You are on page 1of 16

GEOPHYSICS, VOL. 67, NO. 5 (SEPTEMBER-OCTOBER 2002); P. 13481363, 18 FIGS. 10.1190/1.

1512781

Tutorial Converted-wave seismic exploration: Methods

Robert R. Stewart , James E. Gaiser , R. James Brown , and Don C. Lawton

ABSTRACT

Multicomponent seismic recording (measurement with vertical- and horizontal-component geophones and possibly a hydrophone or microphone) captures the seismic waveeld more completely than conventional single-element techniques. In the last several years, multicomponent surveying has developed rapidly, allowing creation of converted-wave or P-S images. These make use of downgoing P-waves that convert on reection at their deepest point of penetration to upcoming S-waves. Survey design for acquiring P-S data is similar to that for P-waves, but must take into account subsurface V P / VS values and the asymmetric P-S ray path. P-S surveys use conventional sources, but require several times

more recording channels per receiving location. Some special processes for P-S analysis include anisotropic rotations, S-wave receiver statics, asymmetric and anisotropic binning, nonhyperbolic velocity analysis and NMO correction, P-S to P-P time transformation, P-S dip moveout, prestack migration with two velocities and waveelds, and stacking velocity and reectivity inversion for S-wave velocities. Current P-S sections are approaching (and in some cases exceeding) the quality of conventional P-P seismic data. Interpretation of P-S sections uses full elastic ray tracing, synthetic seismograms, correlation with P-wave sections, and depth migration. Development of the P-S method has taken about 20 years, but has now become commercially viable.

BACKGROUND

The primary method of hydrocarbon exploration remains P-wave seismic reection surveyingand for good reason. Among all the elastic waves, compressional waves arrive rst, usually have high signal-to-noise ratios, have particle motion that is close to rectilinear, are easily generated by a variety of sources, and propagate in uid environments. We expect that P-wave reection surveying will be the dominant exploration method for some years to come, but several questions are germane. Can we improve P-wave pictures? Can we generate additional or augmenting images using the seismic method? When P-wave surveying fails, can we use other seismic techniques? Perhaps the most straightforward answer to these questions is, Yes, try multicomponent recording and analysis. There is considerable promise in using multicomponent recordings to improve P-wave sections themselves. However, the thrust of

this paper is to provide a tutorial on using multicomponent data to create images from another mode of seismic energy: the converted shear waves. In using the term converted-wave exploration, we imply a particular conversion: a downward-propagating P-wave, converting on reection at its deepest point of penetration to an upward-propagating S-wave (see Figure 1). There are many other types of energy conversion that may occur as seismic waves move through the earth. In relatively high-velocity layers (such as permafrost, volcanics, or salt), there may be an S-wave leg generated in the downgoing path that converts from and then back to a P-wave. In the marine case, we might have a P-to-S downgoing transmission at the ocean bottom that will eventually reect as an S-wave and reconvert to a P-wave at the ocean oor (Tatham, 1982). Nevertheless, modeling and eld measurements show that transmitted or multiple conversions generally have much lower amplitudes than the primary

Manuscript received by the Editor February 2, 2000; revised manuscript received February 28, 2002. CREWES Project, University of Calgary, Department of Geology and Geophysics, 2500 University Drive N.W., Calgary, Alberta T2N 1N4, Canada. E-mail: stewart@geo.ucalgary.ca; dclawton@geo.ucalgary.ca. WesternGeco, 1625 Broadway, Denver, Colorado, 80202. c 2002 Society of Exploration Geophysicists. All rights reserved. 1348

Converted-wave Seismic Exploration: Methods

1349

P-to-S reection, that is, P down, S up (Rodriguez, 2000). We will focus only on this P-S conversion. We note here (Figure 1) that there is an asymmetry in the P-S raypath as described by Snells law: sin / V P = sin / VS , where and are the P- and S-wave angles of incidence and reection, respectively, and V P and VS are the P- and S-wave velocities, respectively. Since VS < V P , is less than , and the S-wave leaves the interface closer to perpendicular than the incident P-wave. Whereas Snells law gives the basic geometry of the raypaths, the Zoeppritz equations provide the amplitude of the reections. In general, a P-wave incident on an interface between two elastic media will generate a shear-wave reection. For the single interface of Figure 2, we can see a quasi-sinusoidal variation with offset in the converted-wave reectivity (for interactive calculation of Zoeppritz P-P and P-S reectivity see http://www.crewes.org/Samples). This energy partitioning is part of the reason why there are amplitude-variation-withoffset (AVO) effects in P-wave data. Aki and Richards (1980) show that the P-S reectivity, R P S , can be approximated (assuming property changes are small) by

R P S = k (1 + )

VS + 2 , VS

(1)

where k = ( tan )/2, = 2(sin2 )/ 2 + 2(cos cos )/ , = VP / VS , = lower upper , VS = VS lower VS upper , = 1 2 (lower + upper ), VS = 1 (VS lower + VS upper ), and V P = 1 (V P lower + 2 2 VP upper ). There are three notable aspects to equation (1): it has no explicit dependence on P-wave velocity change, it is negative up to moderate angles for positive parameter changes and, generally, it indicates that there is P-S energy, at moderate angles, comparable to the corresponding P-P energy (Figure 2). So, we now have the two basic aspects of P-S wave propagation: asymmetric ray paths according to Snells law, and sinusoidal amplitude variation with offset as described by the Zoeppritz equations. Concentrated work in P-S analysis has been proceeding for about 20 years; several commercial acquisition and processing groups are now active. We might compare this with the de-

velopment of 3-D seismology using P-waves. The 3-D seismic concept and early experimentation began in the 1960s, theory and processing were largely worked out in the 1970s, and their application came in the 1980sabout 20 years from concept to common practice. P-S surveying was proposed and tried in the late 1970s, with processing fundamentals developed in the 1980s to early 1990s. Assuming that P-S surveying is on a similar track to that of 3-D seismic analysis, we would expect it to become common practice in the next several years. Exciting results are coming in, but the overall evaluation of the method is still in progress (Duey, 2001). We might inquire why multicomponent, and especially P-S, methodologies havent been widely used in hydrocarbon exploration in the past (as opposed to common multicomponent usage in earthquake studies). There would seem to be several reasons for this. In the acquisition world of several years ago, multicomponent sensors were not available in large numbers. On land, there were challenging logistics in planting and cabling them, and in recording their output, which requires three times the normal number of channels per station. Fourcomponent marine sensors (three orthogonal geophone elements and a hydrophone) were rare. Converted-wave survey design was not well understood, and design software not commercially available. Processing P-S waves was also problematic. Much of the theory for processing the data had not been worked out, and there were subsequent delays in development of software and experienced personnel. One of the processing stumbling blocks was the variation of the P-S reection points in depth for any given source-receiver offseta result of unequal P-incidence and S-reection angles (Figure 3). Gathering, mapping, and binning for P-S data were (and are) more complicated than in P-P analysis with its straightforward common-midpoint assumption. Furthermore, interpreting P-S data was difcult because few S-wave velocity logs had been acquired and P-S synthetic seismograms were not available. On top of all this, P-P and P-S sections displayed events, from the same reector, at different

FIG. 1. A converted-wave (P-S) reection at its conversion point (CP) compared to a pure P-wave reection at its midpoint (MP). Note the CP is shifted toward the receiver. The P-wave angle of incidence and S-wave angle of reection are given by and , respectively. Directions of positive phase, as shown by arrowheads, are according to Aki and Richards (1980).

