Sie sind auf Seite 1von 8

GEOPHYSICS, VOL. 68, NO. 4 (JULY-AUGUST 2003); P. 11611168, 14 FIGS., 1 TABLE.

10.1190/1.1598108

Amplitude and AVO responses of a single thin bed

Yinbin Liu and Douglas R. Schmitt


reflections from a thin layer are concerned with seismic resolution, detection, and amplitude variation with offset (AVO).
Resolution and detectability for a thin layer have been studied
by Widess (1973), Neidell and Poggiagliolmi (1977), Koefoed
and de Voogd (1980), Kalweit and Wood (1982), de Voogd and
Rooijen (1983), Gochioco (1991), Chung and Lawton (1995,
1996), Liu and Schmitt (2001), and others.
In exploration geophysics, the generally accepted threshold
for vertical resolution of a layer is a quarter of the dominant
wavelength (Yilmaz, 1987). In this paper, the layer is called
a thin layer when 1 < /d 4, and an ultra-thin layer when
/d > 4, where is the dominant wavelength within the layer
and d is the layer thickness.
Widesss classic paper (1973) studied the normal pulse reflections from the top and bottom of a thin layer with equal amplitude and opposite polarity, and noted that reflections from
very thin beds are not always small. Neidell and Poggiagliolmi
(1977) and Kallweit and Wood (1982) studied the reflections of
a thin layer and a wedge model, and showed that the thickness
information is encoded in the amplitude and shape of the reflected wavelet when the thickness is less than the tuning thickness of layer. De Voogd and Koefoed (1980) and Gochioco
(1991) demonstrated that the seismic response of coal seams
as thin as /20 to /50 can give rise to a distinct reflection signal.
Chung and Lawton (1995, 1996) studied the reflection characterizations for different sedimentary formations and showed
that the amplitude dependence on the thickness is nonlinear.
Almoghrabi and Lange (1986), Lange and Almoghrabi (1988),
and Juhlin and Young (1993) studied the AVO response of a
thin layer by considering multiples, and showed the AVO response of a thin layer may differ significantly from the AVO
response of a simple interface.
The simplified methods used by the above authors gave
a clear description of the interference between the top and
bottom interfaces, especially for seismic resolution. However,
these methods only apply to either normal incidence reflections (Widess, 1973; Neidell and Poggiagliolmi, 1977; Koefoed
and de Voogd, 1980; Kalweit and Wood, 1982; de Voogd and
Rooijen, 1983; Gochioco, 1991; Chung and Lawton, 1995, 1996)

ABSTRACT

The seismic reflection characterizations of a thin layer


are important for reservoir geophysics. However, discussions on the reflection for a thin layer are usually restricted to precritical angle incidence. In this work, an exact analytical solution is derived to model the reflection
amplitude and amplitude variation with offset (AVO) responses of a single thin bed for arbitrary incident angles.
The results show that the influence of an ultra-thin bed
is great for opposite-polarity reflections and is small for
identical-polarity reflections. Opposite-polarity precritical reflection amplitudes first decrease in magnitude with
the wavelength/thickness ratio to a local minimum, then
increase to a maximum, and finally decrease gradually to
zero as the layer vanishes. Opposite-polarity postcritical
reflections monotonically decrease from near unity to
zero, proportional to the thickness of the layer. Identicalpolarity precritical reflection amplitudes first increase in
magnitude with the wavelength/thickness ratio to a local maximum, then decrease to a minimum, and finally
increase to the amplitude of a single bottom reflection
when the layer vanishes. Identical-polarity postcritical
reflections have magnitudes near unity. The AVO responses for both opposite and identical-polarity acoustic thin beds gradually increase with angle. The influence of the Poissons ratio of the thin bed is small for
either small incidence angles or thicknesses less than
7% of the seismic wavelength, but is large for high incidence angles or thicknesses greater than 13% of the
wavelength. A decrease of Poissons ratio causes a pronounced AVO response that reaches its maximum at the
quarter-wavelength tuning thickness.

