You are on page 1of 13

Effect of Diffusion on Dispersion

Raman K. Jha, Steven L. Bryant, and Larry W. Lake, The University of Texas at Austin

Summary Thus, the essential phenomena giving rise to hydrodynamic


It is known that dispersion in porous media results from an inter- dispersion observed in porous media are (1) stream splitting of the
action between convective spreading and diffusion. However, the solute front at every pore, causing independence of particle veloci-
nature and implications of these interactions are not well under- ties purely by convection; (2) velocity gradient in pore throats in
stood. Dispersion coefficients obtained from averaged cup-mixing the direction transverse to flow; and (3) diffusion. Taylors disper-
concentration histories have contributions of convective spreading sion in a capillary tube accounts only for the second and third of
and diffusion lumped together. We decouple the contributions of these phenomena, yielding a quadratic dependence of dispersion on
convective spreading and diffusion in core-scale dispersion and Pclet number. Plug flow in the bonds of a physically representative
systematically investigate interaction between the two in detail. We network accounts only for the first and third phenomena, resulting
explain phenomena giving rise to important experimental observa- in a linear dependence of dispersion on Pclet number. When all the
tions such as Fickian behavior of core-scale dispersion and power- three phenomena are accounted for, we can explain effectively the
law dependence of dispersion coefficient on Pclet number. weak nonlinear dependence of dispersion on Pclet number.
We track movement of a swarm of solute particles through a
physically representative network model. A physically representa- Introduction
tive network model preserves the geometry and topology of the Traditionally, dispersion coefficient is quantified from the flow-
pore space and spatial correlation in flow properties. We developed averaged (cup-mixed) effluent-concentration history obtained for a
deterministic rules to trace paths of solute particles through the core sample (Lake 1989). The dispersion coefficient has contribu-
network. These rules yield flow streamlines through the network tions (in inseparable form) from (1) convective spreading, caused
comparable to those obtained from a full solution of Stokes equa- by variations in path lengths and velocities of solute particles
tion. Paths of all solute particles are deterministically known in the traveling along different streamlines, and (2) molecular diffusion.
absence of diffusion. Thus, we can explicitly investigate purely The nature and consequences of interaction between convective
convective spreading by tracking the movement of solute particles spreading and diffusion are often not well understood. For exam-
on these streamlines. ple, longitudinal dispersion in porous media is often modeled as
Then, we superimpose diffusion and study dispersion in terms
of interaction between convective spreading and diffusion for a DL = Do +  L v , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (1)
wide range of Pclet numbers. This approach invokes no arbitrary
parameters, enabling a rigorous validation of the physical origin of where L is a fundamental property of the medium called dispersiv-
core-scale dispersion. In this way, we obtain an unequivocal, quanti- ity. This equation expresses dispersion as a sum of diffusion and
tative assessment of the roles of convective spreading and diffusion convective spreading.
in hydrodynamic dispersion in flow through porous media. Eq. 1 implies independence of diffusion and convective spread-
Convective spreading has two components: stream splitting and ing. Under what conditions is this approximation valid? Because
velocity gradient in pore throats in the direction transverse to flow. convective spreading is orders of magnitude larger than diffusion,
