Sei sulla pagina 1di 41

A tutorial on AVO and Lamé constants for rock parameterization and fluid detection.

Bill Goodway; PanCanadian Petroleum Limited,Calgary,Canada.

Introduction
For more than 3 decades, industry has known that shear seismic waves (S-waves) contain different rock
information than do our standard compressional seismic waves (P-waves). Periodically, efforts to record S-
waves, even on the seabed, have attracted industry attention with increasing success. Separate efforts to
analyse conventional P-waves for the S-wave information contained or missing within them, have had more
success and popularity as a function of significantly lower recording cost and relatively simpler analysis.
These methods termed AVO (amplitude versus offset) are a logical and quantifiable petrophysical extension
into pre-stack seismic data from the confusing and simplistic interpretation of stacked amplitudes. Since the
mid 90’s new AVO inversion methods have been gradually succeeding in the Western Canadian Sedimentary
Basin (WCSB), most notably in the exploration and development of gas pools within both clastic and
carbonate plays. The improved approach described in this paper, combines conventional seismic AVO with
petrophysical log analysis through the understanding of the fundamental Lamé parameters of rigidity Mu (μ)
and incompressibility Lambda (λ). These parameters form a common theme between the varied aspects of
AVO and log analyses considered and necessitate the use of some equations that are relatively simple to
follow and hopefully reveal a deeper insight into the connection between rock physics and quantitative AVO.
Standard industry AVO methods exploit anomalous variations between seismic compressional wave velocity
(Vp) or impedance (Ip) and shear wave velocity (Vs) or impedance (Is), to indicate changes primarily in pore
fluid, as well as lithologic properties (Gassmann 1951, Pickett 1963, Tatham 1982, Castagna 1993a). Other
methods using AVO measurements derive Poisson’s ratio (Ostrander 1984, Shuey 1985, Verm & Hilterman
1995) or P and S reflectivities, i.e. impedance contrasts (Gidlow et al.1992, Fatti et al.1994, Wallace & Young
1996). The emphasis on seismic velocity and density arises from the Knott-Zoeppritz equations for continuity
of displacement (u) and stress (σ) across interfaces between different lithologies for a propagating seismic
wave. Displacement and stress are usually derived from a plane wave solution of the acoustic wave equation;
u = Aeiω(t−x/V). However the underlying physics in the wave equation; d2u/dx2 = ρ/M (d2u/dt2) does not involve
seismic velocities, but instead the ratio of density (ρ) to modulus (M).
This paper will show that understanding velocity or impedance measurements in Lamé parameter terms of
rigidity μ and incompressibility λ, offers new insight into the rock property factor ρ/M that governs wave
propagation. In what follows, λ is considered to be pure incompressibility and not the bulk modulus κ, as λ is
the only modulus involved in both the hydrostatic stress-strain relationship and acoustic wave propagation for
a fluid (i.e. where rigidity μ vanishes). In addition to these Lamé parameters λ and μ two new attributes
LambdaRho λρ and MuRho μρ (Lamé impedances) are obtained from AVO by using moduli and density
relationships to impedance. An improved identification of reservoir zones is possible by the enhanced
sensitivity to pore fluids from pure incompressibility λ or λρ. Furthermore a better understanding of lithologic
variations independent of fluid effects, such as sand/shale ratio can be obtained by analysing fundamental
changes in rigidity μ, incompressibility λ, and density ρ parameters as opposed to a mixture of parameters
within seismic velocities or impedances.

Theory and methods.


The dependence of seismic wave propagation on moduli is not obvious in standard AVO methods that rely on
intercept and offset gradient, P-wave and S-wave reflectivity (Rp, Rs) or velocities Vp/Vs and Poisson ratio
variations. Some authors point out the need for a more physical insight afforded by rigidity μ (Wright 1984,
Thomsen 1990, Castagna et al. 1993b). Castagna also indicates that the link between velocity and rock
properties for pore fluid detection, is through the bulk modulus κ that is embedded in Vp. However as λ relates
uniaxial and lateral strain to uniaxial stress, it is primarily a longitudinal measure and hence orthogonal (i.e.
independent) to μ, a quantity that relates a shearing stress to strain. This is not the case for κ since it implicitly
relates to a measure of shear strain due to the volume’s change in shape resulting from triaxial or hydrostatic
stress. This confusion can be seen in the following relationships; Vp2 = (λ + 2μ)/ρ = (κ + (4/3)μ)/ρ (where κ =
λ + (2/3)μ) and Vs2 = μ/ρ and is further developed in the sections explaining moduli. Figures 1 and 2, show
how the mixed parameters within velocity (Vp) for the compressional wave experiment result in an ambiguity

1
in discriminating porous sand from shale. The equivalent and simpler shear experiment measures shear
velocity (Vs) that is proportional to rigidity μ and thereby has less ambiguity discriminating between sand and
shale. However a measure of rigidity alone is unable to provide a direct estimate of fluids within the sand pore
volume.

Shear Wave Compressional P Wave

=>

Undeformed schematic Shear Velocity Compressional Velocity


rock volume of porous (Dipole Log) (Sonic Log)
sandstone
μ Vp = λ + 2μ
Vs =
ρ ρ
Figure 1.Modes of seismic wave propagation

Background encasing rock volume

Sandstone
Porosity Shaliness
λ decrease μ decrease

Impedance IP , Velocity VP and ρ decrease


for both sanndstone porosity and shaliness
P Impedance x VP = ρ VP2 = λ + 2μ

Figure 2. Sandstone porosity versus shaliness ambiguity.

2
Hooke’s Law explained in Lamé moduli terms.
Hooke’s law relates the rank 2 tensors of stress (σ) to strain (e) with indices ‘j’ and ‘l’ being face normals and
‘i’ and ‘k’ the directions of stress or strain in a Cartesian system of x, y and z. The particle displacement vector
u also has indices ‘j’ and ‘l’ denoting x, y or z directions.
z 33

Hooke' s law; σ ij = cijkl ekl 23


13 32
1 ⎛ du du ⎞ 1
where strain ekl = ⎜⎜ k + l ⎟⎟ = (u k ,l + u l ,k )
31
22
2 ⎝ dxl dx k ⎠ 2 21 12
y
11
as represented in derivative indical notation x
(notation as used in appendix for wave equations)

Stiffnesses cijkl quantify the stress to strain ratio and are rank 4 tensors (i.e. relating the rank 2 stress σij to strain
ekl tensors) given as;
c ijkl = λδ ij δ kl + μδ ik δ jl + μδ il δ jk
This expression contains the Lamé parameters or moduli of a single incompressibility Lambda (λ) and two
rigidity Mu (μ) terms. These Lamé moduli are combined for specific directions of stress to strain tensor
relationships, as constrained by the Kronecker delta function rules. For the λ term these constraints are δij
requiring that i = j and likewise δkl requiring that k = l. So we can see that as the stress and strain face normals
and directions are in the same x, y or z direction, then this λ modulus relates to longitudinal or compressional
quantities. However the μ terms are not so simple. As the two terms have the requirements of δik δjl and δil δjk,
so strain indices k, l are replaced by i, j. This means that the stress and strain indices have the same i and j
directions and so these μ rigidity moduli appear in Hooke’s law relating compressional stress to strain i.e. i = j.
This can be seen in the derivation for axial or compressional stress by substituting the expression for cijkl in
Lamé moduli into Hooke’s law;
σij = cijkl e kl = (λδ ijδ kl + μδik δ jl + μδilδ jk )e kl where δijδ kl ⇒ i = j, k = l
⇒ λδijδ kl e kl ⇒ λδ ije kk
for the first μ term δik δ jl ⇒ i = k , j = l, with the second μ term δil δ jk ⇒ i = l, j = k
⇒ (μδik δ jl + μδil δ jk )e kl ⇒ μeij + μe ji = 2μeij
⇒ σij = λδ ije kk + 2μeij where for axial stress i = j
⇒ σii = λe kk + 2μeii ⇒ in the x direction σ xx = (λ + 2μ )e xx + λe yy + λe zz
This confusing result shows that the rigidity modulus μ is involved in the stress-strain relationship for axial
compression. A physical explanation to better understand this, is that a compressional stress that alters the
volume of a body must also change the body’s shape thereby invoking a resistance as a function of both the
incompressibility modulus λ, as well as the rigidity modulus μ of the body’s material.
By contrast the pure shear or rotational stress-strain relationship is clearly a function of rigidity μ only i.e.
from the change in the body’s shape alone. This is seen below by substituting the expression for cijkl in Lamé
moduli into Hooke’s law to give;
σ ij = c ijkl e kl = (λδ ij δ kl + μδ ik δ jl + μδ il δ jk )e kl where for shearing stress, strains i ≠ j, k ≠ l, ⇒ λδ ij δ kl e kl = 0
and the first μ term δ ik δ jl ⇒ i = k , j = l, with the second μ term δ il δ jk ⇒ i = l, j = k
⇒ (μδ ik δ jl + μδ il δ jk )e kl ⇒ μe ij + μe ji = 2μe ij
⇒ σ ij = 2μe ij where for shear stress in the z direction on the " x normal" face
1 ⎛ du du ⎞ du du
σ zx = 2μe zx = 2μ ⎜ z + x ⎟ and as x = 0 so ⇒ σ zx = μ z
2 ⎝ dx dz ⎠ dz dx

3
Defining seismic impedance and the relationship of modulus, density and propagation velocity from
the wave equation and its plane wave solution.
Seismic impedance is a fundamental quantity forming the basis for all reflection methods. However by being
the product of propagation velocity and density, impedance is not an obvious quantity for the average
practising geophysicist, unlike propagation velocity and hence arrival times. As AVO attribute analysis
involves extracting Rp and Rs reflectivities or impedance contrasts and as the new Lamé impedances λρ and
μρ stem from impedance inversion, so the following is a derivation of impedance and the relationship of
modulus to velocity, from the wave equation.
dσ dν
Starting with the1D acoustic wave equation; =ρ
dx dt
du iω (t − x ) iω (t − x )
( where particle velocityν = = iωAe V with particle displacement u = Ae V
dt
as a plane wave solution of the wave equation and σ is stress with V as propagation velocity)
Integrate both sides with respect to distance x;
dσ dν
∫ dx
dx = ρ ∫
dt
dx

⇒σ = ρ
d
dt
(∫ ) d
νdx = ρ ⎛⎜ iωAe
dt ⎝
iω (t − x )
∫ V dx ⎞ = ρ


d ⎛ − iω

dt ⎝ iω
VAe
iω (t − x ) ⎞
V

d ⎛ iω (t − x V ) ⎞
⎟ = - ρV ⎜ Ae
dt ⎝


du d ⎛ iω (t − V ) ⎞x
next substituting ν = = ⎜ Ae ⎟
dt dt ⎝ ⎠
σ
⇒ σ = - ρVν ⇒ I = ρV = -
ν
gives the definition of Impedance I (I = ρV) as being the negative ratio of stress to particle velocity
(Aki &Richards1980, p137)
Also as pressure is defined as P = - σ so P = ρVν
Continuing with this result the original wave equation may be written in terms of stress as;
dσ dν dσ dσ −σ
=ρ ⇒V =− (fromν = )
dx dt dx dt ρV
So the wave equation can be simply stated as
" Propagation velocity V is equal to the negative ratio of the time to space derivatives of stress"
Finally rewriting σ = - ρVν in terms of displacement where M is the medium's modulus
du
and using Hooke's law σ = M gives;
dx
du du
M = − ρV
dx dt
integrating both sides w.r.t. x gives;

∫ dt dx ⇒ Mu = − ρV dt (∫ udx) = − ρV dt ⎜⎝ − iω Ae
du d d⎛ V (
iω t − x )⎞ (
iω t − x )
Mu = − ρV V
⎟ replacing u = Ae V

( )⎞
d⎛ Ae
iω t − x
V
(
iω t − x ) iρV ⎜⎝
2 ⎟
⎠ (
iω t − x ) (
iω t − x )
⇒ MAe V =− ⇒ MAe V = ρV 2 Ae V
ωdt
leading to the expression relating modulus, density and propagation velocity as M = ρV 2

One of the consequences of the relationship of stress to the product of particle velocity and impedance, is that
for the same initiating stress or pressure a seismic wave will have a lower particle velocity in a high impedance

4
material than a low impedance material. Though this is strictly an elastic effect, an explanation for anelastic
attenuation can be deduced as the low impedance material particle must travel further per frequency cycle (i.e.
faster) than the high impedance material particle, so it will suffer more energy loss through frictional forces.
This gives rise to the common observation that sound or seismic waves suffer less high frequency attenuation
in high impedance materials such as steel or water when compared to low impedance material such as wood or
air. Further insight through an understanding of the Lamé parameters in relationships involving the elastic
wave equation and anisotropy, are given in the appendices.

Understanding common moduli and moduli ratios in Lamé terms.