FIG. 2. P-P and P-S plane-wave reection coefcients as a function of P-wave angle of incidence. Note that the absolute value of the P-S reection coefcient is plotted. The S-wave velocities for the upper and lower layers are 1750 m/s and 2650 m/s, respectively. Density is constant.

1350

Stewart et al.

times! Sometimes, events seen on one section were not seen on the other or had different character. Finally, there were few denitive case histories. In other words, the method was in its early infancy. Now, however, almost all of this has changed. P-S surveydesign software exists, geophones and recording channels are generally available, and many logistical difculties have been overcome. Several commercial processing packages and contractors handle P-S processing. In addition, S-wave logging is more commonly conducted, and interpretive software is available. We understand better the differing P and S responses and are building some impressively successful case histories (Caldwell, 1999). A broad goal of seismic exploration is to create a 3-D depth image of rock type, structure, and saturant. For numerous reasons, we are unlikely to be able to do this with P-waves alone. For example, different rocks can have similar P-wave properties, giving rise to low-reectivity interfaces. The S-wave properties may show greater variation, providing a signicant contrast and reection. Thus, given the need, how can we obtain S-wave properties from surface seismic measurements? We can attempt to nd S-wave information via AVO analysis of P-P data. AVO is useful in some areas, but is often limited due to its higher order dependence of P-wave reectivity on S-wave velocity. In addition, there are signal-to-noise and processing problems with AVO analysis (Arnold and Chiburis, 1998). Other issues include calibration, anisotropy, attenuation, thin-bed tuning, and multipathing. We also use limited and perhaps oversimplied models to relate the AVO response to rock properties. Recall too, that to conduct AVO analysis, we need a background V P / VS value that must be determined elsewhere. Furthermore, the most fundamental of seismic measurements is that of event arrival times. AVO analysis does not give us S-wave traveltimes. P-S surveys do. The P-S time structure and isochron maps can often be used with P-P sections to extract robust (if lower resolution) V P / VS values. Pure-shear surface-seismic surveys are also possible and have been used. They directly measure S-wave parameters. Unfortunately, S-S seismic sections are often noisy and can have lower resolution, partially because of the two-way propagation path of the S-waves through the usually attenuative, lowvelocity, and heterogeneous near surface. Furthermore, S-wave

sources have been expensive, scarce, and not easily applicable in some environments (e.g., marine settings, transition zones, muskeg, and environmentally sensitive areas). However, new generations of marine shear sources (Clark, 2000) and horizontal land vibrators may help. S-S listening times in recording are about double or triple those of P-waves and some 30% to 50% longer than for converted waves. All told, pure S-wave surveys usually cost more than P-S acquisition (Kendall and Davis, 1996). Modeling also suggests that S-S reections can undergo phase changes at relatively short offsets that can complicate their stacking (Guevara, 2000). Thus, expense and data quality have limited the applicability of S-S surveys to date. Thomsen (2001) suggests that, Converted-wave surveys will come to replace shear-source surveys, wherever the exploration problem demands a shearwave solution. P-S surveys are a relatively inexpensive, broadly applicable, and effective way to obtain S-wave information. Lets look at some of the tools used in converted-wave work.
TOOLS OF THE TRADE

Core measurements Various drilling and coring tools can retrieve rock samples from the subsurface. These rock plugs or cores can then be subjected to a variety of measurements in a rock-property laboratory. The apparatus for elastic-property measurement often consists of an ultrasonic source and receiver (piezoelectric transducers), xed on either end of the core and emplaced in a pressure vessel. P- and S-wave transmission times are measured across the core (which has a known path length) to derive velocities (King et al., 2000). These transmission tests are typically conducted at varying conning and pore uid pressures, gas saturations, or temperatures. The full waveform of the transmitted pulse is usually recorded, which allows the calculation of attenuation, or Q , the quality factor (Toksoz and Johnston, 1981). Other methods of making elastic core measurements include resonant-bar techniques and stress-strain analysis. Empirical and theoretical relationships among rock properties can be established from these measurements (e.g., Gueguen and Palciauskas, 1994; Mavko et al., 1998; Schon, 1998). We need to remember, however, that ultrasonic measurements around 1 MHz may give different absolute velocities and Q values than lower frequency (roughly 15 kHz) well-log observations or seismic measurements at approximately 50 Hz (Stewart et al., 1984). Elastic-wave logs Full-waveform sonic logs give information about both P-wave and S-wave velocities (Mari et al., 1994). Transmitted tube waves from a monopole source, exural waves from dipole sources, and screw modes from quadrupole sources can all be used to estimate formation shear velocity under various borehole and formation conditions. Cross-dipole logs (where sources and receivers are grouped in parallel and perpendicular combinations) have been used to estimate anisotropy (Mueller et al., 1994; Xu and Parra, 1998). The basis of much of the lithologic work in P-S exploration relates to anomalous changes in VS with respect to V P . A changing VP / VS value is often closely tied to a changing lithology

FIG. 3. The location of the P-S conversion point moves from the receiver toward the asymptotic conversion point (ACP) with increasing depth. The P-S reectivity curve is shifted from the midpoint (MP) toward the receiver.

Converted-wave Seismic Exploration: Methods

1351

(Rider, 1996) and perhaps pore geometry. For example, Miller and Stewart (1990) use data from full-waveform sonic logs in the Medicine River eld in Alberta, Canada, to analyze V P / VS values from pure lithologies (sandstone, shale, limestone; Figure 4a). Similar to values reported by other authors, their V P / VS values for sandstone cluster around 1.6, with higher values near 1.9 for shale and limestone. When the lithologies become mixed or complex, the V P / VS values change to more intermediate quantities (Figure 4b). Miller (1996) also shows well-log values from a eld in southern Alberta (Figure 5). The Glauconitic sandstone reservoir shows signicantly higher VS values than do the off-reservoir shales at the same depth. Thus, we expect a decrease in V P / VS from regional shales to reservoir sand. We can also plot these data as V P / VS versus gamma-ray values. Generally, there is a small increase in V P / VS for sands with more clay or shaliness. In another analysis of an array sonic log [from the Davey well (3-13-34-29W4) in central Alberta, Canada] Miller (1996) plotted V P / VS values, photoelectric factor (PEF) curves, and anhydrite fraction. The other primary lithology was dolomite. Recall that the PEF or photoelectric absorption is strongly dependent on the average atomic number of the formation and by inference the lithology (Rider, 1996). Miller (1996) noted that V P / VS values from the array-sonic track the PEF curve and volume of anhydrite quite well: as the anhydrite-versus-dolomite fraction increases, so do the V P / VS and PEF logs. Anhydrite

V P / VS values are about 2.0, dolomite about 1.85. Similarly, Figure 6 shows the correlation of V P / VS with dolomite (1.85) and limestone (1.95). Again, as the limestone fraction versus dolomite increases, so do the V P / VS and PEF logs. General empirical relationships or calibrated rock models may be useful to determine VS if only V P or other nonelastic logs are available (Castagna et al., 1985; Jrstad et al., 1999). P- and S-log values also form the basis for generating synthetic seismograms, which assist in the analysis of P-P and P-S sections. Lawton and Howell (1992) developed a procedure for calculating one-dimensional P-S synthetic seismograms (by ray tracing through a layered medium with use of Zoeppritz energy partitioning). Examples of P- and S-velocity logs and their corresponding P-P and P-S synthetic seismograms from the Blackfoot area of southern Alberta are shown in Figure 7. We recall that elastic-wave velocities and impedances can be related to various combinations of the Lame parameters and density. For example, P-wave impedance, I P , can be writ2 = ( V P )2 = ( + 2) , and S-wave impedance, I S , as ten as I P 2 = ( VS )2 = , where , and are the Lame parameters IS 2 2 2 IS , and is the density. Goodway et al. (1997) write = I P and suggest that analysis of and can provide an indication of pore uids. Other authors (e.g., Duffant et al., 2000) group the terms in equation (1) to dene a shear-wave elastic impedance (SEI). The SEI value is of a generalized impedance form SEI( ) = VSa b , where a and b are functions of the P-wave

FIG. 4. (a) Well-log V P / VS values versus V P for pure lithologies of the Medicine River eld. (b) V P / VS value versus V P for mixed lithologies (from Miller and Stewart, 1990). SS = sandstone, SH = shale, and LS = limestone. Nordegg, Detrital, and Shunda are lithostratigraphic units.