We show that, if plug flow occurs in the pore throats (accounting only dispersion often is modeled as a purely convective process. Is
for stream splitting), all solute particles can encounter a wide range dispersion in porous media predominantly an artifact of convective
of independent velocities because of velocity differences between spreading (Coats et al. 2004), or are there serious implications of
pore throats and randomness of pore structure. Consequently, plug its interaction with diffusion?
flow leads to a purely convective spreading that is asymptotically The objective of this paper is to explain core-scale dispersion
Fickian. Diffusion superimposed on plug flow acts independently from pore-scale physics. We decouple the contributions of convec-
of convective spreading (in this case, only stream splitting), and, tive spreading and diffusion on core-scale dispersion and system-
consequently, dispersion is simply the sum of convective spread- atically investigate the interaction between the two in detail.
ing and diffusion. In plug flow, hydrodynamic dispersion varies We investigate dispersion by tracking motion of a swarm of
linearly with the pore-scale Pclet number when diffusion is small solute particles through a granular porous medium. We extracted
in magnitude compared to convective spreading. a physically representative network model from a dense random
For a more realistic parabolic velocity profile in pore throats, par- packing of spheres to represent the pore space accurately. We
ticles near the solid surface of the medium do not have independent developed deterministic rules to trace a solute particles path from
velocities. Now, purely convective spreading (caused by a combina- the inlet of the medium to its outlet. Because the rules are deter-
tion of stream splitting and variation in flow velocity in the transverse ministic, the paths of solute particles are completely known in
direction) is non-Fickian. When diffusion is nonzero, solute particles absence of diffusion. It also ensures that the flow paths are revers-
in the low-velocity region near the solid surface can move into the iblethat is, upon reversal of the flow direction, each particle will
main flow stream. They subsequently undergo a wide range of inde- exactly retrace its path back to the inlet, a characteristic of purely
pendent velocities because of stream splitting, and, consequently, convective spreading. Upon flow reversal, convective spreading
dispersion becomes asymptotically Fickian. In this case, dispersion gets canceled and echo dispersion is zero (Hiby 1962; Taylor 1972;
is a result of an interaction between convection and diffusion. This John et al. 2008; Jha et al. 2009). Thus, our algorithm allows us
interaction results in a weak nonlinear dependence of dispersion to investigate convective spreading explicitly and rigorously in a
on Pclet number. The dispersion coefficients predicted by particle realistic pore space. To the best of our knowledge, this has not
tracking through the network are in excellent agreement with the been reported previously.
literature experimental data for a broad range of Pclet numbers. Next, diffusion is superimposed and movement of solute par-
ticles because of the combined effects of convection and diffusion
is monitored. Dispersion is quantified from spatial and temporal
Copyright 2011 Society of Petroleum Engineers
statistics of solute particles. With this framework, we can quanti-
This paper (SPE 115961) was accepted for presentation at the SPE Annual Technical tatively investigate the influence of increasing diffusion on disper-
Conference and Exhibition, Denver, 2124 September 2008, and revised for publication.
Original manuscript received for review 7 July 2008. Revised manuscript received for sion. We explain the origin of core-scale dispersion as a result of
review 11 February 2010. Paper peer approved 3 May 2010. interaction between convection and diffusion.

