You are on page 1of 12

Assessment of Mud-Filtrate-Invasion

Effects on Borehole Acoustic Logs and


Radial Profiling of Formation
Elastic Properties
Shihong Chi,* SPE, Carlos Torres-Verdín, SPE, Jianghui Wu,** SPE, and Faruk O. Alpak,† SPE,
U. of Texas at Austin

Summary for the processing and interpretation of sonic logs. A common


Despite continued improvements in acoustic-logging technology, model used in the open literature assumes that a sharp interface
sonic logs processed with industry-standard methods often remain exists between the altered zone and the undisturbed formation
affected by formation damage and mud-filtrate invasion. Quanti- (Baker 1984). The term “stepwise” is used to describe this type of
tative understanding of the process of mud-filtrate invasion is nec- mud-filtrate-invasion model. Linear-gradient models have been
essary to identity and assess biases in the standard estimates of described for syntheses of acoustic waveforms as well (Stephen
in-situ compressional- and shear-wave (P- and S-wave) velocities. et al. 1985). Actual radial distributions of elastic wave properties
We describe a systematic approach to quantify the effects of mud- resulting from invasion can be complex and are dependent on the
filtrate invasion on borehole acoustic logs and introduce a new specific petrophysical properties of the rock as well as on the static
algorithm to estimate radial distributions of elastic properties away and dynamic properties of the fluids involved. We divide the in-
from the borehole wall. Radial saturation distributions of mud vaded zone into a set of concentric radial layers to represent an
filtrate and connate formation fluids are obtained by simulating the actual invasion profile and call it a “multilayered” model. Subse-
process of mud-filtrate invasion. Subsequently, we calculate radial quently, we describe a procedure for calculating the radial distri-
distributions of the elastic properties using the Biot-Gassmann butions of formation elastic properties with the Biot-Gassmann
fluid-substitution model. The calculated radial distributions of for- fluid-substitution model starting from numerically simulated radial
mation elastic properties are used to simulate array sonic wave- distributions of flow saturation.
forms. Finally, estimated P- and S-wave velocities for homoge- Theoretical and experimental studies (Geertsma and Smit 1961;
neous, stepwise, and multilayered formation models are compared Biot 1956a, 1956b; Toksöz et al. 1976; Domenico 1976; Dutta and
to quantify mud-filtrate-invasion effects on sonic measurements. Ode 1979; Murphy 1982) have shown that rocks saturated with
We use a nonlinear Gauss-Newton inversion algorithm to es- hydrocarbons and water can be differentiated with acoustic veloc-
timate radial distributions of formation elastic parameters in the ity measurements. Using time-lapse acoustic logging, bypassed
presence of invaded zones using normalized spectral ratios of array zones can be identified, and fluid movement in rock formations
waveform data. Inversion examples using synthetic and field data can be monitored in open and cased holes. If the invasion depth is
indicate that physically consistent distributions of formation elastic beyond the radial length of investigation of a logging tool, the
properties can be reconstructed from array waveform data. In turn, measured velocities will reflect those of the invaded zone, and log
radial distributions of formation elastic properties can be used to corrections will be necessary to estimate virgin-formation velocities.
construct more-realistic near-wellbore petrophysical models for The current industry practice of acoustic-data processing makes
applications in reservoir simulation and production.
use of the slowness-time coherence (STC) method to estimate
Introduction in-situ velocities of rock formations. This type of processing yields
values of P- and S-wave velocities, both of which are intricate
During and after drilling, the near-wellbore formation is often functions of formation and borehole properties, frequency, and
altered by stress buildup and release, mud-filtrate invasion, chemi- number of time samples used in the time window of the STC
cal reactions, and many other factors. These alterations cause the algorithm. When invasion into a formation is significant, the STC
physical properties in the near-wellbore region to be different from
method still gives one P-wave velocity and/or S-wave velocity at
those of the virgin rock formation. Stress concentration around a
each depth of measurement, which may reflect only properties of
wellbore may cause near-wellbore damage and induce formation
the invaded zone. However, waveforms recorded by modern
anisotropy on P- and S-wave velocities. The stress-induced anisot-
acoustic tools may contain valuable information on the radial dis-
ropy can be identified by dispersion analysis (Plona et al. 2002).
Positive radial velocity gradients focus the elastic waves propa- tribution of near-wellbore formation properties.
gating away from the wellbore back toward the borehole wall. In an effort to estimate radial distributions of near-wellbore
Such a phenomenon can be identified easily from high-amplitude formation properties, Hornby (1993) developed a tomographic re-
acoustic arrivals. In this paper, we focus our attention on mud- construction technique that yielded formation P-wave velocities
filtrate-invasion effects only. from sonic travel times. The estimation was performed with a
It is well known that formation properties inferred from wire- series of linear inversions followed by ray tracing. However,
line logging measurements may not reflect true properties of virgin travel-time tomography using ray tracing requires accurate picking
formations. A realistic description of the invaded zone is important of the refracted arrivals of each wave train and the use of an
accurate high-frequency ray-trace approximation.
Few studies of seismic full-waveform inversion (Sen and Stoffa
1991; Zhou et al. 1997; Pratt 1999a, 1999b) show that the use of
* Currently with ConocoPhillips. amplitude data can improve the resolution of 1D vertical distribu-
** Currently with Baker Atlas.
† Currently with Shell Intl. E&P. tions of velocity and density. The amplitude and phase behavior of
full-waveform data are sensitive to the petrophysical properties of
Copyright © 2006 Society of Petroleum Engineers
formations supporting the propagation of elastic waves. Therefore,
This paper (SPE 90159) was first presented at the 2004 SPE Annual Technical Conference full-waveform analysis has significant potential in acoustic log-
and Exhibition, Houston, 26–29 September, and revised for publication. Original manuscript
received for review 4 June 2004. Revised manuscript received 29 May 2006. Paper peer
ging to estimate petrophysical properties of invaded zones and
approved 25 July 2006. virgin formations. The reconstruction of formation slowness away