FIG. 5. (a) VS versus V P values from well logs for a sand reservoir and off-reservoir shale. VS is signicantly higher in sand than in shale in these wells (Miller, 1996). (b) VP / VS value versus gamma-ray value from well logs as in (a). V P / VS shows a slight increase with shaliness in sand. Purer shales have distinctly higher V P / VS values.

1352

Stewart et al.

incidence angle and an average V P / VS value. The SEI may be computed from logs and used to interpret P-S data or guide inversions. Sonic reection imaging, using an array-sonic logging tool and processing reections or diffractions instead of direct arrivals, shows considerable promise for reservoir imaging. Esmersoy et al. (1998) show an example of migrating data from Schlumbergers Dipole Shear Sonic Imager in a horizontal well to create both P- and S-wave sections. They concluded that reectors at a range of 3 m were imaged, with the possibility of imaging up to 10 m. Vertical seismic prole (VSP) surveys VSP surveys generally use three-component (3-C) geophones and have done so for many years (Toksoz and Stewart, 1984; Hinds et al., 1996; Hardage, 2000). This is partially a result of using offset sources that generate clear P and S energy on both horizontal and vertical channels (Figure 8). Competent deeper rocks usually have higher velocities that allow waves to propagate at large angles from the vertical direction. This is different from the surface seismic case where the generally low-velocity near surface bends raypaths to the vertical. At the surface, this places P-waves largely on the vertical geophone and S-waves on the horizontal elements. In the VSP case, multicomponent analysis with 3-C recording is usually required to disentangle the P- and S-wave energy on all three channels. Interestingly, VSP processing has had an impact on P-S surface analysis: P-wave reection points between a given source

FIG. 7. From the left: P-wave impedance log, P-P synthetic seismogram, S-wave impedance log, and corresponding P-S synthetic seismogram for the 08-08-23-23W4 well in the Blackfoot eld, southern Alberta (from Stewart et al., 1996). The P-wave synthetic seismogram uses a 8-12-75-85-Hz Ormsby wavelet; the S-wave seismogram has a 8-12-45-55-Hz Ormsby wavelet.

FIG. 6. From the left: V P / VS , VLS (volume limestone versus dolomite; greater limestone fraction plots to the right), and PEF (photoelectric logs) plotted in depth for the Davey well (3-13-34-29W4) in central Alberta. Note the consistent tracking of V P / VS with the limestone volume (from Miller, 1996).

Converted-wave Seismic Exploration: Methods

1353

and a specic VSP receiver at different depths form a curved trajectory in depth; they do not fall on a straight commonmidpoint line (Figure 9a). Note that the location of P-S reection points in the VSP geometry are also curved and displaced toward the receivers (Figure 9b). Such is also the case with the P-to-S conversion for a surface source and receiver (Figure 3). The VSP case helped researchers understand how to do common-conversion-point (CCP) mapping for P-S data in a surface geometry. A basic aim of the VSP survey is to determine seismic interval velocities. From the traveltimes of rst-break energy in depth, we can calculate velocities. An example from the Hamburg area of central Alberta (Coulombe et al., 1996) is shown in Figure 10. We note the low V P / VS value of the porous dolomite with respect to the surrounding limestones. In addition, the seismic velocities are generally lower than sonic values as we would expect from attenuation-dispersion arguments. We can also assemble well logs, VSP data, synthetic seismograms, and surface seismic into a compelling display called a composite plot or L plot. This is a very useful compendium of correlated data that often allows a more condent interpretation. An example is the composite plot from the Rolling Hills of southern Alberta shown in Figure 11 (Geis et al., 1990). Note the good correlation of events across the various data types. Throughout the recording and processing sequence, it is critical to keep polarities consistent. This can be accomplished

FIG. 8. VSP geometry with an offset source. Both transmitted and reected P-P and P-S waves are shown.

by using SEG recommended eld polarities and the Aki and Richards (1980) energy-partitioning equations. When we do this, a particular reector will generally show the same polarity on P-P and P-S eld data in the normal case where V P ,VS and all change in the same direction (Brown et al., 2002). Close to the point of reection (in depth), P-S and P-P events have about the same frequency content (Figure 12), implying that P-S images should have higher spatial resolution than the associated P-P sections. In fact, this is often observed in VSP images. In the shallow surface-seismic case, P-S resolution may also be superior to that of the P-P data. However, the opposite is generally true for the deeper section. Unfortunately, by the time the deeper P-S events are recorded at the surface, their frequency content has often decreased relative to that of the P-wave (Figure 13). Greater attenuation of the shorter P-S wavelengths is likely a major contributor to this loss. Deffenbaugh et al. (2000) indicate that for cases where Q P > Q S , there is a resolution crossover depth below which the P-P mode gives better resolution than the P-S. If Q S > Q P , then P-S resolution should be better than the P-P case. However, other factors, such as large variation in the stratigraphy of the shallow section (with attendant changes in velocity, impedance, anisotropy, and attenuation) may complicate S-wave propagation and lead to a decrease in its bandwidth. Another example of P-P and P-S VSP data comes from the Willesden Green region of central Alberta (Figure 14). We see much more reection activity in the shallow part of the P-S section as compared to the P-P section (Stewart et al., 1995). In addition to the higher spatial resolution of P-S events measured in situ, this may be the result of larger relative changes in S-wave impedance relative to the P-wave impedance. In this case, both sections have been plotted in P-wave time (effectively, both sections are processed to depth and then converted back to P-wave time). Coulombe et al. (1996) used VSP measurements to analyze AVO effects in a carbonate section. They found that P-P and P-S AVO effects were in evidence and could be modeled (Figures 15 and 16). In the VSP data, the P-S section gave a higher resolution picture in which the top and bottom of the porous zone could be identied. Daures et al. (1999) use 3-C VSP data to guide the processing and interpretation of a four-component (4-C) ocean-bottom seismic (OBS) dataset from the Balder eld, North Sea. They found P and S velocities from the VSP to be useful in calibrating sonic logs,

FIG. 9. Schematic diagram of (a) P-P reection trajectories and (b) P-S conversion points for the VSP case of a surface source and subsurface receivers. The midpoint (MP) and asymptotic conversion point (ACP) for a surface source are indicated.

1354

Stewart et al.

FIG. 10. Interval velocity from VSP and blocked sonic logs. The porous dolomite reservoir shows a signicant V P / VS drop relative to surrounding limestones (from Coulombe et al., 1996). The seismic velocities are somewhat lower than the sonic values.

