You are on page 1of 8

SPE 96897

Mapping Fluid Flow in a Reservoir Using Tiltmeter-Based Surface-


Deformation Measurements
J. Du, SPE, Pinnacle Technologies Inc.; S.J. Brissenden and P. McGillivray, Shell Canada Ltd.; S. Bourne and P. Hofstra,
Shell Intl. E&P; and E.J. Davis, SPE, W.H. Roadarmel, SPE, S.L. Wolhart, and C.A. Wright, SPE, Pinnacle
Technologies Inc.

Copyright 2005, Society of Petroleum Engineers


analytical analysis. The most common monitoring techniques
This paper was prepared for presentation at the 2005 SPE Annual Technical Conference and used in oil and gas fields are:
Exhibition held in Dallas, Texas, U.S.A., 9 – 12 October 2005.
1. Optical instrument leveling surveys or Global
This paper was selected for presentation by an SPE Program Committee following review of
information contained in a proposal submitted by the author(s). Contents of the paper, as
Positioning System (GPS) surveys1. These are
presented, have not been reviewed by the Society of Petroleum Engineers and are subject to conducted continuously or periodically to determine
correction by the author(s). The material, as presented, does not necessarily reflect any
position of the Society of Petroleum Engineers, its officers, or members. Papers presented at changes in position of monuments across the field.
SPE meetings are subject to publication review by Editorial Committees of the Society of
Petroleum Engineers. Electronic reproduction, distribution, or storage of any part of this paper
2. Interferometric Synthetic Aperture Radar (InSAR) 1.
for commercial purposes without the written consent of the Society of Petroleum Engineers is This enables mapping of surface displacement along
prohibited. Permission to reproduce in print is restricted to a proposal of not more than 300
words; illustrations may not be copied. The proposal must contain conspicuous the satellite line of sight over large areas.
acknowledgment of where and by whom the paper was presented. Write Librarian, SPE, P.O.
Box 833836, Richardson, TX 75083-3836, U.S.A., fax 01-972-952-9435.
3. Tiltmeter-based surface deformation monitoring2,3.
High precision tiltmeters are placed near the earth’s
Abstract surface to measure the displacement gradient (tilt)
Surface deformation measurements have been used for years induced by field operations such as fluid injection
in oilfields to monitor production, waterflooding, waste and production.
injection, steam flooding, and Cyclic Steam Stimulation
(CSS). They have been proven to be a very effective way to Each technique has advantages and disadvantages, and in
monitor the field operations and save money for operators some cases two or even all three can be used in combination to
wishing to avoid unwanted surface breeches, casing failures get the necessary combination of precision, spatial coverage
and excessive subsidence due to production. This paper and temporal resolution. In the case history shown here, only
demonstrates that more information can be extracted from tiltmeter data is used and the inversion process calculates and
surface deformation measurements by inverting the surface compares measured and theoretical tilt, but only minor
deformation for the volumetric deformation at the reservoir changes are needed to perform the same calculations with
level, so the areal distribution of volumetric deformation can displacement.
be identified. First, a poroelastic model is presented to After acquiring the surface deformation data, how would
calculate the deformation due to the volumetric change in the we use it to image the reservoir process? Surface
reservoir. Then, a linear geophysical model is formulated to displacement and tilt have been used to monitor injected steam
invert for the reservoir volumetric deformation from the migration using the dislocation model4 and to study formation
measured surface deformation (or tilt). Constraints are added shear as well as dilation5. A dislocation model is usually used
into the procedure as necessary to better resolve the inversion to model fractures and faults where there are clear
problem. After each inversion, the theoretical surface displacement discontinuities. Segall’s poroelastic model6 was
deformation (displacement, tilt, reservoir compaction and applied, using only the Green’s function due to the center of
volumetric strain) can be calculated from the inverted dilatation source7 for observation points located on the
volumetric deformation distribution which best fits the surface. This was inverted for subsurface volume changes in a
measured deformation data (or tilt) at the surface. The fractured zone at the depth of 30 m (100 ft) during a pump test
technique of mapping fluid flow using surface deformation and during a grout injection into porous gravel at an extremely
was applied to real data from a cyclic steam injection project. shallow depth of 3-4 m8. A poroelastic model9, with the
solution written in terms of pressure changes in the reservoir,
Introduction was adopted to invert the surface displacement for the pressure
Through the decades, many oil companies and individual changes in the reservoir10. To model the deformation due to
researchers have studied reservoir compaction and its the fluid flow through the porous media, where the
associated surface subsidence. Two techniques are used: deformation is dominated by the matrix flow, the dislocation
forward modeling for prediction and direct measurements (or model might not be appropriate. In this paper, the poroelastic
monitoring). The forward modeling includes numerical model6 is adopted to be the forward model to predict the
analysis using Finite Element Method and analytical or semi- deformation (displacement or tilt) due to the changes in the
2 SPE 96897

