You are on page 1of 12

H o n o r a r y L e c t u r e

3UDFWLFDOVHLVPLFSHWURSK\VLFV7KHHIIHFWLYHXVHRIORJGDWD
IRUVHLVPLFDQDO\VLV
TAD M. SMITH, Apache Corporation
Downloaded 03/01/13 to 141.117.79.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

A quick scan of the SEG web site shows that the phrase
“seismic petrophysics” has been used explicitly in the
title of a paper or abstract six times, the earliest of which was
with the justification that many of these edits are below seismic
resolution, and are therefore not important. In many instances
this may well indeed be true, but given that we are in search of
by Williams et al. in 1996 (“The Hugoton cross-well survey; ever more subtle seismic responses, it is my opinion that these
A direct look at stratigraphy, seismic petrophysics and shale edits should be performed with care anytime we are generating
anisotropy”). However, the first attempt at a definition and an log-based seismic models or using log data to interpret seismic
expanded description of “seismic petrophysics” was published responses. Failure to properly edit log data can lead to erroneous
by Wayne Pennington in The Leading Edge in 1997 (“Seismic assumptions and expectations from seismic amplitudes.
petrophysics: An applied science for reservoir geophysics”).
He defines seismic petrophysics as follows: “…the purposeful Identification of spurious log data
application of rock physics theory, as calibrated by laboratory Well log data can be faulty for any number of reasons, the most
and well measurements, to the interpretation of seismic data…” common of which are poor wellbore conditions (e.g., washouts)
The discussion of “seismic petrophysics” in Pennington’s pa- and problems with the measurements (e.g., cycle skips on sonic
per, as well as most subsequent uses of the phrase in the pub- data). Examination of caliper log data is the most common tech-
lished literature, focused largely on what today is simply referred nique for identifying areas in the wellbore that potentially con-
to as “rock physics.” The 2004 paper by Kittridge et al. (“Seis- tain bad data. This is especially relevant for density data, as the
mic petrophysics for clean sandstones: Integrated interrogation density tool is a pad type of device and needs to be in contact
of lab- and well-based data for improved rock physics model- with the wellbore wall (Figure 1).
ing”) specifically mentions the importance of fully integrating It is important to recognize that not all wellbore washout will
geological and petrophysical data into the rock physics workflow, result in bad data. Conversely, lack of washout in the wellbore
although this is also implied in the paper by Williams et al. This
is an important aspect of seismic petrophysics that I subscribe
to; that is, the need to understand the underlying geology and
petrophysical properties of the reservoir (and the bounding lay-
ers) prior to developing any rock physics or AVO models. An
important corollary is that rock physics modeling should not be
considered a substitute for careful analysis of log or core measure-
ments. Application of the wrong rock physics model may allow
in-situ velocities to be predicted, but it will not yield any insights
into the underlying controls on the elastic properties of the sub-
surface rocks and, worse, may lead to incorrect forward models.
This could lead to potentially misleading and expensive mistakes,
as I will attempt to demonstrate throughout this document.
Because of the breadth and multidisciplinary nature of seis-
mic petrophysics, it is not possible in an overview paper to cover
in any detail all the various topics that could be discussed. In-
stead, I have decided to focus on four aspects of seismic petro-
physics that I consider to be the “workhorses” of the discipline:
(1) log editing and conditioning, (2) shear velocity prediction,
(3) fluid replacement modeling, and (4) porosity modeling.
While others may disagree with my assessment of what consti-
tutes the “workhorses” of seismic petrophysics, it is my opinion Figure 1. Washout effects on density. Caliper data are often the
that flawed application of these four fundamental processes will best indicator of potentially bad data in well logs. In this example,
compromise all subsequent rock physics and AVO modeling. the density excursion to low values across the washout zone is not
supported by evaluation of other curves (e.g., deep resistivity). Also
Data editing and conditioning notice the deviation of the density response from the shale background
trend at a depth of approximately 7200 ft. It is important to recognize
Careful conditioning of well-log data prior to the application that not all wellbore washout will result in low-quality density data.
of a rock physics or geophysical model is an often overlooked, Notice that zone A has some washout as indicated by the caliper
or over-simplified, aspect of the rock physics workflow. Editing log (approximately 1–1.5 inches of washout), yet the density data
frequently requires depth shifting, estimation of pseudodata to are consistent with the density trend in this well and for the area.
Conversely, analysis of the caliper log across zone B shows over six
replace bad log data, and invasion corrections (if necessary). Un- inches of washout. The low densities across zone B will affect the
fortunately, geophysicists often oversimplify the editing process, modeled AVO response, and should be replaced with pseudodata.

1128 The Leading Edge October 2011


H o n o r a r y L e c t u r e

does not guarantee that the log data are useable. It is therefore
also useful to examine multiwell depth-trend plots and crossplots
for identification of anomalous data (Figure 2a and 2b). In most
subregions within a basin, multiwell trend-plot analysis will
show relatively consistent sonic and density behavior with depth.
Large deviations from this expected trend can therefore be useful
Downloaded 03/01/13 to 141.117.79.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

indicators of questionable data. However, deviations from the


“normal” trend may also be due to unexpected lithologies, or
complexities associated with the structural configuration, across
an area. It is therefore important that some knowledge of the
local geology be incorporated into this type of analysis. Useful
tools to incorporate in this analysis are mudlogs and structural
geology maps.
For multiwell trend-plot analysis, it is important that the
correct depth reference be utilized. In poorly consolidated and
compacting offshore basins, the most useful depth reference is
depth below mudline (DBML). This is because DBML is a use-
ful proxy for effective stress in these environments. In basins with
consolidated rocks (e.g., onshore Paleozoic or Mesozoic basins),
it is usually necessary to use a key stratigraphic horizon or acous-
tic marker as the depth reference.