October 2006 SPE Reservoir Evaluation & Engineering 553


from the borehole wall can provide valuable calibration factors for
measurements acquired with shallow-reading devices. Corrections
provided by full-waveform inversion may be of primary impor-
tance for the robust and accurate computation of synthetic seis-
mograms. There is, however, one major difficulty to overcome in
full-waveform inversion. In all field applications, the effective
source wavelet; the coupling among source, borehole, and forma-
tion; and the coupling between receivers and the formation depend
on the in-situ borehole conditions. The practical difficulty in esti-
mating the sonic-source signature is probably the reason why very
few attempts have been made to use the information borne by the
full waveform of acoustic-logging data.
Cheng (1989) described an indirect method for determining
S-wave velocities from full-waveform acoustic logs by means of
inversion. This method makes use of the spectral ratio of the Fig. 1—Schematic of the process of mud-filtrate invasion
P-wave trains at two source/receiver separations and simulta- in rock formations penetrated by a mud-filled and overbal-
neously inverts for both S-wave velocity and P-wave attenuation. anced borehole.
In that paper, Cheng’s approach based on spectral-ratio data was
extended to full-waveform arrays to overcome the practical diffi- that mud filtrate penetrates to a radial distance of 0.5 m after 4 days
culty of the unknown source-output spectrum. Lee and Kim (2003) of invasion.
estimated P-wave velocity distributions from synthetic seismic Because of relative permeability, in the example case consid-
data by use of a spectral-ratio method. Frazer et al. (1997) and ered here, oil-based mud invades the gas-bearing sandstone for-
Frazer and Sun (1998) described an inversion method for process- mation at a much slower rate than for the case of water-based
ing array full-waveform data. Their approach does not require an mud-filtrate invasion. Fig. 4 shows the time evolution of oil-based
exact source function, although it is a necessary part of the inver- mud-filtrate saturation in the radial direction of a low-porosity (15
sion. Because of this, the chosen source function can bias the p.u.) and low-permeability (30 md) formation. Oil-based mud fil-
inversion results. trate reaches a radial distance of approximately 0.2 m after 4 days
Our inversion approach first transforms array waveforms into of invasion. Fig. 5 shows that mud filtrate primarily concentrates
the frequency domain to construct a set of normalized array sonic within a distance of 0.1 m away from the borehole wall.
data. The normalized wavefield is independent of the spectrum of
the source; hence, the proposed inversion method allows one to Radial Distributions of Density and P- and S-Wave Velocities
make use of the full-waveform content of sonic data without re- in an Invaded Zone. Once radial profiles of invasion are obtained
quiring knowledge of the source signature. Previous studies of from numerical simulations, radial distributions of the elastic prop-
full-waveform borehole acoustic measurements assumed a homo- erties of a formation can be calculated using the Biot-Gassmann
geneous and isotropic formation and a model for borehole wave fluid-substitution equations. Following the tutorial presented by
propagation without surface irregularities. In this paper, we make Smith et al. (2003), a calculation is first performed for the basic
the assumptions of a radially multilayered formation model and a formation and fluid properties. Subsequently, for each point in the
cylindrical borehole. saturation profile we calculate (a) fluid properties, (b) saturated
Methodology bulk modulus of the rock, (c) bulk density of the rock, and (d) P-
and S-wave velocities.
Numerical Simulation of Mud-Filtrate Invasion. Mud-filtrate When computing fluid and formation properties after invasion,
invasion is regarded as a water- or oil-injection process into a gas a careful selection of homogeneous or patchy saturation models is
or oil reservoir wherein two-phase immiscible fluid flow is as- necessary. Over geologic time, fluids in the pores of the rocks
sumed in the simulations. To study the effects of different types of become homogeneously distributed. This is one of the important
mud filtration on gas or oil reservoirs, the following cases are
selected: (a) water-based mud invades oil and gas reservoirs pen-
etrated by open holes and (b) oil-based mud invades gas reservoirs
penetrated by open holes. In each of these cases, it is assumed that
a vertical well penetrates horizontally layered rock formations and
that the rock formation considered does not communicate hydrau-
lically with its upper and lower shoulder beds. To study the sen-
sitivity of mud-filtrate invasion to formation petrophysical prop-
erties, we consider a low-permeability (30 md) and low-porosity
(15 p.u.) reservoir and a high-permeability (300 md) and high-
porosity (30 p.u.) reservoir for Cases (a) and (b).
Details of the numerical simulation of the process of mud-
filtrate invasion in open holes can be found in Wu et al. (2004,
2005) and George et al. (2004). In the simulations, both the dy-
namic growth of the mudcake and the dynamic decrease of the
mudcake permeability are coupled to formation properties. This
results in a dynamic monotonic decrease of flow rate into the
formation. After a short initial spurt of mud-filtrate invasion, the
invasion process reaches a steady state. Cycles of mudcake ruboff
and buildup can also affect the process of mud-filtrate invasion.
Fig. 1 is a schematic of the process of mud filtrate invading a
permeable rock formation. The invasion front moves deeper into
the formation with time. Fig. 2 shows the time evolution of water-
Fig. 2—Time evolution of mud-filtrate invasion for the case of
based mud-filtrate saturation in the radial direction of a low- water-based mud invading a 15%-porosity, oil-bearing rock for-
porosity, low-permeability, and oil-bearing sandstone formation. mation. From left to right, the first curve is the water-saturation
The water-based mud filtrate reaches a radial distance of approxi- profile 1 day after the onset of mud-filtrate invasion. The time
mately 0.8 m after 4 days of invasion. Mud-filtrate invasion is increment is 1 day, and the rightmost curve describes water
shallower for the corresponding high-porosity case. Fig. 3 shows saturation at the 16th day.