INTRODUCTION

An ultra-thin gas-sand layer has often a detectable seismic reflection response (e.g., Schmitt, 1999). Generally, the

Manuscript received by the Editor March 25, 2002; revised manuscript received February 4, 2003.

Formerly University of Alberta, Institute for Geophysical Research, Department of Physics, Edmonton, Alberta T6G 2J1, Canada; presently
517-4739 Dalton Dr. NW, Calgary, Alberta T3A 2L5, Canada. E-mail: yinbin@telus.net.
University of Alberta, Institute for Geophysical Research, Department of Physics, Edmonton, Alberta T6G 2J1, Canada. E-mail: doug@
phys.ualberta.ca.

c 2003 Society of Exploration Geophysicists. All rights reserved.


1161

1162

Liu and Schmitt

or precritical incidence angles (Almoghrabi and Lange, 1986;


Lange and Almoghrabi, 1988; Juhlin and Young, 1993). Moreover, these methods are all based on ray theory and should
be restricted to the case in which the wavelength is small relative to the dimension of the structure. Thin bed reflections can
be accurately described only by wave theory. In this paper, the
amplitude and AVO responses of thin or ultra-thin beds are
quantitatively studied by an exact analytical solution.
THEORY AND ALGORITHM

Model
Real reservoir structures and reflection seismic data are very
complex. Despite this, simplified geometries (e.g., a layered
model) with simplified media approximation (e.g., acoustic media) may help us to extract physical essences from complex
background. Our models consist of an acoustic thin layer (models I and II) embedded between two half-spaces (Figure 1). An
elastic model (model III) is also discussed in order to study the
influence of Poissons ratio.
The parameters and properties of these three models are
listed in Table 1. Parameters c1 = sin1 (1 /2 ) and c2 =
sin1 (1 /3 ) are the P-wave critical angles on the top and
bottom interfaces relative to the incident wave, respectively.
Model I uses Widesss (1973) calculated parameters and denotes a thin high-velocity layer. Model II denotes a thin
transition layer, whereas model III represents the Wabasca
gas formation in the Western Canadian Sedimentary Basin
(Schmitt, 1999). The densities for models I and II are uniform. Models I and III produce opposite-polarity P-wave reflections, whereas model II produces identical-polarity P-wave
reflections. The opposite- (identical-) polarity reflections mean
that wavelet reflections from the top and bottom of a thin
layer are opposite- (identical-) polarity (Kallweit and Wood,
1982). Generally, a high- or low-impedance bed (2 2 > 1 1
and 3 3 , or 2 2 < 1 1 and 3 3 ) produces opposite-polarity
reflections, whereas increased or decreased impedance bed

FIG. 1. Thin bed model, reflection rays, and the corresponding


time delay t = (2d/2 ) cos 2 .
Table 1.
1 (m/s)
Model I
Model II
Model III

3050
3050
2200

1 (g/cm )
3

2.7
2.7
2.3

2 (m/s)
6100
4575
1500

(1 1 > 2 2 > 3 3 or 1 1 < 2 2 < 3 3 ) produces identicalpolarity reflections.


Method
A plane monochromatic wave with unit amplitude illuminates the thin layer (Figure 1). For acoustic media, the reflection coefficient can be written as (see Appendix A)

R() =

i Z 1 Z 3 Z 22 sin(k z2 d) + (Z 1 Z 2 Z 2 Z 3 ) cos(k z2 d)

,

i Z 1 Z 3 + Z 22 sin(k z2 d) + (Z 1 Z 2 + Z 2 Z 3 ) cos(k z2 d)
(1)
where Z i = (i i )/cos i , and k z2 = (/2 ) cos 2 . Parameters i ,
i , and i (i = 1, 2, 3) are the densities, velocities, and incident
or refracted angles, respectively; d is the thickness of the thin
layer. For the postcritical angle incidence (1 > c1 and c2 ), k z2
becomes imaginary (Snells law), and the waves propagated
within the layer are evanescent waves (also called inhomogeneous waves). When k z2 d = (2 d/2 ) cos 2 = m , equation (1)
becomes