As a consequence of combining the Lamé constants of rigidity μ and incompressibility λ in the stress-strain
relationship for a body so μ and λ form the basis for more familiar moduli definitions. These are dependent on
the form of the body and the nature of the stress i.e. bound insitu material (rock in the subsurface), an unbound
bar (core) or bulk triaxial measurements (fluids under compression). Useful insight into common moduli and
moduli ratio definitions can be obtained by conversion to basic Lamé terms.
a) Compressional P-wave modulus; bound uniaxial compression M = λ + 2μ. This is the modulus involved in
P-wave propagation in the Earth for both surface seismic and sonic logging.
b)Young’s modulus or unbound uniaxial compression; Y = λ(1-2σ) + 2μ or Y = M - 2λσ. The unbound
nature of this measurement (e.g. a core) reduces the stiffness of the body as seen by the last equality where
bound modulus M > Y by a factor of 2λ times Poisson’s ratio. The factor of two is a result of λ acting in both
unbound lateral directions.
c) Bulk modulus; κ = λ + (2/3)μ or κ = M - (4/3)μ. Where the reduced unbound rigidity of the bulk modulus
compared to the bound M modulus can be seen from M > κ by a rigidity factor of (4/3)μ. The bulk modulus
contains a rigidity term (2/3)μ as a consequence of a triaxial stress changing not just the volume of a body (i.e.
a compressional resistance as a function of incompressibility λ), but also its shape (i.e. a shearing resistance as
a function of rigidity μ). This is my objection to calling the bulk modulus incompressibility (in favour of
Lambda) as κ contains an element of rigidity.
Alternatively the bulk modulus can also be seen as a “triaxial Young’s modulus” from 3κ = Y/(1-2σ).
d) Poisson’s ratio; σ = λ /(2λ + 2μ) and Vp/Vs ratio = √(2+λ/μ).
As the various definitions for moduli and moduli ratios contain mixtures of λ and μ so there are some
inconsistencies in considering negative λ’s with positive μ’s for sedimentary rocks. For example the bulk
modulus κ is still positive for negative λ’s until -λ = (2/3)μ beyond which it is negative. However Poisson’s
ratio is negative for negative λ’s until at -λ = μ Poisson’s ratio is infinite and discontinuous, returning to being
positive for -λ > μ. These impossible and non-linear variations of bulk modulus and Poisson ratio with the
fundamentally unambiguous λ/μ ratio is shown in the graph below (figure 3a). Further discussion and potential
explanation of the confusion is given in the following section on improved sensitivity using Lamé attributes.
Poisson’s ratio, σ = λ/(2λ+2μ) or 2σ/(1−2σ) = λ/μ and (Vp/Vs)2 = (λ/μ)+2, as given above are close to the
most sensitive moduli ratio of λ/μ. Unfortunately the non linear complexity of σ in Lamé terms (Thomsen
1990) and the constant 2 in (Vp/Vs)2, reduce the full impact of the relative λ/μ ratio change between
lithologies or fluids within lithologies. This can be seen for positive values in figure 3a, as a reduction below
the 1:1 line in the slopes or relative ranges of Poisson’s and Vp/Vs ratios versus the larger λ/μ ratio range. The
relationship λ/μ = 2σ/(1−2σ) is interesting as it shows that for a change in Poisson ratio, the numerator
increases as twice the σ change while the denominator is reduced by 1-2σ thereby enhancing the λ/μ change to
a maximal asymptotic approach for σ = 0.5 where λ/μ is infinite.

5
Fig.3a, Variation of Vp/Vs, Poisson and (Bulk Modulus K)/Mu Ratios with
Lambda/Mu Ratio

4.0
3.5
Vp/Vs, Poisson, Bulk K/Mu Ratios (-ve

3.0
2.5
2.0
values in red)

1.5
1.0
0.5
0.0
3.6 3.2 2.8 2.4 2.0 1.6 1.2 0.8 0.4 0.0 0.4 0.8 1.2 1.6 2.0 2.4 2.8 3.2 3.6
0.5
1.0
1.5
2.0
Lambda/Mu Ratio (-ve values in red)

Vp/Vs Poisson ratio Bulk Modulus K/Mu

Lamé constants used in improved AVO attribute sensitivity for fluid/lithology discrimination.
The basic proposal in this paper is to use moduli/density relationships to velocities V or impedances I, given
as:
Ip2 = (Vp. ρ)2 = (λ + 2μ)ρ and Is2 = (Vs. ρ)2 =μρ.
These relationships enable extraction of the orthogonal Lamé parameters λ and μ from logs with measured
density ρ, or λρ and μρ from seismic without density. The simple derivations are:
λ = Vp2.ρ − 2Vs2.ρ, μ = Vs2.ρ, and λρ = Ip2 − 2 Is2, μρ = Is2.
Incompressibility λ is not directly measurable in rocks unlike rigidity μ, but the extraction can be seen as a
form of stripping off the μ sensitive rock matrix to reveal the most sensitive pore fluid indicator λ.
TABLE 1 Vp(m/s) Vs(m/s) ρ(g/cc) Vp/Vs (Vp/Vs)2 σ λ+2μ μ λ λ /μ
Shale 2898 1290 2.425 2.25 5.1 0.38 20.37 4.035 12.3 3.1
Gas Sand 2857 1666 2.275 1.71 2.9 0.24 18.53 6.314 5.9 0.9
Avg.change 1.4 % 25 % 6.4 % 27 % 55 % 45 % 9.2 % 44 % 70 % 110 %
(moduli λ, μ are in GPa’s)

Table 1, shows the justification and power of the method in petrophysical analysis. Actual Vp, Vs and ρ values
from a shallow gas well (not in the study area discussed later) have been combined to give various rock
property values and average % change i.e. contrast at the interface for detectability. The unusual behaviour of
a very limited Vp change (1.4%) compared to Vs (25%) for this thick, good quality gas sand zone requires
some explanation, as most standard measurements concentrate on this non-responsive Vp change. The answer
lies in comparing the last four columns, where it is apparent that because of Vp’s dependence on both λ and μ,
the effect of decreasing λ as a direct response of the gas porosity, is almost completely offset by an increase in
μ in going from capping shale to gas sand. However by breaking out λ from Vp and combing it into a λ/μ
ratio, average changes of 70% for λ and 110% for λ/μ are by far the most sensitive to the variation in rock
properties between shale and gas sand. The best conventional measurements from Poisson’s ratio (σ) of 45%
and (Vp/Vs)2 of 55% are far less sensitive.
With the increased sensitivity in discriminating shale from gas sand of the Lamé parameters in Table 1, an
analysis of errors must be considered as follows. Starting from equations for λ, μ, λρ and μρ above, an error
analysis can be determined from the partial derivatives for λ and μ in terms of Vp, Vs and ρ.

6
For λ: dλ=2ρVpdVp, dλ= -4ρVsdVs and dλ=(Vp2-2Vs2)dρ, and for μ: dμ=ρ2VsdVs and dμ=Vs2 dρ (1)
If errors in Vp, Vs and ρ are uncorrelated and random then the total error ε in λ and μ is given as the sum of the
partial derivatives or errors squared;
ε = √(Σ ε2)⇒ for λ; ε(λ) = √Σ{(2ρVp dVp)2 + (-4ρVsdVs) 2 + [(Vp2-2Vs2)dρ]2} (2)
which can be expressed in terms of fractional errors;
ε(λ) = dλ/λ = √Σ{[(2Vp2/(Vp2-2Vs2))dVp/Vp]2 + [(-4Vs2/(Vp2-2Vs2))dVs/Vs)]2 +(dρ/ρ)2}
⇒ dλ/λ = √Σ{ [(2Vp2dVp/Vp)2 +(4Vs2(-dVs/Vs))2]/(Vp2-2Vs2)2 + (dρ/ρ)2} (3)
and similarly for μ; dμ/μ = √Σ{(2dVs/Vs) 2 + (dρ/ρ)2} (4)
Assuming errors in Vp, Vs and ρ of 1% each with shale values for Vp = 2898m/s and Vs = 1290m/s from table
1, results in a dλ/λ error of 3.7 % from equation 3, and a dμ/μ error of 2.2 % from equation 4. These errors are
within an acceptable range and confirms the observation of dλ/λ errors in practice (Ganguly and Goodway,
2001 PCP internal TEK conference). However if the errors in Vp and Vs are correlated and the dVs/Vs error is
2.5 times larger than dVp/Vp (Cambois, 2000; Herrmann and Cambois, 2001) then in equation 3 the sign for the
dVs/Vs term is retained, despite being squared, and equation 3 can be rewritten as;
dλ/λ = √Σ{ [(2Vp2dVp/Vp)2 -(4Vs2(2.5dVp/Vp))2]/(Vp2-2Vs2)2 + (dρ/ρ)2} (5)
Using the same Vp and Vs shale values as above from table 1, along with Vp and ρ errors of 1% gives a dλ/λ
error of 1.1 % from equation 5 and this is now similar to the original Vp and ρ errors and much less than both
Vs and μ errors.
Distinguishing porous sands within varied channel facies (highlighted in yellow) can be difficult to impossible
using common AVO attributes such as Vp/Vs and Poisson ratio as seen in table 2 below.
TABLE 2 Vp Vs Rho P Impedance S Impedance LambdaRho MuRho Bulk.Mod. Vp/Vs Poisson Lambda/Mu
Ostracod Limestone 5464 2186 2.70 14752.80 5902.20 147.97 34.84 63.41 2.50 0.40 4.25
Ostracod Shale 2851 2115 2.30 6557.30 4864.50 -4.33 23.66 4.98 1.35 -0.11 -0.18
Detrital 4372 2429 2.58 11279.76 6266.82 48.69 39.27 29.02 1.80 0.28 1.24
Channel Shale Plug 4372 2342 2.60 11367.20 6089.20 55.06 37.08 30.68 1.87 0.30 1.48
Regional Shale/Silt 4098 2186 2.55 10449.90 5574.30 47.05 31.07 26.58 1.87 0.30 1.51
Porous Channel Sand 4098 2342 2.35 9630.30 5503.70 32.16 30.29 22.28 1.75 0.26 1.06
Tight Channel Sand 4684 2623 2.55 11944.20 6688.65 53.19 44.74 32.55 1.79 0.27 1.19
AVG. % CHANGE POROUS/TIGHT SAND 21.45 19.44 49.27 38.51 37.48 2.03 5.32 11.30
(Vp,Vs in m/sec, density in gm/cc, impedance moduli λρ, μρ in GPa.gm/cc, Bulk Modulus in GPa.).
Table 2 compares log derived values for typical Glauconite incised valley and regional Ostracod facies in the
WCSB. The improved discrimination between highlighted porous versus tight sand shown by impedance
attributes of λρ (49%) and μρ (39%) is significant by comparison to Poisson’s ratio of only 5% and (Vp/Vs) of
2%. Note that P impedance has a reasonable 21.5% change between the same channel sand facies and this is
routinely exploited by standard interpretation of stacked seismic amplitudes. However P impedance alone shows
a small 13.4% discrimination between porous channel sand and channel shale plug, whereas λρ has a 38%
discrimination. Following the moduli comparisons in the previous section, further confusion between positive
and negative moduli and ratios can be seen in table 2, highlighted in green for the Ostracod shale. The log
measurements give a negative Lambda and Poisson ratio but positive bulk modulus. Why is this so inconsistent?
On the one hand Lambda and Poisson’s ratio suggest an impossible material where a compressive longitudinal
stress results in longitudinal and lateral contraction in strain. By contrast there is nothing too unusual for this
shale with positive bulk (and Young’s) modulus, suggesting that a normal triaxially compressive stress would
result in a volumetric contraction. The standard argument for the positive bulk modulus is that for negative
Lambda a material offers resistance to triaxial compression as a function of its rigidity Mu, but would still have
a permissible negative Poisson ratio causing a lateral contraction for uniaxial longitudinal compression in an
unbound state. The implication is to believe the moduli over the ratio, either the positive bulk or Young’s
moduli. However by considering Lambda as a real and fundamental modulus for incompressibility, these
conflicting values for both bulk and Young’s moduli must be just mathematical inconsistencies as a
consequence of the elastic experiment used to obtain them resulting in mixed and confusing combinations of
Lambda and Mu.