Generation of pseudodata
Numerous empirical transforms are available in the published
literature to assist with the generation of pseudo density or ve-
locity data (shear-wave prediction will be discussed in a later
section). For density, commonly used models include those of
Gardner et al. (1974), Castagna et al. (1991), and Wang (2000).
For sonic data, some examples of empirical models include those
by Wyllie et al. (1956), Raymer et al. (1980), Han (1986), Eb-
erhart-Phillips (1989), and many others (also see Avseth et al.,
2007). It is important to recognize that all published coefficients
are lithology specific, and do not apply on a global scale. Figure 2. Effective use of trend plots for identifying bad density data.
Although these empirical transforms provide useful guides (a) Evaluation of multiwell trend plots is an effective technique for
for understanding velocity and density behavior, they generally quickly identifying spurious data (sonic and density). In this example,
do a poor job of predicting velocity and density logs for the pur- note the anomalously high density values from approximately 5000–
5100 ft. These high values could be due to an unexpected lithology
pose of seismic analysis (Figure 3). The most rapid and robust (e.g., carbonates); yet comparison to other logs in the same well (e.g.,
approach for log prediction, in my experience, is via multilinear resistivity) and examination of mudlog data show no unusual rock
regression (MLR; Figure 4). During the MLR training process it types. (b) Examination of the caliper and DRHO data also indicate
is important that the user not incorporate bad data, or data that that there was an unexplained problem with the density measurement
have been affected by near wellbore damage. This applies to both across this zone (mudballs?). In this particular example, the sonic data
may also be affected by poor borehole conditions (track 7). Failure to
the curve being predicted, as well as to the curves being used for examine the density data within the framework of a multiwell analysis
the prediction. could lead the user to miss this problem area, potentially leading to
The logs that have the highest correlation with velocity typi- problems with seismic well ties.
cally are the following:
4) Gamma Ray (or some indication of lithology; e.g., Vclay)
1) Neutron porosity
2) Resistivity The order in which the above logs will best correlate with
3) Density velocity or density will vary, in part, on the degree of consolida-
4) Gamma ray (or some indication of lithology; e.g., Vclay) tion in the subsurface and the composition of the rock. It is also
important to recognize that MLR is best applied over zones of
The logs that have the highest correlation with density typi- limited depth range in order to avoid compaction effects.
cally are the following:
Shear-velocity prediction
1) Neutron porosity (if available) It is common practice in rock physics to estimate a shear log for
2) Velocity the purposes of AVO modeling. The need to estimate a shear
3) Resistivity velocity is often necessary because no measured shear velocity

October 2011 The Leading Edge 1129


H o n o r a r y L e c t u r e
Downloaded 03/01/13 to 141.117.79.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 3. Application of empirical trends. It is common practice to apply various empirical trends to estimate density and sonic logs. In this plot,
density is calculated using Gardner et al. (1974) and Castagna et al. (1991). Neither model adequately predicts density, highlighting the risk
of uncalibrated application of these models. It is possible, however, to calibrate the “Gardner coefficients” on a local basis to better estimate density.

is available in the well, or because the quality of the measured than the shear-wave velocity predicted using the coefficients of
shear-wave data is too low for modeling purposes. It is also use- Greenberg and Castagna (1992); most of this large deviation oc-
ful to estimate a shear velocity for the purposes of evaluating the curs in the shale, whereas the sandstone shear velocities closely
quality of a measured shear log, as well as guiding the editing conform to the predicted values (Figure 5b). Although these
process of a measured shear log. large deviations are not direct evidence that the data are in error,
such large deviations typically do not occur with high quality
Recognizing bad data shear-velocity data. As a general approximation, if the measured
There are some simple techniques for assessing the quality of a shear velocity deviates from the estimated shear log by more
measured shear log, the most effective of which is an examina- than 15%, the data quality should be carefully assessed.
tion of the measured data for anomalously low or high VP/VS The quality of an S-wave log is probably best assessed by ex-
ratios (Figures 5 and 6); negative Poisson’s ratios (VP/VS < 1.414) amination of VDL displays of stacked semblance plots, if avail-
are clearly in error and need to be corrected (Figure 6). In many able (Figure 7). This provides the user with perhaps the most
instances, comparison of the measured shear data to an esti- robust tool for recognizing spurious shear data. If raw waveform
mated shear log can provide useful guidance on shear-log qual- data are available, it is possible (desirable) to reprocess the data in
ity. For example, examination of the shear-wave data in Figure an attempt to improve the results.
5 shows that the measured shear log is upwards of 25% slower

1130 The Leading Edge October 2011


H o n o r a r y L e c t u r e

Estimating a shear log


Techniques for estimating S-wave logs include (1) application
of effective medium models (e.g., Kuster and Toksöz, 1974), (2)
application of heuristic models (e.g., Xu and White, 1994), and
(3) application of one of the many empirical models published
in the literature (e.g., Greenberg and Castagna, 1992). Based
Downloaded 03/01/13 to 141.117.79.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

on my experience, the coefficients of Greenberg and Castagna


are the quickest and easiest to apply, and work reasonably well
on a global scale. The effective medium and heuristic models
are also effective techniques for shear velocity prediction, but
they are more difficult to implement than empirical techniques,
and require the application of both a frame property model and
Gassmann’s equation (1954). They also require that the user
make assumptions about the pore geometry of both sand and
shale. Given the considerable uncertainty in the shear velocity
estimation process, it is my opinion that these heuristic models
are no more accurate than empirical models.