R() =

Z3 Z1
.
Z1 + Z3

(2)

The reflection coefficient in this case looks as if the layer


were absent. The reflection coefficient will be equal to zero if
the two half-spaces have the same impedances (Z 1 = Z 3 ).
For an ultra-thin layer (k z2 d 1), we take the first order
of approximation for R() and have cos k z2 d 1, sin k z2 d =
(2d/2 ) cos 2 . For opposite-polarity reflection, Z 1 = Z 3 , and
the R()in equation (1) can be written as

i Z 22 Z 12 d/2 cos 2
R() =
.
Z1 Z2

(3)

Therefore, the reflection coefficient of an ultra-thin layer for


opposite-polarity reflection is approximately proportional to
the layer thickness.
For identical-polarity reflections, or in any case where
|Z 1 Z 3 Z 22 |(2 d/2 ) cos 2 Z 2 |Z 1 Z 3 |, the reflection coefficient R() of an ultra-thin layer (k z2 d 1) in equation (1) is
almost identical to equation (2). The influence of an ultra-thin
layer on the reflection coefficient can then be neglected.
Figure 2 shows an example for the angular reflection
coefficient spectrum for model I. The thickness of the layer
was varied from d = 2 to d = 2 /100. The curves are computed
in steps 1 = 0.5. It can be seen that the reflection coefficients
increase with increasing angle of the incidence for 2 /d > 2.
The normal reflection coefficients are zero at 2 /d = 1 and 2
because the layer thickness is an integral of half-wavelength
(2 /2) (Brekhovskikh, 1980). There is a null at incident angle
1 = 25.7 for d = 2 . This is because 1 = 25.7 corresponds
to a refracted angle 2 = 60 , which has an apparent thickness
d = 2 /2[t = (2d/2 ) cos 2 ] and so also satisfies the condition

Model parameters.
2 (g/cm3 )
2.7
2.7
2.2

3 (m/s)
3050
6100
2500

3 (g/cm3 )
2.7
2.7
2.35

c1

c2

30
41.8

30
61.6

Thin Bed AVO

of an integral of half-wavelength. Note that the zero appears


only when the layer thickness exceeds 2 /2.
For elastic case, the analytical expressions for reflection coefficients can only be given by matrix form because of the coupling effects of P-wave and SV-wave. For multilayered mediums, the solution is described by the propagator matrix method.
Interested readers may find the discussions about propagator
matrices in, for example, Rokhlin et al. (1999) and Ursin and
Stovas (2002).
For exploration geophysics, the source waveform is not
monochromatic. The reflection impulse response can be written as

(x, z, t) =

G()R()ei(k x xkz zt) d,

(4)

where G() is the wavelet spectrum, and R() is the monochromatic reflection coefficient of the composite layer. In the following, a 50-Hz Ricker wavelet is used for all simulations. To
study the influence of the SV-wave on the reflection amplitude, we compare the reflection field from both acoustic (i.e.,
no shear wave) and elastic thin layer reflections. Figure 3 shows
the calculated reflection waveforms from model III when the
top is an acoustic half-space and the thin layer and bottom
half-space are elastic with Poissons ratio of = 0.25 (solid)
and when the thin layer and two half-spaces are all acoustic
(dashed) at 1 = 20 . It can be seen that the amplitude responses
for the elastic case have a slightly smaller amplitude than those
for the acoustic case. This is because in the elastic case, part of
the energy converts into shear waves or Lamb waves and radiates into the bottom half-space. In the following analysis, we
first deliberately ignore the influence of SV-waves, but in the
final section of this paper the elastic case will be discussed.
EFFECTS OF BED THICKNESS AND INCIDENT ANGLE
ON REFLECTION AMPLITUDE