volumetric deformation (volumetric strain) at the reservoir


level.
After constructing the poroelastic forward model, an
inversion model incorporating a penalty function method11
with a smoothness constraint and with sign constraints is
formulated to invert the tilt data measured on the surface for
the volumetric strain source at the reservoir level. When the
best-fit solution is found with the forward model, the total
volumetric strain is calculated, in addition to the reservoir
compaction.
In this paper, we will first give a description of the Peace
River field and the tiltmeter-based monitoring background.
Then we will present the forward model and inversion
formulation, followed by an application to the Cyclic Steam
Stimulation process monitoring on one pad of horizontal wells
in the Peace River field.

Peace River Field Description and Monitoring


Background
The Peace River field is located at northern Alberta as shown
in Figure 1. Shell Canada’s Peace River development12 covers
an area of approximately 37,000 hectares with an estimated of
8-10 billion barrels of ~8 API bitumen in place. The reservoir Figure 1. Map of Alberta showing the location of the Peace River
lies at a depth of about 600 m, and has 25 m of sandstone12 Field.
with a porosity ranging from 25% to 30%, and permeability
from 200 md to 1650 md13.
Shell Canada has developed several pads of horizontal
wells in Peace River. Tiltmeters have been used at Pad-A and
Pad-B to monitor the cyclic steam process. There are ten
wells in Pad-A running North-South and eight wells in Pad-B
running East-West (see Figure 2). Each well has three lateral
extensions approximately 500 meters in length. A total of
fifty tiltmeters are used to monitor half of Pad-A and seventy
tiltmeters are used to monitor all of Pad-B.
The wells in both pads are under cyclic steam injection. Pad-B
During the injection cycle, steam is injected at a rate up to 800
tons per day cold water equivalent (CWE) for 2-3 months14.
The wells are then put on production until production decline Pad-A
and economics dictate the need for more steam. For example,
the first cycle for Pad-A started in late September of 2002 with
injection continuing until mid-December. Production began in
the latter half of December and continued until June 2003.
The objectives of the long-term tiltmeter-based reservoir
monitoring at Peace River are:
1) Identify and characterize any hydraulic fracturing that
is contributing to the injection mechanism into the
reservoir.
2) Identify the measured surface deformation (tilt) for the
volumetric changes at the reservoir level. Tiltmeter- Figure 2. Map showing the location of pads in the Peace River
based deformation monitoring provides a direct Field from SPE 83505.
measurement of the areal coverage of the reservoir
dilation and reservoir volume reduction. Hydraulic fracture growth (indeed, all dislocation modes:
opening, shear, and twisting) produces a very unique pattern in
By achieving the objectives listed above, deformation- the deformation fields they generate. This unique pattern
based, reservoir-level monitoring has proven helpful in betrays the presence of fracture growth and is frequently used
ongoing efforts to optimize such variables as the length of as a diagnostic tool for the determination of fracture growth
well laterals, injection rates, lateral spacing and cycle times. parameters such as orientation and in some instances depth,
asymmetry and shape. Deformation caused by vertical and
sub-vertical fractures is thus easily discernible from that
SPE 96897 3