554 October 2006 SPE Reservoir Evaluation & Engineering


Fig. 5—Time evolution of mud-filtrate invasion for the case of
oil-based mud invading a 30%-porosity, gas-bearing reservoir.
Fig. 3—Time evolution of mud-filtrate invasion for the case of From left to right, the first curve is the water saturation profile 1
water-based mud invading a 30%-porosity, oil-bearing reser- day after the onset of mud-filtrate invasion. The time increment
voir. From left to right, the first curve is the water-saturation is 1 day, and the rightmost curve describes oil saturation at the
profile 1 day after the onset of mud-filtrate invasion. The time 16th day.
increment is 1 day, and the rightmost curve describes water
saturation at the 16th day.
In the synthesis of array waveforms, we assume a representa-
assumptions of the Biot-Gassmann fluid-substitution model. How- tive source/receiver configuration. For monopole logging, the cen-
ever, mud-filtrate invasion in the near-wellbore region changes the tral frequency of the source is 10 kHz. The minimum source/
connate-fluid distribution. From the numerical-simulation results receiver distance is 9 ft, with the receiver array consisting of eight
described in this paper, it is clear that the equilibrium of fluid receivers spaced at 6-in. intervals. Dipole logging makes use of a
phases in the near-wellbore region is not established within a few 1.5-kHz source and a minimum source/receiver separation of 11.5
days after the onset of the mud-filtrate-invasion process. There- ft. In the lower dipole mode, the eight source/receiver offsets vary
fore, it is more appropriate to use a patchy saturation model to from 3.51 m (11.5 ft) to 4.57 m (15 ft) with a receiver spacing of
compute the corresponding bulk modulus (Hill 1963). 0.15 m (6 in.). We simulate sonic array waveforms for a homo-
geneous virgin formation as well as for a saturated formation after
Numerical Simulation of Time-Lapse Borehole Acoustic Mea- 4 days of mud-filtrate invasion.
surements. On the basis of saturation-profile and fluid-
substitution calculations, we obtain a discretized and concentric Processing and Interpretation of Monopole and Dipole Log-
multilayered rock model. The radial discretization is sufficient to ging Data. The sensitivity of acoustic measurements to different
ensure accurate representation of the continuous radial distribu- invasion models and radial saturation profiles can be assessed by
tions in the calculation of elastic properties. A corresponding step- comparing the synthetic waveforms for homogeneous, stepwise,
wise-invaded-zone model also is used for the numerical simula- and multilayered radial formation models. Arrival times and am-
tion. The waveforms simulated for the multilayered and stepwise plitudes of P- and S- wave modes are the main waveform features
models are compared, with the objective of determining when the used to assess the effect of mud-filtrate invasion on borehole
stepwise models are adequate to represent the invaded zone. acoustic logging measurements.
We use an industry standard STC method (Kimball and
Marzetta 1984) to determine the P- and S-wave velocities of array
sonic waveforms. Radial depths of investigation for sonic tools
depend on formation elastic properties, transmitter-to-receiver
spacing, wavelength and wave mode considered, and frequency,
among other factors. Because we chose a typical sonic-tool con-
figuration and source frequencies used by the logging industry, we
can focus our study of invasion effects on velocity measurements.
After P- and S-wave velocities are determined, the depth of inves-
tigation can be ascertained from the inspection of radial velo-
city profiles.
When correlating seismic data with acoustic logs, it is often
found that synthetic seismograms do not match the measured seis-
mograms. As a result, corrections to the acoustic logs by means of
Biot-Gassmann fluid substitution are common in the seismic in-
dustry to eliminate mud-filtrate-invasion effects from sonic logs.
In this correction procedure, it is assumed that the measured ve-
locities are those of the invaded zone saturated with mud filtrate.
For the case of oil-based mud, filtrate mixing with gas may have
to be considered before fluid substitution. By displacing the satu-
Fig. 4—Time evolution of mud-filtrate invasion for the case of
oil-based mud invading a 15%-porosity, gas-bearing reservoir.
ration fluid in the invaded zone with the connate formation fluid
From left to right, the first curve is the water saturation profile 1 and by applying the Biot-Gassmann fluid-substitution equation,
day after the onset of mud-filtrate invasion. The time increment new velocities are obtained and taken as virgin-formation veloci-
is 1 day, and the rightmost curve describes oil saturation at the ties. We investigate the validity of this practice through several
16th day. case studies.

October 2006 SPE Reservoir Evaluation & Engineering 555


Full-Waveform Inversion. In our inversion of radial elastic prop- tions are simulated for 4 days of invasion. The latter profiles are
erties, we assume a cylindrical borehole surrounded by concentric used to calculate radial distributions of elastic properties using the
multilayered formations. We first transform array sonic waveforms Biot-Gassmann fluid-substitution model.
into the frequency domain to construct a set of normalized array Fig. 2 shows the time evolution of mud-filtrate invasion for the
sonic spectra. The normalized wavefield is independent of the case of water-based mud invading the oil-bearing sandstone res-
spectrum of the source; hence, the proposed inversion method ervoir of 15% porosity. Fig. 6 shows the corresponding radial
allows one to make use of the full-waveform content of sonic data distributions of density and P- and S-wave velocities calculated
without requiring knowledge of the source signature. The normal- with the fluid-substitution parameters described in Table 2. As
ized wavefield (or spectrum) of full waveforms constitutes the shown in Fig. 6, in the invaded zones the P-wave velocity is higher
measurement data input to the full-waveform inversion algorithm. than that of the virgin formation, while the S-wave velocity be-
For estimation of the radial distributions of elastic properties, den- comes lower than the original velocity. P-wave velocities esti-
sity and P- and S-wave velocities are assigned to each concentric mated from the array waveforms (Fig. 7a) simulated for stepwise
layer in the near-wellbore region. Subsequently, we simulate the and multilayered models are approximately 1% lower than those of
measured wavefields with the assumed properties. A cost function the virgin formation. STC processing can introduce 1% uncertainty
that enforces the quadratic difference between the simulated and in the estimation of P-wave velocity. Therefore, the 1% lower
measured normalized wavefields is used for inversion. We make P-wave velocity may result from STC processing. This implies
use of a Gauss-Newton fixed-point iteration search to determine a that the P-wave is not sensitive to the invaded zone, which exhibits
stationary point at which the cost function attains a minimum. The a higher P-wave velocity. On the other hand, S-wave velocities
stationary point yields the radial distribution of elastic properties. obtained from STC processing of monopole and dipole waveforms
Details of the full-waveform inversion are given in the Appendix. are 2% lower than those of the virgin formation. This behavior
agrees with the velocity decrease described in Fig. 6 and indicates
Case Studies that S-wave propagation is sensitive to the invaded zone.
Sandstone Oil Reservoir Invaded With a Water-Based Mud. In Fig. 8 shows the radial distributions of density and P- and
onshore drilling activities, water-based mud is still widely used S-wave velocities for the high-porosity and high-permeability case
because of its efficiency to balance formation pressure without after 4 days of invasion. The corresponding STC results show that
substantially increasing costs. Elastic properties of oil in rock for- the P-wave velocity is the same as that of the virgin formation
mations are much closer to those of mud filtrate derived from a within the processing error bound of 1%. This behavior indicates
water-based mud. Compared to the case of mud filtrate invading a that P-wave propagation is sensitive to the radial region beyond
gas reservoir, it may be more difficult to distinguish mud-filtrate- the invaded zone. The S-wave velocity obtained for this invaded
invasion effects on the measured velocities. Moreover, fluid-flow model is approximately 2.6% lower than that of the virgin zone.
properties of oil and gas are very different from each other. To This relatively larger reduction in S-wave velocity compared to the
study the sensitivity of the measured velocities to porosity and low-porosity case is expected because more oil is replaced by
permeability, this section first considers a low-porosity (15 p.u.) water-based-mud filtrate.
and low-permeability (30 md) reservoir. Table 1 summarizes the In summary, P-wave propagation is sensitive to the virgin zone,
petrophysical and fluid properties used in the simulation of water- whereas S-wave propagation is sensitive to the invaded zone when
based mud-filtrate invasion. Subsequently, we focus our attention the radial length of mud-filtrate invasion is approximately 0.5 to
to the case of a high-porosity (30 p.u.) and high-permeability (300 0.8 m for the two sandstone reservoirs considered in this section.
md) reservoir. Saturation profiles of mud filtrate and oil in forma-
Sandstone Gas Reservoir Invaded With an Oil-Based Mud.
Our study indicates that, in general, oil-based mud causes shal-
lower invasion than water-based mud and, hence, less formation
damage. Because oil-based mud is also chemically less active
compared to water-based mud, it is more effective in inhibiting
shale swelling. This is one of the reasons why oil-based mud is
widely used to drill expensive offshore wells.