Caveats with Greenberg and Castagna


One of the problems with the Greenberg and Castagna coef-
ficients is that they will produce VP/VS ratios that are too low
for poorly consolidated rocks (Figure 8a). Vernik et al. (2002)
correct for this by adding a “tail” to the coefficients when the
P-wave velocity is slower than approximately 2500 m/s (Figure
8a). VP/VSs ratios for “slow” formations can also be modeled us-
ing the heuristic models of Krief et al. (1990) and Nur et al.
(1995; Figure 8b). It is interesting to note that when using ei-
ther heuristic models or effective medium models, VP/VS ratios
for consolidated quartz sands (P-wave velocity > approximately
2500 m/s) are lower than those predicted using Greenberg and
Castagna (Figure 8b). In order to match the VP/VS ratios pre-
Figure 4. Problem density at casing points. (a) Casing points are
dicted by Greenberg and Castangna, small concentrations of commonly washed out, leading to spurious density data (“casing point
brine-saturated microcracks must be added to the rock matrix. prospects”). In this example, multilinear regression was utilized to
Perhaps most importantly, regardless of the approach taken generate pseudodensity across the casing point, resulting in a density
when estimating an S-wave log, it is imperative that the user un- log that is consistent with other measurements (LWD resistivity and
derstand the underlying lithology. This is because rock compo- P-wave velocity). (b) Offset synthetic illustrating the effects of the
density quality on the modeled seismic response. While incorrect
sition provides a first-order control on dry frame VP/VS ratios, density data will not affect time-depth relationships, they will affect
especially for consolidated rocks. This is illustrated on Figure 8a the ability of the interpreter to use amplitudes to tie well logs to seismic
and 8b by the addition to the plot of the fused glass beads by data. Failure to recognize the bad density data in this well will clearly
Berge et al. (1995). Even though glass beads are not naturally degrade the quality of the well-to-seismic tie.
occurring materials, they can be used to illustrate how applica-
tion of Greenberg and Castagna coefficients, in the absence of eling and quantifying the various fluid scenarios which might
proper lithologic calibration, can lead to the wrong VP/VS ratio give rise to an observed AVO or 4D response. Details of the
(especially for the higher-velocity samples). Gassmann workflow will not be covered in this note, as they
Figures 9 and 10 illustrate the effect on a seismic model are documented in detail in numerous other publications (e.g.,
when the subsurface lithology is not understood during the Mavko et al., 1998; Smith et al., 2003; Avseth et al., 2005; and
shear velocity estimation process. In this example, the reservoir many others). For a discussion on the sources of uncertainty and
consists of complex lithology (lithic arenite), and contains less pitfalls associated with the application of Gassmann, the reader
than 30% quartz (Figure 9b; Helmold et al., 2006). Treating this is referred to comments by Mavko, and Katahara and Smith
reservoir as quartz sand will result in an AVO gradient response (both in the CSEG Recorder, 2005).
that is more negative than that calculated using the measured One of the more problematic areas in the application of
data (Figure 10). This potentially establishes false expectations Gassmann is in the calculation of the dry frame bulk modulus,
for the interpretation of prestack seismic data. or Kdry (Figure 11). Explicit calculation of the dry frame bulk
modulus using Gassmann’s equation frequently results in Kdry
Fluid substitutions values that are noisy and frequently negative (Figure 11). Us-
Fluid substitution is an important part of seismic attribute ing these Kdry values for forward modeling various fluid scenarios
work, as it provides the interpreter with a reliable tool for mod- will result in P-wave velocities that over estimate the fluid effects

1132 The Leading Edge October 2011


H o n o r a r y L e c t u r e
Downloaded 03/01/13 to 141.117.79.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 6. Negative Poisson’s ratios are commonly encountered when


computed from P- and S-wave sonic data. There are numerous reasons
for this (artifacts associated with data acquisition, quality, and
processing, as well as anisotropy in near-wellbore subsurface stresses),
but because negative PR values probably do not occur in naturally
occurring materials in the subsurface, they need to be removed. In most
instances, it is most effective to edit the shear-velocity data using one
of the techniques discussed in the text, although negative PR values
sometimes are due to faulty P-wave data.

