Anda di halaman 1dari 20

Sonic Investigations In and

Around the Borehole

J.L. Arroyo Franco Sonic measurements have come a long way since their introduction 50 years ago.
M.A. Mercado Ortiz
Pemex Exploración y Producción The latest advancement in sonic technology delivers the highest quality data seen
Reynosa, Mexico to date, allowing acoustic measurements to characterize mechanical and fluid
Gopa S. De properties around the borehole and tens of feet into the formation.
Chevron Energy Technology Company
San Ramon, California, USA

Lasse Renlie
Statoil ASA
Stjørdal, Norway
Finding and producing hydrocarbons efficiently Formation properties often vary directionally,
and effectively require understanding the rocks so to be completely described, they must be
Stephen Williams
Norsk Hydro ASA and fluids of a reservoir and of surrounding measured in three dimensions. The borehole has
Bergen, Norway formations. Three basic oilfield measurements— a natural, cylindrical 3D coordinate system:
electromagnetic, nuclear and acoustic—have axial, or along the borehole; radial, or perpen-
For help in preparation of this article and in acknowledge- been devised to achieve this end. With advances dicular to the borehole axis; and azimuthal, or
ment of their contributions to the development of the Sonic
Scanner acoustic scanning platform and applications,
in tool design and in data acquisition, processing around the borehole. Variations around and away
thanks to Sandip Bose, Jahir Pabon and Ram Shenoy, and interpretation, each measurement type has from the borehole depend on many factors,
Cambridge, Massachusetts, USA; Tom Bratton and
Adam Donald, Denver, Colorado, USA; Chung Chang, Tarek
evolved to produce more and different information. including the angle the borehole makes with
Habashy, Jakob Haldorsen, Chaur-Jian Hsu, Toru Ikegami, None, perhaps, has evolved more than the sedimentary layering. Axial variations are typical
David Johnson, Tom Plona, Bikash Sinha and Henri-Pierre
Valero, Ridgefield, Connecticut, USA; Steve Chang,
acoustic, or sonic, measurement. of vertical boreholes in horizontal layers, and can
Takeshi Endo, Hiroshi Hori, Hiroshi Inoue, Masaei Ito, In their early days, sonic measurements were indicate changes in lithology, fluid content,
Toshihiro Kinoshita, Koichi Naito, Motohiro Nakanouchi,
Akira Otsuka, Vivian Pistre, Atsushi Saito, Anthony Smits,
relatively simple. They began as a way to match porosity and permeability. Radial rock- and fluid-
Hitoshi Sugiyama, Hitoshi Tashiro and Hiroaki Yamamoto, seismic signals to rock layers.1 Today, sonic property variations arise because of nonuniform
Sagamihara, Kanagawa, Japan; Rafael Guerra and Jean-
Francois Mengual, Rio de Janeiro, Brazil; Dale Julander,
measurements reveal a multitude of reservoir stress distributions and mechanical or chemical
Chevron, Bakersfield, California, USA; Larry O’Mahoney, and wellbore properties. They can be used to near-wellbore alteration caused by the drilling
Chevron, New Orleans, Louisiana, USA; Marcelo Osvaldo
Gennari, Reynosa, Mexico; Pablo Saldungaray, Veracruz,
infer primary and secondary porosity, permea- process. Azimuthal variations can indicate aniso-
Mexico; Keith Schilling, Bangkok, Thailand; Kwasi Tagbor bility, lithology, mineralogy, pore pressure, tropy, which is caused by layering of mineral
and John Walsh, Houston, Texas; Badarinadh Vissapragada,
Stavanger, Norway; Canyun Wang, Beijing, China;
invasion, anisotropy, fluid type, stress magnitude grains, aligned fractures or differential stresses.
Erik Wielemaker, The Hague, The Netherlands; and and direction, the presence and alignment of Improved characterization of compressional
Smaine Zeroug, Paris, France.
fractures and the quality of casing-cement bonds. and shear slownesses in terms of their radial,
Array-Sonic, CBT (Cement Bond Tool), DSI (Dipole Shear
Sonic Imager), ECS (Elemental Capture Spectroscopy), Improvements in sonic measurements are azimuthal and axial variations is now possible
FMI (Fullbore Formation MicroImager), HRLA (High- enhancing our ability to determine some of these with a new sonic tool, the Sonic Scanner acoustic
Resolution Laterolog Array), LSS (Long-Spaced Sonic Tool),
MDT (Modular Formation Dynamics Tester), OBMI (Oil- properties. Accuracy is improving in the basic scanning platform. High-quality waveforms and
Base MicroImager), Platform Express, Sonic Scanner, TLC measurements, which consist of estimates of advanced processing techniques lead to more
(Tough Logging Conditions) and Variable Density are marks
of Schlumberger. compressional- (P-) and shear- (S-) wave slow- accurate slowness estimates, even in unconsoli-
1. Léonardon EG: “Logging, Sampling, and Testing,” in nesses.2 Variations in slowness are also becoming dated sediments and large boreholes, and also to
Carter DV (ed): History of Petroleum Engineering. New
York City: American Petroleum Institute (1961): 493–578.
more fully characterized, leading to an improved reliable through-casing slowness measurements.
2. Slowness, also called interval transit time, is the understanding of how formation properties These improvements result in better characteri-
reciprocal of speed, or velocity. The common unit of change over distance and with direction. zation of subsurface rock and fluid properties,
slowness is microseconds per foot (µs/ft).
meaning more stable wellbores, longer-lasting
completions and enhanced production.

14 Oilfield Review
This article describes the advances in tool
design and resulting data quality of the Sonic
Scanner tool. Examples from the USA, Norway
and Mexico highlight applications that include
determining shear velocities in ultraslow
formations, radial profiling for optimizing
drilling, completion and sampling operations,
fluid-mobility logging, fracture characterization
and imaging away from the borehole.

Engineering Success
More so than electromagnetic and nuclear
logging tools, a sonic tool’s very presence in a
borehole can introduce a bias to the
measurements it acquires. The steel tool housing
is extremely efficient at propagating sonic waves.
Sonic-logging tool designers have minimized this
unwanted effect by isolating the transmitters
from the receivers with insulating materials or
by milling slots and grooves into the steel
sonde (see “History of Wireline Sonic Logging,”
page 32). These efforts were aimed at delaying
undesirable signals and making the tool as
transparent to the measurement as possible.
The Sonic Scanner tool is completely
different from other tools. Its design, material
composition and components were engineered so
that the effects of its presence could be modeled.
These effects can then be incorporated into
predicting the complete tool-borehole-formation
response. These theoretical predictions have
been verified by experimental results in a test
well having known formation properties. As a
result, tool effects can be predicted accurately in
isotropic homogeneous formations, and real-time
corrections can be made at the wellsite.
The transmitter-receiver (TR) geometry and
functionality of the new tool were carefully
designed to provide P-, S-, Stoneley- and flexural-
wave slowness measurements at varying radial
depths of investigation (for a review see “Borehole
Acoustic Waves,” page 34). These modes are
acquired at a logging speed of 1,800 ft/h
[549 m/h]. For the typical scenario with formation
compressional and shear speeds increasing with
distance from the borehole, this is achieved by
increasing TR spacing to probe deeper into the
formation. The Sonic Scanner tool combines this
long-spaced approach with the close TR spacing of
a borehole-compensated arrangement, and also
adds azimuthally distributed receivers. The tool
features 13 axial stations in a 6-ft [1.8-m] receiver
array. Each station has eight receivers placed
every 45° around the tool, for a total of 104 sensors

Spring 2006 15
on the tool.3 A monopole transmitter sits on Each of the three Sonic Scanner monopole tool. This mechanism generates a high-pressure
each end of the receiver array, and another transmitters produces a stronger pressure pulse dipole signal without inducing vibration in the
monopole transmitter and two orthogonally than transmitters in previous sonic tools. With a tool housing. The shaking source can be driven in
oriented dipole transmitters are located farther sharp “click,” they generate clear P- and S-waves, two modes: the traditional dipole source in pulse
down the tool (below). the low-frequency Stoneley mode and the high- mode produces a deep “click”; the new source
frequency energy needed for cement evaluation. also produces a “chirp” with a frequency sweep
Each of the two dipole transmitters is a (bottom left). The chirp mode sustains each
shaking device consisting of an electromagnetic frequency for a longer duration than narrow-
motor mounted in a cylinder suspended in the band dipole sources, providing more dipole
energy to the formation.
As in earlier sonic tools, such as the DSI
Dipole Shear Sonic Imager, the two dipole
Upper Lower Far
sources are oriented orthogonally. One vibrates
monopole monopole monopole in line with the tool reference axis, and the other
Electronics Receiver section Isolator Far transmitter section at 90° to the axis. These devices generate strong
flexural modes—waves that gently shake the
R13 R1 entire borehole the way a person might shake a
X and Y tree from its trunk. Flexural modes propagate up
10 ft dipole and down the borehole and also into the
formation to different depths that depend on
their frequencies. The frequency content—
> The Sonic Scanner tool, with 13 axial stations in a 300 Hz to 8 kHz—of the new chirp dipole source
6-ft receiver array. Each station has eight azimuthally excites flexural modes in all borehole and
distributed receivers, giving the tool 104 sensors. formation conditions, including slow formations,
Three monopole transmitters allow acquisition of and ensures maximum signal-to-noise ratio.
long-spaced and short-spaced data for borehole
The new sonic tool delivers P, S, Stoneley and
compensation at varying depths of investigation.
Two orthogonal dipole transmitters generate flexural-mode waveforms with unprecedented
flexural waves for characterization of shear-wave quality. An example from a typical fast formation
slowness in slow and anisotropic formations. offshore Norway shows waveforms acquired from
the monopole and dipole transmitters (next
page, top). At high frequencies, the monopole
source generates clear P-, S- and Stoneley waves,
while at low frequencies, it generates
predominantly Stoneley waves. The X- and Y-
dipole transmitters generate flexural waves. The
dispersion curves show slowness versus frequency
for the nondispersive shear, slightly dispersive
Stoneley and highly dispersive flexural arrivals.
The low-frequency limit of the flexural-wave
dispersion curve is in line with the slowness of
> the shear head wave and the true shear slowness
The frequency sweep of the Sonic Scanner 100
dipole transmitter. The strong chirp creates a of the formation. The two flexural curves match,
80 indicating absence of azimuthal anisotropy.
wide-band response (inset) that is flat from
Magnitude, dB

about 300 Hz to 8 kHz. Waveforms from the same sources in a slow


60
formation in the USA display evident differences
40 compared with fast-formation results (next page,
400
20
bottom). The high-frequency monopole source