Opposite-polarity reflection

1163

wavelet at normal and 20 incidences for d = 2 to d = 2 /8


(Figure 4a) and d = 2 /10 to d = 2 /100 (Figure 4b). The waveforms for normal incidence are similar to those of Widess
(1973) computed by time delay approximation. It can be seen
that two reflection wavelets from the top and bottom of the thin
bed overlap, and that the time delay t for R1 and R2 can be
approximately calculated by ray theory as t = (2d/2 ) cos 2
(see Figure 1), where 2 is the refracted angle and 2 is the
P-wave velocity of the layer. Obviously, as the bed thins or
the incident angle expands, the delay time t decreases and
worsens the overlap. Oblique incidence is equivalent to thinner layers, in terms of delay time. The influence of an ultra-thin
bed on reflection amplitudes is relatively large for oppositepolarity reflections. For example, say the reflection coefficients
are about 0.2 at 1 = 0 and 0.22 at 1 = 20 for 2 /d = 20. A
bed with 2 /d = 40 still reflects 11% of the amplitude of the
incident wave.
The critical angle at the top interface of model I is c1 = 30 .
Figure 5 is the reflection of model I for d = 2 to d = 2 /100
at two postcritical incidence angles of 1 = 40 (solid lines)
and 1 = 60 (dashed lines). The reflection amplitudes decrease
with decreasing thickness. The Zoeppritz equations (Aki and
Richards, 1978) predict the reflection and transmission of a single interface, and show that the reflection for postcritical angles
has a magnitude near unity. Therefore, the Zoeppritz equations
are not suitable to study the amplitude and AVO responses for
thin-layer problem.
The maximum magnitude of each reflection waveform at
incident angle 1 and layer thickness d can be calculated by
taking the absolute maximum value of the waveform. Figure 6
shows the maximum reflection magnitude as a function of the
wavelength/thickness ratio (2 /d) at several incident angles.
The curves are computed in steps d of 0.0052 . The precritical amplitude response for opposite-polarity reflections
first decreases to a local minimum (at 2 /d 2 for normal
incidence) indicating destructive interference, then increases
to a maximum (at 2 /d 4 for normal incidence) indicating

Figure 4 shows the composite reflection waveforms of model


I calculated by the analytical solution for a 50-Hz Ricker

FIG. 2. Reflection coefficient spectrum (single frequency) for


different wavelength/thickness (2 /d) in model I.

FIG. 3. Comparison between elastic (Poissons ratio = 0.25)


and acoustic thin layer reflections in model III for the layer
thickness changes from d = 2 to d = 2 /100. Solid and dashed
lines denote the elastic and acoustic layers, respectively.

1164

Liu and Schmitt

constructive interference at the tuning thickness, and finally


gradually decreases to zero as the layer vanishes. The greater
the angle of the incidence or the thinner the layer thickness,
the smaller the reflection amplitudes become. The minima and
maxima for high incidence angles appear at smaller thickness
than those for low incidence angles because layers appear to be
thinner at oblique angles. The maximum absolute amplitudes
for the postcritical incidence angles (1 30 ) monotonically
decrease from near unity (total reflection) to zero. This is
because the wave within the thin layer is evanescent, resulting
in amplitude attenuation.
Figure 7 shows the AVO response of model I where the
thickness of the layer was varied from d = 2 to d = 2 /100.
The maximum absolute amplitudes increase with increasing
angle of incidence or offset. The amplitude changes are large
for d = 2 and d = 2 /2 (1 < c1 ) and for d = 2 /4 to d = 2 /10,
but are slight for 2 /d > 20 when 1 is less than about 40 .