March 2011 SPE Journal 65


invalid in realistic pore space, even in relatively simple porous
media. Mellor (1989) showed that the topology of the pore-space
network in Finneys dense, random packing of equal spheres is
completely disordered. Picking bond radii randomly from a dis-
tribution disregards spatial correlation in the bond conductances.
Neglecting spatial correlation may fail to account for physically
significant features of porous media and affects macroscopic prop-
erties of the networks (Bryant et al. 1993).
z

We adopt the approach of physically representative network


models, which replicate the pore space more closely. In this work,
we use a computer-generated, dense, random, periodic packing of
10,000 spheres as a model porous medium (Thane 2006). Such
packings capture essential geometric and topologic features of
sediments. The periodicity eliminates artifacts from boundary
walls. The radius and the center coordinates of each sphere are
known, which completely determines the microstructure of the
y x medium (Fig. 1). The radius of each sphere is 2.1918104 m.
The medium is approximately 34 sphere diameters long in the z
Fig. 1A dense, random periodic packing of 10,000 spheres. direction and 17 sphere diameters long in the x and y directions.
The radius of each sphere is 2.1918104 m. The coordinates The porosity of the medium is 36%.
on the x, y, and z axes are in units of 104 m. From these data, we extract a physically representative network
model that has pore bodies located at the same spatial positions
as the pores in the actual medium. Moreover, the bonds con-
Many previous studies [see, for example, Safmann (1959) necting neighboring pores have the same conductances as in the
and Sahimi et al. (1986)] likewise have shown strong arguments actual medium. Thus, a physically representative network model
for the fundamental importance of this interaction. In that sense, preserves the geometry, topology, and spatial correlation in flow
our conclusions are not surprising, but they are reassuring. If one properties. It is 3D and unstructured.
accounts for solute-transport mechanisms rigorously in the pore
space of a disordered granular material, the correct dispersion Evaluating Bond Conductances and Obtaining Flow Rates.
coefficient emerges when the pore-scale results are averaged. Delaunay tessellation is an unambiguous way of dividing the
Our predictions of dispersion coefficients are consistent with the sphere packing into cells by grouping together sets of four nearest
experimental results reported in the literature for a broad range of spheres. Thus, a Delaunay cell is a tetrahedron (Fig. 2a) whose
Pclet numbers, including near independence of dispersion coef- vertices lie at the centers of the four spheres forming that cell.
ficient on Pclet number for a diffusion-dominated regime. Models The interior of the cell encloses a region of void space identi-
more sophisticated than ours [e.g., a direct solution of the Stokes fied as the pore body. The geometric center of the Delaunay cell
(or Navier-Stokes) equation in the pore space of a granular mate- can be considered as the pore center. Each face of a Delaunay
rial coupled with a solution of the convection/diffusion equation] cell is a plane of maximum constriction and represents a narrow
would also predict core-scale dispersion correctly. We propose that entrance to the pore body (Fig. 2b). Because each cell is a tetra-
ours is the simplest model that captures the essential physics. hedron, every pore has four faces or throats connecting it to four
neighboring pores.
Model Development Two pores are connected by a path of converging/diverging
Physically Representative Network Model. Pore-network mod- cross section (shown schematically in Fig. 3a). The flow conduc-
eling is an important tool that provides a link between continuum tance of the path is governed by the narrowest constriction in the
(core) -scale properties of a porous medium and the pore-scale path. Fig. 3b shows a cell face shared by two neighboring cells
physics. It has been used widely to estimate dispersion coefficient (pores). The cell face is the narrowest constriction in the flow path
[see, for example, Bruderer and Bernabe (2001), Bijeljic et al. connecting the two pores. Fig. 3b also shows two circles that can
(2004), and Acharya et al. (2007)]. In a network model, the pore approximate the narrowest constriction. rc is the radius of the larg-
space is discretized into a set of pores (nodes) connected by pore est circle that can be fit in the narrowest constriction, and re is the
throats (bonds). Because it is very difficult to capture the details radius of the circle having the same area as the narrowest constric-
of pore geometry explicitly, most of the network models reported tion. If the constriction radius is approximated by rc, the path (or
in the literature make some simplifying assumptions. Common bond) conductance is underestimated because some area available
assumptions include same length for all the throats and regular for flow in the constriction is not taken into consideration. On the
network lattice. Throat radii often are picked randomly from an other hand, if the constriction radius is approximated by re, the
assumed distribution. All of these assumptions are found to be path (or bond) conductance is overestimated because, for a given

(a) (b)

Fig. 2(a) Delaunay tessellation grouping together four nearest neighboring spheres. The vertices of the tetrahedra correspond
to the centers of the spheres. (b) Delaunay cell in a pack of equal spheres defines pore bodies (cell interior) and throats (void
area in cell faces). S, T, U, and V are the centers of four spheres. The pore body is centered at X, while the throat between grains
U, V, and T is centered at W.

66 March 2011 SPE Journal


Pore throat

Pore 2
rc
Pore 1 re

(a) (b)

Fig. 3(a) A converging/diverging path connecting two neighboring pores. (b) The faces are areas of narrowest constriction
(throats) that connect the cell to neighboring cells. rc is the radius of the largest circle that can fit in the constriction. re is the
radius of the circle having the same area as that of the void space. The arithmetic average of rc and re is a good estimate of the
equivalent radius of the bond that describes its hydraulic conductivity.

