Você está na página 1de 5

Multiscale modeling of flow and transport in fractured porous

media via the lattice Boltzmann method

Dongxiao Zhang
Mewbourne School of Petroleum and Geological Engineering, University of Oklahoma, Norman, USA

ABSTRACT: Our physical observations and theoretical treatments of flow and transport in fractured porous
media are usually associated with various length scales: pore- (microscopic), local- (macroscopic), fracture-,
and field (megascopic)-scales. Dominant processes and governing equations may vary with scales. Extending
from one scale to another requires upscaling (or downscaling). This allows the essence of physical processes at
one level to be attributed at a coarser (or finer) level. In this conference proceedings paper, we review some of
our recent work in modeling flow and transport in fractured porous media at coupled scales by the method of
lattice Boltzmann and in upscaling from the microscopic to the macroscopic scale.

1 INTRODUCTION In this study, we quantify such interactions with


detailed, pore scale simulations of flow and solute
Flow and solute transport in naturally fractured porous transport in porous media with a single fracture of
media have been studied extensively due to their varying apertures with the lattice Boltzmann (LB)
importance in water resources management, contam- method. The LB simulation can provide us with
inant migration in aquifers, oil/gas production, and detailed flow and solute transport quantities at vari-
radioactive waste repositories [e.g., Bear et al., 1993]. ous scales and allow us to examine the continuous or
Owing to statistically complex distribution of geolog- double-porosity representation of flow and transport
ical heterogeneity and the multiple length and time in fractured porous media.
scales involved, three approaches are commonly used
in describing fluid flow and solute transport in natural
fractured porous formations: (1) discrete fracture mod- 2 MODEL AND THEORY
els, (2) continuum models, including double porosity
models, using effective properties to represent the The LB method is a tool for simulating fluid flows
effects of fractures, and (3) hybrid models that com- and modeling physics in fluids. Unlike conventional
bine the above two. In the discrete fracture approach, numerical schemes based on discretizations of macro-
the effect of porous matrix on flow and transport is scopic continuum equations, the LB method is based
usually neglected (e.g., Smith and Schwartz, 1993; on microscopic description and mesoscopic kinetic
Berkowitz and Scher, 1997). A recent work of Kang equations. This feature gives LB method the advan-
et al. (2002a) on flow in a fractured porous system tage of studying non-equilibrium dynamics, especially
reveals that this may not be a good assumption when in the problem of acid stimulation involving inter-
the matrix permeability is finite. In this case, the matrix facial dynamics and complex geometries. Since its
flow has significant effects on the fractured system and introduction as a tool, the LB method has been
the assumption that the matrix is impermeable does proven to be competitive in studying a variety of flow
not hold, and hence the use of the cubic law to calcu- and transport phenomena (Chen and Doolen, 1998),
late the fracture permeability may cause a significant including single- or multi-phase flow (Gunstensen and
error. When solute transport is considered in such a Rothman, 1993; Zhang et al., 2000) and chemical
fractured porous system, it is more important to take dissolution/precipitation (Kang et al., 2002b, 2003)
into account the contribution of the porous matrix, and in porous media. The LB approach is, at present, the
it is essential to represent the interactions between the only existing computational method that is capable of
factures and the matrix blocks in the continuum and capturing the details of porous media geometry, and
hybrid approaches (e.g., Warren and Root, 1963;Tsang all of the relevant physical and chemical processes,
et al., 1996; Lee et al., 2001). thus providing necessary data at both the microscopic

443
Copyright 2005 Taylor & Francis Group plc, London, UK
similar to that of fi . It has been shown using the
Chapman-Enskog expansion technique that such a LB
transport scheme recovers the convection-diffusion
equation successfully.
The LB method can be modified to include sur-
face reactions for simulating dissolution/precipitation
in 2D porous media (Kang et al., 2002a, 2003). In
this approach, the surface reactions are accounted for
as boundary conditions to the LB transport equation.
The resulting evolution of pore geometries is coupled
with the LB flow model by updating the geometries at
each time step. Then, the new flow field is passed to the
transport equation.As such, the coupling of flow, trans-
port, and reactions is achieved, leading to the dynamic
evolution of pore geometries.