The influence of the critical angle of the top interface on the


reflection coefficients and AVO is obvious for 2 /d < 2 (total
reflection), but cannot be observed for 2 /d > 4. The AVO response smoothly passes the critical angle (c1 = 30 ) and only
at near grazing incidence (1 90 c1 ) does the effective
reflectivity approach unity.
Identical-polarity reflection
Figure 8 shows the calculated reflection waveforms of model
II at normal and 20 incidences for various 2 /d ratios. Figure 9
shows the maximum absolute amplitudes as a function of the
wavelength/thickness ratio for several incident angles. The precritical amplitude responses for identical-polarity reflections
first increase to a local maximum (at 2 /d 2 for normal incidence) indicating constructive interference, then decrease to a

FIG. 5. Opposite-polarity reflections in model I for d = 2


to d = 2 /100 at two postcritical incidence angles. Solid and
dashed lines denote = 40 and 60 , respectively.

FIG. 4. Opposite-polarity reflections in model I. The layer


thickness changes from d = 2 to d = 2 /8 for (a) and from
d = 2 /10 to d = 2 /100 for (b). Solid and dashed lines denote
the normal and 20 incidences, respectively.

FIG. 6. Maximum absolute amplitudes of opposite-polarity reflections in model I as a function of 2 /d for several incident
angles.

Thin Bed AVO

1165

minimum (at 2 /d 4 for normal incidence) indicating destructive interference, and finally increase to the amplitude of the
single bottom-reflection wavelet without the thin layer. The
maxima and minima shift to smaller values of 2 /d for the
larger incident angles. Figure 9 shows that thinner layers result
in larger reflection amplitudes in the precritical case because
of constructive interference between identical-polarity reflections. The maximum absolute amplitudes for 2 /d greater than
about 20 are basically invariant to the wavelength/thickness
ratio. This means that the amplitude differences with and without thin layers are small. The amplitude responses of identicalpolarity reflections are not sensitive to an ultra-thin layer. A
single ultra-thin layer appear to be a single interface. The postcritical reflection amplitude for identical-polarity reflections is
near unity, which is similar to the total reflection of a single
interface.
Figure 10 shows the AVO response for identical-polarity reflections for d = 2 to d = 2 /100. The reflection amplitudes

increase with the increasing angle of incidence for 1 < c2 and


have a near unity value for 1 c2 (total reflection). The influence of the critical angle on reflection amplitude for identicalpolarity reflections is similar to the case of a single interface.

FIG. 7. Maximum absolute AVO for opposite-polarity reflections for different wavelength/thickness (2 /d) in model I.

FIG. 9. Maximum absolute amplitudes of identical-polarity reflections in model II as a function of 2 /d for several incident
angles.

FIG. 8. Identical-polarity reflections in model II for the layer


thickness changes from d = 2 to d = 2 /100. Solid and dashed
lines denote the normal and 20 incidences, respectively.

EFFECTS OF POISSONS RATIO ON AVO RESPONSES

Now let us turn our attention toward the reflection from an


elastic layer and consider the influence of shear waves. We assume an identical Poissons ratio of 0.25 for both the thin sand
and the bottom half-space. The critical angle on the bottom
interface is c2 = 61.6 . Figures 11 and 12 show the amplitude
and AVO responses for various 2 /d. It can be seen that the
amplitude and AVO responses appear as a composite effect of
opposite and identical-polarity reflections because of the different impedances (1 1 6= 3 3 ) but have a little bit more complex
structures. The elastic AVO responses exhibit opposite-polarity
behavior for precritical angles and identical-polarity behavior

FIG. 10. Maximum absolute AVO for identical-polarity reflections for different wavelength/thickness (2 /d) in model II.

1166

Liu and Schmitt

for postcritical angles. AVO responses basically increase with


increasing angle of the incident angle for the precritical angles
and are complex for the postcritical angles (bottom interface)
because of the effects of the SV-wave and the different material properties between the top and bottom half-spaces. Note
the very small sharp peaks and valleys on the curves are due
to numerical noise in the calculation.
Wet sands have relatively high Poissons ratios compared
with gas sands (Rutherford and Williams, 1989; Hilterman,
2001). We choose different Poissons ratios to study the AVO
responses of thin gas sands. The Poissons ratio of the bottom half-space was fixed at 2 = 0.25, while the thin layer was
given two Poissons ratios: 2 = 0.1(gas sand) or 2 = 0.4(wet
sand). Figures 13 and 14 show the influence of Poissons ratio
on the AVO responses for d = 2 to d = 2 /8 (Figure 13) and

FIG. 11. Maximum absolute amplitudes of the thin-sand reflections as a function of 2 /d for several incident angles. The
sand layer and bottom half-space are elastic with Poissons ratio
= 0.25.