FIG. 11. Composite plot showing well logs, VSP in depth and two-way time, synthetic seismograms, P-wave surface seismic, and VSP sections. Data are from southern Alberta. Note the great reection activity (and noise) in the converted-wave section (from Geis et al., 1990). The P-P and P-S VSP extracted traces or corridor stacks are labeled P-VET and PS-VET, respectively.

Converted-wave Seismic Exploration: Methods

1355

gathering the OBS data, and processing it. In addition, the VSPextracted traces (VETs) for P-P and P-S waves were important assets in correlating the various seismic sections. Leaney et al. (1999) use multioffset VSP data to build elastic (and anelastic) models for converted-wave simulations. They found that it was important to take anisotropy and attenuation into account when attempting to produce realistic converted-wave seismograms. Multicomponent surface-seismic acquisition An early discussion concerning the surveying, analysis, and interpretation of converted waves was published by Garotta et al. (1985). They used a two-element (vertical and horizontal) geophone for acquisition. The data recorded were quite good. Nevertheless, using a 3-C geophone is now recommended so that off-line effects, misorientation, anisotropy, crooked lines, and 3-D shooting can be handled. Lawton and Bertram (1993) tested four 3-C geophones, using a source at various azimuths to the receivers. The goal of this survey was to determine if the output response (particle motion plot or hodogram) of the horizontal elements would indicate the direction of the source from the receiver. That is, would the geophone response faithfully reproduce anticipated particle motion polarization? They found that the three geophones commercially available at that time all performed well. This was important for further establishing the vector delity of the eld measurements before proceeding to more sophisticated processing. Guevara (2000) conducted similar analysis for the polarization of two-dimensional, three-component (2D3C) data recorded from an ofine source in Blackfoot eld, Alberta. In this case too, the azimuths of the data hodograms compared well to those of the source-receiver eld geometry. Especially useful were the azimuths determined from refracting S-waves and ground roll. Eaton and Lawton (1992) analyzed the fold of a 2-D P-S section and found it to be oscillatory under certain conditions. Careful design is required to avoid regions of low fold. Vermeer (1999) and Lawton and Hoffe (2000) also discuss issues in the design of 3-D surveys, including the raypath asymmetry and depth-targeted techniques. The raypath asymmetry will push reection points toward the receiver lines giving different coverage from that expected from P-wave data (Ronen et al., 1999). Staggering receiver lines (or shooting slightly off the orthogonal) can smooth P-S fold. Shooting on the half-station may also help fold distributions. In 3-D surveys, there may be some operational and processing advantages to shooting parallel to the receiver lines instead of orthogonal to them. However, subsequent analysis methods that rely on a wide range of azimuths (e.g., azimuthal anisotropy determination) would normally require some kind of crossline shooting (Rosland et al., 1999). Processes such as adjacent-trace averaging, depth- and velocity-variant binning, and dip moveout (DMO) will aid in smoothing the nal fold. Acquisition limitations FIG. 13. Frequency coherency spectra from the Carrot Creek, Alberta, surveys for (a) P-P data and (b) P-S data. From the depth of interest, the P-P and P-S dominant frequencies recorded at surface are 32 and 21 Hz, respectively (from Eaton et al., 1991). The recording of land and marine multicomponent data is still somewhat constrained by several factors. Threecomponent geophones and 4-C cables have sometimes been in short supply. As demand continues to increase, more

FIG. 12. Spectra from the P-wave VSP and P-S VSP data of Figure 11. The similar temporal frequency content of P- and S-wave data, averaged over depths from 520 m to 1840 m, indicates a higher spatial resolution of the S-wave data (from Geis et al., 1990).

1356

Stewart et al.

FIG. 14. P-S and P-P VSP sections from the Willesden Green surveys (Stewart et al., 1995). Note the greater reection activity in the P-S section in the shallow regions. Both sections are plotted in two-way P-wave traveltime.

FIG. 15. AVO traces for VSP data. The P-wave and S-wave gathers are from offsets ranging from 80 m to 2500 m (from Coulombe et al., 1996).

Converted-wave Seismic Exploration: Methods

1357

multicomponent sensors are being built. Several types of autoleveling surface geophones (mechanical and solid-state) are currently being deployed which promise to increase planting efciency and delity. There are still issues about the use of arrays or single phones that need to be resolved with careful eld work and processing. However, several tests on land indicate that by the nal migrated-section stage, there is little difference in P-wave data from a single vertical element of the 3-C geophone and conventional vertical-element arrays. There is evidence that 3-C arrays may actually damage the P-S data due to large and rapidly varying S-wave (receiver) statics (Hoffe et al., 2002). Podded geophones or receivers clustered in a radius of about 1 m may provide signal-to-noise ratio enhancements without the associated static degradation. Buried geophones are sometimes said to record a better signal with less noise. We have often used 3-C geophones in augered holes about 0.3 m deep to minimize wind-noise effects, freezing/thawing conditions, and other cultural noise. To test the effect of a deeper plant, we analyzed two test lines with 3-C geophones buried up to depths of 18 m. Somewhat surprising to us, we have not found an improvement in data quality (Cieslewicz, 1999). This may be a consequence of factors such as difculty in obtaining a good 3-C geophone plant at the bottom of a drilled hole, near-surface multipathing, and the loss of signal amplication otherwise realized on the free surface. Other solutions being sought in P-S data acquisition include the source type and parameters to be used for optimal generation of converted waves and the overall design of recording boxes, telemetries, and data harvesting. The cost of a multicomponent survey continues to be an issue, but more experienced crews and plentiful equipment are helping to improve data quality while lowering costs. Marine 4-C equipment and procedures are rapidly advancing as well. Topics of current attention include seaoor geophone

coupling, vector delity (Tree, 1999), the effects of geophone gimbaling, and receiver statics. For 3-D P-S wave data, shear waves can be polarized in any direction, and it is therefore important that the earth response of the two horizontal geophones be identical. When their response differs, data from the components can not be combined optimally in a vector sense for 3-D processing. Gaiser (1998) extended the concept of surface consistency to multicomponent receivers to correct for differences in geophone coupling between inline and crossline detectors from ocean-bottom cable (OBC) data. Vector deconvolution operators are designed by minimizing transverse energy, and are applied to crossline and verticalcomponent data, resulting in multicomponent spectra that are well balanced. Figure 17 shows an example of portions of two crossline receiver gathers before and after compensation. It illustrates how undesirable 8-Hz resonance (Figure 17a) can be attenuated and the bandwidth of P-S reections improved (Figure 17c). Correcting for variations in coupling improves the crossline and inline components for 3-D vector processing and makes the crossline response more consistent for all receiver gathers. P-S data processing As shown in Figure 1, the reected S-wave returns to the surface more vertically than the incident P-wave. Thus, the reection or conversion point is not midway between the source and receiver. Furthermore, this conversion point location moves toward the receiver for shallower reections and larger V P / VS values (Figure 3). Several authors have presented analysis of the asymmetric reection-point trajectory (Chung and Corrigan, 1985; Tessmer and Behle, 1988) and its importance in P-S imaging. Garotta (1985) and Frasier and Winterstein (1990) outline procedures for handling P-S data. Stewart (1991) extended Chung and Corrigans (1985) work to

FIG. 16. AVO traces for VSP data. The P-wave (a) modeled synthetic and (b) eld data show AVO effects, as do the P-S data [(c) synthetic and (d) eld]. The gathers are from source offsets ranging from 80 m to 2500 m (from Coulombe et al., 1996). The top and bottom of the interpreted porous dolomite layer are indicated. Note that there are some character differences among the various P-P and P-S data.