Fig. 6—Radial distributions of density and P- and S-wave veloc-


ities for the case of water-based mud invading a 15%-porosity,
oil-bearing reservoir after 4 days of invasion.

556 October 2006 SPE Reservoir Evaluation & Engineering


We selected two synthetic cases to study mud-filtrate effects on
the measured velocities of sandstone formations exhibiting low
and high porosities and permeabilities.
Fig. 9 shows the radial distributions of density and P- and
S-wave velocities for the case of a gas-bearing sandstone reservoir
of 15% porosity invaded with an oil-based mud.
P-wave velocities estimated from array waveforms simulated
for a stepwise and a multilayered model are approximately 1.0%
higher than that of the true formation P-wave velocity, but ap-
proximately 3.0% lower than that of an invaded zone saturated
with 90% mud filtrate. Estimated S-wave velocities are approxi-
mately 0.2% lower than that of the true formation S-wave velocity,
while the S-wave velocity for the invaded zone is 1.2% lower than
that of the true formation. Therefore, P- and S-wave velocities are
slightly influenced by mud-filtrate invasion and remain primarily
sensitive to the radial region beyond the invaded zone.
For the 30%-porosity case, Fig. 10 shows the radial distribu-
tions of density and P- and S-wave velocities. It is observed from
the monopole waveforms (Fig. 11) that the presence of invaded
zones does not affect the P-wave arrival time, but it does delay the
S-wave arrival time. On the other hand, P-wave amplitudes do not
change appreciably, whereas S-wave amplitudes increase. In the
dipole waveforms, the change in S-wave arrival is negligible. Di- Fig. 7—Simulated (a) monopole and (b) dipole waveforms for
pole waveforms for the stepwise model are significantly different the homogeneous, stepwise, and multilayered formation mod-
from those simulated for the multilayered formation model. els shown in Fig. 6.
P-wave velocities estimated from data for homogeneous, mul-
tilayered, and stepwise models are the same. This behavior indi-

Fig. 8—Radial distributions of density and P- and S-wave veloc- Fig. 9—Radial distributions of density and P- and S-wave veloc-
ities for the case of water-based mud invading a 30%-porosity, ities for the case of oil-based mud invading a 15%-porosity,
oil-bearing rock formation after 4 days of invasion. gas-bearing rock formation after 4 days of invasion.

October 2006 SPE Reservoir Evaluation & Engineering 557


Fig. 10—Radial distributions of density and P- and S-wave ve-
locities for the case of oil-based mud invading a 30%-porosity,
gas-bearing rock formation after 4 days of invasion.

cates that P-wave propagation is sensitive to the radial region past


the invaded zone. S-wave velocities estimated for the stepwise and
the multilayered models are approximately 1% lower than that of
the virgin formation. According to the Biot-Gassmann fluid-
substitution model, a rock formation saturated with 90% mud fil-
trate should exhibit a 4.3% S-wave velocity decrease compared to
that of the virgin formation. This indicates that the measured
S-wave velocities are primarily sensitive to the virgin formation
but remain influenced by mud filtrate in the near-wellbore re-
gion. In all the cases studied in this section, both P- and S-waves
are not sensitive to the invaded zone when the oil-based mud-
filtrate invasion reaches a radial length of approximately 0.3 m in Fig. 11—Simulated (a) monopole and (b) dipole waveforms for
sandstone reservoirs. the case of homogeneous, stepwise, and multilayered forma-
tion models invaded with oil-based mud after 4 days of invasion
Elastic Radial Profiles of a Fast Rock Formation. This inversion in a 30%-porosity gas reservoir.
case makes use of realistic elastic properties for a fast-formation
model (Schmitt 1988), which consists of six radial layers, includ- Inversions performed with no assumption of an increasing
ing a fluid layer in the borehole. Table 3 describes the actual fluid value of elastic properties away from the borehole wall converge
and formation properties and the radial discretization grid used for to local minima without exception. Fig. 13 shows the radial dis-
the inversion. Amplitude spectra of the simulated waveforms are tributions and crossplots of actual and inverted elastic properties
used in the estimation to accelerate the convergence of the inver- after 20 iterations for a slow formation under the assumption of
sion algorithm. monotonically increasing values of elastic properties. The inver-
Fig. 12 shows radial distributions and crossplots of actual and sion stops at a local minimum, as can be observed from the evo-
inverted elastic properties after 26 iterations of the inversion al- lution of the data misfit and cost functions shown in Figs. 13a and
gorithm. The inverted radial distributions of elastic properties 13b, respectively. Global correlation coefficients for the inverted
show a good agreement with the actual properties. Global corre- and actual velocities are relatively high (0.83 and 0.83 for P- and
lation coefficients calculated between the inverted and the actual S-wave velocities, respectively). However, for bulk density, the
values of elastic properties indicate that all the inverted elastic correlation coefficient is −0.45. This indicates the relatively low
properties exhibit high global similarity with the original model sensitivity of the measurements to radial variations of density in
properties (0.932, 1, and 1 for P- and S-wave velocities and den-
sity, respectively).

Elastic Radial Profiles of a Slow Rock Formation. This example


is an extension of the slow formation model in the simple borehole
cases described by Cheng (1989) within the context of marine
sediments. Table 4 describes the model, which consists of six
radial layers, including one fluid layer in the borehole. The tool
and source configurations are the same as those discussed previ-
ously for simple borehole cases. Given that the low-frequency
content of the full waveforms is more important in soft formations
than in fast formations, data from the frequency band 2–5 kHz are
used for inversion. Inversion results from this example not only
estimate the virgin-formation elastic properties but also give a
description of near-wellbore damage.