caused by volumetric changes at reservoir level by examining show as significant events on the tiltmeters but do not affect
both the shape of the deforming regions and the temporal the long-term tilt trend used for deformation analysis.
characteristics of the tilt responses being measured. An Figure 5 shows a conceptual diagram of a one-dimensional
explanation of this technique is provided in Wright’s paper15. subsidence array. The tilt measurements define the slope of
For this Peace River example no hydraulic fracture growth the surface movement. Vertical displacement contours are
was detected within the monitored area and the potential determined by integrating the slope measurements from
impact of such growth on reservoir-level processes is not several tiltmeters. Since a constant of integration is required,
considered here. one location in the array can be defined as the “zero” point – a
In this paper, we focus on the second objective of the steady reference against which all other elevation changes are
monitoring, which is providing the reservoir deformation measured. To obtain a true “zero” reference, the array can
(volumetric strain) during the Cyclic Steam Stimulation either be extended into a quiet area or the zero point can be
cycles. calibrated with an accurate GPS or level survey reading. The
results are generally presented in a video format; Figure 6
Surface Tiltmeter Monitoring shows subsidence in a field over a one-year period measured
Reservoir monitoring involves installing an array of tiltmeters from an array of tiltmeters.
over the zone to be mapped. The data can be sent in real-time
via wire, radio or satellite to another location for immediate
analysis, or the data can be collected and analyzed
periodically. This is a robust mapping technique that is
relatively independent of rock lithology and formation San Salvador
13-Jan 10:33 (CST)
properties. There are currently thirteen permanent arrays in 7.6 Magnitude

place for reservoir monitoring with approximately 600 India


25-Jan 21:16 (CST)
tiltmeters total installed. 7.8 Magnitude

X Tilt Signal
Y Tilt Signal
Figure 4. X-Y tilt data for one tiltmeter site in eastern Texas
showing large tilts corresponding to earthquakes.

A total of 0.52 uR
measured subsidence
over one week.

Figure 3. Signals from a surface tiltmeter over one month shows


subsidence induced movement as well as twice-daily earthtides.

Tiltmeters measure the change of the displacement Figure 5. Example of a tiltmeter array designed to detect a
gradient, which is far easier to measure than the displacement subsidence slope.
itself. Figure 3 shows tilt measurements from one instrument
over a one-month period. Twice-daily earthtide movements
are evident in the data, but do not impact the measured tilt
over a period of many days. The high precision of the tilt
measurement results in a high resolution of elevation change
after integration and vertical surface displacements as small as
0.0005 cm are considered to be large signals.
In addition to the earthtides other physical phenomena can
be seen in tilt data. Figure 4 shows tilt from one site located in
eastern Texas. Large events are observed on January 13, 2001
(magnitude 7.6 earthquake in San Salvador) and January 25,
2001 (magnitude 7.8 earthquake in India). These earthquakes
4 SPE 96897

In the forward modeling, the reservoir, having a given


depth and thickness, is divided into many reservoir blocks.
Each reservoir block has a constant volumetric strain ∆v ,
which means that we assume the volumetric deformation in
each reservoir block is uniform inside of the block. The

Subsidence (in)
theoretical surface deformation (tilt) on the surface is
calculated by summing the tilt due to the volumetric changes
in all the reservoir blocks together. We should point out that
tiltmeters at the surface only measure the surface deformation.
Whether the deformation is caused by the fluid migration in
the reservoir, thermal expansion in the reservoir or some other
process, tiltmeters would not be able to differentiate between
them.