Skelt, 2004; Katahara, 2004). However, examination of various


heuristic models suggests a simple approximation for generating
reliable fluid substitution models in difficult environments. The
heuristic models of Krief et al. (1990) and Nur et al. (1995) tell
us the following useful equivalency:
Figure 5. A quick and useful technique for assessing shear-velocity
quality is to compare the measured shear log to an estimated shear
log. (a) Log display of data, and (b) crossplot display of the data. The
light gray lines are lines of constant VP/VS ratios (note that a VP/VS
ratio of 1.414 is a Poisson’s ratio of zero).The coefficients of Greenberg
and Castagna were used in this example for S-wave prediction. Small
differences between the measured and the estimated data (< 10%) Kdry is the dry frame bulk modulus of the porous rock, μ is
may be due to uncertainties in the lithology, whereas larger differences the shear modulus of the porous rock, K0 is the bulk modulus
may be due to problems with the measurement or data processing. of the mineral matrix, and μ0 is the shear modulus of the min-
Examination of (a) and (b) shows that the measured S-wave velocities eral matrix. An equivalent statement is that the Poisson’s ratio of
across the shales are consistently slower than the predicted values;
shear velocities across the sandstone layers more closely conform to the porous dry frame is equal to the Poisson’s ratio of the min-
expectations. The relatively large differences between the measured and eral matrix (Nur et al, 1995). This has also been demonstrated
predicted S-wave velocities in this example, as well as experience in experimentally by Spencer et al. (1994). Thus, the composition
this basin, suggest that the measured S-wave data are faulty. Failure of the mineral matrix is a first-order control on the ratio Kdry/μ
to recognize the low quality of these shear data will result in invalid ratio; that is, the higher the K0/μ0 ratio of the mineral matrix, the
AVO models, and possibly establish unrealistic expectations for prospect
risking. higher the Kdry/μ ratio of the porous dry frame. Quartz is a rare
mineral in that it has a Kdry/μ ratio less than 1. Thus, the addition
of most minerals other than quartz to the rock matrix will cause
and generate negative Poisson’s ratios (Skelt, 2004). This is com- the dry frame PR to increase well above the value of 0.08 (the
mon in shaly sandstones, low-porosity sandstones, and intervals value for pure quartz; Figure 12). It is therefore unreasonable to
where the lithologic model is in error. Kdry values can also be in expect that all reservoir sands will have a dry frame Poisson’s ratio
error if the wrong saturation model is used. The most common of 0.08. Another common assumption in rock physics is for the
example of this would be when the water saturation is assumed special case when Kdry = μ (PRdry = 0.12). This will clearly not be
to be 100%, yet the reservoir contains low-saturation gas. the case for complex rocks with low quartz content (Figure 9). In
Various techniques and models have been proposed for per- many instances, PRdry values for these sands will be greater than
forming fluid substitutions in these difficult environments (e.g., 0.20. These relatively high dry frame Poisson’s ratios values can
October 2011 The Leading Edge 1133
H o n o r a r y L e c t u r e
Downloaded 03/01/13 to 141.117.79.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 8. VP/VS crossplots. For both (a) and (b), the blue and green
lines are computed for sands and shales using the Greenberg and
Castagna coefficients. (a) Three data sets are posted on this crossplot
to illustrate some pitfalls associated with application of Greenberg
and Castagna coefficients. As long as the rocks consist primarily of
clean quartz sands (e.g., the Han samples; these are the 10 samples
in this data set with no clay) and clay-rich shales, the application of
Greenberg and Castagna coefficients is often an effective predictor
of shear-wave velocity and VP/VS ratios. If, however, the rocks are
Figure 7. Anomalously high Poisson’s ratios are often observed from composed of minerals other than quartz, Greenberg and Castagna
P- and S-wave sonic data (also see Figure 5). (a) Note the high PR will typically predict VP/VSratios that are too low. The glass beads of
values at approximately 13,100 ft. Occasionally, this can be due to Berge et al. (blue circles) are included on this plot simply to illustrate
unusual lithologies (e.g., ash beds and some coals). The high PR values the important effect of composition on the resultant VP/VS ratios. For
in this example are not due to an unusual lithology, but due to a weak velocities less than 3.5 km/s, Greenberg and Castagna would predict
flexural wave across this interval and the automatic tracking of a later VP/VS ratios that are close to the measured values. For velocities greater
event. (b) In the absence of the raw waveforms, it would be difficult to than approximately 4 km/s, Greenberg and Castagna would predict
determine the reason for the PR values observed across this zone. VP/VS ratios that are too low. Although glass beads are not a naturally
occurring material, they are useful for illustrating the important effect
of composition on VP/VS ratios. Similarly, application of sandstone
coefficients to a limestone interval will generate VP/VSs ratios that are
potentially have an large impact on the computed AVO gradient far too low. Failure to understand the composition of the rock can
response. lead to the prediction of the wrong VP/VS ratios, and consequently
Equation 1 becomes a powerful tool for generating a pseudo misleading AVO models. Finally, for poorly consolidated sediments
Kdry during the fluid substitution process. If a reliable shear mod- (VP < ~2.5 km/s), Greenberg and Castagna predict will predict VP/
VS ratios that are too low. The unconsolidated sands included on this
ulus can be computed from measured data, or if one can be rea- crossplot are from synthetic sand mixtures, with velocities measured at
sonably estimated (from a pseudo shear velocity), then a pseudo various effective stress states (data courtesy of Manika Prasad, Colorado
Kdry can be computed by simple rearrangement of Equation 1: School of Mines). Vernik et al. have proposed a modified model to
account for these low-velocity rocks (a). (b) Various heuristic models
can also be used to better understand the fundamental controls on VP/
VS ratios (see text for discussion). In this plot, quartz sand models
are generated by combining Gassmann’s equation with Krief et al.
This equation can be combined with Gassmann’s equation to (1991) and Nur et al. (1995). For pure quartz, both models predict
solve some of the difficult problems frequently encountered dur- VP/VS ratios that are lower than those predicted using Greenberg
ing the fluid substitution process (shaly sands, invasion correc- and Castagna. The higher VP/VS ratios predicted by Greenberg and
tions, and negative PR values). In order to do this, the user must Castagna may be due to the presence of brine-saturated microcracks in
the rock matrix.