558 October 2006 SPE Reservoir Evaluation & Engineering


Fig. 12—Simultaneous inversion of radial distributions of elastic properties for a six-layer fast formation using noise-free normal-
ized spectra of array waveform data. Panel (a) shows the array waveform in the time domain, and Panel (b) shows the data residuals
yielded by the inversion. In Panel (c), the inverted radial distributions of elastic properties are identified with solid lines and open
circles; additional lines identify original model properties. Panel (d) shows that the correlation coefficient, r2, is the correlation
coefficient between the inverted and actual elastic properties.

the 2- to 5-kHz frequency band. Overall, the inversion algorithm with the Dipole Sonic Imager (DSI)* tool in the depth interval
provides a reliable way to estimate radial distributions of elastic from 13,000 to 13,050 ft within a well penetrating a tight-
properties in soft rock formations. This is a valuable tool for de- sandstone gas reservoir (Anderson Well No. 2). Core data and log
tailed analysis of acoustic-logging data in offshore wells. measurements indicate that the porosity of the fine-grained sand-
stone formation is below 9%. Gas saturation ranges from 80% to
Elastic Radial Profiles From Field Data. This example assesses 95%. Fig. 14a displays the array waveform data acquired at the
the feasibility and performance of the inversion algorithm when depth of 13,030 ft. The bottom four traces show very similar
applied to field data. Full-waveform acoustic data were acquired characteristics in the number of wave modes and in the amplitudes
of each wave mode. The top four traces, however, exhibit a com-
pletely different character. Fig. 14b displays the amplitude spectra
of the array waveform data. The main energy of the waveforms is
contained in the frequency band from 9 to 14 kHz. Input data to the
inversion algorithm are chosen to be the normalized frequency
data in the band of 11–14 kHz, given that previous examples show
that data from a narrower frequency-data band improve the con-
vergence of the algorithm. The STC processing method yields
formation P- and S-wave slowness values of 66.33 and 108.01
␮s/ft, respectively, which are used as the initial properties for the
inversion. Likewise, the density log reads a value of 2.504 g/cm3
for bulk formation density. The mud density at this depth in the

* Mark of Schlumberger.

October 2006 SPE Reservoir Evaluation & Engineering 559


Fig. 13—Radial distributions and crossplots of the actual and inverted elastic properties for a six-layer slow formation. The
evolution of the data misfit and cost function with iteration number is shown in Panels (a) and (b), respectively. In Plot (c), the
inverted radial distributions of elastic properties are identified with open circles. Additional lines identify original model properties.
Panel (d) shows the correlation coefficient, r2, between the inverted and actual elastic properties.

borehole is 14 lbm/gal (1678 kg/m3), whereas the acoustic velocity hole wall and increases the P-wave amplitude (Winkler 1997;
of the mud is approximately 1186 m/s. Mud-filtrate-invasion stud- Chen et al. 1996). The radial distribution of density shows a de-
ies (Salazar et al. 2005) indicate that mud filtrate reaches a radial creasing trend from the borehole wall into the formation. Such a
distance of approximately 1.5 ft in the flow unit of interest after 24 behavior can be interpreted as due to mud-filtrate displacing gas in
hours of drilling. Thus, we use a concentric five-radial-layer model the near-wellbore region. This exercise confirms that the inversion
to describe the near-wellbore invasion zone. The inner radii of the algorithm is reliable to estimate radial distributions of formation
radial formation layers are 0.07, 0.15, 0.30, 0.40, and 0.45 m, elastic properties in the near-wellbore region.
respectively, and the outmost layer is assumed to be unbounded in
the radial direction. Within each radial zone, formation elastic Conclusions
properties are assumed constant. In the cases of water-based mud invading oil-bearing sandstone
First, normalized frequency data from Traces 1 through 4 in the reservoirs, P-wave propagation remains sensitive to virgin zones,
band from 11 to 14 kHz are used as input data for the inversion. whereas S-wave propagation is affected by invaded zones when
The data misfit decreases to 4% after 20 iterations. Fig. 14c indi- the radial length of invasion of mud filtrate is approximately 0.5 to
cates that the formation P- and S-wave velocities increase and bulk 0.8 m. In this latter situation, log corrections are not necessary for
density decreases in the radial direction. It is known that formation P-wave velocities, whereas detailed saturation-profile calculations
damage caused by drilling decreases the formation velocities and are needed to correct the measured S-wave velocities for mud-
that mud-filtrate invasion in the near-wellbore region increases the filtrate-invasion effects. For the cases of oil-based mud invading
bulk density. The inverted radial distributions indicate that a dam- sandstone reservoirs, both P- and S-wave propagations are insen-
aged zone exists in the near-wellbore region even though the bore- sitive to the presence of mud filtrate in the invaded zones.
hole is in excellent condition. Such a result also agrees with the We developed a new full-waveform inversion algorithm that
high-amplitude P-wave components that are observed from Traces makes use of the normalized frequency spectra of sonic wave-
1 through 4 of the array sonic waveforms. A radial profile of forms. The reliability of the inversion algorithm was tested suc-
monotonically increasing P-wave velocity focuses the elastic cessfully with radially multilayered 1D synthetic models. In addi-
waves propagating away from the wellbore back toward the bore- tion, we obtained petrophysically consistent results when the in-

560 October 2006 SPE Reservoir Evaluation & Engineering


Fig. 14—Simultaneous inversion of radial distributions of elastic properties using array waveform data acquired in the Anderson
Well No. 2 and penetrating a tight gas reservoir. Panel (a) shows the array waveforms in the time domain, and Panel (b) shows the
corresponding amplitude spectra. Panel (c) shows the homogeneous-formation model used to initialize the inversion and the
inverted radial distributions of formation elastic properties.

version algorithm was applied to field data acquired in a tight M ⳱ number of frequency-domain measurements
gas reservoir. N ⳱ number of radial layers
Inversion exercises performed on synthetic data indicate that NFREQ ⳱ number of frequencies used for each trace
the low-frequency content of full-waveform data is sensitive to NREC ⳱ number of receivers
formation elastic properties radially away from the borehole wall. r1 ⳱ radius of the first radial formation layer
Thus, high-frequency components of array sonic data may be pref-
R ⳱ borehole radius
erable for estimating the radial distribution of elastic properties in
the near-wellbore region. +− ⳱ reflectivity between the fluid and the first radial
R̂(1)
layer of the formation
Re ⳱ real
Nomenclature Sj ⳱ spectra of normalized waveforms
C(m) ⳱ cost function S(␻) ⳱ effective-source-output spectrum
d(m) ⳱ measurement vector numerically simulated for T ⳱ transpose
specific values of m T12 ⳱ transfer function between the two receivers
dobs ⳱ measurement vector ui ⳱ upper bound
i ⳱ index or imaginary unit Vf ⳱ fluid acoustic velocity
Im ⳱ imaginary Vp ⳱ P-wave velocity
j, k ⳱ indices Vs ⳱ S-wave velocity
J(m) ⳱ Jacobian matrix Wd·WTd ⳱ inverse of the measurement covariance matrix
li ⳱ lower bound Wm·WTm ⳱ inverse of the model covariance matrix
kf ⳱ fluid wave number z1, z2 ⳱ receiver locations
kp ⳱ radial wave number of fluid ␹ ⳱ prescribed value of enforced data misfit
f
m ⳱ size-N vector of unknown parameters ␭ ⳱ Lagrange multiplier or regularization parameter
mR ⳱ size-N reference vector ␳ ⳱ density