Inversion Formulation
Figure 6. Two-dimensional map of subsidence-induced earth
movement over a one-year period. To formulate the inversion model to infer the volumetric
changes in the reservoir from the surface deformation
measurements, the reservoir is divided into many rectangular
While the concept is simple, the magnitudes of the induced blocks. Assuming constant volumetric change within each
surface deformations are quite small and require highly block, Eq. (1) and (2) are numerically integrated over each
sensitive measurement. These minute tilts are measured with block to obtain displacement and tilt. The surface deformation
highly sensitive tiltmeters that operate on the same principle as is calculated by superimposing the deformation due to the
a carpenter’s level. Tiltmeters are metal cylinders roughly 76 volumetric change of every block.
centimeters long and 5 centimeters in diameter, which The first step in the inversion process is to build the
measure their own tilt on two orthogonal axes. As the discrete Green’s function G using the forward model. The
instrument tilts, a gas bubble contained within a conductive- discrete Green’s function G is also called the data kernel in the
liquid-filled glass casing moves to maintain its alignment with general geophysical inversion. It relates the deformation (tilt
the local gravity vector. Precision electronics detect changes on the surface) to the deformation source (volume changes in
in resistivity between electrodes mounted on the glass sensor the reservoir). Each column of matrix G is the tilt at the
that are caused by motion of the gas bubble. The latest surface tiltmeter sites due to a unit volumetric strain of one
generation of high-resolution tiltmeters15 can detect tilts of less reservoir block. The inverse problem we are solving is the
than one nanoradian. Each tiltmeter site has an instrument following:
surrounded by sand within pipe (8 to 23 centimeter diameter) d = Gs . (4)
that is cemented in a relatively shallow (5 to 12 meters deep) where d is the measured tilt data on the surface location. It is
borehole. a vector with the dimension (2n) of twice of the number of
surface tiltmeter sites (n). The unknown combined reservoir
Forward Model volumetric strain ( ∆v ) we are solving for is s. Its dimension
The displacement due to the volumetric changes6,8 is (m) is the number of reservoir blocks. Each column of matrix
calculated as the following: G is the tilt at the surface tiltmeter sites due to a unit
K
u i ( x) = u ∫ (B∆v )g i ( x, ς )dVζ (1) volumetric strain of one reservoir block. After building the
µ V discrete matrix G, we are ready to formulate the inversion
where µ is the shear modulus. Ku is the undrained bulk problem.
modulus. B is Skempton's pore pressure coefficient. ∆v is the The inverse problem shown in Eq. (4) is usually ill-posed
and requires some special action in order to get a robust
ratio of the mass change per unit of bulk volume ( ∆m ) to the
solution. In this paper, two inversion methods are adapted.10,11
fluid density at the reference state ( ρ 0 ). g i (x, ς ) is the The first is the Penalty Function with Smoothness (PFS).
Green’s function for a center of dilation source in a half- The penalty function is written as:
space.4 2
F ( s ) = Gs − d + β 2 Hs
2
(5)
Tilt (Ti, i can be 1 or 2, which represent the two tilt
where G is the discrete Green’s function, d is a vector for the
components), which is displacement gradient, is calculated as:
measured tilt at the surface sites, s is a vector for the combined
∂u ( x ) K u
Ti ( x ) = i (B∆v ) ∂g i ( x, ς )dVς (2) volumetric strain in the reservoir blocks, β2 is the smoothness
µ ∫V
=
∂x3 ∂x 3 factor and H is the finite difference approximation of the
The first term ( B∆v ) inside of the integration in Eq. (1) and Laplace operator.
Eq. (2) is actually a volumetric strain term. The total After minimizing the error function in equation (5), the
volumetric strain6 is calculated as: inverted volumetric strain distribution in the reservoir is:
σ kk (3) ∆vest = s est = (G T G + β 2 H T H ) −1 (G T d ) (6)
ε kk = + B∆v
3K u
SPE 96897 5

The second is the Penalty Function with Smoothness and After performing the inversion, the best-fit inverted
Positive/Negative Constraints11 (PF+S+P/N). The system of volumetric distribution is plotted in Figure 8, which shows
equations for the geophysical inversion becomes the following that two patches of volumetric deformation could be recovered
 G  d  using the inversion of the synthetic surface tilt data set.
 2  s =   (7)
β H  0
subject to si ≥ 0 for injection si ≤ 0 for production cycle.
The Eq. (7) is solved using the non-negative least-squares