1134 The Leading Edge October 2011


H o n o r a r y L e c t u r e
Downloaded 03/01/13 to 141.117.79.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 9. Effect of composition on shear-velocity prediction and VP/


VS ratios. In this example, the shear velocity across the zone of interest
(ZOI) is modeled as a pure quartz sandstone (red curves in tracks 8
and 9). Although there are no carbonate minerals in this reservoir, note
from track 3 that this interval has a neutron-density response similar
to that of a limestone. This should be an indicator that a quartz sand
model for shear-velocity prediction will not be correct. (a) Comparison
of the measured data (blue curves, tracks 5 and 6) to the modeled
result (red curves, tracks 5 and 6) shows that a shear-velocity predictor
based on pure quartz sandstone will result in VP/VS ratios that are too
low. (b) Thin section photomicrograph from a representative sample
from this formation (not from this well). Many of the clear grains and
cements in this thin section are analcime (a type of zeolite). (c) Ternary
diagram of point-count data from this area. Examination of available
petrographic and compositional data shows that this interval has less
than 30% quartz, and is dominated by lithic fragments. It is the lack
of a rigid quartz framework (b) and the high volume fraction of lithic
fragments that give rise to the relatively high VP/VS ratios observed in
the log data. Failure to recognize the complex composition of this rock
will result in predicting VP/VS ratios that are too low (a).

first estimate K0 and μ0 of the mineral matrix and then compute


the pseudo Kdry using Equation 2. This Kdry is combined with
Gassmann’s equation to compute a P-wave velocity for the in-
situ case (Figure 13). Assuming the in-situ P-wave velocity is not
affected by invasion, small adjustments can be made to K0 and
μ0 in order to match the measured P-wave velocity. If large ad-
justments are required to K0 and μ0, the lithologic model may be
incorrect, or invasion of mud filtrate may have affected the mea-
sured P-wave velocity (Figures 14). For most quartz-rich sands,
I find that K0 values can vary from 35–40 GPa and μ0 values
can vary from 35–44 GPa. The reason for this variability is two-
October 2011 The Leading Edge 1135
H o n o r a r y L e c t u r e
Downloaded 03/01/13 to 141.117.79.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 11. Explicit calculation of Kdry and PRdry from Gassmann’s


equation. Note the noisy Kdry and abundance of negative PRdry values
throughout the modeled interval. This is often due to problems with
the compositional model, the porosity, the saturation state, or local
inconsistencies with the log data. Regardless of the cause, these problems
need to be addressed prior to proceeding with fluid substitution
analysis.

Figure 10. AVO modeling results from the well shown in Figure
10. (a) AI versus PR crossplot. Application of the estimated shear
velocity results in a larger contrast in PR than that calculated from
the measured data. (b) Half-space model using data from (a). Note
that the AVO gradient response is enhanced when using the estimated
shear log. (c) Offset synthetic response. Note the following: (1) the AVO
gradient response is enhanced when using the estimated shear-velocity
log, and (2) at the seismic bandwidth, the AVO response is complex,
and not representative of the underlying rock properties. Rock property
plots (a and b) would predict a class I to IIP response, whereas an
apparent class III response is observed on the offset synthetic models (c).

fold: (1) most sands typically contain minerals other than quartz,
and (2) most framework grains contain microstructural defects
(i.e., cracks) that will lower the effective bulk modulus and shear
modulus of the mineral matrix.
A key element of the approach outlined above is the neces- Figure 12. Simple mixing model (Voigt-Reuss-Hill) for dry frame
Poisson’s ratio, assuming that the PRdry = PRmatrix. Note that the smaller
sity for understanding the underlying composition of the rock. the amount of quartz in the system, the higher the dry frame PR. Some
Applying the properties of quartz to a nonquartz reservoir may minerals, such as plagioclase feldspar, have petrophysical properties
result in fluid substitutions (or estimated shear logs) that over- very similar to quartz, and are therefore difficult to recognize from
estimate the predicted AVO effects (Figure 9). conventional logging suites. Note, however, that PRdry for a sandstone
with abundant plagioclase feldspar will be considerably higher than
Porosity modeling that for a pure quartz sandstone. It should not be expected that all
reservoir rocks have PRdry values of 0.12 and lower.
Porosity modeling in sedimentary rocks is usually done via ap-
plication of empirical or heuristic models (Han, 1986; Eberhart-
Phillips, 1989; also see Avseth et al., 2007 and 2010 for more a between porosity and velocity in sedimentary rocks, due in
detail discussions). Although these are useful approaches, they large part to variable pore geometries as well as the presence of
may not accurately describe the underlying velocity-porosity highly variable concentrations of microcracks in the rock ma-
systematics for many reservoirs, especially over large ranges in trix (Smith et al., 2010). In addition, for many reservoir rocks,
porosity. This is because there is often no unique relationship changes in porosity often correspond to changes in clay content

1136 The Leading Edge October 2011


H o n o r a r y L e c t u r e
Downloaded 03/01/13 to 141.117.79.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 13. Frame property model and predicted P-wave velocity using Equation 2 (see text). In this approach Kdry is calculated using Equation
2, which is then used in Gassmann’s equation to compute a P-wave log for the in-situ conditions (red curve in track 4). Note the close match
between the measured and predicted log. Application of Equation 2 is an effective technique for correcting difficult problems during the fluid
substitution workflow.

and/or the amount and types of diagenetic cements and diage- will extrapolate back to some value representative of wet shale or
netic alteration. Consequently, application of empirical or heu- clay at 0% porosity, and not a P-wave velocity of pure quartz.
ristic porosity models may impose on the reservoir a velocity- On the other side of the spectrum, velocity behavior in low-
porosity relationship that does not exist in the subsurface. Prior porosity rocks (porosity < 10–12%) is often dominated by the
to application of any porosity model, the controls on porosity presence of variable concentrations of microcracks in the rock
behavior should be assessed and utilized in the development of matrix (Smith et al., 2010; Ruiz and Cheng, 2010). As a result,
the porosity model. In addition, examination of local well data, velocity-porosity relationships can be poorly defined in these sys-
as well as geological considerations, should also provide insights tems, with the porosity having only a negligible effect on the P-
into the range over which porosity is modeled. wave velocity (Smith et al., 2010a and 2010b). In these systems,
For many poorly consolidated sand and shale sequences, velocities often extrapolate to values lower than those predicted
the P-wave velocity for shale is similar to the P-wave velocity for by the rock matrix at 0% porosity.
brine-saturated sandstones (e.g., Smith and Sondergeld, 2001). Regardless of the reservoir being modeled, it is important
In these scenarios, porosity degradation often occurs simply by to build an understanding of the local controls on velocity, and
the replacement of sandstone layers with shale. As a result, there the causes for porosity degradation or enhancement. This is pos-
are frequently only small changes in P-wave velocity with changes sible only when core data are available and can be analyzed. For
in porosity (Figure 15a). This is analogous to the “sorting” trend the more common scenario when only log data are available, I
of Avseth et al. (2010). When modeling velocity and porosity in find that the most effective means for evaluating the underlying
such systems, it is important to recognize that P-wave velocities controls on velocities is to examine the manner in which Kdry

