You are on page 1of 18

SPE 118703

Stress and Rock Property Profiling for Unconventional Reservoir Stimulation


R.D. Barree and J.V. Gilbert, Barree & Associates LLC, and M.W. Conway, Stim-Lab, Inc.

Copyright 2009, Society of Petroleum Engineers

This paper was prepared for presentation at the 2009 SPE Hydraulic Fracturing Technology Conference held in The Woodlands, Texas, USA, 19–21 January 2009.

This paper was selected for presentation by an SPE program committee following review of information contained in an abstract submitted by the author(s). Contents of the paper have not been
reviewed by the Society of Petroleum Engineers and are subject to correction by the author(s). The material does not necessarily reflect any position of the Society of Petroleum Engineers, its
officers, or members. Electronic reproduction, distribution, or storage of any part of this paper without the written consent of the Society of Petroleum Engineers is prohibited. Permission to
reproduce in print is restricted to an abstract of not more than 300 words; illustrations may not be copied. The abstract must contain conspicuous acknowledgment of SPE copyright.

Abstract
Sonic log data and core measurements are often used to develop models of in-situ stress profiles and rock elastic properties
for use in hydraulic fracture design treatments in unconventional reservoirs. “Unconventional” reservoirs, for the purposes of
this paper, include shale-gas, coalbed methane, and tight-gas projects. In unconventional reservoirs, many of the assumptions
underlying rock property estimation and stress profiling in conventional reservoirs may not apply. In fact, the result of using
conventional approaches may be an incorrect and often misleading stress profile. Fracture geometry predicted using
conventionally derived rock properties and stresses might also be inaccurate.
Methods of deriving rock properties from log and core measurements and the effect of various parameters on resulting
moduli and stress estimates are examined. The paper also discusses the effects of rock anisotropy and inhomogeneity on
static and dynamic properties. The impact of organic materials and trapped gas on sonic logs and conventional mechanical
properties interpretation and the use of synthetic sonic logs are also presented. For static and dynamic measurements on core
samples, the effects of the condition of recovered core and applied laboratory procedures on measurement results are also
considered.
Potential errors resulting from the use of inappropriate mechanical properties for stress profiling and fracture geometry
prediction can be significant. The paper identifies common pitfalls in core and log interpretation. A recommended procedure
to determine useful and accurate rock mechanical properties for stress profile prediction and fracture design is presented.

Introduction
Full-waveform sonic logs and core measurements are commonly used to derive “calibrated” models of in-situ stress profiles.
The results are used as input to hydraulic fracture design simulators, which are used to determine optimum perforation
placement, job size, pump rate, fluid and proppant requirements, and other design variables for optimum reserve recovery.
The methods used to determine rock mechanical properties and stresses assume that the raw input measurements are valid
and that the conditions of measurement are appropriate to conditions during hydraulic fracturing. In unconventional
reservoirs, these assumptions may not be valid.
Because of the potential for large reserves, there is an increasing interest in unconventional reservoirs. For the purpose of
this discussion, unconventional reservoirs are considered to be tight and ultra-tight gas sands (less than 0.01 md effective
permeability), gas-shales, and coalbed methane (CBM) reservoirs. These reservoirs have several characteristics in common:
They all have low, very low, or nearly immeasurable “matrix” permeability. They may be self-sourcing and can contain
organic carbon within the hydrocarbon maturation window, and may be actively generating hydrocarbons at the time of
discovery and development. Many have abnormal pore pressures, relative to a hydrostatic gradient. Many occur in regions
with significant tectonic stress or strain overprints, hence anisotropic stress fields. Production from these reservoirs is
commonly enhanced through the presence of some kind of fracture or micro-fracture network.
These complexities affect the behavior of core and log measurements. Interpretation of core and log data may require
different paradigms and assumptions than those commonly applied in more conventional reservoir systems. Using
conventional assumptions when dealing with measurements in unconventional reservoirs can lead to significant errors in the
derived rock mechanical properties and estimated stress profile. These errors can lead to incorrect predictions of fracture
containment and overall geometry, conductivity, and post-frac performance.

Estimation of In-Situ Minimum Horizontal Stress


The hydraulic fracture closure pressure is assumed to represent the minimum of the three principal stresses acting on the rock
affected by the fracture. In conventional reservoirs at moderate depths, the minimum stress is typically assumed to be
2 SPE 118703

horizontal and to follow some form of the uniaxial strain model. Equation 1 gives a slightly expanded form of the uniaxial
strain estimate of minimum horizontal stress.
ν
Pc = σ h = [σ ]
− α v Pp + α h Pp + ε h E + σ t (1)
(1 − ν )
v

The observed fracture closure pressure (Pc) is assumed to equal the minimum horizontal stress (σh) in this model. The
magnitude of σh is assumed to be controlled by the vertical uniaxial strain with externally applied horizontal tectonic stress
and strain offsets. As discussed by Thiercelin (1994), the model is very simplistic considering the complex deposition,
diagenetic, and deformational history of most reservoir systems. Warpinski (1998) suggests that the two poroelastic constants
(αv and αh) are equal for isotropic materials. This is incorrect and misinterprets the physical meaning of the pore pressure
terms in Equation 1. The first term in brackets, involving αv, is the vertical net effective stress causing compaction of the rock
framework. In this term, the internal pore pressure acts against the externally applied overburden stress to reduce net
intergranular stress. The pore pressure term must be corrected for cementation, consolidation, and other poroelasticity effects.
The second pore pressure term, involving αh, does not involved net intergranular stress but refers to internal fluid pressure
only. The pore pressure acts equally in all directions and is in direct hydraulic communication with the fracturing fluid. In
this case, no poroelastic effect should be applied, and the αh term should be set to unity or removed from the equation.
The observed minimum stress commonly differs from that calculated from the uniaxial strain assumption. The differences
may be caused by errors in estimation of elastic properties (α, ν and E), pore pressure, and overburden stress, or through
externally applied horizontal stresses and strains. In Equation 1, only stresses and strains normal to the plane of minimum
stress are considered, although transverse stresses may be generated through other applied strains. Since stresses and strains
other than the minimum cannot be resolved, they will not be considered in detail here. The maximum stress induced by an
applied lateral strain (ε) is limited by the shear failure envelope of the material, as discussed by Thiercelin (1994). Accurate
determination of the shear limit requires knowledge of the complete in-situ net stress tensor and rock strength (in shear) along
the potential plane of weakness. Typically, none of these can be determined from field data. Instead, a practical maximum
strain limit can be set for use in calibrating stress profiles. The practical limit can be derived from an assumed material
cohesion and friction angle along with an estimate of vertical net effective stress. Using the applied strain boundary condition
has been shown to give more accurate derived stress profiles than a constant stress offset (Blanton, 1999). The stress offset
may be useful to describe residual background stress after inset of shear failure in an active fault environment.
Application of Equation 1 requires representative values of rock mechanical properties. Values of Young’s Modulus (E)
and Poisson’s Ratio (ν) can be determined from core samples (both static and dynamic acoustic tests) and from well log
(dynamic acoustic) measurements. The accuracy and applicability of these derived properties has a direct impact on the
calculated stress profile, even when it is “calibrated” to a single point of measured closure stress. When rock properties and
environmental or test conditions affect the apparent mechanical properties, errors in stress and fracture geometry can result.
Unconventional reservoirs offer many opportunities for these errors to occur. In addition, Equation 1 implies that the stress
profile affecting fracture propagation and geometry is the result of an equilibrium deformation state for an elastic medium.
Unconventional reservoirs, especially coal and shale, may behave as ductile or plastic materials rather than elastic ones.
During hydraulic fracturing, the rock must respond to rapidly induced deformation. In very low permeability systems, a rapid
strain rate can outrun the ability to dissipate internal pore pressure. This can result in deformation behavior controlled by
“undrained” moduli instead of “drained” moduli. Many more complex factors must be considered in deriving a stress profile
and fracture geometry in these complex reservoirs.