7
To emphasise this point, if one considers the hydrostatic case for a fluid where the bulk modulus (κ) is most
applicable (Grant and West 1965 “Interpretation Theory in Applied Geophysics” p25-26) with no rigidity (μ =
0), then the argument for a negative Lambda which would give a negative bulk modulus and positive Poisson’s
ratio (σ = 0.5) would be considered impossible and there is none of the confusion associated with a solid
(Feynman 1963 “Lectures on Physics” Vol.2, p38-3). Figure 3a, also shows that as the value of negative
Poisson’s ratio approach -1 and λ/μ ⇒ -2/3 while κ/μ ⇒ 0 remains permissibly positive, results in Vp/Vs ratios
that approach unity (Vp/Vs ⇒ 1.155 or √4/3). This is not consistent with a general seismological observation
that at any scale of measurement seismic shear waves do not arrive together with compressional waves i.e. S-
waves never have nearly the same velocity as P-waves.
Further arguments against negative or zero Lambda (λ) or Poisson’s ratio (σ) can be made by considering figure
3b as follows. Imagine a starting volume of a hypothetical fluid with the same 10 G.Pa. value for λ and κ as μ =
0 i.e. no solid matrix. Next add a solid matrix with increasing rigidity μ while maintaining a constant κ
(direction of red arrow in fig. 3b). At a value of μ = 15 G.PA., λ has dropped from the fluid staring point of 10
G.Pa. to zero! This is obviously impossible as by adding an increasing solid matrix rigidity with inherent
additional incompressibility to the initial fluid, cannot reduce its original incompressibility of 10 G.Pa. Instead
the path that makes more physical sense for an addition of increasing matrix rigidity and incompressibility, is
shown by the yellow arrow. In this case there is an increase in bulk modulus to 20 G.Pa. for μ = 15 G.Pa. for a
constant Lambda value of 10 G.Pa. that can thought of as maintaining the original fluid’s incompressibility.
For permissible ranges of Mu (μ) > 0 and bulk modulus kappa (κ) > 0 in figure 3b, a line of constant negative
Lambda (e.g. λ = -10) with increasing μ and κ, crosses radial lines of constant σ in an increasing direction. This
is inconsistent with the normal expectation for a line of constant Lambda (e.g. λ = +10) with increasing μ and κ
that crosses radial lines of constant σ in a decreasing direction. The latter normal case has the correct physical
sense in that a constant incompressibility with increasing rigidity implies a decrease of σ and hence Vp/Vs ratio.

-1.00
Fig. 3b, -0.90
Kappa vs Mu with lines of constant Poisson Ratio (radial) and constant Lambda (parallel slopes) -0.80
50 -0.70

constant Poisson Ratio


-0.60
-0.50
0
-1

45 -0.40
=

-0.30
+1
=
λ
t.

-0.20
=
ns

se o =

40
co

-0.10
t.
ea ti
cr ra

ns
ng

0.00
co
in n
lo

0.05
μ sso
se e a

35 0.10
on
i
ea s

an Po
cr ea

al

0.15
κ nt
in cr

se e
d

0.20
ea as
th ta
μ n

wi ons
d io i

30
cr re

0.25
in c
an rat

0.30
C

e
Mu (G.Pa.)

d io d
th o n

0.35
an at
wi oiss

25 0.40
μ
r
κ

th o n

0.45
P

wi iss

0.50
κ
Po

20 -10
constant Lambda G.Pa.

0
10
15 20
30
40
10
50
60
70
5
80
90
100
0
110
0.0 5.0 10.0 15.0 20.0 25.0 30.0 35.0 40.0 45.0 50.0
120
fluid, no solid matrix Kappa (G.Pa.) 130
140

8
From these theoretical arguments Lambda is a better fundamental measure of a material’s incompressibility and
the results for the Ostracod shale in table 2, are more a function of erroneously low Vp sonic log measurements
in this shale.
Continuing this log based analysis of various moduli and AVO attribute sensitivity, the following graphed
figures 4 and 5, compare Castagna and Smith’s (1994) published world-wide logs plotted as P impedance versus
S impedance compared to the equivalent LambdaRho versus MuRho cross-plot.

Fig. 4, P Impedance vs S Impedance


(Castagna & Smith Worldwide Log Data 1994)

10
9
S Impedance (km/sec.gm/cc)

7
6 Gas Sands
5 Shales
4 Brine Sands

3
2
1
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14
P Impedance (km/sec.gm/cc)

Fig. 5, LambdaRho vs MuRho (Castagna and Smith Worldwide Log Data 1994)

80
75
70
65
60
MuRho (Gpa.gm/cc)

55
-ve Poisson
50 Gas Sands
+ve Bulk
45
+ve Young's Shales
40
35 Brine Sands
30
25
20
15
10
5
0
-10 -5 0 5 10 15 20 25 30 35 40 45 50 55
+ve Poisson
LambdaRho (GPa.gm/cc)
-ve Bulk, -ve Young's

A number of interesting observations can be made by comparing the cross-plots:


1) There is clearly a better separation of the gas sands from brine sands and shales in the λρ, μρ cross-plot
as all the space is now available for lithology and fluid characterisation. This is not the case for the

9
equivalent impedance cross-plot that shows a strong narrow correlation between Ip and Is for all rock
types giving rise to relationships such as Castagna’s “Mudrock Line” (dark lines show cut-offs for gas
sand separation on both cross-plots). The reason for this strong Ip to Is correlation stems from the same
ambiguity in the Vp, Vs velocity relationships that share rigidity Mu as mentioned above. By contrast the
λρ, μρ values are fundamentally more orthogonal giving rise to both tight clusters of similar lithologies
(lower left quadrant gas sands as circled) as well as distinctly separated linear relationships for
background brine sands and shales (dashed orange line).
2) There is a more intuitive interpretation of these rock types in λρ, μρ cross-plot space, where the circled
low μρ gas sands are most likely younger, less consolidated sands than the three gas sand values that
appear well separated in the upper left quadrant (high μρ range). However, once again these three
isolated points have either zero or negative Lambdas and Poisson ratios, but positive bulk and Young’s
moduli. So the values could be used for modelling by Gassmann fluid substitution without any thought to
the possibility that the logs are in error. Even further confusion would arise if any of the low μρ gas
sands were to fall in the negative Lambda region shown by the lower left orange triangle, then Poisson’s
ratio would be positive and believable while the bulk modulus, Young’s modulus and Lambda would be
negative. My conclusion from the λρ, μρ cross-plot mapping of the confusing positive to negative non-
linear behaviour for bulk, Young’s moduli and Poisson’s ratio within the clearly consistent negative
Lambda region, is that Lambda alone represents the material’s true incompressibility and the other
parameters disguise measurement errors.
Following this cross-plot analysis for the same log data set, figure 6 below, shows an interesting cross-plot
(derived by T. Chen, W. Zhang and myself) of the λ/μ ratio versus the λρ-μρ difference. The two green
triangles that meet at the (1,0) ‘origin’ show that the gas sands align within the lower left triangle while the
brine sands tend to align in the upper right triangle. This alignment into lithologic “sandstone” lines with
linearly varying fluid properties is explained in the next section on interpretation templating of Lamé
attributes from log analysis from the Western Canadian Sedimentary Basin.

Fig. 6, LambdaRho-MuRho Difference vs Lambda/Mu Ratio


(Castagna and Smith Worldwide Log Data 1994)

10.0

8.0
Lambda/Mu Ratio

6.0
Gas Sands
Shales
Brine Sands
4.0

2.0

0.0
-10 -6 -2 2 6 10 14 18 22 26 30 34
LambdaRho-MuRho Difference

Log analysis motivation and interpretation template for Lamé parameters λ, μ, λρ and μρ.
Figures 7a and 7b, show a gas well sonic and dipole logs in depth where a conventional impedance Ip versus Is
analysis has limitations in clearly discriminating between all the various lithologies and gas sand zones (e.g.
only gas zone A is discernible from a background cut-off on the Vp/Vs curve in fig.7b). Because Ip and Is
share both rigidity and density the log curves in figure 7a, tend to track each other and never crossover, with

10
only the lowest impedance shale or highest impedance carbonate zones being clearly distinguishable. By
contrast the λρ and μρ curves (fig.7a) have similar value ranges, that do crossover with λρ < μρ for gas zones.
Beyond this crossover there is the benefit of immediate diagnostic information for gas zone A showing a better
poro-perm than zone B due to the larger curve separation in zone A. Conversely the λρ > μρ crossover is an
excellent indicator of thin, tight shale breaks as seen separating gas zone A from B, as well as the capping
shale above gas zone B. Figure 7b, compares just the λρ with Vp/Vs log tracks for gas sand detection
sensitivity. Cut-offs for both log tracks were chosen based on the lowest values for the overlying background
shales and silts. The upper gas zone B is clearly distinguished on λρ, unlike Vp/Vs, with λρ in agreement with
the correct gamma log depth for the gas sand.
In general seismic AVO analysis involves some form of cross-plotting such as reflectivity based
intercept/pseudo-Poisson ratio (Verm and Hilterman 1995) or intercept/gradient (Castagna and Swan 1997).
Further log analysis for AVO interpretation has lead to a “fluid factor” separation (Smith & Gidlow 1987)
from cross-plots of reflectivity Rp versus Rs or Ip versus Is. Figure 8 compares the Ip, Is to λρ, μρ cross-plots
for the gas sand log tracks in figure 7, with λρ, μρ showing a significant advantage in isolating both lithologic
properties, e.g. sand, shale, and carbonate facies, as well as the gas zone’s fluid effects. The Ip, Is points in
figure 8, cluster in a tight linear relationship (i.e. Castagna’s “Mudrock Line”) with shale having the lowest
values on both axes. By contrast the lowest λρ values in the λρ, μρ cross-plot, discriminate the best gas sand
zone A, along with higher μρ values than the background shales. Simply put, for the Ip, Is cross-plot all rock
types plot to the upper right direction from the lowest shale values. Conversely for the λρ, μρ cross-plot, the
anomalous gas sands plot in the upper left hand quadrant from the lowest μρ shales, while other more
competent or tight lithologies (silts, cemented sands) plot in the opposed, upper right quadrant relative to the
shales. As mentioned before, the reason for the separation improvement in the λρ, μρ cross-plot compared to
Ip, Is, is that the λρ versus μρ axes are orthogonal with regard to Lamé parameters or moduli, unlike Ip versus
Is, thereby making the cross-plot more discriminating.

Figure 7a. Lower Cretaceous gas sand well logs from Alberta.
g
P.S Impedance LambdaRho, MuRho
metres metres

1240 1240 Silt

1260 1260
Shale Shale
Gas Sand B
1280 1280 Tight Streak

Gas Sand A
Carbonates Carbonates
1300 1300

1320 1320
160
0

40

120
2000

6000

10000

80
14000

18000

MuRho
S Impedance
P Impedance LambdaRho

11
Figure 7b. Comparison Between LambdaRho vs
Vp/Vs
Vp/Vs for Lower Cretaceous Alberta Log Data
Cutoff Cutoff
Gas Sand Shale Gas Sand Shale
1.4 1.6 1.8 2 2.2 0.00 5.00 10.00 15.00
1220

1230

1240

1250 Silt
Depth m

1260
Shale
Shale
1270 Shale
Gas Gas Sand Zone B
1280 Gas Sand Tight Shale Break
Sand
1290 Gas Sand Zone A
1300
Carbonates

1310 Lambda-rho

1320 Vp/Vs
2005031a-3

Figure 8. Cross-plot interpretation template;


P versus S impedance compared to LambdaRho versus MuRho for Alberta gas sand well log in figure 7a.
g
P Impedance v's S crossplot for LambdaRho vs MuRho crossplot of Gas Well Log Data
Gas Well Log Data 80 xx
S Impedance (Vs x Density)

9000 XX Cemented
x
x
Tight x
Cemented X X Carbonates
"MuRho" Rigidity X Density

Gas x
X X Sands
x x
X X
X Carbonates x
8000 Gas Sand X
XTight XX
XXX X Sand x x
x x
x x
x
X XX X
X Sand
X
X 60 x
x x
x x x
x
X
7000 X X X x x
Shaly Gas Sand XX
XX
Silt x x
6000 X Threshold cutoff for xx
Threshold cutoff
Clastics-Carbonates 40 Shaly Gasx x
Sand x Clastics-Carbonates
5000 Shale Legend Legend
Gas Sand Shale Gas Sand
4000 Threshold cutoff for Shaly Gas Sand Threshold cutoff porous Shaly Gas Sand
20
porous Gas Sand Shale Gas Sand Shale
3000
7000 9000 11000 13000 15000 17000 19000
0 20 40 60 80 100 120 140 160
P Impedance (Vp x Density)
"LambdaRho" Incompressibility X Density

Continuing with this log based template interpretation, figure 9 shows a λρ versus μρ cross-plot of logs for a
Mannville gas sand interval with two very different sand porosities (Lithic and Glauconitic) as well as the
background Ostracod shale. There is a clear clustered separation of the gas sands representing a difference in
porosity as shown by the increasing porosity trend toward the origin. A sand/shale lithology change is seen to
occur in the opposed direction normal to the porosity trend thereby clearly isolating the Ostracod shale from
varying sand porosities. These porosity and sand/shale trends would tend to be more confusingly aligned in the
same direction on the equivalent impedance cross-plot of Ip versus Is as compared in figure 8 above.

