Sie sind auf Seite 1von 7

Effects of fluid flow on wave velocity and attenuation in porous rocks

C. Wei
Department of Civil and Environmental Engineering, University of Vermont, Burlington, USA

K.K. Muraleetharan
School of Civil Engineering and Environmental Science, University of Oklahoma, Norman, USA

ABSTRACT: Fluid flow induced by a stress wave dissipates wave energy in fluid-saturated porous rocks,
resulting in intrinsic wave attenuation and velocity dispersion (velocity depending upon frequency). Here the
effects of fluid flow at different scales are analyzed, and a porodynamic model is presented. The model depends
only on measurable poroelastic parameters and the time for fluid pressure to equilibrate. The effects of local
fluid flow are taken into account by virtue of the concept of dynamic compatibility conditions on the interfaces
among coexisting individual phases. The model explains the measured acoustical data for both fully and partially
saturated rocks over a broad range of frequencies and offers a unique way to evaluate the frequency-dependent
behavior of fluid-saturated porous rocks.

1 INTRODUCTION generates fluid-pressure gradient between the peaks


and troughs of the wave, driving macroscopic fluid
Fluid flow induced by a stress wave dissipates energy flow across the averaging volume. At pore scales,
in fluid-saturated porous rocks, resulting in intrin- due to existence of grain microcracks (OConnell &
sic wave attenuation and velocity dispersion (velocity Budiansky 1974, Mavko & Nur 1979) and broken
depending upon frequency) (Winkler & Nur 1979, grain contacts (Murphy et al. 1986), the fluid is
Winkler 1985, Murphy 1982, Jones 1986). A stress squeezed out of the microcracks and contacts into
wave induces fluid flow in three different regions, the nearby rigid pores, leading to microscopic (squirt)
defined by the distance over which the generated fluid fluid flow (Mavko & Nur 1979). The lithologic com-
pressure attempts to equilibrate by diffusion (Pride position in an averaging volume can be heterogeneous,
et al. 2003). Macroscopic fluid flow results from the some regions being very compliant while others being
fluid-pressure gradient developing between the peaks very stiff. Higher fluid pressure is generated in the
and troughs of the stress wave (Biot 1956a, b). Fluid compliant regions in response to the passing stress
flow also occurs at a scale much smaller than the wave- wave, inducing fluid flow at a scale (meso-scale)
length, such as microscopic (pore) scales (Biot 1962, comparable to the size of the averaging volume.
OConnell & Budiansky 1974, Mako & Nur 1979) Mesoscopic flow can also occur in the rocks with het-
and mesoscopic scales (Akbar et al. 1994, Pride et al. erogeneously distributed fluid content (Akbar et al.
2003). Earlier investigations were based on the details 1994, Cadoret et al. 1998).
of flow-related local structure of rocks that are diffi-
cult to quantify (Dvorkin & Nur 1993, Johnson 2001,
Pride & Berryman 2003). Here we analyze the effects
of fluid flow in various regions, and present a poro- 2 DYNAMIC COMPATIBILITY CONDITIONS
dynamic model that depends only upon measurable ON INTERFACES
poroelastic parameters and the time for fluid pressure
to equilibrate. Our model explains the measured seis- To quantify the effects of fluid flow at various scales,
mic data for both fully and partially saturated rocks we first analyze the micro interactive forces in an aver-
over a broad range of frequencies and offers a unique aging volume. Due to the motion of fluid with respect
way to evaluate the frequency-dependent behavior of to solid grains, a viscous boundary layer begins to
fluid-saturated porous rocks. develop in the pores at a certain frequency, producing
Consider a representative averaging volume for micro drag forces tangentially acting on the boundary
a rock with a size much smaller than the wave- of the solid grains. The frequency for the onset of the
length and much larger than the pore size. When a boundary layer is known to approximately equal the
compressional (P) wave propagates through a rock, it characteristic frequency at which maximum energy