Use of Sonic Transit Time for Mechanical Properties


Full-waveform sonic logs, yielding both shear and compressional wave travel time (DTS and DTC respectively) are
commonly used to derive dynamic elastic properties. The assumption is that the acoustic velocities are related to the rock
elastic properties. Conventional equations for Poisson’s Ratio (ν) and Young’s Modulus (E) are given as Equations 2 and 3:

ν=
(R − 2 ) (2)
(2 R − 2)

E = 13447 ρ b
(3R − 4)
(DTC 2 R(R − 1)) (3)

In these equations, the observed formation bulk density (in g/cm3) is given by ρb, and R is the square of the travel-time
ratio:
DTS 2
R= (4)
DTC 2
SPE 118703 3

The travel times (DTC and DTS) are the reciprocals of the compressional and shear acoustic wave velocities (Vp and Vs)
respectively. Equations 2-4 are equivalent to the dynamic moduli definitions of Lama and Vutukuri as cited by Warpinski
(1998). The mechanical properties derived from sonic measurements are assumed to correlate to “static” measurements made
on core samples. Unfortunately, things other than variations in elastic rock properties significantly affect the sonic velocities.
These other factors include fractures and laminations, external stress, temperature, borehole conditions (breakouts, mud
weight, borehole size, and tool eccentricity), pore pressure, and pore fluid saturation. In many cases, there is no correction or
adjustment made to the observed sonic transit times for any of these effects.

Error Propagation in Dynamic Moduli Calculation


Compressional and shear sonic velocities are intrinsically difficult to measure due to the nature of the wave train and the
borehole environment. Various methods are used to reduce the error in detecting and measuring the compressional and shear
arrivals, such as Slowness Time Coherence (STC) plots. The inherent noise in the system, however, makes eliminating all
error impossible. The equations above are used to calculate the dynamic Poisson’s Ratio and Young’s Modulus from the
shear and compressional arrivals. The reliance of these forms on R (equation 4), the ratio of the shear slowness squared over
the compressional slowness squared, leads to large errors being propagated through to the mechanical properties. An analysis
has been performed considering a +/- 5% error in both the shear and the compressional velocity measurements, shown by the
first two bars in Figure 1, to show how that error propagates through into the dynamic Young’s Modulus and Poisson’s Ratio
calculations. Assuming that the density measurement is error free, the Poisson’s Ratio has an inherent error of +/- 20% and
the Young’s Modulus has a final error of +/- 26%. When calculating stress, the ratio of Poisson’s Ratio to one minus
Poisson’s Ratio is used, increasing the error to +/- 28%. Figure 1 summarizes the error analysis.

30.0%

25.0% Properties from DTS/DTC

Properties from DTCO and Lithology


Error Analysis +/- %

20.0%

15.0%

10.0%

5.0%

0.0%
DTSM DTCO R PR YME PR/(1-PR)
Figure 1: Propagation of errors with a 5% error in sonic velocities

Errors of this magnitude obviously have a large effect on the final stress profile calculated using these mechanical
properties. Calculating the same mechanical properties using a correlation based on compressional velocity alone (no shear
input) leads to a reduction in the potential error. Dynamic Young’s Modulus can easily be obtained using compressional
velocity only. Figure 2 shows a regression through 327 tight-gas and shale-gas core samples that gives an excellent estimate
of the dynamic modulus measured under controlled laboratory conditions. Using a lithology volume fraction weighted
average value of ν (based on the curves presented in Mullen, 2008) can give a good estimate of Poisson's Ratio. For a
4 SPE 118703

sand/shale lithology with an assumed +/- 5% error in the compressional velocity, the Poisson’s Ratio error is 9%, with a total
error of 11% for the Young’s Modulus (if the correct shale fraction is known). The error in the ν/(1-ν) term is 14%, or half
the error resulting from using both the shear and compressional velocities in equation 4.

18

16
-2.1557
y = 54857x
14 2
R = 0.9665

12

10

0
40 60 80 100 120 140 160
Vp Transit Time, micro-sec/ft
Figure 2: Estimation of dynamic Young’s Modulus from Vp alone

Generating a stress profile based on the shear/compressional ratio results in the magnification of any inherent errors in the
measurements of both shear and compressional wave slowness. Both the shear and compressional slowness values from
acoustic borehole logs can be subject to significant error of measurement and interpretation. Using a diagnostic injection test
to measure minimum in-situ stress can help to ensure the validity of the calculated stress profile. Using a correlation not
based on shear velocity, and correcting compressional velocity data for saturation and environmental effects also results in
less error prone mechanical properties.

Effect of Gas Saturation on Sonic Logs and Stress Estimates


The impact of hydrocarbon saturation on sonic velocity has been studied extensively (Merkel, 2001 and Holmes, 2004).
Holmes (2004) shows that the presence of 20% gas saturation can change the observed compressional travel-time for a
sandstone reservoir by 25% compared to a water saturated sample. Using measured water velocity data from Merkel (2001)
and calculated gas acoustic velocity (hyperphysics.phy-astr.gsu.edu/hbase/sound/souspe3.html), the volume average DTC
can be estimated using the Wyllie Time Series Acoustic Equation (Homes, 2004) given here as Equation 5:

DTC = φDTFL + (1 − φ )DTMA (5)

In Equation 5, DTFL is the volume-average fluid acoustic velocity for a gas-water mixture and DTMA is the matrix
transit time. Figure 3 shows the computed change in apparent Poisson’s Ratio (ν) using an assumed DTC at 100% Sw of 64.5
μs/ft (DTMA=53 μs/ft and DTFL=197 μs/ft) and a constant DTS of 126 μs/ft. The assumption of constant DTS is based on
the (probably incorrect) belief that the acoustic shear wave is transmitted only through the solid matrix and is unaffected by
pore fluid. A change in gas saturation from 0% to 60% causes ν to drop from 0.32 to 0.03 for these assumptions. This
magnitude of change is greater than the typical contrast between clean sand and plastic shale, and it drastically affects the
estimated stress profile. Laboratory measurements of sonic velocity in water at various pressures, and in oil at and below the
saturation pressure (Merkel, 2001), show that velocity changes of this magnitude can occur in gas-water systems and in
water-oil systems. In oil reservoirs, the acoustic velocity changes dramatically at the saturation pressure when free gas is
generated. Similar effects can be expected in organic rich shale in the gas generation (maturation) window.
SPE 118703 5

0.35

0.3

Apparent Poisson's Ratio
0.25

0.2

0.15

0.1

0.05

0
0 0.1 0.2 0.3 0.4 0.5 0.6 0.7
Gas Saturation 
Figure 3: Effect of gas saturation on Poisson’s Ratio for variable DTC with constant DTS