1358

Stewart et al.

describe converted waves where the source and receiver had unequal elevations. Garotta et al. (1985), in an early and insightful paper, processed P-P and P-S data by handling vertical and horizontal channels separately with different statics and velocities. They also introduced the concept of what was later called asymptotic binning. In asymptotic binning, the whole trace is put at the location dened by a reector depth that is large compared to the source-receiver offset. For a single-layer case, this would be X a = X /(1 + VP / VS ), where X a is the conversion point offset from the receiver position and X is the source-receiver offset. Thomsen (1999) showed how an equation of this form could also be used for the case of vertical transverse isotropy (VTI). Using one conversion point location allows gathering for velocity analysis and avoids the complication of trace mapping prior to stack. We note that Garotta (1999) comprehensively revisits converted-wave analysis as the 1999 Distinguished Instructor of the Society of Exploration Geophysicists. Because of the low S velocity in the near surface, receiver statics in the P-S survey can be large. Lawton (1990) found receiver static shifts of about 70 ms for a case in the Alberta plains. Cary and Eaton (1993) found receiver statics of 100 ms, as did Isaac (1996) in a case from Cold Lake, Alberta. Cary and Couzens (1998) also discussed the problems in separating the effects of structure from those of statics in a 4-C case from the Mahogany eld, Gulf of Mexico. Clearly, the near surface has a marked inuence on P-S data as a result of these large and variable statics, but it also has another undesirable effect: attenuation. This remains a limitation of surface P-S analysis. Further work on Q ltering will undoubtedly help. Separation of P and S events on the full 3-C record has been approached using various techniques. In VSP analysis, methods based on Dankbaars (1985) scheme of polarization

analysis and f -k ltering have proven useful. In the marine environment, the responses of ocean-bottom receivers are analyzed using formulations based on a uid-solid interface (e.g., Donati and Stewart, 1996). Other work using 3-C array forming to reduce noise as well as to estimate direction and wave type looks promising. Yuan et al. (1998) showed some excellent 4-C data from the North Sea, where they concluded that the P-S sections were of better quality than the P-P sections. They used a match-lter approach to remove P-S reections from the vertical geophone data. In transmission through an azimuthally anisotropic material, shear waves generally split into fast and slow polarizations. Thus, two polarized converted waves (P-S1 and P-S2 , fast and slow, respectively) arrive at the receiver. These potentially superposed events need to be separated. There are various algorithms for doing this (e.g., Alford, 1986; Harrison, 1992) that rely on searching for two similar events projected onto a new set of rotated horizontal channels that are time delayed and orthogonally polarized. Lou et al. (2000) scan through a series of rotation angles and time delays in 4-C data from the Valhall eld, North Sea, to estimate the S1 and S2 directions and anisotropy magnitude. They nd P-S1 and P-S2 time delays of about 30 ms at the target level and an anisotropy direction consistent with other seismic and well data. Slotboom (1990) considered the velocity analysis problem. He derived a shifted hyperbola equation for NMO correction that can correct the offset traveltimes more accurately than a normal hyperbolic velocity analysis. Including a fourthorder term in the moveout equation can be helpful. Stewart and Ferguson (1996) presented a method to nd an S-wave interval velocity from P-S stacking velocities using the Dix assumption that the stacking velocity is equal to the rms velocity.

FIG. 17. Crossline responses from portions of two different 3-D common-receiver gathers before compensation (a, b) and after compensation (c, d). Gather (a) is an example of a poorly coupled receiver exhibiting 8-Hz resonance, and gather (b) is an example from a receiver of average coupling. After correction, the two receiver gathers are more consistent in their frequency content, indicated by bandwidth and by relative amplitudes of reections and sediment-water interface waves (Gaiser, 1998).

Converted-wave Seismic Exploration: Methods

1359

After NMO, it is important to understand the Fresnel zone (or the averaging aperture in a stacked section) and the potential of P-S data to be migrated. Eaton et al. (1991) derived the P-S Fresnel zone and found that, for the same frequencies, the P-S Fresnel radius is smaller than the corresponding P-P case. However, with the lower P-S frequencies often observed at the surface, P-P and P-S radii work out to be about the same. They also showed that P-S data could be migrated after stack in a kinematic (traveltime) sense. Harrison and Stewart (1993) considered the migration problem further and derived a P-S migration velocity for a horizontally layered material. As with P-waves, P-S reections are shifted in regions of dip. Thus, some procedures must be undertaken in the stacking process to put reections at their correct zero-offset positions. Harrison (1992) developed an equation and procedure for the DMO correction of converted-wave data. Several authors have investigated methods to enhance imaging via prestack time migration in isotropic (e.g. Nicoletis et al., 1998) and anisotropic (e.g., Riste and Fjellanger, 2000) materials. Depth migration of P-S data can be accomplished (Kendall et al., 1998; Hoffe and Lines, 1999), which, in addition to possibly better imaging, creates P-P and P-S sections sharing the same depth axis (as opposed to raw P-P and P-S time axes). The use of positive and negative source-receiver offsets can help with improved velocity analysis and imaging (Dai et al., 2000). Van Dok et al. (1997) outline a 3C-3D processing ow for data acquired in the Wind River Basin, Wyoming. Key steps in their sequence are full P-P processing of the vertical geophone data (especially for source statics and velocities to be used with the P-S data), edits of the horizontal components, coherent noise ltering, surface-consistent statics, deconvolution, CCP binning, Alford rotation, anisotropic layer stripping, P-S NMO, stack, and 3-D f -k migration. With respect to S-wave birefringence (splitting), Gaiser et al. (1997) note that there is considerably more room in P-S analysis for enhancement of the data and assessment of fractures and stress state. Zhu et al., (1999) emphasize the need for careful preprocessing of multicomponent data and the importance of the velocity elds to the nal prestack imaging. Continuing with the analysis of our reectivity section, we may be interested in extracting an estimate of actual rock properties (e.g., impedance, velocity). Stewart (1991) derived a method for converting S-wave reectivity to a shear-velocity log. This method is similar to seismic inversion via the Seislog method, where we basically integrate and exponentiate the seismic trace to provide an estimate of the velocity. Valenciano and Michelena (2000) used a small-offset approximation to equation (1) to develop an inversion procedure for stacked P-S data. Jin et al. (2000) applied an AVO inversion technique to an OBS data set from the North Sea. They illustrated the practical use of AVO inversion for S-wave velocity and density for seismic reservoir characterization; in particular, for uid contact detection and pore uid change. Interpretation techniques P-S data can be interpreted on their own, but generally will be analyzed in conjunction with P-wave sections or volumes. Thus, a primary aspect of interpretation is correlating the P-P and P-S events. There are a number of ways to accomplish this. As indicated earlier, Lawton and Howell (1992) developed a