200 0 3. Pistre V, Kinoshita T, Endo T, Schilling K, Pabon J,


10 3 10 4 Sinha B, Plona T, Ikegami T and Johnson D: “A Modular
Pressure, Pa

Frequency, Hz Wireline Sonic Tool for Measurements of 3D (Azimuthal,


0 Radial, and Axial) Formation Acoustic Properties,”
Transactions of the SPWLA 46th Annual Logging
Symposium, New Orleans, June 26–29, 2005, paper P.
–200
Pistre V, Plona T, Sinha B, Kinoshita T, Tashiro H,
Ikegami T, Pabon J, Zeroug S, Shenoy R, Habashy T,
–400 Sugiyama H, Saito A, Chang C, Johnson D, Valero H-P,
0 2 4 6 8 10 12 14 16 18 20 Hsu CJ, Bose S, Hori H, Wang C, Endo T, Yamamoto H
Time, ms and Schilling K: “A New Modular Sonic Tool Provides
Complete Acoustic Formation Characterization,”
Expanded Abstracts, 75th SEG Annual Meeting,
Houston (November 6–11, 2005): 368–371.

16 Oilfield Review
High Frequency Low Frequency
13 13

11 11
P

Receiver number
9 9
S
400
7 7
Stoneley
5 5
300 Stoneley
3 3

Slowness, µs/ft
1 1
0 2 4 0 5 10 15 20 200
Time, ms Time, ms Flexural
X Dipole Y Dipole
13 13 100 Shear
11 11
Receiver number

9 9 0
0 2 4 6 8
7 7
Frequency, kHz
5 5

3 3

1 1
0 5 10 15 20 0 5 10 15 20
Time, ms Time, ms
> Waveforms (left) from a fast formation offshore Norway. Monopole transmitters (top) at high frequencies (left) generate clear
P-, S- and Stoneley waves, and low frequencies (right) generate mostly Stoneley waves. Flexural waveforms generated by the
dipole transmitters (bottom) are recorded on the X (left) and Y (right) receivers. Dispersion analysis (right) shows slightly dispersive
Stoneley data, highly dispersive flexural data and nondispersive shear data. The compressional wave is excited only at frequencies
higher than 8 kHz in this formation, and is not shown on the dispersion curve. [Modified from Pistre et al, reference 3 (SEG).]

High Frequency Low Frequency


13 13

11 11
Receiver number

900
9 9 Stoneley
Slowness, µs/ft

7 7 800
Slow shear
5 5 700
3 3 Flexural Fast shear
600
1 1
500
0 2 4 6 0 10 20 30 0 1 2 3
Time, ms Time, ms Frequency, kHz
Fast Flexural Slow Flexural 210
Leaky
11 compressional
Slowness, µs/ft

11
190
9 9
Receiver number

Compressional
7 7 170

5 5 150
0 2 4 6 8
3 3 Frequency, kHz
1 1
0 10 20 30 0 10 20 30
Time, ms Time, ms
> Waveforms (left) from a slow formation in the USA. The high-frequency monopole source (top left) generates no shear wave
and smaller Stoneley waves than in the fast-formation case. At low frequency, the monopole source (top right) generates
predominantly Stoneley waves. The X- and Y-dipole transmitters generate low-frequency flexural waves compared with the fast
formation. Anisotropy in this formation causes flexural-wave splitting, creating a fast and slow flexural wave (bottom left and
right, respectively). The low-frequency dispersion data (right) include the Stoneley mode and two flexural modes. Higher
frequency dispersion analysis of the P-wave data reveals dispersion—labeled leaky compressional—at higher frequencies.
[Modified from Pistre et al, reference 3 (SEG).]

Spring 2006 17
generates no direct shear wave but does generate Analysis of flexural-wave dispersion curves concentrations.6 These simplified relationships
leaky-compressional waves. At low frequencies, from the Sonic Scanner tool classifies formations between dispersion curves are valid when only
the monopole source again generates Stoneley according to anisotropy type by comparing one physical mechanism controls wave behavior.
waves, but, in addition, there is a strong leaky- observed dispersion curves to those modeled When multiple mechanisms are involved, such as
compressional wave generated. The X- and Y- assuming a homogeneous isotropic formation if both stress-induced and intrinsic anisotropy
dipole transmitters generate flexural waves with (below). In a homogeneous isotropic formation, are present, the curves can be different.
the characteristic low-frequency response of a shear waves do not split into fast and slow In addition to acquiring openhole measure-
slow formation. The dispersion data include the components, so the two observed flexural-wave ments in isotropic, anisotropic, homogeneous and
slightly dispersive Stoneley mode and the leaky- dispersion curves have identical slowness-versus- inhomogeneous formations, the Sonic Scanner
compressional wave, but no shear head wave, as frequency signatures, and overlie the modeled tool provides high-quality results behind casing.
expected in a slow formation. In the absence of a curve. In cases of intrinsic anisotropy, such as The improved tool design records waveforms
shear head wave, the shear slowness is estimated shales or fractured formations, the fast and slow through casing with high signal-to-noise ratio.
from the low-frequency limit of the flexural mode. shear-wave dispersion curves are separate Powerful transmitters and large bandwidth allow
The flexural mode is not as dispersive as in a everywhere and tend to the true slowness at zero acquisition of formation slowness data through
fast formation, but more dispersive than that frequency.5 In formations that have undergone casing and cement of varying thickness.
expected from a homogeneous, isotropic forma- drilling-induced damage and are near failure but The ability to measure formation properties
tion. At low frequency, the two flexural-wave are otherwise homogeneous and isotropic, the through casing allows companies to monitor the
dispersion curves level off at different slow- two dispersion curves are identical but show mechanical effects of production on the produc-
nesses, indicating azimuthal anisotropy. The much greater slowness at high frequencies than ing formation. Many formations undergo compac-
flexural waveforms have been mathematically the modeled dispersion for a homogeneous tion, weakening or other changes with time as a
rotated into fast and slow shear-wave directions.4 isotropic formation. In formations with stress- result of pressure depletion or water injection.
induced anisotropy, the fast and slow shear-wave In an example from a Statoil well in the North
dispersion curves cross. This characteristic Sea, Sonic Scanner data were acquired in both
feature is caused by near-wellbore stress 8.5-in. open hole and behind 8-in. OD casing
before any production (next page). The openhole
logs in the zone of interest indicate a slower,
softer formation between X,296 and X,305 m.
Homogeneous Isotropic Inhomogeneous Isotropic The caliper log flags a washout in this interval.
Damaged formation,
When compared with the openhole logs, the
near failure cased-hole compressional and shear slownesses
Slowness

Slowness

are markedly similar, even through the washed-


out zone. The dispersion curves in the two cases
are also similar.
Fast shear Fast shear
In the Middle East, the Sonic Scanner tool
Frequency Frequency has been used multiple times to acquire slowness
through 133⁄8-in. casing in hole sizes larger than
Homogeneous Anisotropic Inhomogeneous Anisotropic
20 inches. In each case, despite poor cement,
Intrinsic anisotropy: Stress-induced good shear-wave slowness data were acquired
shales, layering, fractures anisotropy over the entire interval, and adequate
Slowness

Slowness

compressional slowness was recorded over at


least half the interval.
Slow shear Slow shear
Fast shear Fast shear
The Sonic Scanner tool not only obtains
slowness results behind casing, but can also
Frequency Frequency simultaneously evaluate the quality of the
> Flexural-wave dispersion curves for classifying formation anisotropy and cement bond and the top of cement. Signals
inhomogeneity. In a homogeneous isotropic medium (top left), observed recorded by receivers 3 and 5 ft [0.9 and 1.5 m]
dispersion curves for flexural waves recorded on orthogonal dipole receivers from the two near monopole transmitters are
(red and blue curves) match modeled flexural-wave dispersion (black
circles). In an inhomogeneous isotropic formation (top right), both observed processed to produce a discriminated attenua-
dispersion curves show greater slowness with increasing frequency than the tion measurement that is free of tool-
homogeneous isotropic model. Greater slowness with increasing frequency normalization fluid effects and pressure and
indicates that the near-borehole region has become slower, a sign of temperature drifts. The results are comparable
damage all around the borehole. In a homogeneous anisotropic medium
(bottom left), such as one with intrinsic anisotropy, the fast flexural-wave to those of the CBT Cement Bond Tool, but are
dispersion curve (red) matches the homogeneous isotropic model (to a first also corrected for casing and cement properties.
approximation), while the slow flexural-wave dispersion curve (blue) has the Evaluation of well integrity and formation
same shape but is translated to higher slowness. In an inhomogeneous
properties in the same tool run can avoid
anisotropic medium (bottom right), the two observed flexural-wave
dispersion curves cross. This phenomenon is a result of near-wellbore stress separate logging runs and reduce rig-time and
concentration, and indicates stress-induced anisotropy. mobilization costs.