12
20
Glauconitic Sand
Lithic Sand
Figure 9. 10% Ostracod Shale

i ty
ros
Po
Lambda vs Mu log template

ing
for porosity and shale content

s
rea
Inc
μ GPa. In
10
26% c
re
a
C o s in
nt g S
e n ha
t le

Modified from T. Chen et al 1998

0
0 10 20 30 40
λ GPa. 2005031a-4

A more detailed interpretation template is shown in the following λρ versus μρ cross-plots in figures 10 a and
b. These logs for a shallow Viking sand interval will be used in a following section on a 3D seismic AVO and
VSP case study. Figure 10a, shows a diagnostic separation of lithologies with the marine shales in black (Base
of Fish Scales overlying the Viking sand) clustering with the lowest μρ’s and highest λρ’s.
The terrigenous shales that surround the porous sand zones and the background silt/tight sands show separate
trends dependent on the relative sand/shale content versus porosity (compare to figure 9). The gas to wet sand
separation is primarily a function of the horizontal decrease in Lambda as a fluid effect seen in both figures
10a and b. However figure 10b, shows that a density effect can be discerned by comparing the upper (λρ
versus μρ) to lower (λ versus μ) pair of gas versus wet sand clusters. The increased density trend slopes for the
upper paired clusters (i.e. varying ρ on both axes) are not as obvious as the larger horizontal
Lambda fluid effect for wet to gas sands, as shown by the red arrow.

Fig. 10a, Case Study (VSP, Surface Seismic) Gas Well Logs; LambdaRho vs MuRho
22
21
20
19
18
17
16
15
14
MuRho G.PA. gm/cc

13 Gas Sand Zone


12 Wet Sand Zone
11 Marine Shale above Gas Zone
10 Tight Sandy Shale between Gas & Wet Sand Zones
9 Background Silt/Sand
8
7
6
5
4
3
2
1
0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30
LambdaRho G.Pa. gm/cc

13
Fig. 10b, LambdaRho vs MuRho compared to Lambda vs Mu for Gas and Wet Sand Zones showing
Density effect (black line trends)

20

19

18

17

16

15
MuRho GPa.gm/cc or Mu GPa.

14

13

12

11

10

9 Fluid Lambda effect only


8

0
0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20

LambdaRho GPa.gm/cc or Lambda GPa.

Lambda Mu; Gas Zone Lambda Mu; Wet Zone LambdaRho MuRho; Gas Zone LambdaRho MuRho; Wet Zone

The final portion of this section on cross-plot template interpretation, deals with the additional discrimination
of fluid and lithology effects from the λ/μ ratio versus λρ-μρ difference cross-plot introduced in figure 6. The
template shows that rocks of consistent lithology will tend to align as lines of constant μρ (actually 1/μρ)
radiating through a ratio, difference “origin” of (1,0). The variation along these lines will be primarily driven
by fluid or saturation effects (i.e. decreasing λρ toward and through the (1,0) origin), with porous gas sands
trending to the lower left quadrant from the (1,0) “origin” (as in figure 6 above).
A template created by Dave Mackidd is shown in figure 11, which includes an overlay of curves of constant P-
impedance that are almost orthogonal to the “lithology” lines of constant 1/μρ. These displays have been
termed “ratio difference” LMR cross-plots. Similar cross-plots have been developed coincidently and
independently by Berryman et al (1999, 2002) from a more log based petrophysical background, that have also
successfully demonstrated the improvement in discriminating porosity and saturation effects.

A proof (derived by D. Cooper 1997) of the constant lithology μρ lines can be shown as follows:

The slope of a line through the (1,0) " origin" of figure 11 for two points 1 and 2 on the line is given as
⎛ λ 2 λ1 ⎞
⎜⎜ − ⎟⎟
⎝ μ 2 μ1 ⎠
(λ 2 − μ 2 )ρ 2 − (λ 1 − μ1 )ρ1
assume at least one point on the line (point 1 in the equation) is fixed at the (1,0) " origin"
so at (1,0) λ = μ, i.e. λ / μ = 1 (∴ Vp/Vs = 3 ) and (λ − μ )ρ = 0, then the slope equation reduces to
⎛ λ2 ⎞
⎜⎜ − 1⎟⎟
μ2 λ2 − μ2 1
⇒ ⎝ ⎠ = = making the slopes proportional to the reciprocal of μρ
ρ 2 (λ 2 − μ 2 ) μ 2 ρ 2 (λ 2 − μ 2 ) μ 2 ρ 2

14
Decreasing Ip

1/μρ

Figure 11, Template for λ/μ ratio versus λρ-μρ difference cross-plot with P and S impedance (Ip and Is)
overlay showing radiating lines of constant μρ or Is2 (yellow) crossed by lines of constant Ip (red)
(Generated by D. Mackidd 1997 and introduced by W. Zhang, T. Chen and myself in 1996)

Figures 12a and b, show the interpretation template for this λ/μ ratio versus λρ-μρ difference cross-plot for
logs from a Glauconitic sand interval. The constant 1/μρ line termed a “sandstone line” shows varying gas to
water saturations along its length, with the higher gas saturation through the (1,0) “origin” in the red area. A
detail of this area is shown in figure 12b, where the gas zone is seen to be better isolated beyond the (0.8,-5)
point rather than the (1,0) “origin”.
The variation of the slope beyond the (1,0) origin with increasing gas saturation (black to white dotted lines)
reveals an unexpected effect of the gas on the rigidity or μρ of the sand. A reasonable explanation is that we
are seeing the direct lower density gas effect on μρ in the denominator leading to a slight steepening of the
“sandstone line” slope. This raises the interesting possibility of estimating density as the rigidity μ within μρ
is unaffected and does not vary with fluid saturation (see following section dealing with the Biot-Gassmann
equation).

Further templates showing the relation between gas, wet sand and coals is seen in the pair of λρ versus μρ
and λ/μ ratio versus λρ-μρ difference cross-plots of figure 13 from Y. Li and J. Downton.
A final simple schematic compilation of these various lithologic and fluid/porosity effects is shown in figure
14 as a general cross-plot guide, that is a better representation of how real seismic data might appear given
the impact of S/N errors that tend to expand the cross-plot cloud into the top left quadrant

15
Figure 12a. Lambda/Mu Ratio vs Lambda - Mu Difference
Crossplot from Sonic Logs
2.6

Shale
2.2
Lambda/Mu Ratio

1.8

Sandstone
1.4 Line
Carbonates

Wet Sand
1.0

Porous
0.6 Gas Sand

-20 -10 0 10 20 30 40 50 60
Lambda - Mu Difference from W. Zhang et al 1996
2005031a-5

9801025-04

1.40
Figure 12b.

Lambda/Mu ratio vs 1.20


Lambda-Mu difference
log template
for fluid discrimination 1.00

λ
μ
0.80

0.60

Glauconitic Gas Sand


Modified from T. Chen et al 1998 Glauconitic Wet Sand
0.40
-15 -10 -5 0 5 10
λ-μ GPa.
2005031a-6

16
Figure 13. Glauconitic Facies Gas Sand Log Data

LambdaRho vs. MuRho LambdaRho vs. Lambda/Mu Ratio Gamma


96.7 6.89
158

84
78.6 5.53
Mu*Rho (GPa.g/cc)

Lambda/Mu ratio
45
san
et
d

60.5 4.16
an

24
W
ss
Ga

13

als
42.4 2.80

Co
7

24.4 1.43 3.5

2
Coal and shale Wet sand
6.30 0.07 Gas sand 1
0.51 33.19 65.88 98.57 131.25 0.51 33.19 65.88 98.57 131.25
LambdaRho (GPa.g/cc) LambdaRho (GPa.g/cc)

from Yongyi Li and Jon Downton


2005031a-7

Figure 14. LambdaRho-MuRho Interpretation Template Crossplot Guide

s
Ga
1

125 1
2:
1:
DO
PO OM

LI
TI STO

100
M
RO ITE
L

GH N
E
.gm/cc
Gpa.gm/cc

US

T E
MuRho Gpa

s
50 as We
t
lcareou s
G Ca Shale
SAND SHALE
tic
Clas s
le
Sha
0
0 50 100 150 200 250
GPa.gm/cc
LambdaRho GPa.gm/cc
2005031a-8

17
Biot-Gassmann equation expressed in Lamé parameter terms.
The following common Biot (1956), Gassmann (1951) and Geertsma (1961) equations for fluid substitution
and modelling are transformed into simpler Lamé parameter terms. This result is used as further motivation
for the potential to directly extract meaningful pore fluid detection attributes such as Δλ and Δλρ, to be
shown in the next section on AVO reflectivity equations and Lamé parameterisation.
The last approximate expression, equation 20, for Δλ as a fluid attribute shows that fluid detection is directly
proportional to the fluid λ as well as the porosity effect in ΔK2 and inversely proportional to the
porosity scaled solid rock bulk modulus squared.
2
⎛ K ⎞
⎜1 − dry ⎟

⎝ K solid ⎟⎠
K sat = K dry + is the Biot - Gassmann equation
⎛ K dry ⎞ ⎛ φ ⎞
⎜1 − φ − ⎟(K solid ) + ⎜
−1
⎟⎟
⎜ K solid ⎟⎠ ⎜K
⎝ ⎝ fluid ⎠
where K is Bulk modulus and " sat" is saturated rock with φ as porosity.
Lame ′ form of Biot - Gassmann equation assuming μ dry = μ sat
λfluid (K solid − K dry )2
λsat = λdry + (18)
λfluid (K solid − K dry ) + φ K solid (K solid − λ fluid )
for detection of changes in fluids where Δλ = λsat − λ dry and substituting ΔK = K solid − K dry
λfluid ΔK 2
⇒ Δλ = (19)
λ fluid ΔK + φ K solid (K solid − λfluid )
2
with a useful approximation based on K solid >> λ fluid and φ K solid >> λ fluid ΔK
λ fluid ΔK 2
⇒ Δλ ≈ (20)
φ K solid 2
As a check on this relationship consider the limits:
1) As porosity φ → 0 so ΔK → 0 therefore Δλ → 0 as expected as λsat = λdry.
2) as porosity φ → 1 so ΔK = Ksolid as Kdry → λdry = λair = 0 therefore Δλ → λsat = λfluid as expected for
100% porosity.
Finally one prediction from this form of the Biot-Gassmann-Geertsma equation, is that Δλ fluid detection is
enhanced for rocks with lower bulk moduli (denominator) and higher porosities (numerator ΔK2) such as
poorly consolidated sands, thereby confirming actual observation (see fluid substitution in figure below).

U. Cretaceous Sandstones;Gassmann substitution showing enhanced LambdaRho,


MuRho separation in Fluids and Rock Properties compared to P and S Impedances

18
Angle dependent reflectivity equations and approximations for AVO seismic or VSP analysis
This section will compare two basic approaches to extracting AVO attributes, one for P-wave and pseudo-
Poisson reflectivity, the other for P and S reflectivities (Rp(0°) and Rs(0°)) which can be inverted into λρ
and μρ. Some AVO inversion schemes incorporate an explicit density term (Stewart et al. 1995, Smith 1996)
to potentially extract moduli and density separately. However as the number of unknowns increase and
exceed the measurable CMP gather quantities (intercept and gradient) so these more complex equations are
less robust and the extracted values more ambiguous. This is especially true in the presence of noise where
even two attributes such as Rp, Rs are in doubt (Cambois 2000).
A common starting point for AVO analysis is the tractably linearized 3 term approximation to the Knott-
Zoeppritz equations given by Aki and Richards (1980) shown below with an explanation of assumptions.
These form the starting point for yet further 2 term approximations used to extract Rp(0°) and Rs(0°) or
pseudo-Poisson reflectivity. The basic linearized assumptions in the Aki and Richards approximation are
small fractional velocity and density changes with 2nd order terms ignored and θp < 10 degrees of critical.