P-S synthetic seismogram program to assist in the correlation process. This modeling algorithm uses the offset-dependent reectivity and ray-traced traveltimes of both P-P and P-S events to create synthetic seismograms. We need an acoustic velocity for this modeling, but can infer or guess densities and S-wave velocities if they are not otherwise available. These resultant seismograms are, in fact, AVO stacks of offset reectivities. Classically, the zero-offset P-S reection coefcient is zero, but the zero-offset P-S section obviously does not attempt to provide this. Rather, the nal P-S stack to zero-offset traveltime gives an average of the offset-dependent P-S reectivities. We can alter the V P / VS value in the P-S synthetic seismogram to rene the correlation. So, in practice, we tie the P-wave synthetic seismogram to the P-P section, the P-P and P-S synthetics together, and the P-S synthetic seismogram to the P-S section. Chan (1998) suggested plotting the P-P and P-S sections in logarithmic time; that is, displaying the sections as ln t P P and ln t P S , where t P P and t P S are the P-P and P-S two-way reection times, respectively. For a nearly constantV P / VS value, just a bulk shift can then assist in tying the sections. Even using a single V P / VS value for shrinking the P-S time scale in a plot can often provide a general correlation to the P-P section. These correlations or V P / VS values can be used to determine a mapping between t P S and t P P for time conversions. Gaiser (1996) developed a robust multicomponent correlation analysis to obtain average and interval V P / VS values. However accomplished, sections or volumes must ultimately be plotted on the same scale (whether it is P-wave time or depth) to provide an understanding of the subsurface. Migration to depth with V P and VS velocity elds is certainly a goal for many nal analyses. Time-structure sections and isochrons are used in P-S interpretation similarly to P-P techniques. However, we can also take the ratio of P-P and P-S isochrons from the same interpreted horizons to generate V P / VS values. In this case, V P / VS = 2(TP S / TP P ) 1, where TP S and TP P are the corresponding P-S and P-P isochrons. Miller (1996) shows a case of this isochronratio analysis from Lousana eld, Alberta (Figure 18). Larson (1996) and Margrave et al. (1998) develop interpretation techniques for the analysis of 3-D multicomponent volumes. Time slices, horizon and isochron maps, and coupled P-P and P-S results can all be useful. One 3C-3D seismic survey can provide three products: the P-P volume and the anisotropic volumes (P-S1 and P-S2 , where S1 is the fast S-wave and S2 the slow S-wave). Thus, there are a number of new, independent sections to compare, contrast, and integrate. Although this presents some wonderful interpretation opportunities, it can also be difcult and time-consuming, as Rutledal et al. (2000) indicate in their interpretation of 4C-3D data from the Oseberg eld, North Sea.
WHATS LEFT TO DO?

Converted-wave exploration has come a long way in the last several years, but there is still plenty of room left for progress. Determining which recording environments allow acquisition of good P-S data and how to carry out that acquisition are areas for further effort. For example, what vibrator sweep parameters will provide the best P-S data has not yet been the subject of much experimentation. In addition, nding an appropriate balance between capturing high-quality P-wave data as well as P-S data challenges our survey design expertise.

1360

Stewart et al.

Much remains to be done in designing marine surveys, oceanbottom recording systems, and the deployment of equipment. Both cables and discrete nodes have advantages and disadvantages. Some novel acquisition techniques have been attempted. Moldoveanu et al. (1998) indicate that in transition zones P-S data can be acquired by ramming 4-C receivers into the mud. Stewart et al. (2001) deployed a ve-level, three-component VSP tool in a shallow lake and winched it across the lake bottom to acquire data from an onshore vibrator. They found that credible P-P and P-S sections could be constructed from

the recorded data. New solid-state, digital geophones [and newer optical motion sensors (see Bostick 2000)] with three and four components promise broader band recordings. Being able to record high-delity data down to the 1-Hz range opens up exciting possibilities for better wavelet denition and more complete inversions, especially in the converted-wave case. Investigating other modes in multicomponent data will allow us to separate undesirable from desirable modes as well as to make other sections. In some cases, a vertical vibrator may also

FIG. 18. P-P section (top) and P-S section (middle) from Lousana eld, Alberta. The V P / VS value (bottom) is extracted from the interpreted P-P and P-S sections. The lower values in the Paleozoic, from about shot points 190 to 215, are coincident with an underlying oil-bearing reef (Miller, 1996).

Converted-wave Seismic Exploration: Methods

1361

generate enough S-wave energy to allow an S-S or S-P section to be created directly. And, of course, if the signal is there we should try to use it. Fyfe et al. (1993) show a case from Saudi Arabia in which P-S and S-S sections were produced from a vertical vibrator source recorded into 3-C geophones. Furthermore, an S-S section may be produced from dynamite data. Nieuwland et al. (1994) give an example of processing a pS-S survey in The Netherlands. In this case, an upgoing P-wave from the dynamite source converts at the surface to a downgoing S-wave that is reected back to the surface from various layers as an S-wave. Credible S-wave sections are produced, along with their P-wave counterparts, and are interpreted for gas effects. A downgoing shear conversion at the ocean oor and its reection (PS-S) may be extractable from marine 4-C data in some cases (although anecdotally this seems rare). Nonetheless, Tatham and Goolsbee (1984) describe a case of the PS-SP event (the upcoming S-wave converting to a P-wave at the ocean bottom) from offshore western Florida. There is promise that the base of high-velocity layers (HVL) can be imaged using mode-converted S-waves by replacing the P-wave velocity in the HVL with an appropriate S-wave velocity (Purnell, 1992). Additionally, Hanssen et al. (2000) suggest that while locally converted modes are too weak for subbasalt (HVL) imaging, the simple P-S mode may be useful. More sophisticated processing techniques will be useful to resolve statics, identify and use anisotropy, estimate velocities, and perform more accurate migrations (especially in faulted and folded areas). We also need to understand a great deal more about how the near-surface damages S-wave data. Improvements to interpretation methods incorporating vector and multimode concepts (Mueller et al., 1999) will be benecial. As we develop more condence in P-S images, it is becoming useful to repeat the 3-D multicomponent survey to allow the possibility of time-lapse (4-D) imaging. In this case, we are looking for the seismic signature of uid change and movementa very economic, but complicated quest. As the pressure and saturation state of the reservoir is altered in the course of hydrocarbon production, the elastic (as well as acoustic) properties of the rock change, demonstrating the ultimate need for full (3C-4D or 4C-4D) multicomponent recording and analysis.
CONCLUSIONS