181
Copyright 2005 Taylor & Francis Group plc, London, UK
dissipation occurs due to macroscopic (wavelength- upon permeability k, fluid viscosity f , porosity n, and
scale) fluid flow (Pride et al. 2003). We therefore saturation (Bear 1972); coefficient Z f = f f and f
suggest that these drag forces are induced by the represents the change in ( p f pS ) induced solely by
macroscopic flow and thus can be characterized using a unit variation of n f at very low frequencies. The fre-
Biots theory (Biot 1956a, b, 1962). In addition, the quency dependence of f has been derived by many
pressure compatibility condition across the bound- authors (Biot 1962, Norris 1986, Johnson et al. 1987,
ary of a solid grain requires that the micro pressure Gist 1994) and shall not be repeated here.
pSm of the grain be equal to the micro fluid pressure
f
pm , where f = W (wetting fluid) and N (non-wetting
fluid) at partial saturation, and f = W at full satura-
tion. These compatibility conditions cannot be directly 3 CONSTITUTIVE EQUATIONS
averaged onto the continuum level to yield p f = pS
(Gray 2000), where pS and p f are the averaged solid 3.1 State equations
and fluid pressures, respectively. Rather, the pressure The total stress tensor of the mixture as a whole can
differences ( p f pS ) are nonzero rate-sensitive quan- be decomposed as (Wei and Muraleetharan 2002a)
tities. These dynamic compatibility conditions on the
interfaces among individual constituents play a key
role in the following derivations.
At very low frequencies, fluid pressures have suffi-
cient time to equilibrate due to local pressure gradients, where 1 is the second-order unit tensor with compo-
and the quantity ( p f pS ) always attains its equilib- nents ij ; S is the intrinsic stress tensor of the solid
rium value. At high frequencies, however, fluid pres- phase, which has two contributions: One is associated
sures do not have sufficient time to equilibrate leading with the compression of the solid grains and the other
to local pressure gradients. The quantity ( p f pS ) is associated with the deformation of the solid matrix.
then tends to relax to an equilibrium value. This cap- Formally, we have
illary relaxation process attenuates wave energy. We
suggest that such a process is governed by mesoscopic
and/or microscopic fluid flow, depending upon the
wave frequency. Accordingly, the capillary relaxation
time f is the same as the time for fluid pressure to where is given by
equilibrate by diffusion. The time for fluid diffusion
is evaluated by 2c /Df , where c is the characteristic
length of (meso or micro) fluid flow and Df is the
relevant fluid-pressure diffusivity.
The rate of energy dissipation due to capillary AS is the free energy density of the solid phase and it
relaxation and viscous dragging is given by (Wei & is assumed to be a function of and S ; where is the
Muraleetharan 2002a) infinitesimal strain tensor of the rock matrix. Other
state equations are given by

where n f is the rate of change of the fluid volume


fraction; w f the relative velocity of a fluid with respect
to the solid phase; f the effective (averaged) drag and
f the nonequilibrium capillary pressure and
force; "

3.2 Linear constitutive equations


Here c is the effective intrinsic mass density of con- Because the acoustical behavior of porous media is of
stituent c (c = s, f ); A f is the free energy density of a main concern here, it is sufficient to consider the linear
fluid and it is assumed to be a function of n f and f . problem. Linearizing (5)(7) yields
In this paper, we assume that porous media are sta-
tistically isotropic. Considering linear dissipation and
according to (1), we assume as a first approximation
that f = f w f and " f = Z f n f , where f depends

182
Copyright 2005 Taylor & Francis Group plc, London, UK
for partially saturated porous media. Coefficient G is
the shear modulus of the solid matrix and G = nS0 S .
Other material coefficients with a caret in (16)(20),
and though not explicitly given here, depend on the bulk
modulus KS of the solid material, the bulk modulus of
the fluid(s) (KW , KN ), the tangent properties (S , S ,
, W , N ), capillary relaxation time (W , N ), and
angular frequency . In general, these coefficients are
where quantities with subscript 0 are the values at complex numbers.
the initial state, which is assumed to be equilibrium.
The linear form of (2) is

3.3 Material parameters


Hence the dynamic (nonequilibrium) pressure differ- For fully saturated porous media, material properties
ence can be written as to be determined include KS , KW , S , S , , W , and
W . For partially saturated porous media, in addition to
these properties, it is necessary to determine KN , N
Assume that state variables have time dependence and N . All these tangent properties are independent
as exp(it), where is the characteristic angular of frequency. Hence, these unknown properties can be
frequency and i2 = 1. It follows from (12) that determined by performing static tests ( 0). Lets
say that shear modulus G and bulk moduli of the solid
and fluid(s) are known. The evaluation of S , , W
(and N ), and W (and N ) is shown below.
For later reference, the linear mass balance of First consider the fully saturated case. Define
constituent c (c = S, f ) is introduced as

Physically, f is the volume of a fluid entering the pores


c
where u is the displacement vector. Particularly, it is in a unit volume of bulk material. Using (3), (16), and
noted that = uS . The initial porosity n0 is given (17), we derive
by