area, a circle provides the least resistive path for viscous flow. The the four bonds originating at the center of a pore necessarily overlap
arithmetic average of the two radii is a good estimate of the effec- (Fig. 5). Overlapping bonds may seem unphysical. However, bonds
tive radius of the bond connecting the two neighboring pores, reff = are not to be considered in a completely literal sense. It is a way to
(re+rc)/2 (Bryant et al. 1993). For the purposes of computing flow, model flow from one pore to another. We track particle movement
we replace the converging/diverging geometry of each throat with through the network of overlapping bonds. We neglect momentum
a cylindrical bond of radius reff. This idealization proves to preserve loss and mixing in the overlapping region and discount bond lengths
the essential features of the local flow field while enormously in the overlapping region in calculation of the local Pclet numbers.
simplifying the computation of flow dynamics. This simple model yields an accurate a priori prediction of the
The conductance of the bond connecting the two neighboring permeability of the sphere packing (Bryant et al. 1993).
pores is given by g =  reff4 8l , where  is the fluid viscosity and In the absence of diffusion, a solute particle moves in a bond
l is the distance between the pore centers. parallel to the bonds axis. After reaching the outlet face of the
A conceptual schematic of the network is shown in Fig. 4a bond at a pore center, the particle will enter one of the out-flowing
as a network of electrical resistances. The location of every pore bonds originating at that pore center. One of the most common sim-
body (bond junction) and conductance of bonds connecting it to plifying assumptions made in particle tracking through a network
its neighbors have been calculated. We apply a potential gradient model is the probabilistic choice of an out-flowing bond. A solute
across the network. The side boundaries of the network are sealed. particle arriving at a pore body (junction of bonds) is assigned to
Then, we write the mass-balance equation at each pore. Imposing an out-flowing bond randomly with a flow-rate-weighted probabil-
steady state (no mass accumulation at any pore) results in a set of ity. However, probabilistic choice of out-flowing bond introduces
linear equations, and we can solve for potential at each pore (Fig. irreversibility of dispersion caused by purely convective spreading.
4b). After knowing the potentials, the flow rate through any bond That is, solute particles will not return to the starting location if
can be calculated easily as the flow direction is reversed. Thus, probabilistic choice of out-
flowing bonds causes mixing, even in the absence of diffusion, that
q = gP. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (2) is unphysical (Jha 2008). Moreover, it ignores spatial correlation
in bond conductances, which is one of the key features of the
Rules for Particle Tracking physically representative network models (Sahimi et al. 1986; Jha
Particle Movement in the Network in the Absence of Diffusion. 2008). Therefore, dispersion from convective spreading cannot be
Bonds in a physically representative network model connect centers modeled correctly unless the choice of out-flowing bond is (1)
of pairs of adjacent pores. Because the bonds have nonzero radii, deterministic and (2) symmetric with respect to flow reversal.

Vout
Vout

15 9
10

14 11
7
No Flux
No Flux

12 5
2

1
3 8

4
13 6

Vin
Vin y x
(a) (b)

Fig. 4(a) A schematic for the physically representative network as a network of resistances. (b) The flow potential at each pore
in the network for steady-state, single-phase flow in the z direction, with no-flow boundaries on the side faces. The coordinates
on the x, y, and z axes are in units of 104 m.

March 2011 SPE Journal 67


If, during any timestep, the particle hits the bond wall, it is
reflected back into the bond.
The algorithm for tracking particle movement through a capil-
Exit on lary tube is validated by comparing simulations in a cylindrical
in-flowing bond tube with results of Taylors theory and experiments (Jha 2008).
A close agreement of simulated results with Taylors theory veri-
fies the algorithm.