FIG. 12. Maximum absolute AVO for the thin-sand reflections


for different wavelength/thickness (2 /d) in model III. The
sand layer and bottom half-space are elastic with Poissons ratio
= 0.25.

d = 2 /15 to d = 2 /100 (Figure 14). This influence is large for


d > 2 /8 at moderate or large angle of incidence (1 > about
15 for model III), but is small for either small angle of incidence or d < 2 /15. The increase of amplitude with incident
angle or offset is stronger for small Poissons ratio (gas sand)
than that for large Poissons ratio (wet sand), which is similar
to the AVO responses of a single interface for class 3 gas sand
(e.g., Rutherford and Williams, 1989; Hilterman, 2001). The
influence of Poissons ratio on the AVO responses is also dependent on the thickness of the layer and reaches its maximum
at the tuning thickness (d = 2 /4). The combination of two increasing AVO effects (interference and small Poissons ratio)
may results in a sizeable increase for AVO response in thin
gas sand. The influence of Poissons ratio on AVO responses
is both large and complex for the postcritical angle incidence
(bottom interface) because shear waves play a dominant role
for large incidence angles (1 > c2 ).

FIG. 13. Maximum absolute AVO for the thin-sand reflections


( = 0.1 for solid and 0.4 for dashed lines) for d = 2 to d = 2 /8
in model III.

FIG. 14. Maximum absolute AVO for the thin-sand reflections


( = 0.1 for solid and 0.4 for dashed lines) for d = 2 /15 to
d = 2 /100 in model III.

Thin Bed AVO

1167

CONCLUSIONS AND DISCUSSIONS

REFERENCES

For exploration geophysics, seismic AVO now is one of the


major criteria for recognizing potential hydrocarbon reserves.
However, traditional AVO analysis is based on the Zoeppritz equations and only contains single-interface information.
Many observed seismic attributes cannot be explained by this
kind of oversimplified model [for example, the low-frequency
shadow that appeared at the bottom reflection from a gas sand
(e.g., Taner et al., 1979)]. On the other hand, seismic stratigraphy (e.g., Payton, 1977; Anstey, 1982) studies the seismic reflection patterns to identify the conditions under which the rocks
were deposited. These mainly contain the volume scattering
information within depositional sequences. The volume information is more useful than just a single interface reflection
for reservoir characterization because it carries stratigraphic
structure, lithological change, and pore fluid information within
depositional sequences. However, the multiple scattering of interfering seismic waves within a depositional sequence or fractured reservoir remains poorly understood (Liu and Schmitt,
2002). The results of a single thin bed given above have significant implications on seismic stratigraphic interpretation using
amplitude. We believe that the integration for well log analysis, seismic AVO, and seismic stratigraphy, as well as giving us
more physical insights into wave scattering within reservoir and
depositional sequences, will help us to perform more accurate
stratigraphic and lithological interpretations.