October 2006 SPE Reservoir Evaluation & Engineering 561


Acknowledgments Pratt, R.G. 1999a. Seismic waveform inversion in frequency domain, Part
The work reported in this paper was funded by the U. of Texas at 1: Theory and verification in physical scale model. Geophysics 64:
Austin’s Research Consortium on Formation Evaluation, jointly 888–901. DOI: http://dx.doi.org/10.1190/1.1444597.
sponsored by Aramco, Baker Atlas, BP, British Gas, ConocoPhil- Pratt, R.G. 1999b. Seismic waveform inversion in frequency domain, Part
lips, Chevron, Eni E&P, ExxonMobil, Halliburton Energy Ser- 2: Fault delineation in sediments using crosshole data. Geophysics 64:
vices, Marathon, the Mexican Inst. for Petroleum, Norsk-Hydro, 902–914. DOI: http://dx.doi.org/10.1190/1.1444598.
Occidental Petroleum Corp., Petrobras, Schlumberger, Shell Intl. Salazar, J.M., Torres-Verdín, C., and Sigal, R. 2005. Assessment of per-
E&P, Statoil, Total, and Weatherford. meability from well logs based on core calibration and simulation of
mud-filtrate invasion. Petrophysics 46 (6): 434–451.
Schmitt, D.P. 1988. Shear wave logging in elastic formations. J. Acoust.
References
Soc. of Am. 84: 2215–2229. DOI: http://dx.doi.org/10.1121/1.397015.
Baker, L.J. 1984. The effect of the invaded zone on full wavetrain acoustic Sen, M.K. and Stoffa, P.L. 1991. Nonlinear one-dimensional seismic wave-
logging. Geophysics 49: 796–809. DOI: http://dx.doi.org/10.1190/ form inversion using simulated annealing. Geophysics 56: 1624–1638.
1.1441708. DOI: http://dx.doi.org/10.1190/1.1442973.
Biot, M.A. 1956a. Theory of propagation of elastic waves in a fluid satu- Smith, T.M., Sondergeld, C.H., and Rai, C.S. 2003. Gassmann fluid sub-
rated porous solid I. Low-frequency range. J. Acoust. Soc. Am. 28: stitutions: A tutorial. Geophysics 68: 430–440. DOI: http://dx.doi.org/
168–178. DOI: http://dx.doi.org/10.1121/1.1908239. 10.1190/1.1567211.
Biot, M.A. 1956b. Theory of propagation of elastic waves in a fluid satu- Stephen, R.A., Cardo-Casas, F., and Cheng, C.H. 1985. Finite difference
rated porous solid II. Higher frequency range: J. Acoust. Soc. Am. 28: synthetic acoustic logs. Geophysics 50: 1588–1609. DOI: http://
179–191. DOI: http://dx.doi.org/10.1121/1.1908241. dx.doi.org/10.1190/1.1441849.
Chen, X., Quan, Y., and Harris, J.M. 1996. Seismogram synthesis for Toksöz. M.N., Cheng, C.H., and Timur, A. 1976. Velocities of seismic
radially layered media using the generalized reflection/transmission waves in porous rocks: Geophysics 41: 621–645. DOI: http://
coefficients method: Theory and applications to acoustic logging. Geo- dx.doi.org/10.1190/1.1440639.
physics 61: 1150–1159. DOI: http://dx.doi.org/10.1190/1.1444035. Toksöz, M.N., Wilkens, R.H., and Cheng, C.H. 1985. Shear wave velocity
Cheng, C.H. 1989. Full waveform inversion of P waves for Vs and Qp. and attenuation in ocean bottom sediments from acoustic log wave-
J. of Geophys. Res. 94: 15619–15625. forms. Geophys. Res. Lett. 12: 37–40.
Torres-Verdín, C. and Habashy, T.M. 1994. Rapid 2.5-dimensional for-
Cheng, C.H., Wilkens, R.H., and Meredith, J.A. 1986. Modeling of full
ward modeling and inversion via a new nonlinear scattering approxi-
waveform acoustic logs in soft marine sediments. Paper LL presented
mation. Radio Science 29 (4): 1051–1079. DOI: http://dx.doi.org/
at the 27th SPWLA Annual Logging Symposium, Houston, 9–13 June.
10.1029/94RS00974.
Domenico, S.N. 1976. Effect of brine-gas mixture on velocity in an un-
Winkler, K.W. 1997. Acoustic evidence of mechanical damage surround-
consolidated sand reservoir. Geophysics 41: 882–894. DOI: http://
ing stressed boreholes. Geophysics 62: 16–22. DOI: http://dx.doi.org/
dx.doi.org/10.1190/1.1440670.
10.1190/1.1444116.
Dutta, N.C. and Ode, H. 1979. Attenuation and dispersion of compres- Wu, J., Torres-Verdín, C., Sepehrnoori, K., and Proett, M.A. 2005. The
sional waves in fluid-filled rocks with partial gas saturation, White influence of water-base mud properties and petrophysical parameters
model: Part I–Biot theory. Geophysics 44: 1777–1788. DOI: http:// on mudcake growth, filtrate invasion, and formation pressure. Petro-
dx.doi.org/10.1190/1.1440938. physics 46 (1): 14–32.
Frazer, L.N. and Sun, X. 1998. New objective functions for waveform Wu, J., Torres-Verdín, C., Sepehrnoori, K., and Delshad, M. 2004. Nu-
inversion: Geophysics 63: 213–222. DOI: http://dx.doi.org/10.1190/ merical Simulation of Mud-Filtrate Invasion in Deviated Wells.
1.1444315. SPEREE 7 (2): 143–154. SPE-87919-PA. DOI: 10.2118/87919-PA.
Frazer, L.N., Sun, X., and Wilkens, R.H. 1997. Inversion of sonic wave- Zhou, C., Schuster, G.T., Hassanzadeh, S., and Harris, J.M. 1997. Elastic
forms with unknown source and receiver functions. Geophys. J. Int. wave equation travel time and wavefield inversion of crosswell data.
129: 579–586. Geophysics 62: 853–868. DOI: http://dx.doi.org/10.1190/1.1444194.
Geertsma, J. and Smit, D.C. 1961. Some aspects of elastic wave propaga-
tion in fluid-saturated porous solids. Geophysics 26: 169–181. DOI: Appendix—Methodology for
http://dx.doi.org/10.1190/1.1438855. Full-Waveform Inversion
George, B.K., Torres-Verdín, C., Delshad, M., Sigal, R., Zouioueche, F., The model of wave propagation in a borehole considered in this
and Anderson, B. 2004. Assessment of in-situ hydrocarbon saturation paper assumes an isotropic, radially multilayered formation with
in the presence of deep invasion and highly saline connate water. no geometrical irregularities along the borehole wall. As with any
Petrophysics 45 (2): 141–156. inversion procedure, the accuracy and reliability of the results will
Gill, P.E., Murray, W., and Wright, M.H. 1981. Practical Optimization. depend on how accurately the forward model describes actual
London: Academic Press. in-situ conditions. In practical field studies, irregularities on the
Hill, R. 1963. Elastic properties of reinforced solids: Some theoretical borehole wall can be quantified with caliper measurements. Con-
principles. J. Mech. Phys. Solids 11: 357–372. DOI: http://dx.doi.org/ sequently, vertical formation intervals can be identified that are
10.1016/0022-5096(63)90036-X. consistent with the assumption of a smooth borehole wall. Along
Hornby, B.E. 1993. Tomographic reconstruction of near borehole slowness the length of the receiver spacing (6 in.), the formation can be
using refracted borehole sonic arrivals. Geophysics 58: 1726–1738. regarded as homogeneous in the axial direction.
DOI: http://dx.doi.org/10.1190/1.1443387. In practice, the effect of borehole/receiver coupling can be
Kimball, C.V. and Marzetta, T.L. 1984. Semblance processing of borehole
assumed negligibly small in comparison to source coupling. The
acoustic array data. Geophysics 49: 274–281. DOI: http://dx.doi.org/
spectral ratio of the pressure responses at two receiver locations, z1
and z2, from a common source can be written as