18 Oilfield Review
Washout
Bit Size
6 in. 16 Waveform Waveform
Caliper STC Coherence Variable Density Log Bit Size STC Coherence Variable Density Log

Depth, m

Depth, m
6 in. 16 0 µs 10,000 6 in. 16 0 µs 10,000
Gamma Ray Shear Slowness Arrival Time Gamma Ray Shear Slowness Arrival Time
0 gAPI 150 80 µs/ft 540 0 µs 10,000 0 gAPI 150 80 µs/ft 540 0 µs 10,000
X,290 X,290

X,300 X,300

X,310 X,310

X,320 X,320

300 300 300 300

250 250 250 250


Slowness, µs/ft

Slowness, µs/ft
200 200 200 200
Amplitude, dB

Amplitude, dB
150 150 150 150

100 100 100 100

50 50 50 50

0 0 0 0
0 2,000 4,000 6,000 8,000 10,000 0 2,000 4,000 6,000 8,000 10,000
Frequency, Hz Frequency, Hz
> Openhole (left) and cased-hole (right) results in a Statoil North Sea well. The Sonic Scanner tool measures P-, S- and Stoneley-wave
slownesses in open hole and behind casing, even where the caliper (Track 1) indicates a washed-out zone (between X,296 and X,305 m)
in the openhole logs. Flexural-mode slowness displayed in Track 2 of each set is more sharply defined, with a narrower color band, in
the cased-hole example than in the openhole logs. In the dispersion curves (bottom), compressional-wave slowness is in dashed green
and shear-wave slowness is in dashed blue.

Extreme Slowness frequency limit, the leaky-P dispersion curve complications that make sonic logging chal-
Some formations are so slow that not only is the tends toward the P-wave slowness, and at the lenging.8 The formation lithology is diatomite
S-wave slowness greater than that of the mud, high-frequency limit, it reaches the borehole- and cristobalite—forms of opalized silica.
but the P-wave slowness approaches that of the fluid slowness.7 Permeability is low, averaging 2 mD. From earlier
mud. In these circumstances, the P-wave loses The Antelope formation in the Cymric oil field studies, compressional-wave slowness in this
energy to the formation, in what is known as a in the San Joaquin Valley, California, is such a formation is known to approach 200 µs/ft, which is
leaky-P mode, and is dispersive. At the low- case, combining extreme slowness with other near the slowness of the mud wave, and shear-

4. The X- and Y-dipole sources are separated by 1 ft. While 5. For anisotropy to be identified in this way, the anisotropy 7. Valero H-P, Peng L, Yamamoto M, Plona T, Murray D and
this avoids electrical cross-talk, it also means that symmetry axis must be perpendicular to the borehole Yamamoto H: “Processing of Monopole Compressional in
waveforms must be shifted by 1 ft before Alford rotation. axis. For example, crossed-dipole logging tools in Slow Formations,” Expanded Abstracts, 74th SEG
This reduces the number of collocated waveforms from vertical wells can detect anisotropy caused by aligned International Meeting, Denver (October 10–15, 2004):
13 to 11. vertical fractures, and in horizontal wells can detect 318–321.
Alford RM: “Shear Data in the Presence of Azimuthal anisotropy caused by horizontal laminations. 8. Walsh J, Tagbor K, Plona T, Yamamoto H and De G:
Anisotropy: Dilley, Texas,” Expanded Abstracts, 56th SEG 6. Sinha BK and Kostek S: “Stress-Induced Azimuthal “Acoustic Characterization of an Extremely Slow
Annual International Meeting, Houston (November 2–6, Anisotropy in Borehole Flexural Waves,” Geophysics 61, Formation in California,” Transactions of the SPWLA 46th
1986): 476–479. no. 6 (November-December 1996): 1899–1907. Annual Logging Symposium, New Orleans, June 26–29,
Winkler KW, Sinha BK and Plona TJ, “Effects of 2005, paper U.
Borehole Stress Concentrations on Dipole Anisotropy
Measurements,” Geophysics 63, no. 1 (January-February
1998): 11–17.

Spring 2006 19
Compressional ∆T
50 µs/ft 360
Stoneley ∆T
200 µs/ft 1,200
Time-Based Anisotropy
Sonde Deviation
100 % 0
-10 deg 90
∆T-Based Anisotropy
Hole Azimuth Slow Waveform
0 % 100
Offline 0 deg 360 0 µs 30,000 Slowness- Slowness-
Energy Anisotropy Flag, % Frequency Analysis Frequency Analysis
Total Azimuth Fast Waveform
Maximum 0 deg 360 0 2 4 6 16 0 µs 30,000
Energy Coherence Energy Coherence Energy
Hole Diameter ∆T-Based Slow Shear Window Stop
0 100
5 in. 20 Azimuth Uncertainty 200 µs/ft 1,200 0 µs 30,000
Minimum
Energy Gamma Ray Fast Shear Azimuth ∆T-Based Fast Shear Window Start Fast Shear Fast Shear Slow Shear Slow Shear
0 100 0 gAPI 150 -90 deg 90 200 µs/ft 1,200 0 µs 30,000 200 µs/ft 1,200 200 µs/ft 1,200 200 µs/ft 1,200 200 µs/ft 1,200
Depth, ft

1,250

1,300

1,350

1,400

1,450

1,500

1,550

1,600

> Shear-wave slownesses computed from flexural-wave logging in the extremely slow Antelope formation, Cymric field, California. In the diatomite zone,
down to 1,500 ft, shear slownesses in Track 3 average 700 µs/ft and approach 900 µs/ft in some intervals. Below that, shear slownesses decrease to about
400 µs/ft. The large separation between minimum and maximum offline energy in the depth track indicates anisotropy. Track 1 shows gamma ray (green),
hole diameter (yellow), hole azimuth (light blue) and azimuth of the continually rotating tool (dark blue). Azimuth of the fast shear wave, shown in Track 2
(red), is relatively constant in the anisotropic zone above 1,500 ft, in spite of continual tool rotation. In addition to fast (red) and slow (blue) shear slownesses
estimated from dispersion analysis, Track 3 shows Stoneley-wave slowness (black), P-wave slowness (green curve), and slowness-based (left edge of
track) and time-based (right edge of track) anisotropies. Track 4 shows the waveforms and time windows used for flexural-wave analysis. Slowness-time-
coherence projections for fast and slow shear are shown in Track 5 and Track 7, respectively. Slowness-frequency-analysis (SFA) projections for fast and
slow shear are shown in Track 6 and Track 8, respectively. (Modified from Walsh et al, reference 8.)

20 Oilfield Review
wave slowness exceeds 800 µs/ft in some sections.9
SFA Energy
Nine-component vertical seismic profiles and
crossed-dipole sonic logs have detected changing
anisotropy magnitude and direction with depth
Waveforms from 15,061 ft Shear Slowness
and around the field.10 Knowledge of acoustic
velocities and of anisotropy can be important for 13 100 µs/ft 400
12
designing fracture stimulations and enhanced oil- 11
10

Waveform number
recovery operations.
9
Measurements with the Sonic Scanner tool 8
provide new insight into the acoustic behavior of 7
6
these complex rocks. Waveforms were recorded in 5
an interval from 972 to 1,650 ft [296 to 503 m] in a 4
3
well near the crest of the Cymric structure. In the 2 15,000
diatomite zone, down to 1,500 ft [457 m], shear 1

slowness estimated from flexural-mode dispersion 0 2,500 5,000 7,500 10,000


processing is at least as great as that found in Time, µs

earlier logging programs, averaging 700 µs/ft and


approaching 900 µs/ft in some intervals (previous 400 300
page). Below that, shear slowness decreases to
350
about 400 µs/ft in the cristobalite zone. 250
15,050
300
Much of the logged interval exhibits
Slowness, µs/ft