Mass densities S and f and volume fraction n f


can be eliminated from (8)-(10), (13) and (14). Using
(4) and (15), we derive where K U is the complex undrained bulk modulus, M 
is the complex fluid storage coefficient, and 
B is the
complex effective stress coefficient. They are given by

for saturated porous media, and

D is the complex undrained bulk modulus,


where K

183
Copyright 2005 Taylor & Francis Group plc, London, UK
and change H as well as the saturation capacity c . For
partially saturated rocks, we can show that

To determine the material properties, static com-


pression tests with different jacketed and drainage and
conditions can be performed. These tests are used
to determine static poroelastic parameters including
drained bulk modulus KD , undrained bulk modulus
KU , effective stress coefficient B , pore compress-
ibility Cn , Bishop pore pressure coefficient B, and Parameters S and can be obtained by solving
unjacketed compressibility CS (Kmpel 1991). These (34) and (35). Again, parameter S is calculated from
parameters are related to each other by the following S = G/nS0 .
relationships (Wei and Muraleetharan 2004): Finally, we evaluate capillary relaxation time W
(and N ). These time parameters are related to local
fluid flow. Indeed, they are equal to the characteris-
tic times of local fluid flow (Wei and Muraleetharan
2004). For example, for fully saturated conditions, if
the local flow is dominated by microscopic squirt flow,
we have

where w is the dynamic viscosity of the fluid and is


the typical aspect ratio of the microscopic apertures.
If the local flow is represented by meso-scale flow,

where Kn = Cn , and Kn = 1/CS . It is very important


to note that in general, because of (29)(32), there are
only four independent parameters among KD , KU , B , where m is the characteristic size of meso-scale
Cn , B, CS and G. For locally homogeneous porous heterogeneity.
media, the number of independent poroelastic parame-
ters reduces to three. In this case, we can show that W
vanishes, and the model predicts that the viscoelastic 4 EXAMPLES
effects associated with local fluid flow are negligible
(Wei and Muraleetharan 2004). This is consistent with Figure 1 illustrates the velocity dispersion and attenu-
experimental observations (Winkler 1985, Gist 1994). ation of P1-wave in the fully saturated Berea sand-
Once the four independent poroelastic parame- stone for three different relaxation times, showing
ters (including G) are obtained, they can be used to that increasing W (or W ) continuously shifts the
calculate the tangent material parameters. Note that attenuation peak to lower frequencies. The rightmost
attenuation peak is attributed to macroscopic flow
(Biot flow). Clearly, our model captures potential
energy loss due to fluid flow in fully saturated rocks at
any frequency from the seismic range to the ultrasonic
Now Eqs. (24)(28) can be solved to yield the values range, assuming that the wavelength is much larger
of S , , and W . Parameter S is calculated from than the pore size.
S = G/nS0 . In reality, rocks can have a capillary relaxation spec-
Next consider the partially saturated case. A trum, corresponding to various mechanisms of fluid
detailed procedure to evaluate parameters W and N flow operating at different frequencies. As shown in
was presented in Wei and Muraleetharan (2002b). Here Figure 2, two peaks were observed in the bulk atten-
it is necessary to determine the constant-suction mod- uation QK1 versus frequency for the fully saturated
ulus KM and the bulk modulus associated with suction Harz quartzite (Paffenholz & Burkhardt 1989). The