冋兰 册
10.1190/1.1441659.

Lee, K.H. and Kim, H.J. 2003. Source-independent full waveform inver- R̂共+1−兲 eikpf r1 eikf z2
sion of seismic data. Geophysics online (published electronically on 20 S共␻兲 i eikz2dk +
May 2003). DOI: http://dx.doi.org/10.1190/1.1635054. 1− R̂共+1−兲 eikpf r1 z2

冋兰 册
−⬁
Murphy, W.F. III. 1982. Effects of microstructure and pore fluids on the T12 = ⬁
, . . . . . . . . . (A-1)
acoustic properties of granular sedimentary materials. Ph.D. thesis, R̂共+1−兲 eikpf r1 eikf z1
S共␻兲 i eikz1
dk +
Stanford U., Stanford, California.
−⬁
1 − R̂共+1−兲 eikpf r1 z1
Plona, T., Sinha, B., Kane, M. et al. 2002. Mechanical Damage Detection
and Anisotropy Evaluation Using Dipole Sonic Dispersion Analysis. where S(␻) is an effective-source-output spectrum, kf is the
(1)
Paper F presented at the 43rd Annual SPWLA Conference, Osio, Japan, fluid wave number, kp is the radial wave number of fluid, R̂+−
f
2–5 June. is related to the reflectivity between the fluid and the first ra-