200
250

Amplitude, dB
azimuthal anisotropy, as indicated by the large
separation between minimum and maximum 200 150

offline energy, and also between the fast and slow 150
100
shear-wave slownesses. Anisotropy magnitude 100
ranges from 4 to 8%, consistent with results of 50
50
previous studies.11 Slowness anisotropy is 15,100
0 0
calculated by dividing the difference between 0 2,000 4,000 6,000 8,000
fast and slow shear slownesses by their average. Frequency, Hz
The azimuth of the fast shear direction is
between N35W and N15W, in general agreement
with previous studies.12
Along with the typical fast and slow shear-
slowness curves and slowness-time-coherence
(STC) projections seen in many sonic-log plots,
displays of Sonic Scanner data include new
quality-control tracks showing slowness-
frequency analysis (SFA). To create SFA plots, a > Construction of a slowness-frequency-analysis (SFA) log for controlling the quality of shear-slowness
dispersion curve is generated at each depth estimation from flexural waves. Dipole flexural waveforms at each depth (top left) are analyzed for
their slowness at varying frequencies. Resulting data are plotted on a slowness-frequency plot
using the recorded dipole flexural waveforms
(bottom left), with circle size indicating amount of energy. Energies are color-coded and projected
(above right).13 The dispersion curve is projected onto the slowness axis. The color strip is plotted at the appropriate depth to create a log (right).
onto the slowness axis, and this projection is The slowness estimate from dispersive STC processing is plotted as a black curve. If this matches the
zero-frequency limit of the SFA projection, the slowness estimate is good.
9. Hatchell PJ, De GS, Winterstein DF and DeMartini DC:
“Quantitative Comparison Between a Dipole Log and
VSP in Anisotropic Rocks from Cymric Oil Field,
California,” Expanded Abstracts, 65th SEG Annual
International Meeting, Houston (October 8–13, 1995):
13–16. plotted in a log versus depth presentation, In this extremely slow formation, the
10. De GS, Winterstein DF, Johnson SJ, Higgs WG and similar to the way an STC projection is monopole source does not excite a compressional
Xiao H: “Predicting Natural or Induced Fracture
Azimuths from Shear-Wave Anisotropy,” paper
constructed. The estimated slowness log from head wave, but rather a strong leaky-P mode.
SPE 50993-PA, SPE Reservoir Evaluation & Engineering 1, dispersive STC processing is overlaid on the SFA Compressional slowness must therefore be
no. 4 (August 1998): 311–318.
projection, and if the estimated slowness estimated from dispersive STC processing,
11. De et al, reference 10.
12. Hatchell et al, reference 9.
matches the low-frequency limit of the SFA analogous to the technique for determining
13. Plona T, Kane M, Alford J, Endo T, Walsh J and Murray D: projection, the quality of the slowness estimate is shear slowness from flexural modes.
“Slowness-Frequency Projection Logs: A New QC high. In azimuthally anisotropic formations, SFA Compressional slowness is estimated at 192 µs/ft
Method for Accurate Sonic Slowness Evaluation,”
Transactions of the SPWLA 46th Annual Logging projections may be plotted for both the fast and
Symposium, New Orleans, June 26–29, 2005, paper T. slow shear directions.

Spring 2006 21
Dispersive STC Processing
Traditional STC Processing 3.5 kHz to 6.5 kHz
Washout Depth = 1,470 ft Depth = 1,470 ft
Gamma Ray Compressional Compressional 225 225
Slowness Slowness Slowness-
Measured depth, ft

0 gAPI 150 Monopole Leaky-P Frequency Time

Slowness, µs/ft

Slowness, µs/ft
Bit Size Processing Processing Analysis
2,000 µs 4,000 200 200
4 in. 14 140 µs/ft 240 140 µs/ft 240 140 µs/ft 240
Waveform
Caliper Coherence Coherence Energy Amplitude
175 175
4 in. 14

150 150
1,470
1,000 2,000 3,000 4,000 1,000 2,000 3,000 4,000 5,000 6,000
Time, µs Time, µs
1,480
Depth = 1,510 ft Depth = 1,510 ft
225 225
1,490

Slowness, µs/ft

Slowness, µs/ft
200 200
1,500

175 175
1,510

150 150

1,000 2,000 3,000 4,000 5,000 1,000 2,000 3,000 4,000 5,000 6,000
Time, µs Time, µs
> Estimating compressional slowness by processing leaky-P dispersion data in the slow Antelope formation (left). Traditional monopole processing in
Track 2 does not give slowness estimates as reliably as does dispersive STC processing (Track 3). STC plots (right) from two different depths show the
improved coherence delivered by dispersive STC processing (right) compared with traditional STC processing (left). Track 4 shows slowness-frequency
analysis (SFA) using leaky-P dispersion data, such as those shown in the dispersion curves (below). (Modified from Walsh et al, reference 8.)

220 300 220 300 understood, and decisions can be made


Depth = 1,470 ft Depth = 1,510 ft
210 210 regarding how to take advantage of or mitigate
250 250
the situation.
200 200
Slowness, µs/ft

In a recent exploration well in the South


Slowness, µs/ft

200 200
Amplitude, dB
Amplitude, dB

Compressional slowness
190 190 from monopole leaky P Timbalier area of the Gulf of Mexico, Chevron
150 150
Compressional slowness successfully penetrated a target sandstone. In
180 from monopole leaky P 180
100 100 other wells, the same formation had presented
170 170
completion challenges, so the logging program in
160 50 160 50
this well included measurements to assess its
150 0 150 0 mechanical properties.
0 2,000 4,000 6,000 8,000 10,000 0 2,000 4,000 6,000 8,000 10,000
Frequency, Hz Frequency, Hz Radial profiles of compressional and shear
slownesses can reveal important information
> Dispersion curves for compressional arrivals in the upper, diatomite zone (left) and the lower,
cristobalite zone (right). Compressional-wave slowness is estimated by the slowness of the
about the state of the formation near the
leaky-P mode at low frequency. [Modified from Walsh et al, reference 8]. borehole. Radial variation in compressional
slowness is revealed by examining the difference
in P slowness detected by the receiver array from
the near and far monopole transmitters. Rays
in the shallow diatomite section and at 175 µs/ft Radial Profiles of Slowness Variation from the near transmitter sample the altered
in the cristobalite section (top). Variations in formation properties may be zone near the borehole, while rays from the far
Following on the initial success of the Sonic natural or induced by the drilling process, and transmitter sample the unaltered zone, also
Scanner tool, Chevron is planning to run the tool may be beneficial or detrimental to the E&P called the far field.
in more wells in this field in 2006. Sonic activity at hand. By fully characterizing P- and A clear picture of radial variation emerges
velocities will support microseismic fracture- S-wave slownesses in a large volume around the when the P-wave data from all three transmitters
mapping techniques.14 borehole, the cause of the variation can be and 13 receivers undergo tomographic recon-
struction.15 This inversion technique uses ray

22 Oilfield Review
tracing to calculate signal arrival times at all the
Compressional
sensors, and updates an initially homogeneous Fast Shear Slow Shear Compressional

Measured depth, ft
180 µs/ft 80 Differential Differential Differential
formation model to create a final model that Hole Diameter Slow Shear
satisfies the observed data. To visualize the 0 % 25 0 % 25 0 % 25
9 in. 19 300 µs/ft 100
Distance from Gamma Distance from Distance from
resulting compressional-slowness radial profile, Density Fast Shear Ray
Borehole Center Borehole Center Borehole Center
the differential percentage between observed 1.7 g/cm3 2.7 300 µs/ft 100 2 ft 0 10 gAPI 110 0 ft 2 0 ft 2
slowness and far-field slowness is color-coded
and plotted against radial distance from the
borehole wall (right).
X,480
Data from this Chevron well showed that the
sandstones of interest exhibited radial variations
in compressional slowness approaching 15% at
the borehole wall and extending up 1 ft [30 cm]
into the formation. However, quantifying the X,490
radial P-wave slowness variation alone does not
identify its cause. Compressional-slowness
variations can be caused by fluid changes, such
as invasion of drilling fluid, or by radial changes
X,500
in stress or formation strength. Additional
information from the shear-slowness radial
profile could help distinguish between these.
Shear-slowness radial profiles are constructed
in a multistep procedure.16 Semblance processing X,510
of flexural waveforms at low frequencies provides
an initial estimate of formation elastic param-
eters. These parameters are used to model a
homogeneous isotropic formation. Differences
X,520
between measured and modeled slownesses at a
large selection of frequencies form the input to
an inversion procedure that yields the actual
flexural-slowness radial profile. The results are
plotted with colors that represent the amount of X,530
difference between observed slowness and the
slowness of the unaltered, far-field formation.
In the South Timbalier case, the shear-
slowness radial profile shows a large difference
X,540
in near-wellbore slowness compared with far-
field slowness. Flexural-wave dispersion curves

14. Bennett L, Le Calvez J, Sarver DR, Tanner K, Birk WS,


Waters G, Drew J, Primiero P, Eisner L, Jones R, Leslie D,
Williams MJ, Govenlock J, Klem RC and Tezuka K: X,550
“The Source for Hydraulic Fracture Characterization,”
Oilfield Review 17, no. 4 (Winter 2005/2006): 42–57.
15. Zeroug S, Valero H-P and Bose S: “Monopole Radial
Profiling of Compressional Slowness,” prepared for
presentation at the 76th SEG Annual International
Meeting, New Orleans, October 1–3, 2006. > Compressional and shear radial profiling in a Chevron Gulf of Mexico well. P-wave data from all
Hornby BE: “Tomographic Reconstruction of Near- three transmitters and 13 receivers are input to tomographic reconstruction based on tracing rays
Borehole Slowness Using Refracted Sonic Arrivals,”
Geophysics 58, no. 12 (December 1993): 1726–1738.
through a modeled formation with properties that vary gradually away from the borehole. The
percentage difference between observed compressional slowness and slowness of the unaltered,
16. Sinha BK: “Near-Wellbore Characterization Using Radial
Profiles of Shear Slowness,” Expanded Abstracts, far-field formation is plotted on color and distance scales to indicate the extent of difference away
74th SEG Annual International Meeting, Denver from the borehole (Track 6). In these sandstones, compressional slowness near the borehole varies
(October 10–13, 2004): 326–331. by up to 15% from far-field slowness, and the variation extends to 1 ft from the borehole wall. Shear-
wave radial profiles appear in Tracks 3 and 5 for the fast and slow shear differences from far-field
slowness, respectively. Large differences, attributed to plastic yielding in the near-wellbore region,
are shown in red, and extend out to about 10 in. from the borehole wall. These differences occur only
in the sandstone intervals, identifiable from the gamma ray log in Track 4.