Algorithm for Particle Tracking Through


the Network
Entrance on
1. Starting positions of a swarm of 15,000 particles at the inlet
out-flowing bond face of the network are decided in advance. Particles are distributed
to the inlet bonds in proportion to their flow rates. In an inlet bond,
particles are distributed uniformly over its inlet face.
2. We track one particle at a time. A particle moves with convec-
tive and diffusive steps through a bond until it reaches a pore body
(bond junction). Its position is mapped to an out-flowing bond on
Fig. 5Intersecting bonds at a pore body. An exit point on an the basis of deterministic rules discussed earlier and described in
in-flowing bond is mapped on an entrance point on an out-flow- Appendix A.
ing bond on the basis of deterministic rules. We used the same rules for particle tracking for all the Pclet
numbers. The position of a particle at the entrance of an outlet
bond depends only the flow configuration and relative flow rates
at a pore center. This rule is strictly valid when solute transport is
In order to model convective spreading realistically, we devel- dominated by convection. A better rule to decide the out-flowing
oped deterministic rules to decide an out-flowing bond and to map bond in a diffusion-dominated regime (NPe < 4) would be based on
the entrance point of the solute particle on this bond. The rules are cross-sectional areas of bonds rather than the flow rates. However,
based on the patterns followed by streamlines and are described the point of the rules is to capture the convective transport from
briefly in Appendix A. A similar concept has been implemented one bond to the next in a deterministic way and thereby account
previously in 2D regular networks (Bruderer and Bernabe 2001; for the contribution of convective spreading. Thus, in this work,
Sahimi et al. 1986). Our algorithm can be seen as a generalization we use the same outflow rules for all values of NPe.
of their approach for 3D, unstructured networks. The principal 3. The particle continues its movement through successive
logical features of the rules are (1) at every pore center, the rules bonds until it reaches the outlet of the network. Particle positions
establish a one-to-one mapping from points on the outlet face(s) are scanned at regular time intervals. We also record the residence
of in-flowing bond(s) to points on the inlet faces of out-flowing time of the solute particle. The dispersion coefficient is calculated
bond(s), and (2) they follow the patterns expected from actual from both spatial and temporal statistics.
streamlines. Thus, when the direction of flow is reversed, a par-
ticle retraces its flow path exactly in the absence of diffusion. Results
A direct calculation of the flow field would eliminate the need for We use the model described to focus on the three physical features
the rules. However, it was not practical to resolve the flow field of pore-level flow and solute transport. The first is stream split-
so accurately for the entire domain, which contains approximately ting of the solute front at every pore. This causes a sequence of
35,000 pores. independent and random velocities of solute particles in successive
After reaching the outlet face of the bond at a pore body, the bonds. The second is velocity gradient in pore throats in the direc-
particle immediately jumps to the inlet face of the next out-flowing tion transverse to flow, the consequence of viscous momentum
bond (decided from the deterministic rules). This causes a small transfer with no-slip boundary conditions. The third is diffusion,
discontinuity in the path of the particle because of overlapping of which results in the random movement of solute particles between
bonds in the region near the pore body. Discontinuities are small streamlines. We examine these phenomena individually and in
compared with the path length. Moreover, the effect of discontinui- combinations. Our deterministic algorithm of particle tracking
ties on particle statistics accumulated over several pores tends to (which is conceptually equivalent to solving Stokes equation for
get cancelled. Thus, the particle statistics are not affected by these the entire domain) in a realistic porous medium enables us to inves-
discontinuities (Jha 2008). tigate systematically the effect of coupling of these phenomena
A particle continues its motion through successive bonds to the and gives an unambiguous explanation of the origin of core-scale
outlet of the network. dispersion. The combination of the first two gives rise to convec-
tive spreading. The combination of the second and the third gives
Particle Tracking in a Bond With Convection and Diffusion. rise to Taylors dispersion in a capillary tube. We show that only
The motion of a solute particle in a bond in presence of diffusion when all three phenomena are accounted for do we obtain model
is divided into small timesteps of equal duration dt. We split each predictions consistent with the experimental observations.
step into a convection-only displacement followed by an instan-
taneous diffusive jump, rdiff = 6 Do dt . The magnitude of convec- Convective Spreading as a Diffusive Process. We first apply the
tion-only displacement is equal to the product of fluid velocity at algorithm to investigate convective spreading caused purely by
that location and the duration of the timestep. The direction of the stream splitting. We impose a plug-flow velocity profile in each
convection-only step is parallel to the bond axis. In the diffusive bond (i.e, all particles in a bond move at the same velocity). The
step, a particle moves in a random direction on the surface of a network flow calculation yields volumetric flow rates through each
sphere with radius r (Bruderer and Bernabe 2001), where Do is bond. The direction of movement is parallel to the bond axis, and
the diffusion coefficient. The particles diffusive displacement in a the speed is v = q /  reff2 . Diffusion is neglected. Thus, only one of
Cartesian coordinate system is given by (Bijeljic et al. 2004): x = the physical features of interest, stream splitting, affects this set
rdiff cos sin, y = rdiff sin sin, and z = rdiff cos, where  is of simulations.
uniformly distributed between 0 and 2 and cos is uniformly Fig. 6a shows paths of five different pairs of particles. The
distributed between 1 and 1. For every timestep, we pick a value particles in each pair initially were close to one another. The
for  and .  is a random number between 0 and 2. For generat- rules determining the exit bond taken by a particle entering a pore
ing , we pick a random number between 1 and 1 and take the constitute a mapping. The mapping is unique for each pore in the
inverse cosine of that. network. At the level of pore throats, the mapping relates geometric