Use of Spectral GR Logs to determine TOC and Vshale


In conventional reservoirs, the producing interval is generally a “clean” sand or carbonate surrounded by “shale”. The shale
bounding beds are considered to be non-reservoir rocks and are commonly expected to provide some degree of fracture
height containment. The volume-fraction “shale” is often computed from a normalized total gamma ray (GR) log using one
of several published relations (Crain, 2008), such as Steiber’s correlation (Steiber, 1975), given as Equation 6:
0.5 I sh
Vsh = (6)
(1.5 − I sh )
In Equation 6, the GR Index (Ish) is the normalized value of the selected GR curve between the value of GR at the
apparent “clean sand” line (GRsand) and at the “100% Shale” line (GRshale), as shown by Equation 7:
(GR − GRsand )
I sh = (7)
(GRshale − GRsand )
The total GR observed response is generated by many species of radioactive material in the rock. A spectral GR tool
differentiates three principal components: thorium, potassium, and uranium. Of these components, thorium is generally
associated primarily with clay minerals. Potassium may be associated with clays, potassium-feldspars, salts, and other
materials that may be rock constituents. Uranium is most often associated with organic matter and precipitated salts and is
generally not related to clay content. Uranium response can often be correlated to total organic carbon (Crain, 2008, La
Gesse, 2008), while clay content can be more accurately related to thorium, and possibly potassium, content.
Unconventional reservoirs may contain relatively large amounts of total organic carbon (TOC), and may be hydrocarbon
source-rocks (especially shales). It is common to find a high total GR response in organic-rich shale while actual clay content
may be only 3 to 20%. Basing the shale volume-fraction (Vshale) estimate on normalized total GR can imply that
hydrocarbon rich, high TOC zones are non-reservoir when in fact they are the primary completion target. Incorrect Vshale
estimates also affect the implied plasticity of the rock and the estimated values of mechanical properties if lithology-based
correlations are applied. Besides the impact of organic material on shale or clay content estimates, the presence of organic
material can also have a significant impact on acoustic velocity measurements, calculated dynamic rock mechanical
properties, and estimated in-situ stress. All these effects are tied to the generation of hydrocarbons (gas and or oil) from solid
kerogen in shale source-rocks.
As kerogen matures and generates free hydrocarbons, the volume of the organic material increases dramatically. Because
of the small pore size, high capillary threshold pressure, and low permeability of shale, the generated hydrocarbons may
remain trapped in the pore system until it reaches a pressure sufficient to generate hydraulic (micro) fractures. For a
hydrocarbon sourcing shale in the gas generation window, this can mean that the internal pore pressure can actually reach the
6 SPE 118703

fracture pressure of bounding rocks or of the shale itself (Rondeel, 2001). Some (unreleased) field data confirms pore
pressures in shale nearly equal to the fracture closure stress in surrounding layers. Net effective stress within the shale can be
very low, even at great depths, making the formation susceptible to shear failure at low differential stress. The presence of
gas in the pore space, along with the elevated internal pressure, results in dramatic slowing of the compressional acoustic
signal and abnormally high observed values of DTC.
As an example, Figure 4 shows the total high-resolution GR (HSGR) curve and a total GR composed of the sum of the
scaled spectral GR input curves (GR_SPEC). The individual spectral GR components are plotted in the far-right track of the
figure, after scaling to API units. Scaling is accomplished by multiplying the thorium signal (ppm) by 4.0, uranium (ppm) by
8.0, and potassium volume fraction (v/v) by 1500. These scaling factors have been found to give comparable total GR curves
in most cases, although other scaling factors have been published (http://server4.oersted.dtu.dk/research/RI/SNG/SNG-
logs.html, 2008). The GR_TH+K curve in Figure 4 is the scaled total GR, without the uranium component. This curve,
sometimes referred to as the Computed GR (CGR) gives a better appreciation of actual mineralogy and clay content for the
organic rich shale.
The log section at far left in Figure 4 shows the measured and derived (synthetic) DTC curves for the well. The DTCO
curve is the measured compressional travel time. It is significantly affected by the organic material and generated gas in the
shale. The result is an abnormally slow DTC compared to the synthetic DTC curves that match the measured DTC above and
below the gas/TOC zone. Using the anomalous measured DTC curve will generate incorrect estimates of dynamic elastic
properties for the rock. Use of a synthetic DTC curve, or DTC corrected for the presence of gas saturation and TOC, may
give a better estimate of rock elastic properties. Improved estimates of rock elastic moduli generate more accurate computed
stress profiles when elastic deformation is expected. Only careful calibration of the computed stress profile to a directly
measured stress allows the best of the many possible DTC interpretations to be selected. Many field cases have been studied
in which the measured DTC clearly leads to incorrect stress profiles while synthetic DTC curves give useful results.

Sonic Slowing due Separation due


to Gas and TOC to Uranium in
Effects Organics

DTC from
Resistivity

DTC from
PHIN and GR

DTC Measured

Measured total GR
and GR_SPEC

Figure 4: Effect of organic material on GR and DTC

Derivation of Mechanical Properties from Non-Sonic Logs


The use of derived synthetic compressional sonic logs has been presented by Mullen, et al (2007). Methods are presented to
derive DTC curves from simple scaling models applied to computed GR (without organic effects), neutron and crossplot
porosity, and deep electrical resistivity (beyond the invaded zone). The methods and models used are an outgrowth of the
SPE 118703 7

Passey (1990) technique, which relates TOC percentage to the sonic and resistivity response. In estimating rock properties, it
is advantageous to use the same response to eliminate the gas and TOC effect.
The synthetic DTC curves in Figure 4 were derived using models similar to those presented by Mullen (2007) and
calibrated to the local log conditions in zones expected to be at 100% Sw. Local calibration or scaling of the derived DTC
curves is recommended to detect small variances in the measured DTC that can show trapped gas in the region investigated
by the sonic tool. In organic-rich shale formations, the effect on the measured sonic is typically large, as in the case shown.
The relative magnitudes of the derived and measured DTC curves are systematic and diagnostic. In formations with high
gas or TOC content, the measured DTC will be slowest. The DTC computed from average total porosity (neutron-density
crossplot porosity) will generally be very close to the measured DTC curve as both neutron and density response will be
affected by gas and TOC. The DTC curves derived only from neutron porosity will give the next lower values of DTC, and
DTC from deep resistivity will give the fastest estimate of transit time slowness. In developing synthetic rock mechanical
properties and stress profiles, use of the DTC from resistivity often gives the best correlation to observed closure stress and
fracture height development.
Elastic moduli (ν and E) can be derived from the synthetic DTC curves using lithology volume fractions, as demonstrated
by Mullen (2007). These moduli can also be derived directly from conventional open-hole log responses using simple
models. For example, useful estimates of Poisson’s Ratio can be derived using power-law functions of resistivity and
computed GR, as shown by Equations 8 and 9:

PR_GR=CPG*GREPG (8)

PR_RESIST=CPR*RESISTEPR (9)

The coefficients in Equations 8 and 9 are obtained by correlation of the derived synthetic curves to values of ν computed
from DTC and lithology, or travel-time ratio (R), in areas of good borehole condition and 100% water saturation (usually
non-reservoir intervals). Similarly, good estimates of E can be derived directly from open-hole log responses using simple
linear scaling, power-law, and exponential equations, as shown in Equations 10-12:

YME_GR=CEG*GR+OEG (10)

YME_RESIST=CER*RESISTEER (11)

YME_PHIA =62*10(-0.0145*DTC_PHIA) (12)

Other useful estimates of mechanical properties can be derived from various combinations of log measurements,
including cased-hole pulsed neutron logs. The key is that dynamic properties derived from measured full-waveform acoustic
logs can be used to develop correlations when the data from the sonic logs are valid. This is only the case in 100% water-
saturated rocks with good borehole conditions and intact rock with minimal open fractures. Even then, the pore pressure must
be nearly constant across the zone used for calibration. As shown by Merkel, et al (2001) both shear and compressional
velocity vary with net stress to the 1/3 power. If pore pressure changes locally, especially in a hydrocarbon sourcing shale,
then all observed velocities will be poor representations of rock elastic properties. Elastic properties computed from
uncorrected (measured) sonic-derived properties will be wrong, and the estimated stress profile will be incorrect due to both
erroneous elastic moduli and incorrect pore pressure assumptions.