Spring 2006 23
800 yielding of grain contacts. The caliper shows no
wellbore enlargement through this zone, so the
damaged material has not yet fallen into the
700
borehole, but the increase in shear slowness near
the borehole wall indicates that it is near failure.
The Sonic Scanner data indicate a wide zone of
Slowness, µs/ft

600
damage that will require extra care when the
time comes to design a well completion.
500 Compressional and shear radial profiles bring
new information not previously available from
any logging tool. Borehole imaging tools and
400 calipers have been able to deliver images or
evidence of drilling-induced borehole irregular-
ities such as breakouts and fractures, but are
0 1,000 2,000 3,000 4,000 5,000
useful only after the borehole shape has
Frequency, Hz
changed. The Sonic Scanner tool probes deep
into the formation to reveal mechanical damage
800
beyond the borehole wall.
Radial profiling may also help to fine-tune
700 ~20% shear alteration programs for acquisition of fluid samples. In an
Slowness, µs/ft

example from the North Sea, Sonic Scanner


600
compressional radial profiles were computed for
500
two intervals from which samples were
subsequently acquired using the MDT Modular
400 Formation Dynamics Tester. Zone A showed little
difference between near-wellbore and far-field
slowness (next page). Two fluid samples were
1 2 3 4 5 6 7
Alteration radius/borehole radius
taken from this interval after pumping times of
> Comparison of flexural-wave dispersion seen in a South Timbalier well
75 and 80 minutes and no sampling problems. In
with modeled results (top). Observed flexural-wave slownesses (red and Zone B, the radial profile indicated formation
blue circles) show much larger dispersion than the model for a homogeneous damage out to 12 in. from the borehole wall.
isotropic formation (blue curve). The large difference at higher frequencies During the attempt to obtain a fluid sample, the
indicates near-wellbore damage. Stoneley-wave slownesses appear as probe on the sampling tool became plugged, and
green circles. In the bottom figure, the difference between observed and
modeled flexural slowness is plotted against distance, in borehole-radius ratio no sample was obtained.
units. The difference between observed and modeled flexural slowness Formation damage does not necessarily mean
amounts to 20% out to a distance equivalent to about two borehole radii. that samples cannot be acquired, but sampling in
these zones may have an increased risk of tool
plugging or sticking. To minimize these risks,
sampling from potentially damaged zones should
also indicate a high degree of near-wellbore The radial heterogeneity in shear slowness be delayed and attempted later in the sampling
alteration (above). The analysis is complicated rules out invasion or other fluid-related causes of program, so that samples from other intervals
somewhat by the addition of anisotropy; the fast near-wellbore alteration, because shear waves can be collected first with less risk.
shear and slow shear waves exhibit distinct are almost insensitive to changes in pore fluid.
differences relative to the unaltered, far-field Fluid-related changes would cause only Characterizing Permeable Zones and Fractures
slowness. In the sandstones, both fast and slow compressional-slowness radial variation. The Petrophysicists and reservoir engineers have long
shear slownesses are up to 20% greater than the measurable radial variation in shear slowness sought a continuous measurement of permeability
far-field slowness in a zone roughly 10 in. [25 cm] indicates that the formation has undergone to optimize well completions and production
from the borehole wall. mechanical damage in the form of plastic scenarios, but continuous permeability is one of
the most difficult properties to measure in an oil
17. Brie A, Endo T, Johnson DL and Pampuri F: “Quantitative 18. Kimball CV and Endo T: “Quantitative Stoneley Mobility
Formation Permeability Evaluation from Stoneley Inversion,” Expanded Abstracts, 68th SEG Annual
well. Using empirical relationships calibrated to
Waves,” paper SPE 49131, presented at the SPE Annual International Meeting and Exhibition, New Orleans core measurements, permeability or mobility—
Technical Conference and Exhibition, New Orleans, (September 13–15, 1998): 252–255.
September 27–30, 1998.
the ratio of permeability to viscosity—can be
Liu H-L and Johnson DL: “Effects of an Elastic
Membrane on Tube Waves in Permeable Formations,” inferred from other measurements such as
Journal of the Acoustic Society of America 101, no. 6 porosity or nuclear magnetic resonance logs.
(June 1997): 3322–3329.

24 Oilfield Review
Direct measurements can be made with wireline
Distance from Distance from
formation testers at isolated points along the well, Borehole Center Borehole Center
or on core, but these require additional logging Caliper 0 ft 3 0 ft 3
6 in. 16 Compressional
runs and coring costs. Stoneley-wave analysis is a Position Slowness Gradient Slowness Differential
Gamma Ray
powerful technique that delivers a direct, of MDT Tool
0 gAPI 100 0 % 10 0 % 10
continuous measurement of mobility along
the well.17
The idea of measuring mobility from the
Stoneley wave was first expressed in the 1970s,
but proved difficult in practice. Many attempts
have been made to develop empirical correla-
tions between permeability and Stoneley
attenuation, but these methods required
calibration with other information and neglected
several important factors, such as mudcake
permeability and the presence of the tool itself.
Approaches that simplified the complex behavior
of Stoneley waves were seldom successful, but an A
inversion method that uses a model derived from
full Biot poroelastic theory reliably determines
pore-fluid mobility from Stoneley waveforms.18
For application with Sonic Scanner data, the full
Biot inversion technique was extended to
incorporate tool response.
The full Biot inversion scheme requires
several borehole, mudcake and formation
parameters to evaluate fluid mobility using
Stoneley-wave data. The list includes: hole
diameter; mud slowness, attenuation and
density; formation P and S slowness, density and
porosity; grain modulus; pore-fluid modulus and
B
density; and mudcake density, bulk modulus,
shear modulus, thickness and membrane
stiffness. The computation outputs fluid mobility
and associated error ranges.
This inversion technique has been available
for several years, but application has not always
been successful because the inversion requires
extremely low-frequency Stoneley waves—down > A compressional-wave radial profile indicating intervals of successful and
to 300 Hz. Data with this frequency content have risky fluid sampling. In interval A, the compressional-wave radial profile
not been available in the past, because earlier (Track 3) shows a small differential between near-wellbore slowness and
sonic tools interacted negatively with low- far-field slowness. There is little near-wellbore alteration in the zone where
the MDT Modular Dynamics Formation Tester successfully collected two
frequency signals and required filtering to formation-fluid samples. In Track 3, the amount of slowness difference
remove frequencies below 1,500 Hz. Now, the between near and far field is indicated by gold and brown color intensity,
wideband sources of the Sonic Scanner tool while depth of alteration is indicated by the horizontal length of the colored
generate strong Stoneley waves with reliable low- area. In interval B, the compressional-wave radial profile shows darker
colors, indicating a higher degree of near-wellbore alteration extending
frequency content for mobility calculations. farther away from the borehole. In this zone, the MDT probe became
An example from a Statoil well in the plugged and was not able to collect any formation-fluid samples. Track 2
Haltenbanken area of the Norwegian Sea shows displays the slowness gradient obtained from the tomographic reconstruction.
good correlation between mobilities calculated The gradient indicates the difference in slowness from one slowness-
model cell to the next, moving away from the borehole in small increments.
from Stoneley waves and those measured by
MDT pretests. Input values of formation and
fluid properties of a zone near the oil/water
contact were determined with logs from the
Platform Express integrated wireline logging

Spring 2006 25
tool. The MDT results from eight drawdown
Porosity Compressional ∆T
pretests and one tight pretest correlate closely
1 m3/m3 0 300 µs/ft 0
with mobilities extracted from Stoneley-wave Gamma Ray Shear Slowness Shale
analysis (right). MDT Mobility
0 gAPI 150 300 µs/ft 100
Sand
The continuous mobility log exhibits high Caliper Stoneley Slowness 1 mD/cP 10,000

mobility inside sand packages and low mobility 6 in. 16 300 µs/ft 200 Mobility Error Bar Bound Water
near shale streaks and at the depth of the tight Shale Volume Reconstructed Stoneley Mobility Error Oil

Depth, m
MDT pretest. Because the Sonic Scanner 0 m3/m3 1 300 µs/ft 200 1 mD/cP 10,000
Water
mobility results are somewhat sensitive to a few Density Mud Slowness Stoneley Mobility
1.96 g/cm3 2.96 240 µs/ft 40 1 mD/cP 10,000 Coal
parameters that are not well constrained by
logging measurements, such as mud slowness,
mud attenuation and mudcake stiffness, tests
were conducted to study the effect of uncertainty
in these parameters on the mobility error bars.
The continuous mobility log shown is the one
with the least uncertainty. X,X00