Figure 1. CMT geometry of sandstone of porosity 16%. 3 MULTISCALE MODELING OF FLOW

and macroscopic scales. Shown in Figure 1 is one of We have developed a unified microscopic/macroscopic
such rock geometries obtained by computed micro- porous media method capable of simulating flow in
tomography at the resolution of microns, on which we various-length-scale porous systems and in systems
have performed detailed single and multi-phase LB where multiple-length scales coexist (Kang et al.,
simulations (Zhang et al., 2000). 2002a). The governing equations, omitted here, are
In the LB method, the primitive variable is the parti- similar to (3)(4) with an additional term to account
cle distribution function, which satisfies the following for resistance in the porous matrix. Application of
evolution equation (e.g., Rothman and Zaleski, 1997; this method to a unidirectional steady flow through
Chen and Doolen, 1998): a homogeneous and a heterogeneous porous medium
recovered Darcys law when the effects of inertial
forces and the Brinkman correction can be neglected.
Simulations performed on a fractured porous system
indicated that as far as the overall permeability of the
where fi is the particle velocity distribution function system was concerned, the current method gave a very
along the i direction, t is the time increment, is good result and that this method is capable of handling
the relaxation time relating to the kinetic viscosity by fractured systems with large length-scale spans.
eq
v = ( 0.5)RT , fi is the corresponding equilibrium Here we show an example of flow through a
distribution function, and e i s are the discrete veloci- bimodal heterogeneous porous medium generated by
ties. The density and velocity of the fluid are calculated using a two-stage procedure. The field is 16 m 16 m,
from the particle distribution function via grid size is 81 81. The permeability field is com-
posed of two permeability populations: one with a
low mean permeability (the red material region),
and the other with a high mean value (the blue
region). The permeability also varies within each
It is well known that, using the Chapman-Enskog
region. Overall, the maximum value of the perme-
expansion, the LB equation (1), recovers the correct
ability is 6.07 1013 m2 , and the minimum value
continuity and momentum equations at the Navier-
is 8.46 1017 m2 . The pressure (head) difference
Stokes level (e.g., Chen and Doolen, 1998),
between the entrance and the exit is 0.5 m. The den-
sity and viscosity of the fluid are 997.81 kg/m3 and
1.00246 103 kg/m/s. There is no analytical solu-
tion for flow through such a porous medium because
of the heterogeneity of the medium, even though the
flow is assumed to be steady and Darcys law is satis-
fied. To verify the validity of the method in simulating
where p = RT is the fluid pressure. flow through such a medium, we compared our simu-
The transport of dissolved chemicals having low lation results with those obtained by using the Finite-
concentrations may be described by another distribu- Element Heat- and Mass-Transfer code (FEHM). In
tion function gi . This satisfies an evolution equation both simulations, grid size is 81 81; pressure (head)

444
Copyright 2005 Taylor & Francis Group plc, London, UK
16
k 1
34.869
34.3579
30.8353 current method
12.3452
12 5.84217 FEHM
0.8
3.64783
1.22104
0.397418

(p-po)/(pi-po)
0.238475
0.237247 0.6
8 0.21838
0.181986
0.170097
0.158534
0.13355 0.4
0.121968
0.10232
4 0.0792314
0.0355692
0.0156639 0.2
0.0132229
0.0127849

0
0 4 8 12 16 0
0 4 8 12 16
x (m)
Figure 2. The permeability distribution and the pressure
contours obtained by current method (solid lines) and FEHM
(dashed lines). The field is 16 m 16 m. The pressure (head) Figure 3. The dimensionless pressure profile at center line
is 0.5 m at entrance, and 10 m at exit. The permeability has the (y = 8 m) obtained by the current method (solid line) and
unit m2 , and its values are multiplied by 1014 . The Reynolds FEHM (dashed line). pi is the pressure at the entrance, and
number in the current method is 5.96 103 . po is the pressure at the exit.

is specified at entrance and exit; and velocity along


the normal direction of the other two boundaries is set
to zero.
Figure 2 shows the permeability distribution and
the contours of pressure (head) obtained by these two
methods. The permeability values are multiplied by
1014 . The solid lines are the results obtained with
the current method, and the dashed lines are results
obtained with FEHM.The Reynolds number in the cur-
rent method is Re = 5.96 103 . Apparently, at this
Reynolds number, the pressure contours of these two
methods agree with each other very well, and they
are both mainly distributed in two regions where the
permeability values are relatively small. This is cor-
rect since in such regions, the resistance to the flow is
larger and, accordingly, greater pressure drop is needed
to achieve the overall flow rate; hence the pressure
contours are denser than elsewhere. Figure 3 shows
the dimensionless pressure profile along the horizon-
tal center of the domain (at y = 8). The results of these Figure 4. A simple fractured porous medium.
two methods have good agreement.
Figure 4 shows a simple fractured system where a of the porous matrix varies from 0.04 to 0.601, the
fracture is embedded in a porous medium of poros- pore width ranges from 0.0625 to 0.5625 mm, and the
ity . In this particular case, the fracture is of width matrix permeability is one to four orders of magnitude
h = 32 (in lattice unit, lu) compared to the total width less than the corresponding fracture permeabilities.
of 600 lu for the whole system, and the matrix poros- The flow in the system where the fracture and the pore
ity is = 0.601. To explore the effects of the matrix scales coexist can be simulated with two approaches.
porosity (and thus the pore diameter and matrix One is to simulate the system with the conventional
permeability) as well as the fracture width on solute LB model which essentially solves the Navier-Stokes
transport, these parameters and the size of the fractured equation, and the other is to use the unified multiscale
porous system are varied. In the first group of simula- LB model of Kang et al. (2002a) where the fracture is
tions, the fracture aperture is h = 32 lu, corresponding recognized as pore space and the porous matrix is aver-
to 1 mm in linear dimension; in the second group, aged and accounted for with a resistance term in the
the aperture is h = 52 lu or 1.625 mm. The porosity same equation. Good agreements have been founded