1138 The Leading Edge October 2011


H o n o r a r y L e c t u r e

and μ vary with changes in porosity (Figure 15b). Implicit in


these empirical observations are the factors governing the rock’s
elastic response to changes in porosity. In the example shown
in Figure 15b, it is clear that both Kdry and μ increase as poros-
ity decreases (the effect on Kdry is clearly larger). Examination
of this z-axis (Vclay) indicates that the observed increase in Kdry
Downloaded 03/01/13 to 141.117.79.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

and μ with decreasing porosity may be due to the increase in


the amount of clay, and/or cementing silt, in the lower-porosity
fraction of the reservoir sand. Also note from Figure 15b that the
Kdry/μ ratio increases as porosity decreases; this is consistent with
higher concentrations of clay/silt in the lower porosity fraction of
the reservoir. Modeling porosity in this reservoir using a constant
value for the Kdry/μ ratio would clearly be inappropriate (e.g., for
a monomineralic system). The changes in the frame properties
observed in Figure 15b can be combined with Gassmann’s equa-
tion to model velocity behavior as a function of changing poros-
ity. This is a bit more time intensive than application of a com-
mercial model-based approach, but it honors the local porosity
systematics governing velocity behavior and leads to models that Figure 14. Application of Equation 2 for correcting P-wave logs for
are consistent with the velocity behavior in the subsurface. invasion effects. In this example, a gas sand appears to be invaded
with drilling mud. Invasion is inferred in this example from the large
Summary comments variations in the PR values computed from the measured P- and
S-wave logs (c.f., Dvorkin et al., 1999). Application of Equation 2
The primary purpose of “seismic petrophysics” is to condition allows invasion corrections to be applied without the calculation of Sxo
and interpret log data in support of a more quantitative interpre- (water saturation in the invaded zone).
tation of seismic data. This aspect of the interpretation workflow
typically resides in the domain of the geophysicist, and is usually
referred to as “rock physics.” It’s been my observation, however, nearly all theoretical rock physics models, as well as by most em-
that most geophysical workflows oversimplify the log editing pirical observation. Unfortunately, many practitioners of rock
and interpretation process, and do not adequately address the physics continue to treat most formations as quartz-rich sand-
issue of reservoir composition. Conversely, most petrophysicists stones. As we have seen (e.g., Figure 10), this can result in un-
focus exclusively on the reservoir zone, and do not do properly reasonable expectations for the AVO response. In the absence of
edit the rest of the well logs for seismic integration. In addition, local knowledge of the underlying composition, it is impossible
many petrophysicsts do not assess whether the petrophysical in- to fully and quantitatively understand the controls on any giv-
terpretation is consistent with the measured elastic properties of en seismic response. Even for those instances when we do have
the reservoir. In many instances, these editing and interpreta- nearby well control, it is my opinion that interpreters often miss
tion shortfalls do not cause problems. However, as we attempt opportunities, or make costly mistakes, simply because so many
to interpret increasingly subtle seismic amplitude responses, it is flawed assumptions are made about the composition and elastic
my opinion that more care can and should taken in the integra- properties of the reservoir.
tion of log data into geophysical workflows; this usually extends Porosity-velocity systematics are complex, and are often not
beyond the typical rock physics workflow. easily described by most (if any) rock physics models (with the
During the editing and interpretation process, it is important obvious exception of those rocks composed of pure mineral end-
that rock physics models not be used as a substitute for rigorous members). This is especially true with unconventional reservoirs
analysis and interpretation of measured data (both log and core (e.g., tight gas sands and shale gas), where the effects of pore
measurements). While rock physics models can greatly facilitate microstructure (e.g., cracks; Smith et al., 2010; Ruiz and Cheng,
the modeling of various parameters (e.g., porosity), model inputs 2010) and organic matter need to be considered in the evalu-
are, in practice, typically treated as fitting parameters. That is, ation of porosity (for an excellent overview of shale petrophys-
adjustments are made to the inputs until a satisfactory match ics, the reader is referred to Sondergeld et al., 2010). Prior to
is found between the modeled velocities and the measured ve- generating a rock physics model that describes porosity-velocity
locities. It is important to recognize, however, that a high-quality relationships, it is therefore important that the requisite time is
match between model results and measured data does not nec- taken to understand the fundamental controls on porosity and
essarily indicate that the model is yielding useful information velocity in the reservoir. Regardless of the reservoir type, this is
about the underlying controls on velocities. best accomplished through a careful analysis of log data and in-
It is my opinion and experience that the two most im- tegration of any available core data. Simple application of a rock
portant, yet overlooked and oversimplified aspects of the rock physics model that predicts the in-situ log response, especially for
physics workflow, are (1) the composition of the rock (both the unconventional reservoirs, may not accurately reflect the under-
reservoir and the bounding layers) and (2) the porosity-veloci- lying rock physics, and may therefore lead to incorrect forward
ty systematics. The importance of composition is predicted by models.

