 13–34.
Jones, S. M., 1995, Velocities and quality factors of sedimentary rocks at Tingay, M. R. P., R. R. Hillis, R. E. Swarbrick, C. K. Morley, and A. R.
low and high effective pressures: Geophysical Journal International, 123, Damit, 2009, Origin of overpressure and pore-pressure prediction in the
774–780, doi: 10.1111/j.1365-246X.1995.tb06889.x. Baram province, Brunei: AAPG Bulletin, 93, 51–74, doi: 10.1306/
Khaksar, A., C. M. Griffiths, and C. McCann, 1999, Compressional and 08080808016.
shear-wave velocities as a function of confining stress in dry sandstones: van Ruth, P., R. Hillis, and P. Tingate, 2004, The origin of overpressure in
Geophysical Prospecting, 47, 487–508, doi: 10.1046/j.1365-2478.1999 the Carnarvon Basin, Western Australia: Implications for pore pressure
.00146.x. prediction: Petroleum Geoscience, 10, 247–257, doi: 10.1144/1354-
Krushin, J. T., 2005, Quantifying shale porosity — A thermodynamically 079302-562.
based, predictive model which includes the effects of mechanical compac- Zhang, J., 2011, Pore pressure prediction from well logs: Methods, modi-
tion, temperature, mineralogy, and chemical diagenesis: Gulf Coast As- fications, and new approaches: Earth-Science Reviews, 108, 50–63, doi:
sociation of Geological Societies Transactions, 55, 401–414. 10.1016/j.earscirev.2011.06.001.

View publication stats

You might also like