68 March 2011 SPE Journal


120 120 120 120
120
120

100 100 100 100


100
100

80 80 80
80 80
80

z
z

z
z

z
60 60
60
60 60
60

40
40 40
40 40
40

20
20 20
20
20
20

30
20 50 30
60 30 45
55 10 15 40 20 60
55
y
55 50
55 50 20 20 10 y 35 30
35
15 20 50
y x y 25
y
10 10
50 60
x x x x

(a) (b)

Fig. 6(a) Paths of five different pairs of neighboring particles, each beginning in the same pore but not at the same math-
ematical point. Flow is from bottom to top, with sealed sides. Bonds have plug-flow velocity profiles (no gradient in velocity in
the transverse direction), and diffusion is zero. Particles split their paths after traveling together for some number of pores. (b)
Paths of several particles starting at different positions in the same pore. Particles paths are independent of each other. If each
convective step is independent and has same global statistics, solute-particle displacements will be normally distributed. The
coordinates on the x, y, and z axes are in units of 104 m.

regions; well-defined segments of the exit faces of in-flowing length in velocity, its total displacement can be considered as the
bonds are connected to well-defined segments on inlet faces of sum of independent and random convective steps. Consequently,
out-flowing bonds. At the level of particles, the mapping connects after a few steps, the spatial distribution of solute particles dis-
a single point within a geometric region on an exit face to a point placements is expected to be normal (Gaussian), as per the central-
within a corresponding region on an inlet face. Thus, if a pair of limit theorem, and convective spreading caused purely by stream
particles reaches an exit face and is in the same geometric region, splitting can be considered like a diffusive process. Cenedese and
it will enter the same out-flowing bond (details in Appendix A). Viotti (1996) and Moroni and Cushman (2001) show by 3D par-
As the pair of particles proceeds through the network, eventually ticle-tracking velocimetry experiments in beadpacks that velocity
it will arrive at an exit face on opposite sides of a line separating components quickly become independent. Correlation lengths are
two geometric regions on that face. of the same order as the grain dimensions. Longitudinal dispersion
When the particles fall on opposite sides of the split stream coefficient becomes Fickian after the solute front has traversed five
on an exit face, they take different paths and their movements to six pore diameters (Manz et al. 1999). Though the experiments
become independent of each other. They are unlikely to come are not diffusion-free, the agreement with our simple model is
together again, and their positions become uncorrelated. This is encouraging.
true even in the absence of diffusion. It is simply a consequence
of the asymmetric splitting and joining of flow paths around ran- Particle Tracking Without Diffusion. With Plug Flow in Net-
domly arranged obstaclesthat is, the grains forming the porous work Bonds. Having seen the role of convective spreading caused
medium. This is the nature of convective spreading. It is useful by stream splitting with small groups of particles, we now track
to contrast the situation in ordered arrangements of grains. There, movement of a swarm of 15,000 particles through the same physi-
the splitting/joining of the flow stream is symmetric and positions cally representative network model. Again, we impose a plug-flow
of a pair of particles remain correlated regardless of their initial velocity profile in each bond of the network and neglect diffusion.
separation (Jha 2008). If plug flow occurs in bonds and there is no The particles move from bottom to top of the domain (positive z
diffusion, then no independence of particle paths occurs by means direction). The average interstitial velocity is 5.12105 m/s, which
of convective spreading in an ordered packing. roughly corresponds to 0.12 grain diameters per second.
Fig. 6b shows paths of several solute particles (moving without The particle positions are scanned at various times. A prob-
diffusion with plug flow in bonds), starting at the same inlet pore. ability-distribution plot of particles spatial positions is shown at
A particles displacement becomes independent of other particles several different times in Figs. 7a through 7c. For comparison, a
very quickly because of frequent splitting of flow passages. The normally distributed probability-distribution plot having the same
randomness of convective spreading in a porous medium is inher- mean and standard deviation as the particle-position statistics is
ent in the morphology of the pore space (Sahimi et al. 1986). also plotted in each case.
In the conventional Fickian representation of dispersion, con- The dispersion coefficient at each time is calculated from
vective spreading is considered to be diffusion-likethat is, a sta- spatial statistics using DL =  z2 / 2t , where  z2 is the variance of
tistically random process. The sufficient conditions for convective solute-particle positions in the z (longitudinal) direction at time t.
spreading because of splitting at pore junctions can be treated like For verification, we also compute the dispersion coefficient using
diffusion in a continuum transport equation and can be stated in the particles residence-time statistics and using a solution to the
terms of the central-limit theorem (Chandrashekhar 1943; Sahimi 1D convection/diffusion equation (Lake 1989). The dispersion
et al. 1986). The central-limit theorem states that the sum of a coefficient calculated from spatial statistics increases with time
large number of independent and identically distributed random (and travel distance) and approaches an asymptotic value (Fig. 7d).
variables will be distributed approximately normally (Bear 1972). The asymptotic dispersion coefficient is very close to that obtained
After a particle has traveled a distance greater than the correlation from temporal statistics in all the cases studied. For consistency,