October 2011 The Leading Edge 1139


H o n o r a r y L e c t u r e
Downloaded 03/01/13 to 141.117.79.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Figure 15. Velocity systematics. (a) P-wave velocity versus porosity. Note that there is only a small change in P-wave velocity with changing
porosity. Constant Vclay lines are from Eberhart-Phillips (1989). Examination of these data implies that porosity changes are due to changes in silt
and clay content (e.g., the “sorting trend” of Avseth et al, 2010). Also note that extrapolating to 0% porosity does not result in the matrix velocity
for quartz. Failure to account for the important role of clay (more specifically, composition) in the porosity-velocity systematics would lead to
application of the wrong porosity-velocity model, and potentially misleading AVO models. (b) Dry frame bulk modulus and shear modulus versus
porosity. Note that the Kdry/μ ratio increases with decreasing porosity. This is consistent with increasing clay content as porosity decreases. Direct
observation of variations in the dry frame properties with porosity is an effective means for better understanding velocity-porosity systematics,
and for constructing porosity models that are representative of the reservoir in the subsurface. Perhaps most importantly, the ability to use a rock
physics model to match the measured velocities does not necessarily imply that it is representative of the underlying controls on velocities.

A constant theme that exists throughout this “opinion piece” pretation: Cambridge University Press.
is the need for the geophysical community (and, for that matter, Avseth, P., T. Mukerji, G. Mavko, and J. Dvorkin, 2011, Rock-physics di-
the petrophysical community) to rigorously incorporate knowl- agnostics of depositional texture, diagenetic alterations, and reservoir
edge of the underlying controls on the elastic response of the res- heterogeneity in high-porosity siliciclastic sediments and rocks—A
review of selected models and suggested work flows: Geophysics, 75,
ervoir and its bounding layers. As we search for ever more subtle
no. 5, 7531–7547.
AVO responses, it is important that we take the time to fully un- Berge, P. A., B. P. Bonner, and J. G. Berryman, 1995, Ultrasonic velocity-
derstand the underlying geological and petrophysical properties porosity relationships for sandstone analogs made from fused glass
of the reservoir, and not assume that our sands are dominated beads: Geophysics, 60, no. 1, 108–119, doi:10.1190/1.1443738.
by quartz, or that they conform to some preferred rock physics Castagna, J. P., M. L. Batzle, and T. K. Kan, 1993, Rock physics-the link
model. The application of rock physics models in the absence of between rock properties and AVO response, in J. Castagna, and M.
this important calibration step can be misleading, and can lead M. Backus, eds., Offset-dependent reflectivity—theory and practice
to potentially expensive mistakes. of AVO analysis: SEG., 131–171.
Dvorkin, J., D. Moos, J. Packwood, and A. Nur, 1999, Identifying
References patchy saturation from well logs: Geophysics, 64, no. 1, 1–5, doi:
Avseth, P., T. Mukerji, and G. Mavko, 2005, Quantitative seismic inter- 10.1190/1.1444681.

1140 The Leading Edge October 2011


H o n o r a r y L e c t u r e

Eberhart-Phillips, D., D.-H. Han, and M. D. Zoback, 1989, Empiri- Smith, T. M., C. M. Sayers, and C. H. Sondergeld, 2010, Rock properties
cal relationships among seismic velocity, effective pressure, poros- in low-porosity/low permeability sandstones: The Leading Edge, 28,
ity, and clay content in sandstone: Geophysics, 54, no. 1, 82–89, no. 1, 48–59, doi:10.1190/1.3064146.
doi:10.1190/1.1442580. Smith, T., C. Sondergeld, and T. Ali Ousseini, 2010, Microstructural
Gardner, G. H. F., L. W. Gardner, and A. R. Gregory, 1974, Formation controls on electric and acoustic properties in tight gas sandstones;
velocity and density—the diagnostic basics for stratigraphic traps: some empirical data and observations: The Leading Edge, 29, no. 12,
Downloaded 03/01/13 to 141.117.79.62. Redistribution subject to SEG license or copyright; see Terms of Use at http://library.seg.org/

Geophysics, 39, no. 6, 770–780, doi:10.1190/1.1440465. 1470–1474, doi:10.1190/1.3525362.