445
Copyright 2005 Taylor & Francis Group plc, London, UK
10

*
4

Figure 5. Snapshots of concentration fields for three scenar- 2 =0.601,h=32


ios of matrix porosity () at two dimensionless times, which 1
=0.453,h=52
=0.308,h=32
are defined as the plume residence time normalized by the =0.191,h=52
=0.069,h=32
characteristic time tc = h2 /Dm , where h is the channel width 0

and Dm is the molecular diffusivity. The Peclet number is 0 0.05 0.1 0.15
t*
51.35.
Figure 6. Time evolution of the mass transfer coefficient
between these two approaches although the latter is at Pe = 51.35. The coefficient and time are normalized with
more efficient computationally. respect to tc .

4 UPSCALING OF TRANSPORT
uniform spatially and integrating the relationship over
Figure 5 shows snapshots of the concentration field space leads to
at two dimensionless times for three scenarios (with
matrix porosity = 0, 0.308, and 0.601 and fac-
ture aperture of 1 mm). Time is made dimensionless
with characteristic time tc , defined as tc = h2 /Dm ,
where h is the channel (fracture) width and Dm is the
molecular diffusivity. The Peclet number (defined as
Pe = Uh/Dm ) is 51.35. When the pore matrix becomes where Mf and Mm are the respective total solute mass
permeable, the plume transfers some of its mass into in the facture and the matrix systems and Mf = dMf /dt
the matrix as it migrates along the facture and some of is the rate of change of mass in the facture.
the mass in the matrix transfers back into the fracture Figure 6 shows the time evolution of the dimension-
after the main plume passes, creating a long tail. It is less mass transfer coefficient = tc for different
seen that both the mass transfer and plume tailing in matrix porosity values and facture apertures. It is seen
the facture are enhanced as increases. As such, the that with time the coefficient goes to a constant for
overall dispersion of a plume increases as the matrix all matrix porosity values under consideration. The
becomes more permeable. The spatial moments of the asymptotic mass transfer coefficient increases as the
plume can be analyzed on the basis of these detailed matrix porosity. The time taken for the coefficient to
concentration fields. reach a constant also increases with the porosity. The
Alternatively, the fractured porous media may mass transfer between the fracture and the matrix is
be conceptualized with dual-porosity and/or double- usually non-negligible even with porous matrix of low
permeability models (e.g., Warren and Root, 1963; porosity (e.g., 4% and 6.9%) and thus low permeabil-
Gerke and van Genuchten, 1993). In such models, the ity, in which the contribution of the matrix flow to the
mass transfer between the facture and matrix systems fractured system may be safely neglected (Kang et al.,
is commonly described by the following first-order 2002a).
model: Figure 7 plots the asymptotic mass transfer coef-
ficients as a function of (h/dg )2 for two fracture
apertures h, where dg is the average grain diameter
(the width of the grain, in our case). It is seen that for
h/dg > 1, the mass transfer coefficient is more or
where cf and cm are the solute concentration in the less a linear function of (h/dg )2 . Our simulations do
facture and the matrix, respectively, and is the mass not cover the region of h/dg < 1, which is not neces-
transfer coefficient. In the double-permeability model, sarily linear. When the fracture is wide compared to
an advection-induced component may be included in the average grain size, one has approximately = c
the mass transfer expression. However, in our simu- (h/dg )2 where c is a dimensionless constant, similar to
lations there is no lateral flow component owing to the formation factor, which depends on porosity, tor-
the specific boundary conditions. Assuming to be tuosity, and other factors. It is seen from Figure 7 that

446
Copyright 2005 Taylor & Francis Group plc, London, UK
2 media is non-Gaussian with long tails. The long tailing
stems from mass transfer between the fracture and the
porous matrix and from the contrast in flow velocities
1.6 in the two media. It was shown that when the frac-
ture aperture is sufficiently large, the mass transfer
coefficient is proportional to the matrix diffusivity and
1.2 inversely proportional to the square of the grain size
of the porous matrix. For such cases, the mass trans-
*

h = 32 fer coefficient is independent of the fracture aperture.