184
Copyright 2005 Taylor & Francis Group plc, London, UK
a 80
4,500
Phase velocity [m/s]

= 0.002 ms
4,000 250 s
10 s 60

3,500 0.1 s

1,000/Q
0.001 s
40
3,000

2,500 20
b 400

300 0
10 s 0.1 s = 0.001 s -3 -2 -1 0 1 2 3 4
1,000/Q

200 Frequency [log(Hz)]

Figure 2. Bulk attenuation QK1 versus frequency for the


100 fully saturated Harz quartzite with two distinct relaxation
Biot loss times.
0
-1 0 1 2 3 4 5 6 7
Frequency [log(Hz)]
distributed in the pores and no significant mesoscopic
liquid flow occurs and W 0 s. If saturation is not
Figure 1. (a) Phase velocity and (b) attenuation versus
frequency of P1-wave in the fully saturated Berea sandstone. extremely high, due to high compressibility and mobil-
ity of the gas, we also expect that N 0 s. This set of
(W , N ) leads to the lower bound of wave velocity,
since at the same saturation the rock with uniformly
model predicts that the attenuation peak at 0.03 Hz distributed fluid has the smallest rigidity. For the rocks
corresponds to W = 250 s and the peak at 1.5 kHz saturated through the drying process, if the wetting
corresponds to W = 0.002 ms. The energy loss at the fluid patches are fully undrained, the rock achieves
lower frequency is attributed to mesoscopic flow (the the largest rigidity. From these remarks, it is expected
so-called Biot-Gardner effect). From (37), we obtain that, at low to moderate saturation, the upper bound
m 12 cm, which is two times larger than the diame- of velocity is obtained by setting (W , N ) = (, 0) s,
ter (5 cm) of the sample. On the other hand, the energy and for high saturation, the upper bound of velocity
loss at the higher frequency is associated with micro- can be calculated with (W , N ) = (, ) s. Figure 3a
scopic flow and the typical aspect ratio of apertures is illustrates that the model accurately predicts the wave
estimated using (36) as 6 104 . velocity bounds for Estaillades limestone (Cadoret
In partially saturated porous media, capillary relax- et al. 1995).
ation can be induced by mesoscopic flow of the The model predicts very well the wave velocity
non-wetting fluid due to existence of heterogeneous in uniformly saturated rocks by setting W 0 s and
fluid distribution in the pores such as the patchy sat- N 0 s, but significantly underestimates the attenua-
uration. The pattern of fluid distribution in a rock tion (not shown here). The discrepancy is attributed to
depends upon the saturating path (Cadoret et al. 1995): the possible intrinsic material heterogeneity existing
a wetting process used to saturate a rock results in in the porous material. At any saturation, reasonable
homogeneous saturation, whereas a drying process values of W and N should be those at which both pre-
produces patchy saturation. If the non-wetting fluid is dicted velocity and attenuation approximately equal
a gas, the gas phase is interconnected in the pore space the measured data. Based on this argument, we cal-
at low to moderate saturation. Therefore at low to mod- culate the saturation dependence of attenuation based
erate saturation, due to the high mobility of the gas, N on the velocity data. Figure 3b demonstrates the good
is very small, and the energy loss induced by the local agreement between the predicted attenuations and the
gas flow is negligible. At high saturation, however, measurements for Estaillades limestone. In particu-
the gas is trapped in the wetting fluid, and the coales- lar, our model correctly predicts an attenuation peak
cence and separation of gas bubbles at meso-scale can at high saturation. The attenuation peak at high satu-
attenuate significant wave energy. In this case, N is ration was commonly observed in partially saturated
nonzero. rocks (Murphy 1982, Yin et al. 1992). At 90% sat-
For homogeneous rocks saturated through the wet- uration, KN 105 Pa, kN 103 , N 10 s, and we
ting process, the moisture content is homogeneously estimate m 1 cm. This size is comparable to that

185
Copyright 2005 Taylor & Francis Group plc, London, UK
a ACKNOWLEDGMENTS
3,200
Upper bound
Financial support for this research was provided by the
Phase velocity [m/s]

3,000
U.S. National Science Foundation under grants CMS-
0112950 and CMS-0301457.

2,800
REFERENCES
Lower bound
2,600 Akbar, N., Mavko, G., Nur, A. & Dvorkin, J. 1994. Seismic
80 signatures of reservoir transport properties and pore fluid
b distribution. Geophysics 59: 12221236.
Bear, J. 1972. Dynamics of Fluids in Porous Media. New
60
York: American Elsevier.
Biot, M. A. 1956a. Theory of propagation of elastic waves in
1,000/Q

40 a fluid-saturated porous solid: I. Low-frequency range. J.