ACKNOWLEDGMENTS

We express our deep appreciation to the sponsors of the CREWES Project for their commitment to the development of multicomponent seismology. Many thanks to the CREWES staff, especially former employees Joanie Whittemore and Chuandong (Richard) Xu, for their help in assembling this paper. This paper beneted considerably from the insightful review by Dr. Mike Mueller of BP. We thank him, the other reviewers, and 20 students in a recent graduate course on Converted-wave Seismic Exploration at the University of Calgary for the generous gift of their comments. This work was further supported by a collaborative research grant from the Natural Sciences and Engineering Research Council of Canada (CRD #223726-99).
REFERENCES Aki, K., and Richards, P. G., 1980, Quantitative seismology: Theory and methods: W. H. Freeman and Sons. Alford, R. M., 1986, Shear data in the presence of azimuthal anisotropy: Dilly, Texas: 56th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 476479. Arnold, A., and Chiburis, E., 1998, AVOCurrent status and the future: 68th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 248250. Bostick III, F. X., 2000, Field experimental results of three-component ber-optic seismic sensors: 70th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 2124. Brown, R. J., Stewart, R. R., and Lawton, D. C., 2002, A proposed polarity standard for multicomponent seismic data: Geophysics, 67, 10281037. Caldwell, J., 1999, Marine multicomponent seismology: The Leading Edge, 18, 12741282. Cary, P. W., and Couzens, R. A., 1998, Processing 4-C data from Mahogany eld, Gulf of Mexico: CREWES Research Report, 10, 29.129.25. Cary, P. W., and Eaton, D. W. S., 1993, A simple method for resolving large converted-wave (P-SV) statics: Geophysics, 58, 429433. Castagna, J. P., Batzle, M. L., and Eastwood, R. L., 1985, Relationships between compressional-wave and shear-wave velocities in clastic silicate rocks: Geophysics, 50, 571581. Chan, W. K., 1998, Analyzing converted-wave seismic data: Statics, interpolation, imaging, and P-P correlation: Ph.D. thesis, Univ. of Calgary. Chung, W. Y., and Corrigan, D., 1985, Gathering mode-converted shear waves: A model study: 55th Ann., Internat. Mtg. Soc. Expl. Geophys., Expanded Abstracts, 602604. Cieslewicz, D., 1999, Near-surface characterization using threecomponent buried geophones: M.Sc. thesis, Univ. of Calgary. Clark, D., 2000, Whither geophysical technology? The views of the vendors: The Leading Edge, 19, 862873. Coulombe, C. A., Stewart, R. R., and Jones, M. E., 1996, AVO processing and interpretation of VSP data: Can. J. Expl. Geophys., 32, 4162. Dai, H., Li, X.-Y., and Mueller, M. C., 2000, Compensating for the effects of gas clouds by prestack migration: A case study from Valhall: 70th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 10471050. Dankbaar, J. W. M., 1985, Separation of P- and S-waves: Geophys. Prosp., 33, 970986. Daures, R., Granger, P.-Y., Vuillermoz, C., 1999, 4C OBS data processing guided by well data, Balder eld case history: Presented at the 61st Ann. Conf., Eur. Assn. Geosci. Eng. Deffenbaugh, M., Shatilo, A., Schneider, B., and Zhang, M., 2000, Resolution of converted waves in attenuating media: 70th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 11781180. Donati, M. S., and Stewart, R. R., 1996, P- and S-wave separation at a liquid-solid interface: J. Seis. Explor., 5, 113127. Duey, R., 2001, Multicomponent acquisition: From side dish to entree?: Harts E&P, 74, no. 1, 6870. Duffant, K., Landr, M., and Rogn, H., and Al-Najjar, N. F., 2000, Shear-wave elastic impedance: The Leading Edge, 19, 1222 1229. Eaton, D. W. S., and Lawton, D. C., 1992, P-SV stacking charts and binning periodicity: Geophysics, 57, 745748. Eaton, D. W. S., Stewart, R. R., and Harrison, M. P., 1991, The Fresnel zone for P-SV waves: Geophysics, 56, 360364.

The reection seismic method has used P-waves for many yearsand with great success. However, there is more to be done in exploration seismology (especially with regard to lithology and uids) by using the other elastic modes that are part of the seismic survey, particularly the P-S converted wave. P-S seismic exploration has been developing for about 20 years. The basic tools of P-S analysis include drill-core evaluation, shear-sonic logs, 3-C VSP, elastic synthetic seismograms, and of course multicomponent land and marine seismic surveys. Many of the fundamental techniques for handling these data have been established. There are currently a number of companies providing multicomponent seismic services and research groups attempting to further develop the method. Recent examples of successful P-S imaging indicate a maturing of the technology.

1362

Stewart et al. 13441347. Leaney, S., Cao, D., and Tcherkashnev, S., 1999, Calibrating anisotropic, anelastic models for converted-wave simulation: 61st Ann. Conf., Europ. Assn. Geosci. Eng., Extended Abstracts, 652. Lou, M., Zhang, Y., and Pham, L. D., 2000, Shear-wave splitting and its implied fracture orientation analysis from PS waves in marine multi-component seismic data: First Break, 18, No. 12, viiix. Margrave, G. F., Lawton, D. C., and Stewart, R. R., 1998, Interpreting channel sands with 3C-3D seismic data: The Leading Edge, 17, 509 513. Mari, J. L., Coppens, F., Gavin, P., and Wicquart, E., 1994, Full waveform acoustic data processing: Editions Technip. Mavko, G., Mukerji, T., and Dvorkin, J., 1998, The rock physics handbook: Tools for seismic analysis in porous media: Cambridge Univ. Press. Miller, S. L. M., 1996, Multicomponent seismic data interpretation: M.Sc. thesis, Univ. of Calgary. Miller, S. L. M., and Stewart, R. R., 1990, The effect of lithology, porosity and shaliness on P- and S-wave velocities from sonic logs: Can J. Expl. Geophys., 26, 94103. Moldoveanu, N., Rink, U., and Van Baaren, P., 1998, Multicomponent acquisition in transition zone environment: An experimental study: 68th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 738739. Mueller, M. C., Boyd, A. J., and Esmersoy, C., 1994, Case studies of the dipole shear anisotropy log: 64th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 11431146. Mueller, M. C., Barkved, O. I., and Thomsen, L. A., 1999, A strategy for vector interpretation of multicomponent ocean bottom seismic data: 61st Ann. Conf., Eur. Assn. Geosci. Eng., poster P067. Nicoletis, L., Svay-Lucas, J., and Prigent, H., 1998, True amplitude prestack imaging of converted waves: 68th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 734737. Nieuwland, F., Marschall, R., Papaterpos, M., and Sharp, D., 1994, An example of the use of shear waves in seismic exploration: J. Seis. Explor., 3, 520. Purnell, G. W., 1992, Imaging beneath a high-velocity layer using converted waves: Geophysics, 57, 14441452. Rider, M., 1996, The geological interpretation of well logs, 2nd ed.: Whittles Publishing. Riste, P., and Fjellanger, J.-P., 2000, Application of 3-D prestack time migration for converted waves: 70th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 11461149. Rodriguez, C., 2002, Advanced marine seismic methods: Ph.D. thesis, Univ. of Calgary. Ronen, S., Bagaini, C., Bale, R., Caprioli, P., and Keggin, J., 1999, Coverage and illumination of sea bed 4C surveys: 61st Ann. Mtg., Eur. Assn. Geosci. and Eng., Extended Abstracts, 618. Rosland, B., Tree, E. L., and Kristiansen, P., 1999, Acquisition of 3D/4C OBS data at Valhall: 61st Ann. Mtg., Eur. Assn. Geosci. Eng., Extended Abstracts, 617. Rutledal, H., Ertresvag, E. T., and Berg, O. E., 2000, Interpretation and analysis of the 1997 3D multicomponent seismic survey covering a part of the Oseberg eld in the North Sea: 70th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 570573. Schon, J. H., 1998, Physical properties of rocks: Fundamentals and principles of petrophysics, 2nd ed.: Pergamon. Slotboom, R. T., 1990, Converted wave (P-SV) moveout estimation: 60th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 11041106. Stewart, R. R., 1991, Rapid map and inversion of P-SV waves: Geophysics, 56, 859862. Stewart, R. R., and Ferguson, R., 1996, Shear-wave interval velocity from P-S stacking velocity: Can J. Expl. Geophys., 32, 139142. Stewart, R. R., Ferguson, R. F., Miller, S. L. M., Gallant, E., and Margrave, G. F., 1996, The Blackfoot seismic experiments: Broadband, 3C-3D, and 3-D VSP surveys: Can. Soc. Expl. Geophys. Recorder, 21, no. 6, 710. Stewart, R. R., Huddleston, P. D., and Kan, T. K., 1984, Seismic versus sonic velocities: A vertical seismic proling study: Geophysics, 49, 11531168. Stewart, R. R., Lu, H., Bland, H., and Mewhort, L. E., 2001, A lakebottom cable seismic survey: Acquisition and processing: Can. Soc. Expl. Geophys. Recorder, 8, 3740. Stewart, R. R., Pye, G., Cary, P. W., and Miller, S. L. M., 1995, Interpretation of P-SV seismic data: Willesden Green, Alberta: CREWES Research Report, 5, 15.115.19. Tatham, R. T., 1982, V p / Vs and lithology: Geophysics, 47, 336344. Tatham, R. T., and Goolsbee, D. V., 1984, Separation of S-wave and P-wave reections, offshore western Florida: Geophysics, 49, 5, 493 508. Tessmer, G., and Behle, A., 1988, Common reection point data stacking technique for converted waves: Geophys. Prosp., 36, 671688.