Inverted Volumetric Strain ∆ v


algorithm.16
In order to obtain the best-fit solution, the inversion is
performed for a series of smoothness factors; then the trade-
off curve is plotted showing the error of fitting the measured
tilt data versus the roughness of the volumetric strain
distribution for different smoothness factors. The best
smoothness factor is chosen to be at the knee point of the
trade-off curve, which means that the best solution has the
smallest error and reasonably smoothing volumetric strain
distribution at the reservoir level.
After obtaining the best-fit volumetric changes in the
reservoir blocks, the displacement, tilt, vertical strain and total
volumetric strain could be calculated using Eq. (1), Eq. (2) and Figure 8. Inverted volumetric strain distribution for the
Eq. (8). synthetic data inversion example.
K ∂g ( x, ς )
ε kk = u ∫ ∆vbest − fit k dVς (8)
µ V ∂x k Mapping Results for Surface and Reservoir
Deformation in Peace River Pads
Inversion Validation using Synthetic Data Results are presented for the second cycle from Pad-B where
The inversion technique8,10,11,17 adopted in this paper has been the average reservoir depth is approximately 565 m and the
successfully used to invert for distribution of fluid volume reservoir thickness is approximately 25 m. The reservoir is
changes, fracture opening, fault slip and pressure changes. divided into 100 m by 100 m reservoir blocks with constant
The validation of the inversion technique17 has been studied thickness of 25 m. Each reservoir block is assumed to have
before and noise has been added into the synthetic data10 to one constant volumetric strain, which is uniform within each
study its impact on the inversion results. In this paper, one block. The inversion process is used to obtain the volumetric
simple synthetic example is shown using the reservoir strain distribution in all the reservoir blocks. The second
geometry and the surface tiltmeter array from the real data injection cycle for Pad-B started on Jan. 16, 2005 and lasted
study. Synthetic tilt data at the tiltmeter sites in the Pad-B about two and half months. It has been under production since
array on the surface is generated using an assumed volumetric the end of March 2005.
strain distribution at the reservoir level. The input volumetric
strain distribution is shown in Figure 7, where there are two
separate patches (Northwest and Southeast) of volumetric
deformation.
01/16-02/12 02/12-03/03 03/03-03/31
Input Volumetric Strain ∆ v

Figure 9. Sample tilt from one site in the Pad-B monitoring


array.
Injection Cycle (01/16/05-03/31/05)
Based on the changes in the tilt data observed during the
course of the injection cycle, the analysis was divided into
three periods; Period-I (01/16/05-02/12/05), Period-II
(02/12/05-03/03/05) and Period-III (03/03/05-03/31/05) as
shown in Figure 9. We can see from the Figure 10 that the
deformation is concentrated at the heel of the northern laterals
Figure 7. Input volumetric strain distribution for the synthetic during the first 27 days of injection; the resulting deformation
data inversion example. is relatively small. With the continuing injection, the
6 SPE 96897

magnitude of the deformation increased, but the deformation


center is still located at the heel of the northern laterals (Figure
11). During the last 28 days of injection (Figure 12), the
deformation center was clearly shifted to the northwestern
laterals. Another observation from Figure 10, Figure 11 and
Figure 12 is that less volumetric deformation is observed in
the area of the southernmost six laterals. This is consistent
with the injection data since the two wells to the very south
did not receive steam during this cycle. The volumetric
deformation distribution plots for three different periods show
how the distribution of volumetric deformation changes in the
reservoir vs time.

Figure 12. Inverted volumetric strain distribution from the


measured tilt data on the surface for Injection Period-III
(03/03/05-03/31/05).

Figure 10. Inverted volumetric strain distribution from the


measured tilt data on the surface for Injection Period-I (01/16/05-
02/12/05).

Figure 13. Inverted volumetric strain distribution from the


measured tilt data on the surface for Production Period-I
(03/31/05-05/09/05).

Figure 11. Inverted volumetric strain distribution from the


measured tilt data on the surface for Injection Period-II
(02/12/05-03/03/05).