March 2011 SPE Journal 69


Concentration Profile at 25 seconds Concentration Profile at 100 seconds
0.1 0.04
Simulated data Simulated data
Normal distribution Normal distribution
0.08
0.03
Probability

0.06

Probability
0.02
0.04

0.01
0.02

0 0
0 5 10 15 20 25 0 10 20 30 40 50 60 70 80
Distance, 0.1 mm Distance, 0.1 mm
(b)
( a)
Concentration Profile at 150 seconds 9 Dispersion Coefficient
x 10
0.035 10

Dispersion Coefficient, m2/s


Simulated data
0.03
Normal distribution
8
0.025
Probability

0.02 6
From spatial
Spatial statistics
Statistics
0.015 From temporal
Temporalstatistics
Statistics
4
0.01

0.005
2
0
0 20 40 60 80 100 120 0 50 100 150 200
Distance, 0.1 mm Time, seconds
(c) (d)

Fig. 7Scanned spatial distribution of solute particles traveling through the physically representative network for plug flow in
the network bonds and Do = 0 m2/s [(a) through (c)]. Normally distributed curves having the same mean and standard deviation
as the actual data are also shown for comparison. (d) Dispersion coefficient as evaluated from spatial statistics (dots) for several
times. The dotted line represents the dispersion coefficient obtained from temporal statistics. The asymptotic dispersion coef-
ficient obtained from spatial statistics is very close to that obtained from temporal statistics.