When the borehole is in good condition,


continuous mobility logs from Stoneley waves
can be used to obtain a quick permeability
estimate for selecting sampling points and
perforation intervals, and may function as a
supplement to core or formation-tester permea-
bility points over an extended interval.
Stoneley waves can also be used to
characterize permeability associated with open
fractures. In the US Rocky Mountains, for
example, hard-rock reservoirs depend on
hydraulically induced fractures for economic
production. However, the highly unequal in-situ
stresses in the region give rise to natural fractures
too. If natural fractures are encountered in a well,
cementing and stimulation designs must be
adjusted to prevent cement from entering the
natural-fracture system. For example, fiber-based
treatments for both cementing and stimulation X,X50
can be used to reduce fluid losses.19 Stimulation
programs need to take into account the
magnitude and direction of the principal stresses.
Optimizing the completion design requires
knowledge of the fracture and stress
characteristics around the wellbore and in
the formation.
An open fracture intersecting a borehole > Comparing fluid-mobility values from MDT pretests with those from
causes Stoneley waves to reflect and attenuate.20 Stoneley-wave processing in a Statoil well in the Haltenbanken area of the
Norwegian Sea. In Track 3, continuous fluid-mobility values (blue curve)
Analysis of Stoneley waveforms quantifies these and uncertainties (gray shading) estimated from Stoneley-wave analysis
changes, which are input to an inversion for correlate well with discrete mobility values obtained from MDT drawdown
fracture aperture.21 However, washouts, borehole pretests (red dots). The two measures of mobility match even at the tight
MDT pretest at X,X42.15 m, where the Stoneley mobility also shows an
rugosity and abrupt changes in lithology also can
extremely low value. Porosity, gamma ray, density, caliper and shale volume
cause Stoneley reflections, and should be are plotted in Track 1. Track 2 shows acoustic slownesses. Track 4 displays
considered in the analysis.22 relative volumes of lithology and fluids.
An example of successful application of this
method comes from Colorado, USA.23 In this gas
reservoir, porosity ranges from 3 to 7% and
permeability is in the microdarcies. Stoneley-
wave analysis quantified the aperture and
permeability of fractures that were also seen on

26 Oilfield Review
∆T Stoneley
250 µs/ft 150
Bit Size
4 in. 14
Offline Fracture Permeability
Energy Caliper
100,000 mD 1010
4 in. 14
Maximum Fracture Porosity
Energy Washout 0.1 ft3/ft3 0 FMI Image
0 100 Modeled Stoneley Resistive Conductive
Fracture Width Fracture Trace Length Stoneley Variable
Minimum 250 µs/ft 150 0 in. 0.5 10 µs/ft 0 Density Log Tezuka Model
Energy Orientation North
0 100 S-Se Stoneley Aperture Stoneley Permeability 0 µs 20,440 0 µs 20,440 0 120 240 360

X,100
Depth, ft

X,200

X,300

X,400

> Identifying permeable fractures in Colorado using Stoneley waves. The fracture aperture, or amount
of opening, computed from Stoneley-wave reflection and transmission is displayed in Track 2. Track 3
shows fracture permeability computed from the Track 2 apertures. Zones containing permeable
fractures correlate with zones in which the FMI logs (Track 6) show fractures. The same zones appear
as anisotropic, with large offline energy differences (depth track), and also show large differences
between measured Stoneley slowness and slowness modeled for an elastic, impermeable formation
(orange shading, Track 1). Track 4 shows measured Stoneley waveforms, with amplitude reduction in
the fractured zones. Track 5 shows waveforms generated by the Tezuka model of reference 22.
(Modified from Donald and Bratton, reference 23.)

FMI Fullbore Formation MicroImager images Shear-Wave Directions in Mexico may be different in one horizontal direction
(above). With the broad range of Stoneley-mode Small directional variations in formation compared with the orthogonal horizontal
frequencies acquired by the Sonic Scanner properties can have a major impact on drilling direction. This phenomenon, called elastic
tool, these open natural fractures can be and completion strategies, but these may be anisotropy, occurs in most sedimentary rocks and
reliably characterized. difficult to measure. For example, sonic velocities is caused by layering, aligned fractures or stress

19. Bivins CH, Boney C, Fredd C, Lassek J, Sullivan P, 20. Hornby BE, Johnson DL, Winkler KH and Plumb RA: 21. Endo T, Tezuka K, Fukushima T, Brie A, Mikada H and
Engels J, Fielder EO, Gorham T, Judd T, “Fracture Evaluation Using Reflected Stoneley Wave Miyairi M: “Fracture Evaluation from Inversion of
Sanchez Mogollon AE, Tabor L, Valenzuela Muñoz A Arrivals,” Geophysics 54, no. 10 (October 1989): Stoneley Transmission and Reflections,” Proceedings
and Willberg D: “New Fibers for Hydraulic Fracturing,” 1274–1288. of the 4th SEGJ International Symposium, Tokyo
Oilfield Review 17, no. 2 (Summer 2005): 34–43. Brie A, Hsu K and Eckersley C: “Using the Stoneley (December 10–12, 1998): 389–394.
Abbas R, Jaroug H, Dole S, Effendhy, Junaidi H, Normalized Differential Energies for Fractured Reservoir 22. Tezuka K, Cheng CH and Tang XM: “Modeling of Low-
El-Hassan H, Francis L, Hornsby L, McCraith S, Evaluation,” Transactions of the SPWLA 29th Annual Frequency Stoneley-Wave Propagation in an Irregular
Shuttleworth N, van der Plas K, Messier E, Munk T, Logging Symposium, San Antonio, Texas, June 5–8, 1988, Borehole,” Geophysics 62, no. 4 (July-August 1997):
Nødland N, Svendsen RK, Therond E and Taoutaou S: paper XX. 1047–1058.
“A Safety Net for Controlling Lost Circulation,” Oilfield 23. Donald A and Bratton T: “Advancements in Acoustic
Review 15, no. 4 (Winter 2003/2004): 20–27. Techniques for Evaluating Open Natural Fractures,”
prepared for presentation at the SPWLA 47th Annual
Logging Symposium, Veracruz, Mexico, June 4–7, 2006.

Spring 2006 27
Time-Based Anisotropy
Sonde Deviation
20 % 0
-10 deg 90 Depth = 1,593.04 m
∆T-Based Anisotropy
Hole Azimuth 350 300
0 % 20
Offline 0 deg 360 300 250
Energy Anisotropy Flag, %

Slowness, µs/ft

Amplitude, dB
Total Azimuth 200
250
Maximum 0 deg 360 0 2 4 6 16 150
Energy
Hole Diameter Slow Shear ∆T 200
0 100 100
Fast Shear Azimuth
5 in. 20 350 µs/ft 1,200 150 50
Minimum -90 deg 90
Energy Gamma Ray Fast Shear ∆T
0 100 0 gAPI 150 Azimuth Uncertainty 350 µs/ft 1,200 0 2,000 4,000 6,000 8,000
Frequency, Hz
Depth,
m Depth = 1,658.87 m
350 300

300 250
A

Slowness, µs/ft

Amplitude, dB
200
1,600 250
150
200
100
C 150 50

0 2,000 4,000 6,000 8,000


Frequency, Hz
1,650
B Depth = 1,665.27 m
350 300

300 250

Slowness, µs/ft

Amplitude, dB
200
250
150
200
100
150 50

0 2,000 4,000 6,000 8,000


Frequency, Hz
> A crossed-dipole log (left) from the Pemex Cuitlahuac-832 well, showing zones with isotropy and with differing amounts
of anisotropy. Zone A, an isotropic zone, has low offline energy (depth track) and equal fast and slow shear-wave
slownesses (Track 3). Anisotropic Zones B and C have nonzero offline energies and different fast and slow shear-wave
slownesses. Anisotropy magnitude, either slowness-based or time-based (edges of Track 3), is about 8% in Zone B and
about 2% in Zone C. The azimuth of the fast shear wave (Track 2) remains constant through the anisotropic intervals,
even though the tool is rotating (Track 1), giving confidence in the anisotropy values. Dispersion curves from the three
intervals (right) show characteristic relationships. In Zone A (top), as in other isotropic formations, the dispersion curves
for flexural waves recorded in the two dipole directions (red and blue circles) overlie each other. At the bottom of Zone
B (bottom), the dispersion curves cross each other. The flexural wave that is fast near the borehole, at low frequencies
(red dots), becomes the slower wave with distance from the borehole (blue dots). This indicates that stress-induced
anisotropy is the dominant mechanism of anisotropy in this section. Shallower in Zone B (middle), the dispersion curves
look as though they could cross, but the high-frequency components of the fast shear wave are lost. At this depth, open,
induced fractures were visible in OBMI Oil-Base MicroImager logs. (Modified from Wielemaker et al, reference 25.)

imbalance.24 Until now, wireline sonic tools have Pemex Exploración y Producción wanted to its designated volume. Knowledge of elastic
been able to quantify the magnitude and evaluate the amount and direction of anisotropy anisotropy orientation and magnitude would help
orientation of elastic anisotropy only if the in tight gas-producing sandstone formations in in the design and application of oriented-
difference in velocities was at least 5%. The high the Burgos basin of northern Mexico. These perforating techniques prior to fracture
quality of data provided by the Sonic Scanner formations have low permeabilities and must be treatments and would also improve the success of
tool allows reliable measurement of anisotropy stimulated to produce gas in commercial infill-drilling campaigns.25
as small as 1%, and also helps interpreters quantities. Optimal development depends on
determine the cause of the anisotropy. correctly orienting hydraulic fractures in the
local stress field so that each vertical well drains