Figure 14. Inverted volumetric strain distribution from the


Production Cycle (03/31/05-06/07/05) measured tilt data on the surface for Production Period-II
During the production cycle, the wells are opened and (05/09/05-06/07/05).
production starts. The volumetric distribution is plotted in
Figure 13 and Figure 14 for the first 39 days and next 29 days As an example of how to choose the best-fit solution
of the production respectively. We can see that most of the among a series of solutions with different smoothness factors,
production-related subsidence is located near the northwestern the trade-off curve for the Period-II production is plotted in
laterals of the pad coinciding with the reservoir dilation Figure 15. The plot shows that the error of fitting the
observed during the injection cycle. measured tilt data on the surface increases as the smoothness
factor is increased; the inverted volumetric distribution is
smoother. The optimal smoothness factor is 1.e4, which is
SPE 96897 7

located at the knee of the trade-off curve. For any smoothness


factors smaller than the optimal value, the inverted volumetric Nomenclature
distribution will be too rough. On the other hand, for those B = Skempton’s coefficient
solutions with larger smoothness factors the optimal value d = measured tilt vector (microradians) or
does not fit the measured data well. displacement vector, m(ft)
1 F(s) = objective function with penalty function,
5.E6
1.E6 microradians2 for tilt data d or m2 for
displacement data d
5.E5 G = discrete Green's function matrix, dimensionless
0.8
gi = displacement Green’s function due to a center of
dilatation source, m-2(ft-2)
H =finite difference approximation of Laplacian
0.6 operator
1.E5
Error

1.E4 Ku = bulk modulus of whole rock measured under


undrained condition, Mpa (psi)
0.4 1000 500 β2=100 m = mass of pore fluid per unit bulk volume, kg m-3
(lbft-3)
Τι = tilt, microradians
0.2 ui = displacement in xi direction, m(ft)
∆v = volumetric strain, dimensionless x1
x1 = Cartesian coordinates, m (ft)
x2 = Cartesian coordinates, m (ft)
0
x3 = Cartesian coordinates, m (ft)
0.0E+00 5.0E-05 1.0E-04 1.5E-04
β2 = smoothness factor, dimensionless
Roughness
δij =Kronecker delta
Figure 15. The trade-off curve for production Period-II showing εij = strains, dimensionless
the error of fitting the measured tilt data vs the roughness of the µ = shear modulus, MPa (psi)
inverted volumetric distribution. ρ = fluid density, kg m-3 (lbft-3)
One observation for the second injection cycle for Pad-B is ρ0 = fluid density in the reference state, kg m-3(lbft-3)
the apparently asymmetrical distribution of volume changes,
where there is more volumetric deformation at the western Acknowledgements
laterals than at the eastern laterals. There are several factors The authors would like to thank Shell Canada Ltd. for
that could contribute to the asymmetric coverage, and the permission to publish the presented data and also thank Zeno
deformation data as well as other geological, monitoring and Philip, Dmitriy Astakhov and Scott Marsic for their useful
surveillance data could be used to better understand what are input.
the controlling factors.
References
1. Brink, J.L., Patzek, T.W., Silin, D.B. and Fielding, E.J., “Lost
Conclusions
Hills Field Trial – Incorporating New Technology for Reservoir
The following conclusions can be made: Management,” paper SPE 77646 presented at the 2002 SPE
1. Tiltmeter-based deformation monitoring is an Annual Technical Conference and Exhibition, San Antonio, TX,
effective reservoir monitoring technique for the 29 September-2 October.
measurement of volumetric changes at the reservoir 2. Davis, E., Astakhov, D., Wright, C., “Precise Deformation
level in cyclic steam stimulation applications. Monitoring by High Resolution Tiltmeters”, Butsuri-Tsana,
2. Tiltmeter-based deformation monitoring provides Society of Exploration Geophysicists of Japan, (2001), Vol. 54,
No. 6, pp. 425-432.
direct measurement of the areal coverage of the
3. Davis, E., Wright, C., Demetrius, S., Choi, J. and Craley, G.;
dilation in the reservoir, as well as, identifying areas “Precise Tiltmeter Subsidence Monitoring Enhances Reservoir
where the compaction (or compression, or volume Monitoring,” paper SPE 62577 presented at the 2000
reduction) occurs. SPE/AAPG Western Regional Meeting, Long Beach, CA, June
3. Inverting the measured surface tilt for the volumetric 19-23.
change at reservoir level improves the ability to 4. Okada, Y.: “Internal Deformation Due to Shear and Tensile Faults
interpret reservoir processes. in a Half-Space”, Bull. of the Seism. Soc. Of America, (1992),
4. Volumetric changes can be non-uniform with some Vol. 82, N 2, pp.1018 – 1040.
pad areas deforming more than others. 5. Bruno, M. S. and Bilak, R. A.: “Cost-Effective Monitoring of
Injected Steam Migration Using Surface Deformation Analysis”,
5. Deformation-based, reservoir-level monitoring has
paper SPE 27888 presented at the 1994 SPE Western Regional
proven helpful in ongoing efforts to optimize such Meeting, Long Beach, CA, March 23-25.
variables as the length of well laterals, injection rates, 6. Segall, P.: “Stress and Subsidence Resulting from Subsurface
lateral spacing, and cycle times. Fluid Withdrawal in the Epicentral Region of the 1983 Coalinga
Earthquake.” J. Geophys. Res., (1985), 90, 6801-6816.
8 SPE 96897