Aki, K., and Richards, P. G., 1978, Quantitative seismology: Theory


and methods, vol. I: W. H. Freeman.
Anstey, N. A., 1982, Simple seismics: International Human Resources
Development Corp.
Almoghrabi, H., and Lange, J., 1986, Layers and bright spots: Geophysics, 51, 699709.
Brekhovskikh, L. M., 1980, Waves in layered media: Academic Press.
Chung, H. M., and Lawton, D. C., 1995, Amplitude responses of thin
beds: Sinusoidal approximation versus Ricker approximation: Geophysics, 60, 223230.
Chung, H. M., and Lawton, D. C., 1996, Frequency characteristics of
seismic reflections from thin beds: Can. J. Expl. Geophys., 31, 3237.
de Voogd, N., and den Rooijen, H., 1983, Thin-layer response and spectral bandwidth: Geophysics, 48, 1218.
Gochioco, L. M., 1991, Tuning effect and interference reflections from
thin beds and coal seams: Geophysics, 56, 12881295.
Hilterman, F. J., 2001, Seismic amplitude interpretation: 2001 Soc. Expl.
Geophys./Eur. Assn. Geosci. Eng. Distinguished Instructor Short
Course.
Juhlin, C., and Young, R., 1993, Implication of thin layers for amplitude
variation with offset (AVO) studies: Geophysics, 58, 12001204.
Kallweit, R. S., and Wood, L. C., 1982, The limits of resolution of zerophase wavelets: Geophysics, 47, 10351046.
Koefoed, O., and de Voogd, N., 1980, The linear properties of thin layers, with an application to synthetic seismograms over coal seams:
Geophysics, 45, 12541268.
Lange, J. N., and Almoghrabi, H. A., 1988, Lithology discrimination for
thin layers using wavelet signal parameters: Geophysics, 53, 1512
1519.
Liu, Y., and Schmitt, D. R., 2001, Amplitude and AVO responses of
a single thin bed: Can. Soc. Expl. Geophys. Conv., Expanded Abstracts, 58.
2002, Seismic scale effects: dispersion, attenuation, and attenuation by multiple scattering of waves, Can. Soc. Expl. Geophys.
Conv., Expanded Abstracts, 13.
Neidell, N. S., and Poggiagliolmi, E. 1977, Stratigraphic modeling and
interpretationGeophysical principles, in Payton, C. E, Eds., Seismic stratigraphyApplication to hydrocarbon exploration: Am.
Assn. Petr. Geol. Memoir 26, 389416.
Payton, C. E., Ed., Seismic StratigraphyApplication to hydrocarbon
exploration: AAPG Memoir 26.
Rokhlin, S. I., Xie, Q., Liu, Y., and Wang, L., 1999, Ultrasonic study of
quasi-isotropic composites, in Thompson, D. O., and Chimeti, D. E.,
Eds., Review of progress in QNDE: Plenum Press, 18, 12491256.
Rutherford, S. R., and Williams, R. H., 1989, Amplitudes-versus-offset
variations in gas sands: Geophysics, 54, 680688.
Schmitt, D. R., 1999, Seismic attributes for monitoring of a shallow
heated heavy oil reservoir: A case study: Geophysics, 64, 368377.
Taner, M. T., Koehler, F. and Sheriff, R. E., 1979, Complex seismic trace
analysis: Geophysics, 44, 10411063.
Ursin, B., and Stovas, A., 2002, Reflection and transmission response
of a layered isotropic viscoelastic media: Geophysics, 67, 324325.
Widess, M. B., 1973, How thin is a thin bed?: Geophysics, 38, 11761180.
Yilmaz, O., 1987, Seismic data processing: Soc. Expl. Geophysics.

ACKNOWLEDGEMENTS

The authors thank Guoping Li, Encana, for his suggestions regarding thin-layer seismic attributes. The comments
and suggestions of the associate editor (H. W. Swan), the assistant editor (J. M. Carcione), and three reviewers (D. C.
Lawton, C. Ribordy, and an anonymous reviewer) improved
the communication of this paper. This work was sponsored
by the Seismic Heavy Oil Consortium in the Department of
Physics, University of Alberta, and by the contributors of
the Petroleum Research Fund, administered by the American
Chemical Society for partial support of this research.

APPENDIX A
COEFFICIENTS OF REFLECTION AND TRANSMISSION FOR A SINGLE LAYER

A plane harmonic wave with unit amplitude illuminates the


thin layer (Figure 1). For acoustic media, the displacement potentials in the top half-space, thin layer, and the bottom halfspace can be written respectively (e.g., Aki and Richards, 1978)
as

1 = ei(k x1 x+kz1 zt) + R()ei(k x1 xkz1 zt) ,

(A-1)

2 = A()ei(k x2 x+kz2 zt) + R()ei(k x2 xkz2 zt) , (A-2)


3 = T ()ei(k x3 x+kz3 zt) .