562 October 2006 SPE Reservoir Evaluation & Engineering


dial layer of formation, and r1 is the radius of the first layer. The Measurement and Model Vectors. In the cost function (Eq. A-2),
ratio of the two source terms in Eq. A-1 is equal to unity, resulting the measurement vector dobs is constructed from the real and
in a final frequency representation independent of the source spec- imaginary parts of the normalized spectra, Re(S) and Im(S), in the
trum. This procedure indicates that, in principle, knowledge of the following organized fashion:
source spectrum is not necessary for full-waveform inversion
dobs = 关Re共S1兲, Im共S1兲, . . . Re共Sj兲, Im共Sj兲, . . . Re共SM兲, Im共SM兲兴T,
when spectral ratios are used as input data. We note that the term
. . . . . . . . . . . . . . . . . . . . . . . . (A-3)
T12 in Eq. A-1 is the transfer function between the two receivers.
For practical applications, the source functions are never fully where j⳱2, 3, . . . M.
known because of the variability of mechanical coupling caused by
In Eq. A-3, M is the number of actual frequency domain mea-
vertical variation of formation properties. The source-independent
surements, and the superscript T indicates transpose. The ampli-
formulation described by Eq. A-1 summarizes the conceptual basis
tude of the spectra also can be used as the measurement vector at
for a robust and efficient algorithm that can be used as the forward
the expense of losing phase information. The ordering procedure
computational model for full-waveform inversion. The transfer
that assigns an index, j, to a given measurement is a function of
function T12 depends on both the formation and borehole proper-
frequency for the spectrum of each trace and receiver location.
ties. Specifically, factors that affect the transfer function include:
Real and imaginary parts of one measurement are arranged next to
(a) elastic properties of the formation, (b) borehole radius R, (c)
each other. If the sonic tool consists of NREC receivers, the nor-
acoustic velocity Vf, (d) density ␳f, and (e) quality factors for
malized wavefield has NREC–1 traces because one trace is used to
formation and borehole fluid. Toksöz et al. (1985), Cheng et al.
eliminate the source effect. If the number of frequencies used for
(1986), and Cheng (1989) have quantified some of these factors
each trace is NFREQ, the actual number of measurements used for
using forward-modeling techniques. Because the primary goal of
the inversion is 2M⳱2*NFREQ*(NREC–1).
this paper is to simultaneously invert radial distributions of density
Similarly, the model vector is assembled as
and P- and S-wave velocities, quality factors will not be subject to
quantitative consideration. The normalized wavefield (or the nor- m = 关Vp,1, Vp,2, . . . , Vp,N, Vs,1, Vs,2, . . . , Vs,N, ␳1, . . . , ␳N兴T
malized spectrum) of full waveforms constitutes the measurements . . . . . . . . . . . . . . . . . . . . . . . . (A-4)
entered into the full-waveform inversion algorithm. A cost func- for the simultaneous inversion of P- and S-wave velocities and
tion that enforces the quadratic difference between the simulated density for each radial layer, where Vp, Vs, and ␳ denote P-wave
and measured normalized wavefields is used for inversion. To velocity, S-wave velocity, and density, respectively, in the radial
appraise the robustness of the inversion algorithm, this paper also layers numbered from 1 to N. By denoting the model parameters as
considers a sensitivity study that quantifies the influence of noise, mi, where i⳱1, 2, . . . , 3N, Eq. A-4 can be written as
normalization trace, and regularization parameters.
For the estimation of the radial distribution of elastic properties, m = 关m1, m2, . . . , m3N兴T. . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-5)
density and P- and S-wave velocities are assigned a constant value
within each radial layer in the near-wellbore region. Let m be the Gauss-Newton Fixed-Point Iteration Search. A Gauss-Newton
size-N vector of unknown parameters that fully describes the radial fixed-point iteration search (Gill et al. 1981) is used to determine
distribution of elastic properties, and let mR be a reference vector a stationary point, m, at which the cost function attains a mini-
of the same size as m that has been determined from some a priori mum. This method considers only first-order variations of the cost
information. The estimation (inversion) of m is undertaken by function in the neighborhood of a local iteration point. The corre-
minimizing a quadratic cost function, C(m), defined as (Torres- sponding iterated formula can be written as
Verdín and Habashy 1994) mk+1 = 关JT共mk兲 ⭈ WTd ⭈ Wd ⭈ J共mk兲 + ␭WmT ⭈ Wm兴−1
2C共m兲 = 兵 㛳Wd关d共m兲 − dobs兴 㛳2 − ␹2其 + ␭ 㛳Wm ⭈ 共m − mR兲 㛳2, ⭈ 兵JT共mk兲 ⭈ WTd ⭈ Wd ⭈ 关d共mk兲 − dobs + J共mk兲 ⭈ mk兴
. . . . . . . . . . . . . . . . . . . . . . . . (A-2)
+ ␭WmT ⭈ Wm ⭈ mR其 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-6)
obs
where d is a size-M vector that contains the noisy measure- subject to
ments, and Wd·WTd is the inverse of the data covariance matrix.
This data-weighting matrix describes the estimated variance for li ⱕ mik+1 ⱕ ui, where i = 1, 2, . . . N. . . . . . . . . . . . . . . . . (A-7)
each particular measurement and the estimated correlation be- In these expressions, the superscript k is used as an iteration count,
tween measurements. The parameter ␹ denotes the prescribed and J(m) is the Jacobian matrix of C(m). Upper and lower bounds
value of the enforced quadratic data misfit. A priori estimates of enforced on mk+1 are intended to have the iterated solution yield
the noise in the measurements are used to determine the magnitude only physically consistent results. The fixed-point iteration search
of ␹. In Eq. A-2, d(m) is the measurement vector numerically for a minimum of C(m) is concluded when the measured data have
simulated for specific values of m; Wm·Wm T
is the inverse of the been fit within the prescribed tolerance, ␹2.
model covariance matrix, used to enforce both a quantitative Relative data misfits computed with the formula
degree of confidence in the reference model, mR, and a priori 㛳Wd ⭈ 关d共mk兲 − dobs兴 㛳2
information about m; and ␭ is a Lagrange multiplier or regular- . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (A-8)
ization parameter. 㛳Wd ⭈ dobs 㛳2
The first additive term on the right side of Eq. A-2 drives the are used in this paper to enforce a convergence criterion for
inversion toward fitting the measurements within the desired ␹2 the inversion.
value. The sole presence of such a term in the cost function C(m)
will yield multivalued solutions of the inverse problem as a result
of both noisy measurements and insufficient and imperfect data
sampling. Enforcing an extremely small data misfit may result in SI Metric Conversion Factors
estimated models with exceedingly large norms (Torres-Verdín cp × 1.0* E−03 ⳱ Pa⭈s
and Habashy 1994). The second additive term on the right side of cycles/sec × 1.0* E+00 ⳱ Hz
Eq. A-2 is used to reduce nonuniqueness and to stabilize the in- ft × 3.048* E−01 ⳱ m
version in the presence of noisy and sparse measurements. In this ft3 × 2.831 685 E−02 ⳱ m3
context, the Lagrange multiplier, ␭, controls the relative weight of
in. × 2.54* E+00 ⳱ cm
the two additive terms in the cost function. The developments
considered in this paper make use of a relatively large regulariza- in.3 × 1.638 706 E+01 ⳱ cm3
tion parameter at the outset, which monotonically decreases ac- psi × 6.894 757 E+00 ⳱ kPa
cording to the number of iterations and the observed reduction of *Conversion factor is exact.
the cost function from iteration to iteration.

October 2006 SPE Reservoir Evaluation & Engineering 563


Shihong Chi joined ConocoPhillips in June 2006 as a staff petro- 1991 to 1997, Torres-Verdín held the position of Research Sci-
physicist, working on petrophysical analysis and quantitative entist with Schlumberger-Doll Research; from 1997 to 1999, he
interpretation of 4D seismic data with rock physics models. e- was Reservoir Specialist and Technology Champion with YPF
mail: Shihong.chi@ConocoPhillips.com. He is interested in the (Buenos Aires). Torres-Verdín holds a PhD degree in engineer-
modeling and interpretation of logging data in deviated wells, ing geoscience from the U. of California, Berkeley. He serves as
particularly for acoustic and resistivity anisotropy applications, a Review Chair for SPEJ. Jianghui Wu is a scientist with Baker
as well as the integration of rock physics, seismic, and reservoir Atlas in Houston. His main research areas are analysis of wire-
engineering in time-lapse seismic inversion and applications. line formation tests and formation fluid sampling and the de-
Previously, Chi interned with Schlumberger in 2000 and Baker velopment of a near-wellbore simulator. Previously, he was a
Atlas/Inteq in 2001. He also worked on borehole acoustics, seis- reservoir engineer with the Research Inst. of Petroleum Explo-
moelectrics, and seismic modeling for fracture characteriza- ration and Development, CNPC, from 1996 to 1999. Wu holds
tion and nuclear explosion monitoring as a post-doctoral fel- BS and MS degrees from the China U. of Petroleum and a PhD
low at the Earth Resources Laboratory, Massachusetts Inst. of degree from the U. of Texas at Austin, all in petroleum engi-
Technology from 2004 to 2006. Chi holds a BS degree in geo- neering. Faruk O. Alpak is a reservoir engineer with Shell Intl.
physics from the U. of Science and Technology of China, an MS E&P in Houston. His research interests include parallel reservoir-
degree in exploration geophysics from the U. of Petroleum, simulation techniques, numerical methods, uncertainty-based
East China, and MS and PhD degrees in petroleum engineer- dynamic modeling, inverse problems, numerical optimization,
ing from the U. of Texas at Austin. Carlos Torres-Verdín has and computational electromagnetics. Alpak is an author or
been an associate professor in the Dept. of Petroleum and coauthor of more than 20 journal and conference publica-
Geosystems Engineering at the U. of Texas at Austin since 1999. tions. He holds a BS degree in petroleum and natural gas en-
He conducts research on borehole geophysics, formation gineering from the Middle East Technical U. and MS and PhD
evaluation, and integrated reservoir characterization. From degrees in petroleum engineering from the U. of Texas at Austin.

564 October 2006 SPE Reservoir Evaluation & Engineering

You might also like