7. Cheng, D.H. and Mindlin, R.D., “Thermoelastic Stress in the


Semi-Infinite Solid,” Journal of Applied Physics, (1950), 21.9,
pp. 931-33.
8. Vasco, D.W., Karasaki, K. and Myer, L.: “Monitoring of Fluid
Injection and Soil Consolidation Using Surface Tilt
Measurements”, Journal of Geotech. And Geoenvironmental
Eng., (January 1998), 124, 29-37.
9. Segall, P.: “Induced Stresses due to Fluid Extraction from
Axisymmetric Reservoirs”, Pure Applied Physics, (1992) 139,
536-560.
10. Du, J. and Olson, J. E.: “A Poroelastic Reservoir Model for
Predicting Subsidence and Mapping Subsurface Pressure
Fronts”, J. of Petroleum Science and Engineering, (2001), 30,
181-197.
11. Du, Y., Aydin, A. and Segall, P.: “Comparison of various
inversion techniques as applied to the determination of a
geophysical deformation model for the 1983 Borah Peak
earthquake”, Bull. Seismol. Soc. Am., (1992), 82, No. 4, 1840-
1866.
12. McGillivray, P. R.: “Microseismic and Time-lapse Monitoring of
a Heavy Oil Extraction Process at Peace River”, presented at the
2004 SEG International Exposition and 74th Annual Meeting,
Denver, CO, Oct. 10-15.
13. Glandt, C. A. and Malcolm, J. D.: “Numerical Simulation of
Peace River Recovery Process”, paper SPE 22645 presented at
the 1991 SPE Annual Technical Conference and Exhibition,
Dallas, TX, Oct. 6-9.
14. Geneau, M. D.: “Selection, Operation, and Evaluation of High
Temperature Oil in Water Monitor for Two-Phase Extra Heavy
Oil (Bitumen) Test Separator Service at Peace River”, paper
SPE 83505 presented at 2003 SPE Western Regional/AAPG
Pacific Section Joint Meeting, Long Beach, CA, May 19-24.
15. Wright, C.A., Davis, E.J., Minner, W.A., Ward, J.F., Weijers, L.
and Scheel, E.J., “Surface Tiltmeter Fracture Mapping Reaches
New Depths – 10,000 Feet, and Beyond,” paper SPE 39919,
presented at the 1998 SPE Rocky Mountain Regional/Low-
Permeability Symposium and Exhibition, Denver, CO, April 5-
8.
16. Lawson, C.L. and Hanson, R.J.: Solving Least Squares Problem,
Prentice-Hall, Englewood Cliffs, New Jersey, (1974) Chapter
23, p. 161.
17. Olson, Jon E., Du, Yijun and Du, J.: (1997). “Tiltmeter data
inversion with continuous, non-uniform opening distribution: A
new method for detecting hydraulic fracture geometry”, Int. J.
Rock Mech. & Min. Sci., Vol. 34, No. 3-4, paper No. 236.

SI Metric Conversion Factors


in × 2.54* E-03 = m
*
ft × 3.048 E-01 = m
psi × 6.894 759 E-02 = bar
*Conversion factor is exact.

You might also like