Acoust. Soc. Am. 28: 168178.
20 Biot, M. A. 1956b. Theory of propagation of elastic waves in
a fluid-saturated porous solid: II. High-frequency range.
J. Acoust. Soc. Am. 28: 179191.
0 Biot, M. A. 1962. Generalized theory of acoustic propaga-
20 40 60 80 100
Degree of saturation [%]
tion in porous dissipative media. J. Acoust. Soc. Am. 54:
12541264.
Cadoret, T., Marion, D. & Zinszner, B. 1995. Influence of
Figure 3. Variation of (a) phase velocity [the thin
frequency and fluid distribution on elastic wave velocities
solid line using (W , N ) = (0, 0) s; the thin dashed
in partially saturated limestones. J. Geophys. Res. 100:
line using (W , N ) = (, 0) s; and the thick solid line
97899803.
using (W , N ) = (, ) s] and (b) attenuation [dashed
Cadoret, T., Mavko, G. & Zinszner, B. 1998. Fluid distri-
line = predicted wetting process; solid line = predicted drying
bution effect on sonic attenuation in partially saturated
process] with saturation for P1-wave at 1 kHz in Estaillades
limestones. Geophysics 63: 154160.
limestone. Data points are calculated from the measured
Dvorkin, J. & Nur, A. 1993. Dynamic poroelasticity: A uni-
velocities and attenuations of extensional and shear waves
fied model with the squirt and the Biot mechanisms.
(Cadoret et al. 1995, 1998; solid stars = wetting process;
Geophysics 58: 524533.
empty stars = drying process).
Gist, G. A. 1994. Fluid effects on velocity and attenuation in
sandstones. J. Acoust. Soc. Am. 96: 11581173.
observed through the X-ray CAT scan (Cadoret et al. Gray, W. G. 2000. Macroscale equilibrium conditions for two-
1995, 1998). phase flow in porous media. Int. J. Multiphase flow 26:
467501.
Jones, T. D. 1986. Pore fluids and frequency-dependent wave
propagation in rocks. Geophysics 51: 19391953.
5 CONCLUSIONS Johnson, D. L., Koplik, J. & Dashen R. 1987. Theory of
dynamic permeability and tortuosity in fluid-saturated
The model presented here consistently describes the porous media. J. Fluid Mech. 176: 379402.
acoustical behaviour of both fully and partially sat- Johnson, D. L. 2001. Theory of frequency dependent acous-
urated rocks over a broad range of frequencies. The tics in patchy-saturated porous media. J. Acoust. Soc. Am.
model calculations point to the importance of atten- 110: 682694.
Kmpel, H.-J. 1991. Poroelasticity: parameters reviewed.
uation data in extracting the information on the local
Geophysics 105: 783799.
fluid distribution and the local structure of rocks. The Mavko, G. & Nur, A. 1979. Wave attenuation in partially
agreement between the theory and experiments sug- saturated rocks, Geophysics 44: 161178.
gests that the new model captures the sensitivity of Murphy, W. F. 1982. Effects of partial water saturation on
attenuation to local material heterogeneity. In addition, attenuation in Massilon sandstone and Vycor porous glass.
the new model depends only upon measurable mate- J. Acoust. Soc. Am. 71: 14581468.
rial parameters and capillary relaxation times. The Murphy, W. F., Winkler, K. W. & Kleinberg, R. L. 1986.
relaxation times can be evaluated through either the Acoustic relaxation in sedimentary rocks: Dependence
characterization of local structure of rocks or the mea- on grain contacts and fluid saturation. Geophysics 51:
757766.
surements of wave velocity and attenuation. Due to
Norris, A. N. 1986. On the viscodynamic operator in Biots
the similarity of porous rocks and other porous media equation of poroelasticity. J. Wave Material Interact. 1:
in living or non-living matters, our model should have 365380.
broad applicability in analysing the seismic behaviour OConnell, R. J. & Budiansky, B. 1974. Seismic velocities
of fluid-saturated porous media that are ubiquitous on in dry and saturated cracked solids. J. Geophys. Res. 79:
our planet. 54125426.

186
Copyright 2005 Taylor & Francis Group plc, London, UK
Paffenholz, J. & Burkhardt, H. 1989.Absorption and modulus Proceedings of the 2nd Biot Conference on Poromechan-
measurements in the seismic frequency and strain range ics, Grenoble, France, August 2002.
on partially saturated sedimentary rocks. J. Geophys. Res. Wei, C. & Muraleetharan, K. K. 2004. Acoustical char-
94: 94939507. acterization of fluid-saturated porous media with local
Pride, S. R., Harris, J. M., Johnson, D. L., et al. 2003. Per- heterogeneities: I. Material properties. Int. J. Solids Struct.
meability dependence of seismic amplitudes. The Leading (in review).
Edge 22: 518525. Winkler, K. W. & Nur, A. 1979. Pore fluids and seismic
Pride, S. R. & Berryman, J. G. 2003. Linear dynamics of attenuation in rocks. Geophys. Res. Lett. 6: 14.
double-porosity dual-permeability materials: I. Govern- Winkler, K. W. 1985. Dispersion analysis of velocity and
ing equations and acoustic attenuation. Phys. Rev. E 68: attenuation in Berea sandstone. J. Geophys. Res. 90:
036603. 67936800.
Wei, C. & Muraleetharan, K. K. 2002a. A continuum theory Yin, C.-S., Batzle, M. L. & Smith, B. J. 1992. Effects of partial
of porous media saturated by multiple immiscible fluids: liquid/gas saturation on extensional wave attenuation in
I. linear poroelasticity. Int. J. Eng. Sci. 40: 18071833. Berea sandstone. Geophys. Res. Lett. 19: 13991402.
Wei, C. & Muraleetharan, K. K. 2002b. A continuum the-
ory of porous media and its variational structure, in:

187
Copyright 2005 Taylor & Francis Group plc, London, UK

Das könnte Ihnen auch gefallen