Use of Unconventional Core Mechanical Properties Data


Many people believe that static core measurements of elastic moduli represent a “ground-truth” for calibration of dynamic
core properties or log-derived dynamic moduli. Unfortunately, in unconventional reservoirs, there are many processes
associated with core recovery, transport, and processing that can make the core that is studied in the laboratory less than ideal
as a representation of in-situ rock properties.
In the reservoir, the core is at stable internal pore pressure and temperature, and supported by anisotropic confining
stresses. During drilling and coring, the overburden stress is released and replaced by the circulating mud weight. Lateral
stresses acting on the rock below the bit face can generate sufficient stress concentrations to cause core disking and other
shear and tensile fracture modes (Li, 1997, Lim, 2006). As the core bit passes the sample, lateral stresses are released
allowing lateral expansion and possible generation or enhancement of fractures to occur. During core recovery, the internal
pore pressure is released. For extremely low permeability samples, the release of pore pressure may be accompanied by
expulsion fracturing or enhancement of existing fractures. Finally, at the surface the sample may experience thermal
contraction, desiccation, and possible oxidation of clay minerals. All these factors can cause mechanical failure, fracture, and
changes in elastic properties. The characteristics of unconventional reservoirs will exacerbate these effects.
The impact of micro- (and macro-) facture generation during coring has long been observed. Both anelastic strain
recovery (ASR) and differential strain curve (DSC) analysis have been used to determine the orientation and relative
magnitude of in-situ stresses. These methods involve the measurement of anisotropic strains resulting from core relaxation
8 SPE 118703

after recovery. In most cases, the observed strains are believed to be caused by opening of fractures normal to the maximum
in-situ stress (Warpinski, 1989, Yassir, 1998). These observations make it a near certainty that a core sample from an
unconventional reservoir will not represent in-situ stress or mechanical properties without some attempt to restore the
sample’s integrity.

Effects of stress and strain relaxation and cycling on static and dynamic moduli
When core samples are subjected to static or dynamic testing in a laboratory, the confining stress, net effective stress,
stress history, pore pressure, temperature, and saturation state may all affect the results. Traditionally, core samples are tested
in the “as received” condition and the primary compaction cycle is used to derive static moduli. If the core has developed
open microfractures during coring and handling, and is partially gas saturated, these measurements may not reflect the
expected in-situ conditions. In general, the core will deform more easily because of the increased compliance of the fracture
network, giving a Young’s Modulus that is too low to represent deformation during fracturing. Applied axial loads will close
fractures and generate less lateral strain than would be the case for an intact sample, leading to abnormally low Poisson’s
Ratio. The absolute magnitude and differential stress state established at the start of the test also has a significant effect on the
observed material properties. Often the confining stress is poorly defined.
DIfferential Stress

E4
E5

E2
E3

E1
Axial Strain
Figure 5: Effect of stress cycling on apparent Young’s Modulus

One way to partially offset the effects of core relaxation and microfracture evolution is to stress-cycle the sample before
taking measurements of stress and strain to describe moduli for stress profiling or fracture geometry estimates. Figure 5
shows a hypothetical stress-strain profile for a core sample under cyclic loading. Five different Young’s Modulus values are
illustrated corresponding to the initial compaction modulus (E1), low net stress tangent modulus (E2), unloading tangent
modulus (E3), high net stress tangent modulus (E4), and high net stress secant modulus (E5). As described by Briaud (2008),
many other modulus values may be defined, and the proper modulus for use in a particular case must be selected to represent
local deformation and stress conditions. In an unconventional reservoir, core E1 is likely to represent a partially failed state
with open fractures, and is not representative of conditions during fracturing. Assuming the rock has not been subjected to
multiple stress cycles over a short time, the unloading modulus (E3) is not representative. The secant modulus (E5) may be
useful to estimate long-term residual deformation after multiple stress cycles, but is not useful for fracture geometry or stress
prediction. This leaves the high or low stress tangent moduli E2 and E4.
During hydraulic fracturing, the rock is subjected to fairly rapid deformation, starting at a stable in-situ stress state. If the
in-situ stress and strain conditions could be maintained throughout the coring and handling process, then measuring the
tangent modulus at the average net effective stress should be the correct conditions (Warpinski, 1998). Because the core must
be re-compacted, a better estimate can more often be obtained using a higher net effective stress than that existing in the
SPE 118703 9

reservoir. If a prolonged linear stress-strain regime exists through several stress cycles, the selection of proper stress state is
simplified and the overall linear trend can be used. In general, the most useful modulus is that which describes the process
being modeled.

Effects of saturation state on static and dynamic moduli


Core acoustic measurements are frequently used to verify and calibrate properties derived from logs. Static mechanical
properties measurements are also used to derive conversions from dynamic to static moduli. The effects of saturation and
stress on acoustic velocity measurements that are observed in well logs also exist in laboratory core measurements, along
with stress cycling effects. Many times these effects are ignored or neglected. In very low permeability systems, it is common
practice to perform static deformation studies in an “as received” or unknown saturation state. Even with small gas
saturations in core samples, the acoustic velocity of the core-fluid system can change significantly. The saturation may
change dramatically with pore pressure and sometimes with external stress, introducing further uncertainty in the
measurements. When core and log sonic velocity measurements agree, the results may only confirm that they are measured
under similar saturation conditions, not that both are correct in their representation of rock elastic properties.
The data shown in Figure 6 are shear and compressional velocity measurements made on cores from a tight-gas reservoir
at a depth of 10,000-12,000 feet. The samples were tested under biaxial load with varying radial confining stress and axial
load. Average net stress varied from 1000 to over 4000 psi for the two samples shown. Acoustic velocity measurements were
made with varying net stress and under two saturation conditions: Dry and 100% brine saturated.

120 0.5

DTS 0.45
110
0.4
DTC A Dry

Apparent Poisson's Ratio


DTS A
DTC B 0.35
100 DTS B
DTC, DTS (μs/ft)

PR A 0.3
PR B

90
ν 0.25

0.2
80
0.15
Dry
0.1
70
DTC
0.05

60 0
0 500 1000 1500 2000 2500 3000 3500 4000 4500
Average Net Stress, psi
Figure 6: Effect of saturation and stress on sonic velocity and apparent Poisson’s Ratio

The initial low-stress velocity measurements on the samples (A and B) were conducted on dry cores. The measurements
were repeated at 100% brine saturation at the same stress. The measured shear slowness, DTS, increases from approximately
112 to 117 μs/ft because of the saturation change alone. Slowing of the shear wave in a liquid saturated medium is not
generally expected, but it is consistently observed in core tests. At the same conditions, the observed DTC values drop from
73 to 66 μs/ft and 70 to 67 μs/ft for the two cores. The combined effect of the saturation change is to increase the apparent
dynamic Poisson’s Ratio (PR in legend) from 0.13 to 0.27 for Sample A and 0.18 to 0.25 for Sample B. The magnitude of the
shift in ν due to saturation change alone is more than the typical difference between clean sand and shale. The change in
acoustic velocity and apparent ν with stress change of 3000 psi is less than the saturation effect.
In addition to saturation, the frequency of the acoustic wave also affects the apparent velocity and calculated dynamic
mechanical properties. The data in Figure 7 show measurements of wet (100% brine saturation) and dry cores over a range of
nearly six orders of magnitude of frequency. The effects of water saturation are consistent with the data in Figure 6: DTS
increases with increasing liquid saturation while DTC decreases slightly with increasing liquid saturation. The apparent
10 SPE 118703

dynamic value of ν increases by 50% on average from the dry to saturated state. As frequency increases, the apparent value
of ν also increases by roughly 40% over the range tested. The frequency range is representative of the difference between
typical laboratory conditions (1000-1,000,000 Hz) and well logging conditions (1-1000 Hz). Clearly, the effects of saturation
are at least as significant as the effects of frequency. Both affect the relationship between laboratory and field (well log)
measurements.