28 Oilfield Review
When the vertical stress is the maximum Wellbore
stress, hydraulic fractures propagate in the
direction of the maximum horizontal stress and
they open in the direction of minimum horizontal
stress. Shear waves travel fastest when polarized
in the direction of maximum horizontal stress
(SH) and slowest when polarized in the direction
of minimum horizontal stress (Sh). This is

with
Inter ic Scann
because additional stress stiffens the formation,

Son
val l
increasing velocity, and reduced stress conversely

ogge er tool
decreases velocity. Measuring the direction of

d
the fast shear waves yields the preferred
direction of fracture propagation.
The directions, or azimuths, of fast and slow
shear waves can be seen in a crossed-dipole log.
A crossed-dipole log from the Cuitlahuac-832
well shows both isotropic and anisotropic zones
(previous page). Zone A, an isotropic zone, is
identified by near-zero offline energies and
equal fast and slow shear-wave slownesses.26
Anisotropic Zones B and C are identified by
nonzero offline energies and diverging fast and
> Geologic cross section with trajectory of a deviated well in which Norsk Hydro acquired
slow shear-wave slownesses.
Sonic Scanner imaging data. The high deviation angle required TLC Tough Logging
The two anisotropic zones have different Conditions wireline logging.
amounts of anisotropy. In Zone B, anisotropy
magnitude is about 8%. In Zone C, the amount of
anisotropy is about 2%. Although 2% is lower than
has been reliably detected by other tools, high-frequency data are missing, so it is Norsk Hydro has used the imaging capability
interpreters have confidence in the value impossible to determine the curve trend or the of the Sonic Scanner tool in a highly deviated
because the waveforms are so clear and because anisotropy type. OBMI Oil-Base MicroImager well in the Norwegian Sea (above). Following
the fast shear azimuth remains constant, images at this depth indicate the presence of acquisition of standard sonic waveforms in one
between 30° and 40° over the interval, even with open, induced fractures, which are the likely TLC Tough Logging Conditions wireline pass, a
the tool continually rotating. cause of the loss of high-frequency data and separate TLC imaging pass recorded waveforms
Knowing the magnitude and azimuth of also strongly suggest stress-induced anisotropy. every 0.5 ft [15 cm] from the three monopole
anisotropy is vital, but this does not identify the The 45° azimuth of fractures seen in OBMI sources firing sequentially to the 104 receivers
cause. The anisotropy may be intrinsic to the rock images correlates well with the 40° azimuth of
24. Elastic anisotropy is sometimes called acoustic
or may be stress-induced; identifying the cause is maximum horizontal stress inferred from the fast anisotropy or velocity anisotropy. It can be expressed in
important for understanding how stable the shear direction. terms of a difference of velocities, slownesses, stresses
or elastic parameters.
drilling process will be and how a borehole will In the Burgos basin, maximum horizontal
Armstrong P, Ireson D, Chmela B, Dodds K, Esmersoy C,
respond to stress. Usually, additional information, stress has traditionally been taken to be parallel Miller D, Hornby B, Sayers C, Schoenberg M, Leaney S
such as borehole images or core analysis, is to the strike of the nearest faults. The results and Lynn H: “The Promise of Elastic Anisotropy,” Oilfield
Review 6, no. 4 (October 1994): 36–56
required to pinpoint the cause of anisotropy. from Sonic Scanner logging in five wells in this 25. Arroyo Franco JL, Gonzalez de la Torre H,
Analysis of flexural-wave dispersion curves basin indicate that local stress direction can Mercado Ortiz MA, Weilemaker E, Plona TJ,
Saldungaray P and Mikhaltzeva I: “Using Shear-Wave
provided by the Sonic Scanner tool helps to vary significantly—up to 20° from the strike of Anisotropy to Optimize Reservoir Drainage and Improve
identify anisotropy mechanisms at three depths nearby faults—accentuating the importance of Production in Low-Permeability Formations in the North
of Mexico,” paper SPE 96808, presented at the SPE
in the Cuitlahuac-832 well using only sonic making localized sonic measurements before Annual Conference and Technical Exhibition, Dallas,
measurements. Dispersion curves at 1,593.04 m, designing perforation, stimulation and infill- October 9–12, 2005.
within Zone A, overlie each other closely and drilling operations. Wielemaker E, Saldungaray P, Sanguinetti M, Plona T,
Yamamoto H, Arroyo JL and Mercado Ortiz MA: “Shear-
match the model for a homogeneous isotropic Wave Anisotropy Evaluation in Mexico’s Cuitlahuac Field
formation. Curves from 1,665.27 m, one of the Imaging Well Beyond the Wellbore Using a New Modular Sonic Tool,” Transactions of the
SPWLA 46th Annual Logging Symposium, New Orleans,
most anisotropic intervals near the bottom of The superior quality of waveforms acquired with June 26–29, 2005, paper V.
Zone B, show the crossover characteristic of the Sonic Scanner tool allows for improved 26. The difference between slownesses is called slowness
anisotropy, and the difference between arrival times is
stress-induced anisotropy. Slightly shallower, at imaging away from the borehole. Sonic imaging called time-based anisotropy.
1,658.87 m, the fast and slow shear dispersion uses reflected P-waves to detect reflectors
curves are separated at low frequencies, but the that are subparallel or at low angle relative to
the borehole.

Spring 2006 29
over a distance of 330 m [1,100 ft]. Waveforms
from each source were processed in a sequence X,240
that started with separating reflected P-waves
from Stoneley and refracted P-waves. The
azimuthal distribution of sensors at each X,260

receiver station allows identification of the


direction to the reflector. Then, traces from each
receiver station were depth-migrated using X,280

formation velocities measured by the Sonic


Scanner logs from the earlier logging pass.27 To

Vertical depth, ft
X,300
account for tool rotation and the azimuthal
distribution of sensors, the image at each
receiver station was reconstructed by depth-
X,320
shifting and stacking images from each
azimuthal channel. Finally, the depth-migrated
images were stacked. Images were obtained
X,340
within 48 hours.
The results show a 5-degree dipping event
that extends at least 13 m [43 ft] into the X,360
formation (right). The dip of the event is in
agreement with the expected geology at the well
location. The high-resolution event can be X,380
correlated with a 1-m [3.3-ft] coal bed at the X,460 X,480 X,500 X,520 X,540 X,560 X,580
same depth position indicated by petrophysical
Source-receiver midpoint, ft
logs (next page). The identification of a 1-m coal
> A gently dipping reflector imaged far from the borehole. The borehole
bed indicates the potential to obtain high-
trajectory is shown in red. The high-resolution event detected by sonic
resolution images from a sonic-imaging survey. imaging can be seen above and to the right of the borehole, near the center
The resolution is far better than can be obtained of the image. The reflector correlates with a coal bed at the same depth
from any surface or borehole seismic survey position indicated by petrophysical logs.
(below right).
Another potential application of sonic
imaging is the detection of vertical fractures
near but not intersecting vertical boreholes.
Current techniques such as borehole image
logging and fracture identification from Stoneley
reflections work only if a fracture intersects the
borehole. In many cases, a vertical well will miss
vertical fractures. Deep imaging with the Sonic
Scanner tool expands the volume of investigation
to enable the identification of features that may Sonic-imaging results on seismic scale
delineate reservoir extent or the state of stress
away from the borehole.

27. Migration is a data-processing step that aims to sharpen,


shift and relocate reflectors to their true locations.

> Comparing high-resolution sonic-imaging data with a surface seismic survey. The 1-m coal bed
resolved by Sonic Scanner imaging (inset) cannot be seen in the surface seismic survey.

30 Oilfield Review
Scanning the Horizon
HRLA1 Resistivity Pyrite
The Sonic Scanner tool is a new development,
Gamma Ray Clay and engineers, geologists and petrophysicists are
0 gAPI 200 HRLA2 Resistivity
Carbonate T2 Cutoff
still finding new ways to use its data. By adding
Caliper Quartz the radial dimension and multiple depths of
HRLA3 Resistivity 0.3 ms 3,000
6 in . 16
investigation to the well-known axial and
Density Siderite Log T2 Mean
HRLA4 Resistivity azimuthal sonic measurements, the Sonic
ECS Sigma
1.95 g/cm 3
2.95 Coal 0.3 ms 3,000 Scanner tool performs enhanced characteri-
60 cu 0 T2 Distribution zation of acoustic properties in inhomogeneous
HRLA5 Resistivity Porosity Density
Depth,
Caving m 0.2 ohm-m 200 0.45 m3/m 3 -0.15 2.5 g/cm3 3.0 0.3 ms 3,000
and anisotropic formations. With this infor-
mation, customers are able to predict how
formations and fluids will behave during drilling,
stimulation and production.
The innovative tool design with predictable
acoustics delivers waveforms of excellent quality
and at a wide frequency range. These capabilities
allow slowness estimation in extremely slow
formations, measuring azimuthal anisotropy as
small as 1 to 2%, and reliable application of low-
frequency Stoneley modes for fluid-mobility
estimation and evaluation of natural fractures.
Advanced quality control with slowness-
frequency analysis adds confidence to slowness
estimates obtained by dispersion analysis.
Complete recording of all data from
monopole and dipole sources to 104 receivers
distributed azimuthally around the tool removes
X,900 uncertainties about formation geometry and
structure and improves through-casing and
cement evaluation. Current capabilities obtain
only monopole compressional and shear data in
cased hole. One area of future advancement will
be to extend current openhole applications to
cased wells.
Additional applications will arise as more
companies gain experience with the Sonic
Scanner tool and the high-quality data it
produces. While it is difficult to predict how the
rest of the oil and gas industry will evolve, sonic-
1- m logging enthusiasts anticipate another 50 years of
coal bed
investigations in and around the borehole. —LS

X,925

> Petrophysical logs from the Norsk Hydro well in the Haltenbanken area of the Norwegian Sea,
showing the 1-m coal bed delineated by sonic imaging. Platform Express resistivity logs (Track 2) and
density and porosity logs (Track 3) are input, along with ECS Elemental Capture Spectroscopy data
to derive mineralogy (Track 4). Nuclear magnetic resonance data appear in Track 5.