(A-3)

where k xi = /c = (/i ) sin i ; k zi = (/i ) cos i ; i = 1, 2, 3; i


and i are the velocities and incident or refracted angles, respectively; and c and are phase velocity and angle frequency,
respectively. R() and T () are the generalized reflection and
transmission coefficients, and A() and B() are the amplitudes of the potential within thin layer. The solution can be

written as the form of the propagator matrix by the continuous


boundary conditions for displacements and pressures on the
top and bottom interfaces (e.g., Rokhlin et al., 1999):

S1 = B S3 ,
B = XD

(A-4)
X

(A-5)

where S1 = (u z1 , p1 ) and S3 = (u z3 , p3 ) are the vertical displacement and pressure vectors at the top and bottom interfaces, respectively. X and D are 2 2 matrixes within the layer,
which can be written as
T

!
2 2
,
X =
2 c 2 2 c 2
!

0
eikz2 d
.
D=
0
eikz2 d

(A-6)

(A-7)

1168

Liu and Schmitt

Substituting equations (A-6) and (A-7) into equation (A-5),


we have

cos(k z2 d)
B=
i2 c2 sin(k z2 d)
2

i2 sin(k z2 d)

2 c 2
,

cos(k z2 d)

(A-8)

q
where 2 = c2 /22 1. The displacements and pressures in the
top and bottom interfaces can be written as

T
S1 = [1 (1 R()], 1 c2 [1 + R()] , (A-9)

T
(A-10)
S3 = 3 T (), 3 c2 T () ,

q
where i = c2 /i2 1(i = 1 and 3). Substituting equations
(A-8), (A-9), and (A-10) into equation (A-4) and solving equation (A-4) for R()and T (), we have

T ()
2Z 1 Z 2

= 2
.
i Z 2 + Z 1 Z 3 sin(k z2 d) + (Z 1 Z 2 + Z 2 Z 3 ) cos(k z2 d)
(A-14)
If the material properties for the top and bottom half-spaces
are identical, we have

i Z 22 Z 12 sin(k z2 d)

,
R() = 2
i Z 1 + Z 22 sin(k z2 d) + 2Z 1 Z 2 cos(k z2 d)
(A-15)

i 22 1 3 1 3 22 sin(k z2 d) + (2 3 1 2 1 2 2 3 ) cos(k z2 d)

,
R() = 2
i 2 1 3 + 1 3 22 sin(k z2 d) + (2 3 1 2 + 1 2 2 3 ) cos(k z2 d)

(A-11)

21 2 1 2

.
T () = 2
2
i 2 1 3 + 1 3 2 sin(k z2 d) + (2 3 1 2 + 1 2 2 3 ) cos(k z2 d)

(A-12)

Assuming Z i = i i / cos i (i = 1, 2, 3) and applying sin 1 /


sin 2 = 1 /2 , sin 1 / sin 3 = 1 /3 , and sin 2 / sin 3 = 2 /3 ,
we have

R()

i Z 22 Z 1 Z 3 sin(k z2 d) + (Z 2 Z 3 Z 1 Z 2 ) cos(k z2 d)

,
= 2
i Z 2 + Z 1 Z 3 sin(k z2 d) + (Z 1 Z 2 + Z 2 Z 3 ) cos(k z2 d)
(A-13)

2Z 1 Z 2

.
T () = 2
i Z 1 + Z 22 sin(k z2 d) + 2Z 1 Z 2 cos(k z2 d)
(A-16)
Equations (A-11A-16) are the different forms for the generalized reflection and transmission coefficients for a single
layer in acoustic case. Brekhovskikh (1980) also derived similar
forms by two kinds of different methods (i.e., input impedance
and multiple superposition).

Das könnte Ihnen auch gefallen