140.0 0.3

120.0
0.25

100.0
0.2

Poisson's Ratio
DT, microsec/ft

80.0
0.15
60.0

DTC_(dry) DTS_(dry) 0.1


40.0
DTC_(wet) DTS_(wet)
0.05
20.0
PR_(dry) PR_(wet)

0.0 0
1 10 100 1000 10000 100000 1000000
Frequency, Hertz
Figure 7: Effect of frequency and saturation on acoustic velocity and apparent Poisson’s Ratio

Drained versus undrained moduli


Static core deformation tests can be conducted in two ways when sample permeability is moderate. Variable external load
can be applied with constant internal pore pressure (drained tests) or external load can be applied with the pore fluid trapped
(undrained tests). In the undrained test, the pore pressure increases as the sample is compacted, with more of the externally
applied load supported by the pore fluid instead of the rock framework. This results in a different set of apparent static
moduli that depend on the compressibility of the pore fluid as much as on the rock properties.
In shale, which may be liquid filled and can have apparent matrix permeability of 10-6 to 10-15 darcy (microdarcy to
femtodarcy), it may be impossible for the pore pressure to dissipate in response to applied external loads during the time-
scale of a hydraulic fracture treatment. If this occurs, then the apparent stress in the shale may be better characterized by an
undrained test. Briaud (2008) indicates that Poisson’s Ratio in a clay (shale) sample can approach 0.5 for an undrained test as
the pore fluid pressure drives lateral deformation as if the sample were nearly incompressible and effectively fluidized. In a
drained test, the same sample could exhibit a Poisson’s Ratio closer to 0.35. It is interesting that these are the same two
values frequently observed when deriving calibrated stress profiles in coal seams.
Well cleated water saturated coals appear to have Poisson’s Ratio values approaching 0.5. This may indicate that the coal
is sealed by bounding shale beds and that pressure in the cleats is trapped. Shear in the complex cleat system allows the coal
to deform as a plastic material and applied load is transmitted to the fluid in the cleats. Other coals, with free gas or possibly
drained or depleted pressure, in many cases have apparent Poisson’s Ratio values closer to 0.35.. Whether a coal or shale
behaves as a drained or undrained material depends on local environmental factors of pore fluid saturation, pore pressure, and
the ability to establish pressure communication (possibly through induced shear fractures) to bounding beds.

Comparison of Static and Dynamic Core Mechanical Properties


To illustrate some of the difficulties associated with calibration of dynamic derived properties to static core
measurements, laboratory results from two large data sets are presented. One data set is dominated by low porosity, low
permeability silt and sandstones (tight-gas reservoirs), and the second is dominated by a wide variety of shales, and shaley
siltstones and limestones (shale-gas reservoirs). ASTM procedures were applied in conducting the laboratory measurements.
ASTM standards, however, offer significant latitude in the manner whereby mechanical properties are derived in the
laboratory. The procedures reported here are believed to represent a common practice in the industry at this time.
SPE 118703 11

Laboratory Procedures
Vertical core samples were obtained from whole core samples, faced to a right cylinder, and tested at the “as received”
saturation. Each sample was inserted into a rubber jacket and a radial Linear Variable Displacement Transducer (LVDT) was
placed around the lateral surface of the sample. The sample was mounted between pistons with ports on the contacting
surfaces for controlling pore pressure, and acoustic crystals for acquiring the sonic travel time. All samples reported here
were tested at 0 psi imposed pore pressure and room temperature. The entire assembly was mounted in a pressure vessel that
allows application of confining pressure and axial stress. The top piston extends through the top of the pressure vessel
enabling the application of axial load. The pressure vessel was then loaded into a computer-controlled load frame where
another LVDT was attached for axial strain measurements. The net reservoir stress was calculated for each sample using the
initial reservoir pressure, overburden stress, and the estimated mechanical properties. The confining and axial pressures were
increased at the same rate to the computed net confining pressure. The low axial stress sonic measurement was obtained at
that point. Data logging was begun and the axial and radial strain were measured as the axial load was increased while
holding the confining pressure constant. At near maximum axial stress, the high axial stress sonic velocity was again
determined.

Static and Dynamic Poisson’s Ratio


Figure 8 shows the variation in Poisson’s Ratio derived using the high axial stress acoustic measurements compared to
those derived at low stress. The dashed line shows the 1:1 relationship. The data show that there is no first order correction
for stress state in the dynamic to static conversion. Regardless of stress conditions, the observed dynamic Poisson’s Ratio is a
close approximation of the static Poisson’s Ratio measured under similar saturation conditions. For the two data sets shown,
the static Poisson’s Ratio is 96% of the low-stress dynamic value and 93% of the high-stress dynamic value. The similarity
between statc and dynamic values of ν has been noted by many other authors such as Morales and Marcinew (1993). The
range of uncertainty in the predicted static value of ν, at constant dynamic ν, is significantly larger than the error introduced
by the conversion from dynamic to static values. Variations in both the static and dynamic values may be caused by an
indeterminate saturation condition and other inconsistencies in the core stress state and mechanical condition.

0.4
Low Axial Stress PR
Hgh Axial Stress PR y = 0.9575x
0.35 Linear (Low Axial Stress PR)
Linear (Hgh Axial Stress PR)

y = 0.9258x
0.3
Static Poisson's Ratio

0.25

0.2

0.15

0.1
0.1 0.15 0.2 0.25 0.3 0.35 0.4
Dynamic Poisson's Ratio
Figure 8: Comparison of core dynamic and static Poisson’s Ratio measurements

Static and dynamic Young’s Modulus


The difference between dynamic and static Young’s Modulus estimates is significantly larger than that observed for
Poisson’s Ratio. The data in Figure 9 show static and dynamic Young’s Modulus measurements on the tight-gas and shale-
gas sample data set measured under “as received” saturation conditions. In all cases, the apparent static modulus is lower than
12 SPE 118703

the dynamic modulus for both high and low axial stress states. Various models have been proposed to correlate dynamic
modulus to a more representative static modulus. Each of the proposed models is based on a specific core data set and
implied rock type and may not be generally applicable to all rock types. The data sets used here are presented to develop
useful correlations specifically for tight-gas and shale-gas reservoirs.

14.000
Low Axial Stress Dyn E
High Axial Stress Dyn E
12.000
Static Young's Modulus, MMpsi

10.000

8.000

6.000

4.000

2.000

0.000
0.000 2.000 4.000 6.000 8.000 10.000 12.000 14.000
Dynamic Young's Modulus, MMpsi
Figure 9: Comparison of core dynamic and static Young’s Modulus values

Eissa and Kazi (1988) are credited with the early work in the area of dynamic to static modulus conversions. They
presented two models for well consolidated, low porosity samples. The first model proposed a linear scaling relationship
given by Equation 13, while the second model proposed a log linear relationship using the product of core bulk density (ρ, in
gm/cm3) and dynamic modulus (Equation 14). In both the Eissa and Kazi models, the dynamic modulus is given in GPa
(approximately 6.895 times modulus in million psi).