Gassmann, F., 1951, Elasticity of porous media: Uber die elastizitat po- Sondergeld, C. H., K. E. Newsham, J. T. Comisky, M. C. Rice, and C. S.
roser medien: Vierteljahrsschrift der Naturforschenden Gesselschaft, Rai, 2010, Petrophysical Considerations in Evaluating and Producing
96, 1–23. Shale Gas Resources: SPE 131768 PP, 34.
Greenberg, M. L. and J. P. Castagna, 1992, Shear-wave velocity estima- Spencer, J. W. Jr., M. E. Cates, and D. D. Thompson, 1994, Frame mod-
tion in porous rocks: Theoretical formulation, preliminary verifica- uli of unconsolidated sands and sandstones: Geophysics, 59, no. 9,
tion and applications: Geophysical Prospecting, 40, no. 2, 195–209, 1352–1361, doi:10.1190/1.1443694.
doi:10.1111/j.1365-2478.1992.tb00371.x. Vernik, L., D. Fisher, and S. Bahret, 2002, Estimation of net-to-gross
Han, D.-H., 1986, Effects of porosity and clay content on acoustic proper- from P and impedance in deepwater turbidites: The Leading Edge,
ties of sandstones and unconsolidated sediments: Ph.D. thesis, Stan- 21, no. 4, 380–387, doi:10.1190/1.1471602.
ford University. Wang, Z., 2000, Velocity-Density relationships in sedimentary rocks, in
Helmold, K., W. J. Campaign, W. R. Morris, D. S. Hasting, and S. R. Seismic and Acoustic Velocities, Z. Wang and D. A. Ebrom, eds., Res-
Moothart, 2006, Reservoir quality and petrophysical model of the ervoir Rocks, vol 3, Recent Developments: SEG Geophysics reprint
Tarn deep-water slope-apron system, North Slope, Alaska: 102nd An- series No. 19.
nual Meeting of the Cordilleran Section, GSA, 81st Annual Meeting Williams, M., V. Leighton, H. Tan, A. Vassiliou, J. Beck, T. Nemeth, C.
of the Pacific Section, AAPG, and the Western Regional Meeting of Savage, P. Gutowski, P. Garossino, and D. Howlett, D., Jackson, R.,
the Alaska Section, SPE. Cox, D., J. Stock, and C. Hoffpauir, 1996, The Hugoton cross-well
Katahara, K. W., 1999, The effect of mineral VP/VS on rock VP/VS: SEG survey; A direct look at stratigraphy, seismic petrophysics and shale
Expanded abstracts. anisotropy: SEG Expanded Abstracts, 86–89.
Katahara, K. W., 2004, Fluid substitution in laminated shaly sands: SEG Wyllie, M. R. J., A. R. Gergory, and L. W. Gardner, 1956, Elastic wave
Expanded Abstracts, 23, 1718–1721, doi:10.1190/1.1851154. velocities in heterogeneous and porous media: Geophysics, 23, no. 1,
Katahara, K. W., and T. M. Smith, 2005, Expert Answers: CSEG Re- 41–70, doi:10.1190/1.1438217.
corder, 8–9. Xu, S. and R. E. White, 1995, A new velocity model for clay-
Kittridge, M. G., T. R. Taylor, N. R. Braunsdorf, L. A. Hathon, and sand mixtures: Geophysical Prospecting, 43, no. 1, 91–118,
L. T. Bryndzia, 2004, Seismic petrophysics for clean sandstones: doi:10.1111/j.1365-2478.1995.tb00126.x.
Integrated interrogation of lab- and well-based data for improved
rock physics modeling: SEG Expanded Abstracts, 1750–1753, Acknowledgments: The number of people who have influenced me
doi:10.1190/1.1845164. during my time engaged in the practice of seismic rock properties are
Krief, M., J. Garat, J. Stellingwerff, and J. Ventre, 1990, A petrophysi- too numerous to mention; this includes personal contacts as well as
cal interpretation using the velocities of P and S waves (fullwaveform
those in the published literature. I would especially like to recognize
sonic): The Log Analyst, 31, 355–369.
Kuster, G. T. and M. N. Toksöz, 1974, Velocity and attenuation of seis-
and thank Carl Sondergeld, Arthur Cheng, Keith Katahara, Manika
mic waves in two-phase media: Part I. Theoretical formulation: Geo- Prasad, and Chandra Rai for their infectious enthusiasm and shared
physics, 39, 589–606. insights on various aspects of rock physics. Much of my approach to rock
Mavko, G., 2005, Expert Answers: CSEG Recorder, 8–9. physics has been influenced by their feedback and criticisms. However,
Mavko, G., T. Mukerji, and J. Dvorkin, 1998, The rock physics hand- the ideas and workflows presented in this paper are mine, and do not
book: Tools for seismic analysis in porous media: Cambridge Univer- necessarily represent their opinions or approaches to similar problems.
sity Press. Carl Sondergeld, Joe Comisky, Dan Ebrom, and Arthur Cheng
Nur, A., 1992, Critical porosity and the seismic velocities in rocks: EOS: provided valuable feedback on earlier versions of this manuscript. The
Transactions - American Geophysical Union, 73, 43–66. manuscript is much improved due to their constructive criticisms. I
Pennington, W., 1997, Seismic petrophysics: An applied science for would especially like to thank the SEG and Apache Corporation for
reservoir geophysics: The Leading Edge, 16, no. 3, 241–246,
the opportunity and the time to participate in the Honorary Lecturer
doi:10.1190/1.1437608.
Raymer, L. L., E. R. Hunt, and J. S. Gardner, 1980, An improved sonic
program. It is a privilege and an honor to be associated with these two
transit time-to-porosity transform: Transcripts of Society of Profes- great organizations.
sional Well Log Analysts 21st Annual Logging Symposium.
Ruiz, F. and A. Cheng, 2010, A rock physics model for tight gas sand: The Corresponding author: Tad.Smith@apachecorp.com
Leading Edge, 29, no. 12, 1484–1489, doi:10.1190/1.3525364.
Skelt, C., 2004, Fluid substitution in laminated sands: The Leading Edge, Editor’s note: This article is based on SEG’s 2011 North America
23, no. 5, 485–493, doi:10.1190/1.1756839. Honorary Lecture.
Smith, T. M. and C. H. Sondergeld, 2001, Examination of AVO respons-
es in the eastern deepwater Gulf of Mexico: Geophysics, 66, no. 6,
1864–1876, doi:10.1190/1.1487130.
Smith, T. M., C. H. Sondergeld, and C. S. Rai, 2003, Gassmann
fluid substitutions: A tutorial: Geophysics, 68, no. 9, 430–440,
doi:10.1190/1.1567211.

October 2011 The Leading Edge 1141

You might also like