Esmersoy, C., Chang, C., Kane, M., Coates, R., Tichelaar, B., and Quint, E., 1998, Acoustic imaging of reservoir structure from a horizontal well: The Leading Edge, 17, 940946. Frasier, C., and Winterstein, D., 1990, Analysis of conventional and converted mode reections at Putah Sink, California using threecomponent data: Geophysics, 55, 646659. Fyfe, D. J., Dent, B. E., Kelamis, P. B., Al-Mashouq, K. H., and Nietupski, D. A., 1993, Three-component seismic experiments in Saudi Arabia: Presented at the 55th Ann. Mtg., Eur. Assn. Expl. Geophys. Gaiser, J. E., Fowler, P. J., and Jackson, A. R., 1997, Challenges for 3-D converted-wave processing: 67th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 11991202. Gaiser, J. E., 1996, Multicomponent VP / VS correlation analysis: Geophysics, 61, 11371149. 1998, Compensating OBC data for variations in geophone coupling: 68th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 14291432. Garotta, R., 1985, Observation of shear waves and correlation with P events, in Dohr, G. P., Ed., Seismic shear waves: Part B: Applications: Geophysical Press, 186. 1999, Shear waves from acquisition to interpretation: Soc. Explor. Geophys. Garotta, R., Marechal, P., and Magesan, M., 1985, Two-component acquisition as a routine procedure for recording P-waves and converted waves: Can. J. Expl. Geophys., 21, 4053. Geis, W. T., Stewart, R. R., Jones, M. J., and Katopodis, P. E., 1990, Processing, correlating, and interpreting converted shear waves from borehole data in southern Alberta: Geophysics, 55, 660669. Goodway, W., Chen, T., and Downton, J., 1997, Improved AVO uid detection and lithology discrimination using Lame petrophysical parameters: , , and / uid stack, from P and S inversions: Presented at the Ann. Nat. Mtg., Can. Soc. Expl. Geophys. Gueguen, Y., and Palciauskas, V., 1994, Introduction to the physics of rocks: Princeton Univ. Press. Guevara, S. E., 2000, Analysis and ltering of near-surface effects in land multicomponent seismic data: M.Sc. thesis, Univ. of Calgary. Hanssen, P., Li, X.-Y., and Ziolkowski, A., 2000, Converted waves for sub-basalt imaging?: 70th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 11741177. Hardage, B. A., 2000, Vertical seismic proling: Principles, 3rd ed.: Elsevier. Harrison, M. P., 1992, Processing of P-SV surface-seismic data: Anisotropy analysis, dip moveout, and migration: Ph. D. thesis, Univ. of Calgary. Harrison, M. P., and Stewart, R. R., 1993, Post-stack migration of P-SV seismic data: Geophysics, 58, 11271135. Hinds, R. C., Anderson, N. L., and Kuzmiski, R. D., 1996, VSP interpretive processing: Theory and practice: Soc. Explor. Geophys. Hoffe, B. H., and Lines, L. R., 1999, Depth imaging by elastic waves Where P meets S: The Leading Edge, 18, 370372. Hoffe, B. H., Margrave, G. F., Stewart, R. R., Foltinek, D. S., Bland, H. C., and Manning, P. M., 2002, Analyzing the effectiveness of receiver arrays for multicomponent seismic exploration: Geophysics, 67, in press. Isaac, J. H., 1996, Seismic methods for heavy oil reservoir monitoring: Ph.D. thesis, Univ. of Calgary. Jin, S., Cambois, G., and Vuillermoz, C., 2000, Shear-wave velocity and density estimation from PS-wave AVO analysis: application to an OBS dataset from the North Sea: Geophysics, 65, 14461454. Jrstad, A., Mukerji, T., and Mavko, G., 1999, Model-based shear-wave velocity estimation versus empirical regressions: Geophys. Prosp., 47, 785797. Kendall, R. R., and Davis, T. L., 1996, The cost of acquiring shear waves: The Leading Edge, 15, 943949. Kendall, R. R., Gray, S. H., and Murphy, G. E., 1998, Subsalt imaging using prestack depth migration of converted waves: Mahogany Field, Gulf of Mexico: 68th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 20522055. King, M. S., Marsden, J. R., and Dennis, J. W., 2000, Biot dispersion for P- and S-wave velocities in partially and fully saturated sandstones: Geophys. Prosp., 48, 10751089. Larson, G. A., 1996, Acquisition, processing, and interpretation of P-P and P-S 3-D seismic data: M.Sc. thesis, Univ. of Calgary. Lawton, D. C., 1990, A 9-component refraction seismic experiment: Can. J. Expl. Geophys., 26, 716. Lawton, D. C., and Bertram, M. B., 1993, Field tests of 3-component geophones: Can. J. Expl. Geophys., 29, 125131. Lawton, D. C., and Hoffe, B. H., 2000, Some binning issues for 4C-3D OBC survey design: 70th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 1720. Lawton, D. C., and Howell, C. T., 1992, P-SV and P-P synthetic stacks: 62nd Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts,

Converted-wave Seismic Exploration: Methods Thomsen, L. A., 1999, Converted-wave reection seismology over inhomogeneous anisotropic media: Geophysics, 64, 678690. 2001, Seismic anisotropy: Geophysics, 66, 4041. Toksoz, M. N., and Johnston, D. H., 1981, Seismic wave attenuation: Soc. Explor. Geophys. Toksoz M. N., and Stewart, R. R., 1984, Vertical seismic proling, Part B: Advanced concepts: Geophysical Press. Tree, E. L., 1999, The vector indelity of the ocean bottom multicomponent seismic acquisition system: 61st Ann. Mtg., Eur. Assn. Geosci. Eng., Extended Abstracts, 619. Valenciano, A. A., and Michelena, R. J., 2000, Stratigraphic inversion of poststack PS converted waves data: 70th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 150153. Van Dok, R. R., Gaiser, J. E., Jackson, A. R., and Lynn, H. B., 1997, 3-D converted-wave processing: Wind River Basin case history:

1363

67th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 12061209. Vermeer, G. J. O., 1999, Converted waves: Properties and 3D design: 69th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 645648. Xu, P., and Parra, J. O., 1998, A parametric study of synthetic dipole logging data in azimuthally anisotropic formations: 68th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 260 263. Yuan, J., Li, X., Ziolkowski, A., and Strijbos, F., 1998, Processing 4C sea oor seismic data: A case example from the North Sea: 68th Ann. Internat. Mtg., Soc. Expl. Geophys., Expanded Abstracts, 714 717. Zhu, X., Altan, S., and Li, J., 1999, Recent advances in multicomponent processing: The Leading Edge, 18, 12831288.

You might also like