0.8 h = 52 Even for low porosity and low permeability matrix
whose contribution to flow in the fractured porous
media can be neglected, mass transfer between the
0.4
fractures and the matrix is usually non-negligible.

0
1 2 3 4
h2/dg2 REFERENCES

Figure 7. Dependency of dimensionless mass transfer coef- Bear, J., Tsang, C.F. & de Marsily, G. 1993. Flow and
ficient on h2 /dg2 . Contaminant Transport in Fractured Rock, Academic,
San Diego, Calif.
Berkowitz, B. & Scher, H. 1997. Anomalous transport in
the slope c is more or less independent of the fracture random fracture networks, Phys. Rev. Let., 79(20), 4038.
aperture h. In dimensional terms, one has Chen, S. & Doolen, G. 1998. Lattice Boltzmann Method for
Fluid Flows, Annu. Rev. Fluid Mech., 30, 329.
Gerke, H.H. & van Genuchten, M.T. 1993. A dual-porosity
model for simulating the preferential movement of water
and solutes in structured porous media,Water Resour. Res.,
29, 305.
where cDm may be regarded as the matrix diffusivity, Haggerty, R.H. & Gorelick, S.M. 1995. Multiple-rate mass
defined as the molecular diffusivity Dm in free water transfer for modeling diffusion and surface reactions in
corrected by the formation factor c. It is thus seen media with pore-scale heterogeneity, Water Resour. Res.,
that when the fracture aperture is sufficiently large, the 31, 2383.
mass transfer coefficient is proportional to the matrix Kang, Q., Zhang, D. & Chen, S. 2002a. Unified lattice
Boltzmann method for flow in multiscale porous media,
diffusivity and inversely proportional to the square Phys. Rev. E., 66, 056307.
of the average grain size of the porous matrix but is Kang, Q., Zhang, D., Chen, S. & He, X. 2002b. Lattice
independent of the fracture aperture. Equation (7) is Boltzmann simulation of chemical dissolution in porous
consistent with the finding of Haggerty and Gorelick media, Phys. Rev. E, 036318.
(1995) for diffusion into layers. Kang, Q., Zhang, D. & Chen, S. 2003. Simulation of Disso-
lution and Precipitation in Porous Media, J. Geophysical
Research Solid Earth, 108(B10), 2505.
5 CONCLUSIONS Lee, S.H., Lough, M.F. & Jensen, C.L. 2001. Hierarchical
modeling of flow in naturally fractured formations with
multiple length scales, Water Resour. Res., 37, 443.
The lattice Boltzmann method is not only suitable for Smith, L. & Schwartz, F.W. 1993. Solute transport through
simulating flow in detailed pore geometries but also fracture networks, in Flow and Contaminant Transport
for flow in various-length-scale porous systems and in in Fractured Rock, edited by J. Bear, C.F. Tsang, and
systems where multiple-length scales coexist. Detailed G. de Marsily, pp. 129167, Academic, San Diego, Calif.
lattice Boltzmann simulations of flow and solute trans- Tsang, Y.W., Tsang, C.F., Hale, F.V. & Dverstorp, B. 1996.
port were performed on simply constructed porous Tracer transport in a stochastic continuum model of
media embedded with a fracture. The effects of the fractured media, Water Resour. Res., 32, 3077.
porosity (and thus permeability) of the porous matrix Warren, J.E. & Root, P.J. 1963. The behavior of naturally
and the fracture aperture on transport were investi- fractured reservoirs, Soc. Petrol. Eng. J., 3, 245.
Zhang, D. & Kang, Q. 2004. Pore scale simulation of solute
gated on the basis of these simulations. Both the flow transport in fractured porous media, Geo phys. Res. Lett.
quantities and the mass transfer coefficients between 31, L12504.
the facture and the porous matrix were analyzed. It Zhang, D., Zhang, R., Chen, S. & Soll, W.E. 2000. Pore scale
has been found that unlike transport in channels or in study of flow in porous media: Scale dependency, REV,
pure porous media, solute transport in fractured porous and statistical REV, Geo phys. Res. Lett. 27, 1195.

447
Copyright 2005 Taylor & Francis Group plc, London, UK

Você também pode gostar