1 ⎛ ΔVp Δρ ⎞ Vs 2 ⎛ ΔVs Δρ ⎞ 1 ΔVp


Rpp(θ ) = ⎜⎜ + ⎟⎟ − 2 2 sin 2 θ p ⎜⎜ 2 + ⎟⎟ + tan 2 θ p
2 ⎝ Vp ρ ⎠ Vp ⎝ Vs ρ ⎠ 2 Vp
ΔIp 1 ⎛ ΔVp Δρ ⎞ ΔIs 1 ⎛ ΔVs Δρ ⎞
with = Rpp(0) = ⎜⎜ + ⎟⎟ and = Rss(0) = ⎜⎜ + ⎟
2Ip 2 ⎝ Vp ρ ⎠ 2Is 2 ⎝ Vs ρ ⎟⎠

where: Vp, Vs velocities, ρ density, are averaged across an interface and angle θp (or just θ) is the average of
incident and transmitted P-wave. ΔVp/Vp, ΔVs/Vs, Δρ/ρ are fractional changes in Vp, Vs and ρ across an
interface and can be considered equivalent to ΔlnVp, ΔlnVs and Δlnρ.
The first common industry method considered is a P-P AVO extraction of zero angle Rp(0) and Rs(0)
“seismic traces” through “weighted stacking” of CMP gather data (Gidlow et. al.1992, Fatti et. al.1994).
Starting from the Aki & Richards equation above with some algebraic manipulation gives;
2 ⎛ 2 ⎞
ΔIp ⎛ Vs ⎞ ΔIs ⎜ 1 2 ⎛ Vs ⎞ 2 ⎟ Δρ
Rpp(θ ) = (1 + tan 2 θ ) − 8⎜⎜ ⎟ sin 2 θ − ⎜ tan θ − 2⎜ ⎟ sin θ ⎟ (a)
⎟ ⎜ Vp ⎟
2Ip ⎝ Vp ⎠ 2Is ⎜ 2 ⎝ ⎠ ⎟ ρ
⎝ ⎠
Equation (a) the “Geogain Equation” from Gidlow et. al., is solved in a least squares sense to extract Rp(0)
and Rs(0) by assuming the 3rd term cancels for small density contrasts (Δρ/ρ), as well as for small angles i.e.
tan2θp = sin2θp, and (Vs/Vp) = 1/2. If the density contrast is not small then this third “error” term can be
significant at large angles and more importantly inconsistent for varying rock properties by being dependent
on both angle and Vp/Vs ratio (see graph of angle dependent error below).
Difference and Error % between Tan^2 and Sin^2

0.90 120.00
0.80
Tan^2 or Sin^2 value

100.00
0.70
Tan^2 and Sin^2
Error % between

0.60 80.00
0.50
60.00
0.40
0.30 40.00
0.20
20.00
0.10
0.00 0.00
4 8 12 16 20 24 28 32 36 40 44 48 52
Angle

sin. sqr tan. sqr error %

19
For Vp/Vs ratios < 2 the error between the approximate 2 term Fatti et. al. equation and the exact Aki &
Richards 3 term curve is small and evenly distributed, but still angle dependent. However for ratios > 2.5 or 3,
this error increases with increasing angle and because the error term is strongly dependent on both angle and
Vp/Vs ratio this restricts the useable range of angles.
In practice, however, this equation is very useful and can be used to fairly large angles (see following AVO
model curve comparisons figures 15a, b) as Δρ/ρ has the smallest variation compared to ΔVp/Vp and ΔVs/Vs,
seen in Gardner’s (1974) relationship as Δρ/ρ ≈ (ΔVp/Vp)/4. If the angle range is restricted to a commonly
quoted 30° due to the ignored error term, then the catch-22 problem is that the same large angle restriction
reduces the separability of sin and tan curves involved in the remaining first two terms of equation (a). If this
angle restriction can be reduced then more robust and independent estimates of ΔIp/Ip and ΔIs/Is are possible
in theory.
The situation is much worse for the second industry method based on a 2 term approximation by Shuey
(1984). The formulation is significantly more complex than the Fatti et. al. equation above and is just
reproduced here for comparison.

⎛ Δσ ⎞ 2
Rpp(θ ) = Rpp(0) + ⎜⎜ ARpp(0) + (
⎟ sin θ + tan 2 θ − sin 2 θ
2 ⎟
)ΔVp
⎝ (1 − σ ) ⎠ 2Vp
(b)
⎛ 1 − 2σ ⎞ ΔVp ⎛ ΔVp Δρ ⎞
where σ = Poisson' s ratio and A = B - 2(1 + B)⎜ ⎟ where B = ⎜⎜ + ⎟
⎝ 1−σ ⎠ Vp ⎝ Vp ρ ⎟⎠

The 2 term approximation for the Shuey equation (b) is to drop the 3rd term in (ΔVp/Vp) and this is where the
infamous 30° angle restriction comes from, because dropping a term involving (ΔVp/Vp) is significantly
worse than a Δρ/ρ term in the Fatti et. al. equation (a) (see Gardner’s relationship above).However a different
approximation with no angle dependent error can be obtained by substituting the following relationships into
the original Aki & Richards equation;

ΔIs 1 ⎛ Δμ Δρ ⎞ ΔVp 1 ⎛ Δ(λ + 2μ ) Δρ ⎞ ΔVs 1 ⎛ Δμ Δρ ⎞


= ⎜⎜ + ⎟⎟ , = ⎜⎜ − ⎟⎟ and = ⎜ − ⎟
2Is 4 ⎝ μ ρ ⎠ 2Vp 4 ⎝ (λ + 2 μ ) ρ ⎠ 2Vs 4 ⎜⎝ μ ρ ⎟⎠

2
Δρ ΔVp ⎛ Vs ⎞
⎟ sin 2 θ
Δμ
=> Rpp(θ ) = + (1 + tan 2 θ ) − 4⎜⎜ ⎟ (c)
2ρ 2Vp ⎝ Vp ⎠ 2μ

Now the approximation in this equation (c) (Goodway 1997), is to drop the small zero offset term Δρ/2ρ
(again from Gardner’s relationship Δρ/2ρ ≈ (ΔVp/Vp)/8). Note that this error is independent of angle unlike
the Fatti et. al. or Shuey 2 term approximations (i.e. equations (a) and (b) without 3rd terms) and hence allows
a more accurate fit to the exact Aki & Richards AVO curve at large angles, but has a “bulk” Δρ/2ρ scalar shift
at all angles.
These three 2 term approximations (equations a, b and c) are compared for accuracy in matching the Aki and
Richards 3 term equation as graphed against incident angle in figures 15a, and b, below. Notice that the Shuey
2 term approximation has significant error past 30° and gives rise to the commonly accepted angle restriction
as mentioned above. By contrast the other 2 term approximations allow for the extraction of elastic and
possibly density information, in the AVO curve shape for angles well beyond 30°. (see equation (h) below and
section following on Lamé Elastic Impedance).

20
Fig.15a, P-P refelctivity vs angle from 3 term Aki & Richard's vs 2 term
equations: Fatti et al '94, and Goodway'97 "Mu, Vp" compared to 2 term
Shuey (Model from Gulf of Mexico logs)

-0.05
P-P reflection coefficient

-0.1

-0.15

-0.2

-0.25

-0.3
0 5 9 13 17 21 25 29 33 37 41 45 49 53 57 61 65
Incident Angle

Aki & Richards reflectivity Rp Goodway 2 term"Mu,Vp eqn"

Rp "Fatti et al" 2 term sin=tan for dRho/Rho Shuey 2 term sin=tan for dVp/Vp

Fig. 15b, Rp reflectivity error vs angle; 3 term Aki&Richards minus 2 term


approximations: Fatti et al '94, & Goodway '97 "Mu, Vp" compared to Shuey
(Gulf of Mexico Impedance Model)

0.05
Reflection coefficient error

-0.05
0 4 8 12 16 20 24 28 32 36 40 44 48 52 56 60 64
-0.1

-0.15

-0.2
Angle
A&R-Goodway "Mu,Vp" 2 term difference A&R-Fatti et al 2 term sin=tan for dRho/Rho

A&R-Shuey 2 term sin=tan for dVp/Vp difference

Yet further interesting reformulations of the original Aki & Richards AVO equation can be obtained in terms
of linear cos2θ only. A simplified AVO attribute extraction or analysis method using the FFT of equation (d)
below in offset variable ‘θ’, gives very specific ΔVp/Vp and Δμ/μ scaled spectral notches/peaks as functions
of cos2θ only;
2
Δρ 1 ΔVp ⎛ Vs ⎞ Δμ
R (θ ) = + −⎜ ⎟ (1 − cos 2θ ) (d)

2 ρ (1 + cos 2θ ) Vp ⎝ Vp ⎠ ⎟ μ
and in Lamé moduli, density terms the Aki-Richards equation simplifies to:
2
1 Δ (λ + 2 μ ) ⎛ Vs ⎞ Δμ 1 Δρ
R (θ ) = (1 + tan 2 θ ) − 2 sin 2 θ ⎜⎜ ⎟
⎟ + (1 − tan 2 θ ) (e)
4 (λ + 2 μ ) ⎝ Vp ⎠ μ 4 ρ
Note there is no density influence on R(θ) at 45° i.e. R(θ) is just a function of moduli as a ratio that is almost
Δ(Vp/Vs)2/(Vp/Vs)2. This concurs with Hilterman’s work on far offset reflectivity (shown in an excellent book
on AVO accompanying the SEG DISC 2001).

21
Next this Lamé parameterised equation can also be represented in cos2θ terms:
2
1 Δ (λ + 2 μ ) ⎛ Vs ⎞ Δμ cos 2θ Δρ
R (θ ) = − (1 − cos 2θ )⎜ ⎟ + (f)
2(1 + cos 2θ ) (λ + 2μ ) ⎜ ⎟ 2(1 + cos 2θ ) ρ
⎝ Vp ⎠ μ

that can be further simplified to;

Δ(λρ ) Δμ Δρ
R (θ )4 cos 2 θ = + 2 cos 2 2θ − 2 sin 2 θ (g)
(λ + 2 μ ) ρ (λ + 2 μ ) ρ

or more usefully as a quadratic equation in terms of cos2θ:


Δλ Δρ Δμ
R (θ )2(1 + cos 2θ ) = + cos 2θ + 2 cos 2 2θ
(λ + 2 μ ) ρ (λ + 2 μ )
(h)
Δλ ⎛ Δρ Δμ ⎞
⇒ R (θ )2(1 + cos 2θ ) = + cos 2θ ⎜⎜ + 2 cos 2θ ⎟
(λ + 2 μ ) ⎝ ρ (λ + 2μ ) ⎟⎠

This final result shows a similar though more interesting insight, to equation (e) above. The insight here is that
at θ = 45° cos2θ = 0 and so the reflectivity R(45) is just a scaled version of the change in the basic fluid
indicator Δλ derived as follows (see motivating Lamé Biot-Gassmann equations in previous section).
Δλ
2R (45) = where λ + 2μ = Ip.Vp (product of averaged P impedance and velocity)
(λ + 2 μ )
ΔIp Δ ln(Ip)
and average Ip can be obtained following the basic definitions above for R (0) = =
2Ip 2
R (0) Δ ln(Ip)
Consider zero angle R (0) as a function of time τ : R 0 (τ ) = lim ⇒ R 0 (τ ) =
Δτ →0 Δτ 2Δτ
⇒ 2∑ R 0 (τ )Δτ = ln(Ip) exponentiating gives ⇒ e 2 ∑ R 0 (τ ) Δτ
= Ip
Average Vp or V(τ ) INT can be obtained from the standard Dix VRMS NMO stacking velocity equation;
2
τ2 d(VRMS )
T0 VRMS = ∫ V(τ ) INT dτ ⇒ V(τ ) INT = T0
2 2

τ1 dτ
2
d(VRMS )
Combining all these results gives; ⇒ R 45 (τ )e 2 ∑ R 0 (τ ) Δτ T0 = Δλ (τ )

Finally the Fatti et. al. 2 term approximation (equation a, above) can be formulated in Lamé terms
(from papers by Smith 1999, Smith and Gidlow 2000)

Rpp(θ )[4(λ + 2μ ) ρ ] = Δ(λρ )(1 + tan 2 θ ) + 2Δ( μρ )(1 + tan 2 θ − 4 sin 2 θ )

Note that there is no Vp/Vs term and on the left hand side Rp(θ) is just scaled by the P impedance squared to
give direct estimates of changes in λρ and μρ.

22
Lamé Elastic Impedance.
Following Connolly’s approach (1999), a Lamé formulation for Elastic Impedance reveals a more intuitive and
useful set of attributes from a combination of angle based reflectivity. Starting from the original Aki and
Richards (1980) linearized approximation to the Zoeppritz equations:

1 ⎛ ΔVp Δρ ⎞ Vs 2 ⎛ ΔVs Δρ ⎞ 1 ΔVp


Rpp(θ ) = ⎜⎜ + ⎟⎟ − 2 sin 2 θp⎜⎜ 2 + ⎟⎟ + tan 2 θp
2 ⎝ Vp ρ ⎠ Vp 2
⎝ Vs ρ ⎠ 2 Vp
Substituting the following Lamé formulations;
1 ⎛ ΔVp Δρ ⎞ 1 ⎛ 1 ⎡ Δ (λ + 2μ ) Δρ ⎤ ⎞
Rpp(0) = ⎜ + ⎟= ⎜ + ⎟
2 ⎜⎝ Vp ρ ⎟⎠ 2 ⎜⎝ 2 ⎢⎣ λ + 2μ ρ ⎥⎦ ⎟⎠
ΔVp 1 ⎡ Δ (λ + 2 μ ) Δρ ⎤ ΔVs 1 ⎡ Δμ Δρ ⎤ Vs 2
and = ⎢ − = ⎢ − as well as =γ
Vp 2 ⎣ λ + 2μ ρ ⎥⎦ Vs 2 ⎣ μ ρ ⎥⎦ Vp 2
1 ⎛ 1 ⎡ Δ (λ + 2μ ) Δρ ⎤ ⎞ ⎛ Δμ ⎞ 1 ⎡ Δ (λ + 2 μ ) Δρ ⎤
⇒ Rpp(θ ) = ⎜ ⎢ + ⎥ ⎟ − 2γ sin 2 θp⎜ ⎟⎟ + tan 2 θp ⎢ −
2 ⎜⎝ 2 ⎣ λ + 2 μ ρ ⎦ ⎟⎠ ⎜
⎝ μ ⎠ 4 ⎣ λ + 2μ ρ ⎥⎦
1 ΔIpθ 1 1⎛1 1 ⎞
⇒ Rpp(θ ) = = Δ ln Ipθ = ⎜ (1 + tan 2 θp)Δ ln(λ + 2 μ ) − 4γ sin 2 θpΔ ln μ + (1 − tan 2 θp)Δ ln ρ ⎟
2 Ipθ 2 2⎝2 2 ⎠
Take angle and (Vs/Vp) 2 = γ terms inside fractional natural logs (Δ ln of (λ + 2μ ), μ and ρ )
1 1
( 2 2 2
⇒ Δ ln Ipθ = Δ ln (λ + 2μ ) 0.5(1+ tan θp ) μ − 4γ sin θp ρ 0.5(1− tan θp )
2 2
)
Next integrate to remove Δ ' s and eponentiate to remove natural logs ln' s ;
2 2 2
θp )
⇒ Ipθ = (λ + 2μ ) 0.5(1+ tan μ − 4γ sin θp
ρ 0.5(1− tan θp )
= Lamé Impedance (LI)
So at 0, 30,45 and 60 degrees;
LI(0) = (λ + 2μ ) 0.5 ρ 0.5
LI(30) = (λ + 2 μ ) 2 / 3 μ −γ ρ 1 / 3
LI(45) = (λ + 2μ ) μ − 2γ
LI(60) = (λ + 2μ ) 2 μ −3γ ρ −1
Assuming a 60 degree refelctivity is meaningful, these angular LI' s can be combined
to give isolated estimates of density, Lamé moduli and Vp/Vs ratio directly as;
LI(30) 3 LI(0) 2 LI(60) 3γ LI(0) 2 LI(60) Vp 2
= ρ , = μ , = ( λ + 2 μ ) 1.5
, LI(0) 2/3
LI( 60 ) 1/ 3
=
LI(0) 2/3 LI(45)LI(60)1 / 3 LI(45) 3 LI(45)1.5 Vs 2γ

From these results, the two interesting observations (beyond those made by Connolly 1999) are;
a) density ρ is independent of the (Vs/Vp)2 ratio or γ at all angles, only rigidity (μ) is affected by γ
b) density ρ disappears in the reflection coefficient or LI only at 45°
The direct estimation of density given in the first equation above (combining various (power scaled) Lamé
Impedances inverted from AVO incident angles 0°, 30°, 45° and 60°) is shown in the 1st graph below for a
BQ gas sand log. The 2nd graph below of density vs. LambdaRho for the same log, shows that a cut-off of
around 37 G.Pa.gm/cc for LambdaRho (red line) is a better indication of the upper and lower BQ channel gas
zones than density, because the lowest density values occur above the gas zones in shales and coals. These
non-pay low densities create ambiguous anomalies in stacked amplitudes and simple post-stack inversions
(hence the need for LMR as exploited by H.Mandler; PanCanadian internal TEK conference 2001).

23
S c a le d lo g b a s e d L a m e ' Im p e d a n c e s a t in c id e n t a n g le s (0 , 3 0 , 4 5 a n d 6 0 ) c o m b in e d to
g iv e D e n s ity c o m p a r e d to a c tu a l lo g D e n s ity (p lo tte d in d e p th )

100000000

10000000
Lame' Impedances (various angles) or

1000000
Density gm/cc (log scale)

100000 L I( 4 5 )
L I( 0 ) ^ 0 .6 6
L I( 3 0 ) ^ 3
10000
L I( 6 0 ) ^ 0 .3 3
R h o fr o m L I( 0 ,3 0 ,4 5 ,6 0 )
1000 A c tu a l R h o fr o m L o g

100

10

1
2030

2034

2039

2043

2047

2051

2055

2059

2063

2067

2071

2075

2079

2083

2087

2091

2095

2099

2103

2107

2111

2115

2119
D e p th (m e tre s )

Density compared to LambdaRho for gas sand vs shale discrimination

2.9 170
160
2.8 150
140
2.7 130
LambdaRho G.Pa x gm/cc.

120
2.6 110
Density gm/cc

100
2.5
90 Rho from log
80 LambdaRho
2.4
70
2.3 60
50
2.2 40
30
2.1 20
10
2 0
2030
2034
2039
2043
2047
2051
2055
2059
2063
2067
2071
2075
2079
2083
2087
2091
2095
2099
2103
2107
2111
2115
2119

Depth (metres)

24
Another interesting observation is to compare LI(30) to LI(45);

LI(45) 2/3
= (μ γ ρ) −1 / 3
LI(30)

From this it can be seen that the ratio between the 2/3 power of LI(45°) and LI(30°) impedances, is
significantly and abruptly enhanced for low rigidity μ (hence low Vs/Vp) shales and coals, as well as
carbonates with low Vs/Vp ratios, by the large 1/(μγ/3) factor (i.e. amplification of low rigidity μ by the
power of low Vs/Vp ratios that are yet further amplified by squaring these ratios and dividing by 3).
This spectacular result can be seen on the following log tracks for an Alberta BQ gas well in the very low
rigidity shale (2063m to 2070m) capping the gas sand zone shown by the LambdaRho and MuRho cross-
over between 2070m and 2094m and the low Vs/Vp carbonate below 2094m.

Fig. 16, LambdaRho, MuRho, and Lame' Impedance Ratio


LI(45) to LI(30) plot vs depth for Alberta gas well
100
90
AVO attributes as marked

80
70
60
50
40
30
20
10
0
2030

2034

2038

2042

2046

2050

2054

2058

2062

2066

2070

2074

2078

2082

2086

2090

2094

2098

2102

2106

2110

2114

2118

2122
depth m

LambdaRho LI(45)^0.6/LI(30) MuRho

25
Case Study Log Analysis.
Petrophysical parameters such as lithology, porosity, and fluid saturations were established from a complete
log suite for the case study as shown in figure 17. The combined resistivity with neutron density crossover
indicates two sands within the zone of interest; a 10m upper gas sand just below 700m and a lower wet sand
(below 725m) with marginal gas saturation (fizz gas) as shown by the neutron density approach. Beyond
identifying good porosity sand zones from low impedance shale, distinguishing economic gas from “fizz
water” is the main challenge for 3D seismic and AVO in the case study area. The improved ability to
distinguish gas from wet sand and lithology using Lamé parameters for this Viking sand case study, is shown
in figures 10a and b, in the previous section on “Log analysis motivation and interpretation templates”. The λ
vs. μ and λρ vs. μρ cross-plots, show clear clusters that identify the various sand, silt and shale lithologies, as
well as showing a clear separation between gas and wet “fizz gas”. Understanding the cross-plot differences
between a Lambda only fluid effect compared to LambdaRho with the advantage of a density influence, shows
how the sand zone gas/water saturations may be estimated to better distinguish “fizz water”. The sonic (P and
S), density and gamma logs are also shown in depth in figure 18. The lowest λρ values as well as a crossover
of λρ < μρ curves (not shown) clearly identify the gas from wet sands, whereas the P and S impedance curves
are less discriminating. These low λρ gas zones are the lowest in the whole logged interval, unlike the P
impedance and so λρ is an excellent match to the density-neutron crossover data for gas identification as
shown in figure 17.

Resistivity Neutron Density

Viking Gas 700m


& Gas Sand

Wet Sand
Zone of
Zones interest Wet
Sand

750m

2005031a-1

Figure 17, Case Study Viking gas and wet sand logs.

26
Viking Logs for Gas Detection (Type 2 or 3 Gas Sand)
Shear
LambdaRho Impedance P-Impedance Gamma Density P Velocity
107 x (m/s x gm/cc) (m/s x gm/cc) (m/s x gm/cc) API (gm/cc) (m/s)
0 2 4 6 0 4000 8000 0 4000 8000 0 100 200 2 3 0 4000

700m
Gas Sand

Zone of Wet
interest Sand

750m

2005031a-2

Figure 18, Seismic sonic, density and gamma logs for Viking sand Case Study.

Case Study Seismic walkaway VSP and 3D AVO Inversion for Lamé Parameters.
The ability to accurately estimate Lamé parameters and hence predict lithology and fluid type from
conventional surface seismic data is greatly affected by the quality of prestack seismic amplitudes. These
amplitudes are often degraded by offset dependent interference, including noise and multiples (Cambois
2001). To what degree prestack processing preserves true amplitudes is generally unknown. Therefore,
independent “true” AVO evidence and measurements are needed to calibrate prestack surface seismic
amplitudes in order to better detect economic gas and discriminate lithology.
A walkaway VSP provides a real seismic data set for calibration of surface AVO, as well as the evidence of a
quantifiable AVO response (Downton, Goodway & Chen 1998, 1999). This surface to borehole VSP seismic
method has a number of advantages in that a VSP is a controlled experiment where the results are directly tied
to well logs in depth and time, so that it is possible to quantify the reliability of the different AVO methods. In
addition the geometry and the angle of incidence is known along with the background Vp/Vs ratio, something
that has a high degree of uncertainty when dealing with surface seismic. As the 3D surface seismic was also
acquired and processed in a similar fashion to the VSP, so a direct comparison to calibrate and quantify the
various AVO inversion results with the VSP data is described below.
Figure 19 shows the basic walkaway VSP geometry designed for the AVO analysis. Source and receiver
locations are schematically shown by flags and triangles respectively, with the source offsets (distance
between the source and the surface location of the wellbore) from 100m to 500m at 100m intervals in addition
to the full level recorded zero-offset. Through check-shot and sonic log guided ray-tracing, the incident angle
at the zone of interest was found to range from 0° to 40° for the largest offset. An amplitude preserving
processing flow was designed and applied to both VSP and surface data, resulting in the VSP “AVO gather”
response shown in figure 19.

27
Walkaway VSP Geometry Walkaway VSP AVO Gather
Negative polarity
(peak=impedance Positive polarity
increase)

200
300

500

400
500
100

400

100
200
300
Offset

Gas AVO
Response 700m

Wet
Low Gas Sg
no AVO ?
750m
Modified from Taiwen Chen et al 1998

Figure 19, Walkaway VSP geometry for AVO gather used in Case Study.

The basic processing sequence followed was to perform an AVO extraction on both the P-wave surface and
VSP seismic data as well as the VSP P-S converted wave VSP data to obtain direct estimates of P and S
impedance reflectivities. A modified equation to extract P and S reflectivities (Gidlow et al. 1992, Fatti et al.
1994) is used with the last term for small ρ contrasts being ignored, as shown in the previous section on
reflectivity equations. This approach was chosen in preference to the various Lamé forms of the Aki and
Richards equation (also shown in the reflectivity equations section), as they are not yet in common industry
practice for AVO attribute extraction. Having extracted the P, S reflectivity sections, the next step is to obtain
Ip and Is through inversion. Finally by using the Lamé moduli to impedance relationships given above, the
extraction and transformation to λρ and μρ is obtained.
The VSP P-wave AVO and surface seismic gathers are compared to log based forward modelling in figure 20.
An excellent calibration of VSP to surface seismic was achieved as can be seen in the rightmost VSP data to
model miss-fit error panel in figure 20 below.

Comparison of VSP and Surface Seismic AVO with Log Modelling


Far Near Far Near Far Near
Reflectivity
VSP Extracted Model
Error
AVO gradients Prediction
0.5

Viking
gas sand
0.6

0.7

0.8
Walkaway VSP Synthetic Log Surface Seismic
AVO Gather AVO Gather AVO Gather
from Jon Downton et al 1999
Figure 20, VSP and surface seismic AVO with log modelling.