E s = 0.74 E d − 0.82 (13)


Log E s = 0.02 + 0.71Log (ρE d ) (14)
Log E s = Log (ρE d ) − 0.55 (15)

The proposed models (Equations 13 and 14) were applied to the tight-gas and shale-gas data sets. To obtain the best fit
with the current data set the coefficients of Equation 14 were adjusted to those given by Equation 15. The coefficients for
Equation 13 were found to give a best-fit straight line for the current data. In addition to the modified Eissa-Kazi models, a
power law fit of the form suggested by van Heerden (1987) was also tested. The best-fit power-law model is given by
Equation 16.

E s = 0.1832 E d1.5795 (16)

Figure 10 shows the static Young’s Modulus values for the entire data set computed from the input dynamic modulus. Of
the three equations tested, the modified E-K Log-Linear model (Equation 15) gives the best results by far over the entire
range of moduli tested. It is possible that some of the data scatter is reduced by using measured core bulk density
measurements made under the same saturation conditions as the corresponding dynamic moduli. While saturation state
strongly affects the dynamic Young’s Modulus, the effects of partial saturation on static modulus are not yet known, but are
expected to be minor.
SPE 118703 13

14.000
Eissa-Kazi Linear
Power-Law Model
12.000
Measured Static Young's Modulus, MMpsi

Modified E-K Log-Linear

10.000

8.000

6.000

4.000

2.000

0.000
0 2 4 6 8 10 12 14
Computed Static Young's Modulus, MMpsi
Figure 10: Conversion of core dynamic to static Young’s Modulus

Stress and Material Anisotropy


At least two different types of anisotropy must be considered in applying laboratory core data to the calibration of field log
data: stress anisotropy and material anisotropy. Many unconventional reservoirs exhibit large-scale inhomogeneity and
intrinsic anisotropy including bedding planes, pre-existing fractures, stylolites, and secondary porosity. Large-scale
anisotropy and inhomogeneity can significantly affect the apparent mechanical properties (Schatz, 1993). Cicotti and
Mulargia (2004) indicate that these large-scale effects generate a bigger difference between core-sample measurements and
field-scale log measurements than any correction from static to dynamic that is based on frequency or other factors. The
volume of rock integrated by a borehole sonic log is inversely proportional to frequency. As a result, most well logs offer an
average DTC and DTS value integrated through many rock layers and include the effects of all fractures, partings, and
bedding planes in the result. Miskimins (2002 and 2003) reports that log-derived average acoustic velocities fail to show the
degree of stratification apparent in whole core samples. Stress profiles derived from these homogenized acoustic logs fail to
adequately predict fracture containment when laminations and fractures are present.
Any large-scale average measurements of rock properties through acoustic measurements must suffer from this problem.
Typically, derived stress and rock property profiles will be overly smoothed and will miss discrete layer properties and
discontinuities that can act as freely sliding shear planes or stress concentrations in tectonically active areas. Until the use of
high-resolution borehole image logs for mechanical properties determination from resistivity or sonic data becomes
established, and until fracture models can accurately handle discontinuity effects on a millimeter scale, any fracture geometry
or stress profile must be considered to be grossly averaged. As a result, fracture height growth and vertical fracture continuity
will be overestimated.

Measurement of anisotropic core properties


A relatively recent development is the measurement of core static and dynamic elastic moduli on adjacent core samples
oriented vertically, horizontally, and at 45 degrees to the whole core (wellbore) axis. In some cases of locally homogeneous
and isotropic media, the results for three adjacent plug samples are reported to be similar. In other cases, differences in
moduli of 2 to more than 5 fold have been reported (verbal communication of proprietary data). Aside from the large-scale
effects on anisotropy, the stress field used in these oriented core tests and its impact on observed deformation must be
considered.
14 SPE 118703

Figure 11 shows a simplified drawing of the three core sample orientations for a core through a horizontally laminated
medium. Sample A in Figure 11 shows a vertical plug with an axial load representing the vertical net overburden stress.
Lateral deformation is driven by the axial compression normal to bedding. Loading in this orientation appears to derive the
appropriate Poisson’s Ratio for use in the modified uniaxial strain equation for horizontal stress estimation (Equation 1). Bi-
lateral anisotropy due to microfractures and grain orientation effects will still cause unequal lateral strains that can be
averaged through radial strain measurement.

σz=σv σr≠σv

σz=σh

σr=σh B σr≠σH
σz≠σv
σr≠σH
σr≠σH

A Optimum angle for


shear along
horizontal bedding
C
σr≠σh
Figure 11: Diagram of oriented core test loading geometry

Analysis of axial compression of a horizontal core plug parallel to bedding (B) is more problematic. In this case, the
applied axial load more closely represents a maximum horizontal stress, as in a thrust-fault environment. The geometry is
proposed to represent displacement of the fracture surface by fracturing fluid pressure. Here, the uniform radial confining
stress must represent both the vertical stress and the second horizontal stress, which will normally be far from equal in field
conditions. For a normal stress state, the radial stress normal to bedding in this laboratory geometry will be far less than the
vertical net stress. This can lead to enhanced shear failure along bedding. A potentially more serious concern exists if the
sample is not homogeneous. Any bedding layers of high modulus will concentrate load and can give a higher than normal
modulus, as compared to an integrated average value. If stress concentration reaches the shear failure limit, the local hard
streaks may fail before the softer surrounding rock, leading to odd mechanical behavior. In dynamic measurements, first-
arrival acoustic signals will be channeled through the hard streaks, altering interpretations of dynamic properties.
In the case of the 45-degree core orientation (C), the relative angle between the core plug axis and bedding may control
failure and apparent moduli. In the orientation shown, the tendency for shear failure along the (potentially) weak bedding
planes is maximized. All these core measurements will show differences in acoustic velocity and apparent static or dynamic
moduli depending on orientation for locally heterogeneous samples. The difficulty lies in applying these measurements to
field predictions of moduli and stress. In most instances, the only wellbore acoustic wave velocities that can be measured will
be vertical, and they will correspond most closely to the orientation of sample A in Figure 11. The large-scale measurements
will inherently be affected by heterogeneities not captured in intact core plug samples. Overall, the measurement of core
properties at different orientations may provide additional information about rock heterogeneity, but these measurements
offer no solution to the problem of predicting fracture geometry or closure stress profile.

Acoustic anisotropy
Acoustic anisotropy can be divided into two main categories: vertical velocity anisotropy and vertical to horizontal
anisotropy. The use of vertically polarized dipole sonic logs has become common in recent years. In theory, a fast and slow
shear wave velocity can be determined from an orthogonally polarized transmitter-receiver pair. Variations in shear velocity
SPE 118703 15

are theoretically related to changes in stress, but can also be caused by changes in saturation, bedding features, fractures, and
other environmental factors. Differentiating vertical to horizontal anisotropy seeks to address the tendency of formations to
fail in horizontal or low-angle planes (as in delamination or disking of cores). Currently, however, no acoustic borehole
logging tools can actually measure horizontal compressional and shear velocities. Vertical to horizontal anisotropy appears to
be interpreted based on reflected and refracted tube and surface waves. If properly interpreted, a high degree of vertical to
horizontal anisotropy should correlate to better fracture height containment and possibly higher treating pressures.