Spring 2006 31
History of Wireline Sonic Logging

In a patent awarded in 1935, Conrad


Schlumberger specified how a transmitter
and two receivers might be used to measure
the speed of sound in a short interval of rock
penetrated by a borehole (right).1 He claimed
that the speed and attenuation of sound
would characterize lithology. His invention
failed because neither logging engineers
nor the technology of the time was able to
detect the short time difference—tens of
microseconds (µs)—between signals traveling
at the speed of sound to receivers separated
by just inches.
During World War II, the necessary > Illustration from the 1935 patent
electronics emerged, making sonic logging on acoustic logging by Conrad
possible.2 According to one account, the first Schlumberger. The field engineer
oilfield application of sonic logging was for (13) was supposed to slide a sleeve
(17) until sound coming from
casing-collar location, in 1946.3 Most other
receivers (3 and 4) appeared to
historical accounts state that the first sonic arrive simultaneously at each ear.
applications appeared after the 1948
experiments by Humble Oil Research,
followed by Magnolia Petroleum Company and
Shell.4 These companies designed devices to
collect sonic-velocity information for time-
depth conversion of surface seismic sections
and for correlating seismic reflections to
lithologic interfaces. The tools featured one
transmitter and one or two receivers
separated from the transmitter by isolating
material. By the mid-1950s, service companies
and oil companies were acquiring sonic-
logging data to generate synthetic seismograms In 1960, field crews testing the VLT borehole irregularities and near-wellbore
for comparison with surface seismic sections.5 response in cased holes in Venezuela noticed alteration. The tool-durability problem arose
In 1957, having licensed the Humble patent, that certain zones caused unreadable, low- because early tools used rubber to isolate
Schlumberger introduced its first sonic tool, amplitude signals. They correctly concluded receivers from transmitters, thereby
the velocity logging tool (VLT), for improving that the anomalous signals could be attributed preventing undesirable sound waves from
seismic interpretation. only to cement condition. Measuring and propagating within the tool and overwhelming
The early Magnolia Petroleum paper had recording signal amplitude in addition to desired signals. However, rubber tended to
hinted at the additional possibility of using arrival time gave birth to an unexpected absorb gas from gas-rich formations, causing
sonic velocities to determine porosity and application, and CBT Cement Bond Tool logs the tool to expand and break apart as it was
lithology, but it was scientists in the research soon replaced the temperature survey for brought to surface. The tool was strengthened
division of Gulf Oil Corporation who first detecting top of cement. by replacing the rubber with steel, but then
published experimental observations By the early 1960s, the first sonic tools had the tool housing had to be shaped so that the
confirming the link.6 Within a short time, acquired tens of thousands of logs, and path of sonic waves traveling through the steel
demand for sonic porosity-logging applications engineers set about designing a second- would be longer than the paths through the
overtook that for seismic applications. generation tool to address three problems: tool formation and back to the receivers. (next
durability, and poor signal in the presence of page). Many modern sonic tools continue to

32 Oilfield Review
feature slots and grooves to slow down the transmitter-receiver (TR) spacing captured
arrival of signals—known as tool arrivals— only waves that propagated in the altered
that travel purely through the tool. zone, leaving the unaltered zone away from
A way around the second problem, poor the borehole unexplored. By increasing the
logs in bad hole, came from the Shell engineer spacing to 8 to 12 ft [2.4 to 3.7 m], the LSS
responsible for that company’s first sonic Long-Spaced Sonic Tool improved log
tool.7 His borehole-compensating arrangement response in altered shales. Sonic velocities of
of receivers and transmitters not only the unaltered formation are more
eliminated the problem of poor signal in representative of the reservoir in its natural
washed-out zones, but also removed the state and yield synthetic seismograms that
effects of tool tilt and eccentering on log better match surface seismic traces.
response. Solving two of the three problems The long TR spacing also stretched the
that plagued the earlier tools, Schlumberger received wavetrain, separating the P-, S- and
incorporated this idea into the all-steel design other waves into recognizable packets of
of the borehole-compensated (BHC) sonic tool energy. Efforts intensified to capture the full
that was introduced in 1964. The BHC tool waveform, leading to the development of tools
contained two transmitters and four receivers. that recorded digital waveforms from an array
Along with BHC technology came the of receivers. The first commercial Schlumberger
ability to view registered waveforms on an version of this technology, introduced in the
1980s, was called the Array-Sonic full-waveform
1. Schlumberger C: “Procédé et Appareillage pour la
Reconnaissance de Terrains Traversés par un
sonic velocity tool. Full-waveform logging gave
Sondage.” République Française Brevet d’Invention rise to a host of new processing techniques.
numéro 786,863 (June 17, 1935). Also Doll L: “Method
of and Apparatus for Surveying the Formations
The late 1980s saw research experiments
Traversed by a Borehole,” US Patent No. 2,191,119 with a second-generation digital sonic tool.
(February 20, 1940) (submitted by the estate of Conrad
Schlumberger).
The DSI Dipole Shear Sonic Imager tool had
> A sonic-logging sonde with slots to slow
2. The terms “sonic” and “acoustic” are used eight sets of four monopole receivers that
interchangeably. down tool arrivals.
could function as orthogonal dipole receivers,
3. Pike B and Duey R: “Logging History Rich with
Innovation,” Hart’s E&P (September 2002): 52–55,
and carried one monopole source and two
http://www.spwla.org/about/Logging-history.pdf oscilloscope in the logging truck. Appearing orthogonally oriented dipole sources. The
(accessed April 28, 2006).
on the screen were not only the primary (P-) dipole sources generated flexural waves,
4. From Humble Oil: Mounce WD: “Measurement of
Acoustical Properties of Materials,” US Patent arrivals, or compressional waves, but also allowing characterization of formation
No. 2,200,476 (May 14, 1940). secondary (S-), or shear waves and later anisotropy and shear slowness in slow as well
From Magnolia Petroleum Company: Summers GC
and Broding RA: “Continuous Velocity Logging,”
arrivals. Recognizing the importance of shear as fast formations.
Geophysics 17, no. 3 (July 1952): 598–614. waves made the mid-1960s a time of intense Also in the late 1980s, Schlumberger
From Shell: Vogel CB: “A Seismic Velocity Logging activity in expanding sonic applications. researchers tested a variety of multireceiver
Method,” Geophysics 17, no. 3 (July 1952): 586–597.
Léonardon, reference 1, main text.
Specialists at Shell proposed using the ratio acoustic tools for their ability to acquire sonic
5. Breck HR, Schoellhorn SW and Baum RB: “Velocity of P to S velocity as a lithology indicator, and images—seismic-like images far from the
Logging and Its Geological and Geophysical also used sonic logs to predict overpressured borehole.9 The first commercial sonic-imaging
Applications,” Bulletin of the American Association
of Petroleum Geologists 41, no. 8 (August 1957): zones.8 Schlumberger engineers and service was run in 1996, but processing was
1667–1682. researchers evaluated use of P and S time- and personnel-intensive.
6. Wyllie MRJ, Gregory AR and Gardner LW: “Elastic
Wave Velocities in Heterogeneous and Porous
amplitudes to locate fractures. Although In 2005, the Sonic Scanner acoustic
Media,” Geophysics 21, no. 1 (January 1956): 41–70. these and other shear-wave applications had scanning platform combined many
Tixier MP, Alger RP and Doh CA: “Sonic Logging,” been proposed, the acquisition systems of the innovations of the past and added radial
Journal of Petroleum Technology 11, no. 5 (May 1959):
106–114. time recorded only the arrival time of the P- measurements to simultaneously probe the
7. Vogel CB: “Well Logging,” US Patent No. 2,708,485 wave. The waveform itself, including P, S and formation for near-wellbore and far-field
(May 17, 1955).
later arrivals, was not recorded. slownesses.10 The tool itself is fully
8. Hottman CE and Johnson RK: “Estimation of
Formation Pressures from Log-Derived Shale Another drawback of the BHC tool was its characterized, with predictable acoustics.
Properties,” Journal of Petroleum Technology 17, no. inability to accurately measure the true The wide frequency range of the monopole
6 (June 1965): 717–722.
9. Hornby BE: “Imaging of Near-Borehole Structure
formation interval transit time in zones of and dipole transmitters delivers excellent
Using Full-Waveform Sonic Data,” Geophysics 54, no. invasion, shale alteration and drilling-induced waveform quality in all formation types.
6 (June 1989): 747–757.
damage. The 3- to 5-ft [0.9- to 1.5-m]
10. Pistre et al, reference 3, main text.

Spring 2006 33

Anda mungkin juga menyukai