28
The P and S reflectivity traces extracted from the VSP and surface seismic gathers were inverted using a
sparse spike inversion algorithm to get P and S impedances. These values were then cross-plotted with the
wireline log information showing a reasonable overall correlation (see Downton et al 1999). Qualitative
comparisons between P, S impedances and λρ and λ/μ ratios for are shown in figures 21 and 22. The leftmost
P impedance panel in figure 21, compares surface AVO data to the equivalent VSP AVO and zero-offset VSP
inversion with the log impedance. Though the overall match is reasonable the actual gas sand zone is not
obvious on any of the panels, being of slightly higher impedance than the background shales. The rightmost S
impedance by contrast, shows a strong increase for the gas zone in both surface and VSP P-P, P-S converted
wave inversions, that is in agreement with the log data. However a clear increase in shear impedance alone at
the Viking level, does not distinguish the gas sand from the lower wet sand zones, as shear-waves have a
minimal response to fluids. Figure 22 on the other hand shows an unambiguous and very well resolved Viking
gas sand zone having the lowest λρ and λ/μ ratios compared to the lower wet sands and encasing shales.
The last set of figures from 23 to 26, compare standard 3D seismic sections and maps of P, S impedance based
AVO inversion to the equivalent Lamé parameters and λρ versus μρ cross-plot polygon based methods, as
previously described. Stacked amplitudes on the standard 3D seismic line at the Viking zone as marked in
figure 23a, do not distinguish the gas sand. The equivalent λρ section in figure 23b, clearly isolates the Viking
gas sand from both encasing shales and the lower wet sand zone in the following trough, shown as the wiggle
overlay of the migrated stacked section. The 3D map display in figures 24a and b, of the Viking gas sand
horizon with the production results of three wells, show extracted AVO attributes based on P, S impedances
(24a) compared to LambdaRho (24b). The impedance based attribute identifies good gas sand porosity (red-
yellow areas) at all three wells and this was the reason these wells were drilled with only one of them (lower
left corner) being economic. The LambdaRho attribute map (figure 24b) however, shows a clear grading
between the three wells with only the economic well having unambiguously low values. Interestingly both the
poor gas well to the top right (watered out within only a month of production) and the D&A well, had
marginally low gas saturations that are clearly distinguishable from each other, as well as the economic
producer.
The final two pairs of figures 25a,b, and 26a,b, show AVO cross-plots for λρ, μρ based polygons compared
with a P, S impedance based linear “fluid factor” cut-off, to isolate only the anomalous gas zones from wet
sands and shales. The best gas zone λρ, μρ “polygon 1” in figure 25a, is based on the log cross-plots shown
before in figure 10 and is more discriminating than the linear impedance based cut-off shown as a white
dashed line. The seismic points identified by this “polygon 1” clearly map onto the economic up-dip gas
accumulation in figure 25b. Figure 26a, shows the same cross-plotting approach for the wet or marginal gas
sand “polygon 2 fizz water” prediction, again based on the logs in figure 10. The points identified from this
“polygon 2” template, map predominantly onto the D&A “fizz water” well and the marginal gas producer in
figure 26b, showing the improved gas sand versus wet sand detection of the λρ, μρ approach for surface
seismic.

Conclusions
1) Improved petrophysical understanding of rock properties using orthogonal Lamé parameters λ (pure
incompressibility) and μ (rigidity) over conventional Vp, Vs or moduli based analysis.
2) Greater physical insight into rock properties for pore fluid versus lithology discrimination by isolating the
Lamé impedances λρ, μρ from the seismic reflectivity response.
3) Easier AVO cross-plot cluster isolation for a more sensitive λρ, μρ lithology stack and for potentially
separating fluid saturation effects.
4) A new λ/μ ratio versus λρ-μρ difference cross-plot showing lines of consistent lithology/porosity with fluid
saturation (gas vs. wet zones) altering the lithology line’s slope.
5) The sensitivity of λρ to fluids has the potential to distinguish economic gas zones from fizz-water as shown by
quantitative inversion and calibration of surface and walkaway VSP data.
6) A direct relationship of AVO derived Lamé parameter λ to the Biot-Gassmann equation.

29
Figure 21.
P-Impedance (Ip) S-Impedance (Is)
Inverted P-P AVO Inverted P-P Inverted zero-offset Filtered Inverted P-P AVO Inverted P-P Inverted P-S Filtered
Surface Data AVO Walkaway VSP Corridor Wireline log Surface Data AVO Walkaway AVO Walkaway Wireline log
VSP Stack VSP

0.5

Viking
gas sand
0.6

0.7

High

0.8

from Jon Downton et al 1999


Low 2005031a-5

Figure 22.
λρ λ/μ
Surface P-P Walkaway VSP P-S Walkaway Filtered Surface P-P Walkaway VSP P-S Walkaway Filtered
Seismic VSP Wireline Seismic VSP Wireline

0.5

Viking
gas sand
0.6

0.7

High

0.8

from Jon Downton et al 1999


Low 2005031a-6

Figures 21 and 22, Qualitative comparisons between P, S impedances and λρ and λ/μ
ratios inverted from walkaway VSP and surface seismic AVO in Case Study area.

30
Figure 23a. E-W Migrated Line from 3D
Gas well

Gas Sand ?
Amplitude Anomaly

2005031a-7

Figure 23b.
LambdaRho (Viking Gas Anomaly) line from 3D
Gas well
Stn. 135 110 85

Strong Gas
ΣΨAnomaly
ΣΨscale
Low Gas Sg
Wet Sand Low
No Anomaly

High
2005031a-8

Figures 23a, and b, comparing seismic 3-D line to LambdaRho inversion from Case
Study area with gas sand well.

31
Figure 24a.
3D Seismic Viking Sand Horizon
P, S Impedance based AVO Attribute Map

Marginal gas Attribute


Poor gas No porosity
saturation well

High gas D&A “Fizz Water”


saturation
Very low gas
saturation
Good
gas well
0 1
Good gas
mile sand porosity

2005031a-9

Figure 24b.
3D Seismic Viking Sand Horizon
LambdaRho Attribute Map
Marginal gas
saturation LambdaRho
Poor gas High
(mid-value λρ)
λρ) well Amplitude

High gas
saturation D&A “Fizz Water”
(lowest λρ)
λρ)
Very low gas
saturation
Good (highest λρ)
λρ)
gas well
0 1
Low
mile Amplitude

2005031a-10

Figures 24 a, and b, horizon maps from 3-D comparing industry standard AVO attribute
to LambdaRho inversion for the Viking sand zone in Case Study.

32
Figure 25a. LambdaRho vs MuRho Seismic Crossplot
42 Gas Zone Polygon 1
36 Fluid factor cutoff
MuRho (Gpa.gm/cc)

y
log
30 LambdaRho
itho High gas saturation High
Polygon 1
dL

Amplitude
24
S an

18

12
Shales
6
Low
Amplitude
0
6 12 18 24 30 36 42 48
LambdaRho (Gpa
(Gpa.gm/cc)
.gm/cc) 2005031a-11

Figure 25b. 3D Seismic Viking Sand Horizon


LambdaRho Attribute Map
Points from λρ, μρ crossplot polygon 1

Marginal gas
saturation Attribute
Poor gas No porosity
(mid-value λρ)
λρ) well
High gas
saturation D&A “Fizz Water”
(lowest λρ)
λρ)
Very low gas
saturation
Good (highest λρ)
λρ)
gas well
0 1 Good gas
mile sand porosity
2005031a-12

Figures 25 a, AVO cross-plot for λρ, μρ based polygon 1 (solid white outline), compared
with a P, S impedance based linear “fluid factor” cut-off (dashed line) to isolate only the
anomalous best gas zones from wet sands and shales, displayed back on horizon map
25 b.

33
Figure 26a. LambdaRho vs MuRho Seismic Crossplot
42 Low Gas Saturation Zone Polygon 2
36 Fluid factor cutoff
Polygon 2
MuRho (Gpa.gm/cc)

30 Very low gas saturation LambdaRho


ogy (“Fizz water”) High
Amplitude
l
itho

24
dL
Sa n

18

12
Shales
6
Low
Amplitude
0
6 12 18 24 30 36 42 48
LambdaRho (Gpa
(Gpa.gm/cc)
.gm/cc) 2005031a-13

Figure 26b. 3D Seismic Viking Sand Horizon


LambdaRho Attribute Map
Points from λρ, μρ crossplot polygon 2

Marginal gas
saturation Attribute
Poor gas No porosity
(mid-value λρ)
λρ) well
High gas
saturation D&A “Fizz Water”
(lowest λρ)
λρ)
Very low gas
saturation
Good (highest λρ)
λρ)
gas well
0 1
Good gas
mile sand porosity
2005031a-14

Figures 26 a, AVO cross-plot for λρ, μρ based polygon 2 (solid white outline), compared
with a P, S impedance based linear “fluid factor” cut-off (dashed line) to isolate only the
high water saturation zones from gas sands and shales, displayed back on horizon map
26 b.

34
Acknowledgements
Much of this work benefited significantly from the ideas of a number of people, along with a lot of fruitfully
stimulating discussion and I acknowledge these key individuals who were involved in the development, over
the past six years, of the themes presented in this paper.
From PanCanadian: Taiwen Chen, Dan Cieslewicz, Dave Cooper, Jocelyn Dufour, Nilanjan Ganguly, Mark
Harper, Irene Kelly, Guoping Li, Andrew Lowe, Drew MacGreggor, Dave Mackidd, Holger Mandler, Alli
Marshall, Doug Neilson, Ann O’Byrne, Brian Rex, Robert Riddy, Mantu Sihota, Ian Shook, Garth Syhlonyk,
Gerhard Tjaden, Gordon Uswak, Paul Young, Weimin Zhang.
From outside of PanCanadian: Rob Stewart, Larry Lines (University of Calgary), Jon Downton and Yongyi Li
(Scott-Pickford), Dave Gray and Penny Colton (Veritas DGC), George Smith (University of Cape Town), and
Brain Russell (Hampson & Russell).

References
Aki K., and Richards P.G., 1979, Quantitative Seismology, W.H.Freeman & Co.
Berryman J.G., Berge P.A., and Bonner B.P., 1999 “Role of λ-diagrams in estimating porosity and saturation
from seismic velocities”: 69th Internat. Mtg. Soc. Expl. Geophys., Expanded Abstracts. 176-179.
Berryman J.G., Berge P.A., and Bonner B.P., 2002 “Estimating rock porosity and fluid saturation using only
seismic velocities” Geophysics 67, 391-404.
Biot M.A., 1956 “The theory of propagation of elastic waves in a fluid saturated solid: I. Lower frequency
range, II. Higher frequency range” J. Acoust. Soc. Am., 28 179-191.
Cambois G., 2000 “Can P-wave AVO be quantitative” TLE November 2000.
Cambois G., 2001 “AVO Processing-Myths and Realities” EAGE Conference Expanded Abstracts #N-14.
Castagna J.P.,1993a “Petrophysical imaging using AVO” TLE March 1993.
Castagna J.P., Batzle M.L., and Kan T.K., 1993b “Rock Physics-The link between rock properties and AVO
response” SEG Investigations in Geophysics #8 “Offset-dependent reflectivity-theory and practice of AVO
analysis”.
Castagna J.P., and Smith S.W., 1994 “Comparison of AVO indicators: a modelling study” Geophysics 59,
1849-1855.
Castagna J.P., and Swan H.W., 1997 “Principles of AVO crossplotting” TLE April 1997, 337-342.
Chen T., Goodway W., Zhang W., Potocki D., Uswak G., Calow B., and Gray D., 1998 “Integrating
Geophysics, Geology and Petrophysics: A 3D seismic AVO and borehole/logging case study” Ann. Int.
Mtg., SEG. Expanded Abstracts, 615-618.
Connolly P., 1999 “Elastic Impedance” TLE April 1999.
Downton J., Goodway W., and Chen T., 1999 “Quantitative comparison of deriving P and S impedances
from PP and PS surface and VSP data” CSEG National Convention Expanded Abstracts.
Fatti J.L, Smith G.C, Vail P.J, Strauss P.J, Levitt P.R,1994 “Detection of gas in sandstone reservoirs using
AVO analysis: A 3D seismic case history using the Geostack technique” Geophysics 59, 1362-1376.
Gardner G.H.F., Gardner L.W., and Gregory A.R., 1974 “Formation velocity and density-the diagnostic
basics for stratigraphic traps” Geophysics 39, 770-780.
Gassmann F,1951 “Elastic waves through a packing of spheres” Geophysics 16, 673-685.
Geertsma J., 1961 “Velocity-log interpretation: the effect of rock bulk compressibility” SPE Jour.1, 235-248.
Gidlow P.M, Smith G.C, Vail P.J,1992 “Hydrocarbon detection using fluid factor traces; A case history”
SEG/EAEG Summer Workshop 78-79.
Goodway W., Chen T., and Downton J., 1997 “Improved AVO fluid detection and lithology discrimination
using Lamé parameters; λρ, μρ and λ/μ fluid stack from P and S inversions” CSEG National Convention
Expanded Abstracts 148-151.
Goodway W., Chen T., and Downton J., 1998 “Joint P-P and P-S inversion of a walkaway VSP for Vp, Vs
and Vp/VS compared to log data and for surface P-P calibration and inversion” SEG Post-Convention
Workshop “Can P-wave AVO be quantitative or do we need multi-component”.
Hilterman F. J., 2001 SEG Distinguished Instructor Short Course Monograph.
Leaney S., 1994 “AVO and anisotropy from logs and walkaways” Schlumberger Interpretation Development
35
Forum, Jakarta.
Ostrander W.J, 1984 “Planewave reflection coefficients for gas sands at non-normal angles of incidence”
Geophysics 49,1637-1648.
Pickett G.R, 1963 “Acoustic character logs and their application in formation evaluation” J. Petr. Tech.,
659-667.
Shuey R.T, 1985 “A simplification of the Zoeppritz equations” Geophysics 50, 609-615.
Smith G.C, 1996 “3-parameter Geostack” Ann. Int. Mtg., SEG. Expanded Abstracts, 1747-1750.
Smith G.C, 1999 “The relationship between Lamé’s constants, lambda and mu, and the fluid factor in AVO
analysis of seismic data” S. African Geoph. Assoc. meeting abstracts 6.3.
Smith G.C, Gidlow P.M, 1987 “Weighted stacking for rock property estimation and detection of gas”
Geophysical Prospecting 35, 993-1014.
Smith G.C, Gidlow P.M, 2000 “A comparison of the fluid factor with λ and μ in AVO analysis” Ann. Int.
Mtg., SEG. Expanded Abstracts, 122-125.
Stewart R.R, Zhang Q, Guthoff F,1995 “Relationships among elastic-wave values; Rpp, Rps, Rss, Vp, Vs, k,
σ, ρ” CREWES Report #7.
Tatham R.H,1982 “Vp/Vs and lithology” Geophysics 47., 336-344.
Thomsen L,1986 “Weak elastic anisotropy” Geophysics 51, 1954-1966.
Thomsen L,1990 “Poisson was not a geophysicist” TLE Dec., 27-29.
Verm, R.W., Hilterman, F.J., 1995, “Lithologic colour-coded sections: The calibration of AVO crossplotting
to rock properties” TLE August 1995.
Vernik L., Nur A., 1992 “Ultrasonic velocity and anisotropy of hydrocarbon source rocks” Geophysics, 57,
727-735
Wallace R, Young R,1996 “Pre-stack Inversion: Evolving the Science of Inversion” CSEG Recorder;
December.
Wright J,1984 “The effects of anisotropy on reflectivity offset” 54th Ann.Mtg.,SEG.Expanded
Abstracts,670-672.