Brittle-Ductile Characterization of Shale


Shale “reservoirs” have been characterized as brittle versus ductile based on log-derived mechanical properties. More brittle
systems are perceived to be easier to fracture while ductile systems deform more plastically and may be more difficult to
treat. The degree of brittleness may also relate to the tendency for generation of complex shear fractures. Brittleness, as
explained by Mullen et al (2008), is based on apparent Young’s Modulus (E) and Poisson’s Ratio (ν). A formation with low
E and high ν is considered ductile, while high E and low ν indicates a brittle formation. The characterization is generally
useful for qualitative evaluation of expected treating conditions. Mullen (2008) suggests different fracture treatment designs
for varying degrees of brittleness, with the intent of creating complex shear-enhanced fracture networks in britte rocks and
planar, high conductivity, fractures in ductile rocks.
It must be remembered, however, that the apparent dynamic moduli used to define brittleness are based on observed
acoustic velocities that are affected by TOC, gas content, fractures, and borehole conditions. Applying raw or uncorrected
sonic log data may lead to inappropriate characterizations. For example, a highly fractured and gas charged zone may show
slow acoustic velocities and be interpreted as a low modulus, high Poisson’s Ratio formation. Its behavior may more closely
be modeled by a shattered brittle rock, in terms of appropriate treatment selection.

Examples of Stress Profiles


Many of the factors discussed can be illustrated in a field example comparing rock elastic properties and resulting stress
profile with actual fracture containment shown by tracer surveys. The log data in Figure 12 cover a range of lithologies. Two
separate fracture treatments, indicated by the blue boxes at right, were conducted in the well. The far-left track shows five
interpretations of Poisson’s Ratio derived from log values. The green curve (POIS) is the reported dynamic value of ν from a
full-waveform dipole sonic log as interpreted by the logging company after complete processing of the data. The blue-gray
curve (PRACT) is the value of ν computed using Equations 2 and 4 with no correction for saturation or borehole conditions.
The PRACT curve is a direct overlay of POIS over the logged interval, indicating that (as usual) no fluid substitution or other
corrections have been applied during “log mechanical properties processing” by the logging company.
The PRPHIA curve in Figure 12 was derived from the crossplot average porosity without use of the measured DTC or
DTS data and gives a near perfect overlay of POIS, as is usual. Inputs to this estimate of ν include the estimated matrix and
fluid sonic transit times and the neutron-density crossplot porosity. The two dissenting curves, PRDTC and PRRESIST, are
derived from compressional sonic travel time and the interpretation of lithology and from deep resistivity independent of
lithology and sonic data respectively. In most unconventional reservoirs, these values give the best prediction of stress profile
and fracture geometry.
16 SPE 118703

Stress DTS/DTC Stress PR_RESIST

POIS and
PRRESIST PRACT

PRDTC
DTC
DTCRESIST

Figure 12: Comparison of stress profiles for different log interpretations

The curves in the second track are the reported fast and slow shear slowness reported from the crossed-dipole shear log. A
maximum shear anisotropy of 3 μs/ft out of 150 μs/ft (2%) is observed in a few spots over the logged interval, while most of
the log shows no discernable shear velocity anisotropy. It is interesting to note that this area is highly faulted and has a
known dominant fracture orientation with significant stress anisotropy. It could be that the rock is hard enough that a change
in horizontal stress of more than 500 psi has no measurable effect on shear velocity. Other local factors (saturation, organics,
and fractures) may have larger impacts. Two stress profiles are presented in Figure 12, computed using the log-derived
mechanical properties and Equation 1. The resulting stress profiles are compared to the radioactive tracer log acquired after
the fracture stimulation treatments. The stress profile derived from the acoustic DTS/DTC ratio fails to predict the contained
fracture observed in the lower perforation set. In contrast, the confined fracture is predicted by the resistivity derived stress
profile. In this case both stress profiles would predict a contained fracture geometry for the upper treatment but the
magnitudes of the predicted stresses differ by nearly 1000 psi. In most cases the properties derived from sonic velocity
measurements fail to give useful predictions of fracture geometry or treating pressure.

Conclusions
Estimation of in-situ stress profiles in unconventional reservoirs is complicated by many factors. Determining pore
pressure when matrix permeability is vanishingly small can be nearly impossible by conventional means such as wellbore
measurements or even pressure transient tests. Pore pressure can vary locally and to a large degree when driven by
hydrocarbon maturation. The local pore pressure is a primary driver of the total closure stress. Pore pressure also affects the
net effective stress in the system and can alter acoustic velocities and apparent dynamic moduli.
Acoustic velocities from sonic logs, and dynamic moduli derived from them, can be affected by gas saturation and TOC
in tight formations and shale where near-wellbore flushing while drilling is ineffective. The rock volume interrogated and
integrated by the sonic log may be at a highly variable saturation state, and the measured sonic velocity may respond more to
changes in saturation than to stress or rock property variations. Fractures, laminations, and averaging of dissimilar lithology
response can also affect velocity and dynamic properties derived from borehole sonic logs.
SPE 118703 17

Saturation, stress history, anelastic strain, and many other factors also affect core samples used for both static and
dynamic measurements of elastic moduli. Core data can be used with care to develop models for deformation and correlation
to dynamic properties. Stress cycling and proper saturation of cores is important in deriving useful data for both static and
dynamic core measurements. Core samples typically represent point values of intact rock and will give indications of higher
strength and isotropy than may actually be present in larger scale samples.
Use of core samples at various orientations can provide useful qualitative information about small-scale anisotropy. This
information should be extended to large-scale estimates of stress and deformation with extreme caution. Stress and strain
boundary conditions in these oriented core laboratory tests may not be appropriate for field application.
Use of synthetic sonic logs and elastic moduli is strongly encouraged in unconventional reservoirs. If the derived data
confirms the direct acoustic measurements, then it is possible to derive some degree of comfort. When the synthetic derived
data and measured sonic data disagree, there is usually a good and definable reason. If the cause of the discrepancy can be
traced to pore fluid, pore pressure, or discontinuity effects, then the synthetic data typically give better and more useful
estimates of the necessary rock properties for both stress profile calculation and fracture geometry evolution.

Nomenclature

CEG : Constant offset for estimation of E from GR


CER : Constant offset for estimation of E from resistivity
CPG : Multiplying pre-factor for estimate of ν from GR
DTS : Shear sonic slowness, microseconds/ft
DTC : Compressional sonic slowness, microseconds/ft
DTC_PHIA : Compressional slowness estimated from crossplot average porosity fraction, microseconds/ft
DTFL : Pore fluid sonic travel time (slowness), microseconds/ft
DTMA : Rock matrix sonic travel time (slowness), microseconds/ft
Ε : Young’s Modulus, million psi (unless otherwise specified)
Es : Static Young’s Modulus
Ed : Dynamic Young’s Modulus
EER : Exponent for estimation of E from resistivity
EPG : Exponent for estimation of ν from GR
EPR : Exponent for estimation of ν from resistivity
GR : Total gamma-ray counts, API units
Ish : Dimensionless GR index
OEG : Constant offset for estimation of E from GR
Pc : Total fracture closure stress or minimum in-situ stress, psi
Pp : Pore pressure, psi
R : Square of sonic travel time ratio, dimensionless
RESIST : Formation electrical resistivity, OHMM
Sw : Water saturation in pore space, fraction
Vp : Compressional acoustic velocity, ft/sec
Vs : Shear acoustic velocity, ft/sec
Vsh : Volume fraction shale (or clay) in formation
YME_GR : E estimated from GR, million psi (unless otherwise specified)
YME_RESIST : E estimated from resistivity, million psi (unless otherwise specified)
YME_PHIA : E estimated from crossplot average porosity, million psi (unless otherwise specified)

αv : Vertical poroelastic constant, dimensionless


αh : Horizontal poroelastic constant, dimensionless
εh : Horizontal tectonic strain normal to minimum stress, microstrains
ν : Poisson’s Ratio, dimensionless
φ : Porosity, fraction
ρb : Bulk density, g/cm3
σh : Minimum horizontal total stress, psi
σt : Tectonic stress offset, psi
σv : Vertical total overburden stress, psi
18 SPE 118703

References

Al-Shayea, N. A.: “Effects of testing methods and conditions on the elastic properties of limestone rock,” Engineering Geology v. 74, pp.
139–156. 2004.
Blanton, T.L., and Olson, J.E.: “Stress Magnitudes from Logs: Effects of Tectonic Strains and Temperature,” SPE Reservoir Evaluation &
Engineering, Volume 2, Number 1, Pages 62-68, February 1999.
Briaud, Jean-Louis, “Intro to Moduli’”, ceprofs.tamu.edu/briaud/, 2008.
Ciccotti, M. and Mulargia, F.: “Differences between Static and Dynamic Elastic Moduli of a Typical Seismogenic Rock,” Geophys. J. Intl.,
v. 157, pp. 474-477, 2004.
Crain, E. R., CRAIN'S PETROPHYSICAL HANDBOOK, http://www.spec2000.net/chapters/Chapter06.htm, Rocky Mountain House,
Alberta, Canada, 2008.
Eissa, A., Kazi, A.: “Relation between static and dynamic Young’s moduli of rocks,” Int. J. Rock Mech. Min. Sci. Geomech. Abstr.
Volume 25, Number 6, Pages 479– 482, 1988.
Holmes, M., Holmes, A., and Holmes, D.: “Modification of the Wyllie Time Series Acoustic Equation, to Give a Rigorous Solution in the
Presence of Gas,” presented at the SPWLA 45th Annual Logging Symposium, Noordwijk, The Netherlands, June 6-9, 2004.
http://Hyperphysics.phy-astr.gsu.edu/hbase/sound/souspe3.html, 2008.
http://server4.oersted.dtu.dk/research/RI/SNG/SNG-logs.html, 2008.
Knight, R., Dvorkin, J. and Nur, A.: “Acoustic Signatures of Partial Saturation,” GEOPHYSICS, v. 63, n. 1, pp. 132–138, Jan-Feb 1998.
La Gesse, J., and Hurley, N.: “Predicting Source Rock Quality using GR Wireline Response and Umaa vs. ρmaa Crossplots in the Lewis
Shale, Green River Basin, Wyoming,” paper SPE 114963 presented at the 2008 SPE Annual Technical Conference and Exhibition,
Denver, Colorado, U.S.A., 21-24 September 2008.
Li, Yongyi and Schmitt, D. R.: “Well-Bore Bottom Stress Concentration and Induced Core Fractures,” AAPG Bulletin, V. 81, No. 11 P.
1909–1925, November 1997.
Lim, S. S., Martin, C. D., and Christiansson, R.: “Estimating In-Situ Stress Magnitudes from Core Disking,” In-Situ Rock Stress – Lu, Li,
Kjerholt, and Dahle (eds), Taylor and Francis Group, London, ISBN 0-415-40163-1, 2006.
Merkel R. H. and Barree, R. D., and Towle, G.: “Seismic Response of Gulf of Mexico Reservoir Rocks with Variations in Pressure and
Water Saturation,” THE LEADING EDGE, pp. 290-299, March 2001.
Miskimins, J. L. and Barree, R. D.: “Modeling of Fracture Height Containment in Laminated Sand and Shale Sequences,” paper SPE
80935 presented at the Production and Operations Symposium held in Oklahoma City, OK, March 22-25 2003.
Miskimins, J. L., Hurley, N., and Graves, R.: “A Method for Developing Rock Mechanical Property Logs using Electrofacies and Core
Data,” paper SPE 77783 presented at the SPE Annual Technical Conference and Exhibition, San Antonio, TX, Sept 29-Oct 2, 2002.
Morales, R. H. and Marcinew, R. P.: “Fracturing of High Permeability Formation: Mechanical Properties Correlations,” paper SPE 26561
presented at the 68th Annual Technical Conference and Exhibition, Houston, TX, October 3-6, 1993.
Mullen, M., Russel Roundtree, R, Barree, R.: “A Composite Determination of Mechanical Rock Properties for Stimulation Design (What
to do When You Don’t Have a Sonic Log),” paper SPE 108139 presented at the 2007 SPE Rocky Mountain Oil & Gas Technology
Symposium, Denver, Colorado, U.S.A., 16–18 April 2007.
Passey, Q. R., Creaney, S., Kulla, J. B., Moretti, F. J., Stroud, J. D.: "A Practical Model for Organic Richness from Porosity
and Resistivity Logs", AAPG Bull., Dec. 1990.
Perreau, P.J., Heugas, O., and Santarelli, F.J.: “Tests of ASR, DSCA, and Core Disking Analyses to Evaluate In-Situ Stresses,” paper
number 17960 presented at the Middle East Oil Show, March 11-14, 1989, Bahrain.
Rickman, R., Mullen, M., Petre, E,, Grieser, B., Kundert, P.: “A Practical Use of Shale Petrophysical Properties for Stimulation Design
Optimization: All Shale Plays are not Clones of the Barnett Shale,” paper SPE 115258 presented at the 2008 SPE Annual Technical
Conference and Exhibition, Denver, Colorado, U.S.A., 21-24 September 2008.
Rondeel, H.E.: “HYDROCARBONS,” Tekst voor de cursus Grondstoffen en het Systeem Aarde (HD 698), pp. 17-30 December 2001.
Schatz, J. F., Olszewski, A. J., and Schraufnagel, R. A.: “Scale Dependence of Mechanical Properties: Application to the Oil and Gas
Industry,” paper SPE 25904 presented at the SPE Rocky Mountain Region/Low Permeability Reservoirs Symposium, Denver, CO,
April 12-14, 1993.
Stieber, S. J., and Thomas, E. C.: “The Distribution of Shale in Sandstones and its Effect upon Porosity,” SPWLA Sixteenth Annual
Logging Symposium, June 4-7, 1975.
Thiercelin, M. J., and Plumb, R. A.: “Core-Based Prediction of Lithologic Stress Contrasts in East Texas Formations,” SPE Formation
Evaluation, pages 251-258, Dec. 1994.
Van Heerden, W.L.: “General relations between static and dynamic moduli of rocks,” Int. J. Rock Mech. Min. Sci. Geomech. Abstr.
Volume 24, Number 6, Pages 381– 385, 1987.
Warpinski, N.R., Peterson, R.E., Branagan, P.T., Engler, B.P., and Wolhart, S.L.: “In Situ Stress and Moduli: Comparison of Values
Derived from Multiple Techniques,” paper SPE 49190 presented at the 1998 SPE Annual Technical Conference and Exhibition held
in New Orleans, Louisiana, 27-30 September 1998.
Warpinski, N.R., Teufel, L.W.: “A Viscoelastic Constitutive Model for Determining In-Situ Stress Magnitudes From Anelastic Strain
Recovery of Core (includes associated papers 19042 and 19892 ),” Journal SPE Production Engineering, Volume 4, Number 3,
Pages 272-280, August 1989.
Yassir, N., Wang, D. F., Enever, J.R., and Davies, P.J.: “Experimental Analysis of Anelastic Strain Recovery of Synthetic Sandstone
Subjected to Polyaxial Stress,” Paper 47238 presented at the SPE/ISRM Rock Mechanics in Petroleum Engineering, Trondheim,
Norway, July 8-10, 1998.

You might also like