36
APPENDICES
The elastic wave equation in Lamé moduli terms.
Starting from the general equation of motion neglecting body forces;
d2ui
σ ij, j = ρ
dt 2
where subscript indices i j k and l for stress and strain represent directions and face normals
dσ ij
and use of a comma between indices in σ ij, j = ( the spatial derivative of σ ij in the x j direction)
dx j
Inserting Hooke' s law σ ij = λδ ij e kk + 2μe ij into the equation of motion
d2ui
σ ij, j = λδ ij e kk , j + 2μe ij, j = ρ
(see Hooke' s law section)
dt 2
1 ⎛ du du j ⎞ 1
Next by replacing strain e ij = ⎜ i + ⎟ = u i , j + u j,i
2 ⎝ dx j dx i ⎟⎠ 2

( )
along with some algebraic calculus manipulation of the indices and using the Kronecker rule δ ij ⇒ i = j
the equation of motion can be rearranged as;
σ ij, j = λe kk ,i + μ(u i , j + u j,i ) , j = λu k ,ki + μ(u i , jj + u j,ij ) = λu k ,ki + μu i , jj + μu j, ji
⇒ σ ij, j = λu j, ji + μu j, ji + μu i , jj (as dummy index k ⇒ j).
Note at this point there is a separation into two μ terms for particle displacement derivatives;
one of which shares the same particle derivative as the λ term and is related to longitudinal stress - strain alone
while the other has different spatial derivatives w.r.t. direction indices and is also related to shear stress - strain
d2ui d2ui
⇒ σ ij, j = (λ + μ )u j, ji + μu i , jj = ρ and in vector divergence form (λ + μ )∇(∇ • u ) + μ∇ 2 u = ρ
dt 2 dt 2
This form of the 3D elastic wave equation has an unfamiliar (λ + μ) modulus associated with the
compressional strain or dilatation term ∇•u, instead of the usual P-wave propagation modulus (λ + 2μ). To
resolve this, consider displacement vector u in the time derivative on the right hand side of the divergence
equation above. It is not obviously a compressional or shear motion so the divergence and curl of the whole
equation must be taken as a “test”, along with representing vector u as a sum of two vectors, one whose
divergence is zero and one whose curl is zero; u = u1 + u2 where ∇•u1= 0 &∇×u2= 0.

37
d 2 (u 1 + u 2 )
Starting from (λ + μ )∇(∇ • (u 1 + u 2 )) + μ∇ 2 (u 1 + u 2 ) = ρ with ∇ • u 1 = 0, ∇ × u 2 = 0
dt 2
and taking the divergence
d 2∇ • u 2 d 2 (∇ • u 2 )
⇒ρ = (λ + μ )∇ 2 (∇ • u 2 ) + μ∇ • ∇ 2 (u 1 + u 2 ) ⇒ ρ = (λ + 2μ )∇ 2 (∇ • u 2 )
dt 2 dt 2
Resulting in the familiar apperance of the compressional wave equation involving dilatation ∇ • u 2
and the compressional stress - strain modulus (λ + 2μ )
Next taking the curl
d 2 (∇ × u 1 )
⇒ρ 2
= μ∇ × ∇ 2 (u 1 + u 2 ) as ∇ × ∇(∇ • (u 1 + u 2 )) = 0
dt
d (∇ × u 1 )
2
⇒ρ = μ∇ 2 (∇ × u 1 )
dt 2
Resulting in the shear wave equation involving rotation or transverse ∇ × u 1 motion and rigidity modulus μ.

Anisotropy and Lamé constants.


Beyond the isotropy assumed above, an example of a deeper physical insight into the particle mechanics for
anisotropic media, can be obtained by converting Thomsen’s delta δ (1986) parameter for VTI into Lamé
parameters. These parameters are “layer perpendicular” rigidity μ⊥ and incompressibilities λ13 and λ33
involved in stiffnesses C13 and C33, that result in the following relationship;
2 2
(C13 + C 44 ) 2 − (C 33 − C 44 ) 2 (λ − λ 33 ) + 2μ ⊥ (λ13 − λ 33 )
δ= ⇒ δ = 13
2C 33 (C 33 − C 44 ) 2
2λ 33 + 6λ 33 μ ⊥ + 4μ ⊥
2

These equations reveal δ as being simply a function of the λ13−λ33 difference. A number of lithologies
including shales (Vernik & Nur 1992,Leaney 1994), have a very small or zero λ13−λ33 difference (see graph
below) i.e. a δ range of 0 to ±0.03. Consequently these media can be characterised by 4 instead of the usual 5
elastic VTI constants, as can be seen from C33 = C13+2C44, where λ13 = C13. The 4 elastic constants for these
special cases of VTI anisotropy are specifically; C13, C44 and C12, C66 or in Lamé terms, a more naturally
intuitive description as “layer perpendicular” λ⊥, μ⊥ and “layer parallel” λ≡, μ≡. These results are compared in
the following isotropic versus anisotropic modulus tensor matrices.

Isotropic Stress-Strain Tensor Matrix;


Voigt notation for Cijkl => Cab

⎛ σ xx ⎞ ⎡ C 33 C 33 − 2C 44 C 33 − 2C 44 0 σS 0 σS 0 σS ⎤ ⎛ e xx ⎞
⎜ ⎟ ⎢ ⎜ ⎟
⎜ σ yy ⎟ ⎢C 33 − 2C 44 C 33 C 33 − 2C 44 0 σS 0 σS 0 σS ⎥⎥ ⎜ e yy ⎟
⎜ σ ⎟ ⎢ C − 2C C 33 − 2C 44 C 33 0 σS 0 σS 0 σS ⎥ ⎜⎜ e zz ⎟
⎜ zz ⎟ = ⎢ 33 44
⎥⋅ ⎟
⎜ σ yz ⎟ ⎢ 0 σC 0 σC 0 σC C 44 0S 0 S ⎥ ⎜ e yz ⎟
⎜ ⎟ ⎢ ⎜ ⎟
⎜ σ zx ⎟ ⎢ 0 σC 0 σC 0 σC 0S C 44 0 S ⎥ ⎜ e zx ⎟
⎜ σ xy ⎟ ⎢ ⎥ ⎜
⎝ ⎠ ⎣ 0 σC 0 σC 0 σC 0S 0S C 44 ⎦⎥ ⎝ e xy ⎟⎠

In Lamé terms;

1 ⎛⎜ du i du j ⎞⎟
From σ ij = λδ ij e kk + 2μe ij and e ij = +
2 ⎜⎝ dx j dx i ⎟⎠

38
⎛ σ xx ⎞ ⎡λ + 2μ λ λ 0 σS 0 σS 0 σS ⎤ ⎛ e xx ⎞
⎜ ⎟ ⎢ ⎜ ⎟
⎜ σ yy ⎟ ⎢ λ λ + 2μ λ 0 σS 0 σS 0 σS ⎥⎥ ⎜ e yy ⎟
⎜σ ⎟ ⎢ λ λ λ + 2μ 0 σS 0 σS 0 σS ⎥ ⎜⎜ e zz ⎟⎟
⎜ zz ⎟ = ⎢ ⎥⋅
⎜ σ yz ⎟ ⎢ 0 σC 0 σC 0 σC μ 0S 0 S ⎥ ⎜ e yz ⎟
⎜ ⎟ ⎢ ⎜ ⎟
⎜ σ zx ⎟ ⎢ 0 σC 0 σC 0 σC 0S μ 0 S ⎥ ⎜ e zx ⎟
⎜ σ xy ⎟ ⎢ 0 ⎥ ⎜
⎝ ⎠ ⎣ σC 0 σC 0 σC 0S 0S μ ⎦⎥ ⎝ e xy ⎟⎠

where “zeros” are;


0[S from no shear stress
0[C from no compressional stress (i.e. Λij = 0 as i ≠ j)
0 S from decoupled uniaxial shear stress and strain

Anisotropic VTI Stress-Strain Tensor Matrix (5 term) where C13 = C31;

⎛ σ xx ⎞ ⎡ C11 C11 − 2C 66 C13 ⎤ ⎛ e xx ⎞


⎜ ⎟ ⎢ ⎥ ⎜e ⎟
⎜ σ yy ⎟ ⎢C11 − 2C 66 C11 C13 ⎥ ⎜ yy ⎟
⎜σ ⎟ ⎢ C C 31 C 33 ⎥ ⎜ e zz ⎟
⎜ zz ⎟ = ⎢ 31
⎥ ⋅⎜ ⎟
⎜ σ yz ⎟ ⎢ C 44 ⎥ ⎜ e yz ⎟
⎜ ⎟ ⎢ ⎥ ⎜ e zx ⎟
⎜ σ zx ⎟ ⎢ C 44
⎥ ⎜⎜ ⎟
⎜ σ xy ⎟ ⎢ C 66 ⎦⎥ ⎝ xy ⎟⎠
e
⎝ ⎠ ⎣

In Lamé terms;

⎛ σ xx ⎞ ⎡λ ≡ +2μ ≡ λ≡ λ 13 ⎤ ⎛ e xx ⎞
⎜ ⎟ ⎢ ⎥ ⎜e ⎟
⎜ σ yy ⎟ ⎢ λ ≡ λ ≡ +2μ ≡ λ 13 ⎥ ⎜ yy ⎟
⎜σ ⎟ ⎢ λ λ 31 λ 33 + 2μ ⊥ ⎥ ⎜ e zz ⎟
⎜ zz ⎟ = ⎢ 31
⎥ ⋅⎜ ⎟
⎜ σ yz ⎟ ⎢ μ⊥ ⎥ ⎜ e yz ⎟
⎜ ⎟ ⎢ ⎥ ⎜ e zx ⎟
⎜ σ zx ⎟ ⎢ μ⊥
⎥ ⎜ ⎟
⎜ σ xy ⎟ ⎢ μ ≡⎥⎦ ⎜⎝ e xy ⎟⎠
⎝ ⎠ ⎣

Where ≡ is measured parallel to layering and ⊥ is perpendicular to layering.


If Thomsen’s delta δ VTI parameter is equal to zero then incompressibilities λ13 = λ33 = λ⊥

39
(see explanation above) giving a four instead of five term Anisotropic VTI Stress-Strain Tensor Matrix;

⎛ σ xx ⎞ ⎡λ ≡ +2μ ≡ λ≡ λ⊥ ⎤ ⎛ e xx ⎞
⎜ ⎟ ⎢ ⎥ ⎜e ⎟
⎜ σ yy ⎟ ⎢ λ ≡ λ ≡ +2μ ≡ λ⊥ ⎥ ⎜ yy ⎟
⎜σ ⎟ ⎢ λ λ⊥ λ ⊥ + 2μ ⊥ ⎥ ⎜ e zz ⎟
⎜ zz ⎟ = ⎢ ⊥
⎥ ⋅⎜ ⎟
⎜ σ yz ⎟ ⎢ μ⊥ ⎥ ⎜ e yz ⎟
⎜ ⎟ ⎢ ⎥ ⎜ e zx ⎟
⎜ σ zx ⎟ ⎢ μ⊥
⎥ ⎜⎜ ⎟
⎜ σ xy ⎟ ⎢ μ ≡⎦⎥ ⎝ xy ⎟⎠
e
⎝ ⎠ ⎣

G raph of Lam bda "33" vs. Lam bda "13" for w orld w ide shales m easured
from w alkaw ay VSP'S (Leaney 1994) and core (Nur/Vernik 1992, Lo 1986,
Jones/W ang 1981)
20% m easurem ent error
16

14

12
Lambda "33" Gpa.

10

0
0 2 4 6 8 10 12 14 16
Lam bda "13" G Pa.

40
41

Potrebbero piacerti anche