the dispersion coefficient obtained from temporal statistics is taken are two peaks in the probability-distribution function. The peak
as the dispersion coefficient. at smaller displacement corresponds to solute particles initially
It can be seen from Figs. 7a through 7c that particles are nor- located in the slower zone near the wall of the bonds (i.e., at r
mally distributed at all the times. The particle velocities become reff). At longer distances traveled, a second peak emerges in the
independent quickly because of splitting of flow stream at each distribution. It corresponds to particles initially located in the
pore and randomness of pore structure. Solute particles become faster-moving core of a bond (i.e., at 0 < r < areff, where a 0.9).
normally distributed as expected from the central-limit theorem. These particles are free to move at early times and have traveled
The convergence of the dispersion coefficient to an asymptotic a sufficiently large distance to experience a wide range of flow
value is governed by correlation in the local pore structure. After velocities (Lebon et al. 1997). At longer times, the second peak
the particles have traveled far enough, dispersion coefficient is closer to normal distribution. However, the first peak persists at
reaches an asymptotic value and convective spreading because the longest time observed.
of stream splitting becomes Fickian. The asymptotic dispersion It may be argued that the dispersion coefficient is converging to
coefficient for purely mechanical dispersion, which is caused by an asymptotic value (Fig. 8d). However, the concentration profiles
stream splitting, is evaluated to be 9.25109 m2/s (for our case, make it clear that the transport process is not Fickian. This behavior
with grain radius of 2.1918104 m and average interstitial velocity is general. If a stochastic velocity field contains regions of zero
of 5.12105 m/s). velocity, there is no purely hydrodynamic mechanism by which
With Parabolic Velocity Profile in Bonds. Using the same solute particles in these regions can reach the rest of the pore space.
steady-state solution for flow in the network as in the preceding Particles in these regions cannot have any random velocity from
subsection, we impose a parabolic velocity profile in each bond the velocity distribution. Therefore, the central-limit theorem is not
of the network. As in the preceding subsection, we neglect diffu- applicable and the dispersion coefficient is not well defined in this
sion. Thus, two of the physical features of interest that give rise case. The effect of diffusion must be considered for the transport
to convective spreading (stream splitting and velocity variation) process to become Fickian, even in the limit of high Pclet number
affect this set of simulations. The velocity profile is taken from (Koch and Brady 1985; Duplay and Sen 2004).
the classical Hagen-Poiseuille analysis of flow in a cylindrical Steady-state single-phase flow in the network is a linear pro-
bond of radius reff: cess. If the overall pressure drop across the network is changed,
the pressure difference between the extremities of each bond of the
network changes in proportion. It follows that the velocities, flow
r 2 rates, and transit times in each bond also change in proportion.
v ( r ) = 2 v 1 , . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (3)
reff Hence, the ratio between flow rates in different bonds that connect
to the same node remain unchanged, and, therefore, the mapping
rule at the pore junction remains unchanged if the average veloc-
where r measures the radial distance of the particle from the bond ity changes. Consequently, the dispersion coefficients caused by
axis. purely convective spreading are proportional to the mean velocity.
Fig. 8 shows the particle-position statistics for a swarm of Therefore, in absence of diffusion (or with negligible diffusion),
15,000 particles moving through the network. At early times, there dispersion depends linearly on velocity (Sahimi et al. 1986).

70 March 2011 SPE Journal


Concentration Profile at 25 seconds Concentration Profile at 100 seconds
0.06 0.03
Simulated data Particles in low velocity layer Simulated data
0.05 Normal distribution 0.025 Normal distribution

Probability 0.04 0.02 Particles in the

Probability
main stream
0.03 0.015

0.02 0.01

0.01 0.005

0 0
0 5 10 15 20 25 30 35 0 20 40 60 80 100
( a) Distance, 0.1 mm (b) Distance, 0.1 mm

Concentration Profile at 150 seconds 8 Dispersion Coefficient


x 10
0.02 2.5
Simulated data

Dispersion Coefficient, m2/s


Normal distribution
0.015 2
Probability

0.01 1.5
Non-Fickian behavior

0.005 1

From spatial statistics


0 0.5
0 20 40 60 80 100 120 0 50 100 150 200
(c) Distance, 0.1 mm (d) Time, seconds

Fig. 8Scanned spatial distribution of solute particles for parabolic flow in network bonds and Do = 0 m2/s [(a) through (c)]. The
first peak in the distribution at small distance corresponds to particles in the slow-velocity regions near the pore walls. Other
particles that are free to move form a second peak. Dispersion is not Fickian in this case. (d) Dispersion coefficient as evaluated
from spatial statistics.

Particle Tracking With Diffusion. With Plug Flow in Network and Pclet number with the power-law coefficient in the range
Bonds. Next, we study the influence of diffusion on particle of 1.1 to 1.3. The simulations yield = 1.
statistics and dispersion coefficient. A plug-flow velocity profile The cause of the incorrect scaling is the plug-flow velocity
is imposed in the bonds of the network. Thus, two of the three profile in the bonds. Under this assumption, there is no stretching
physical features of interest (stream splitting and diffusion) affect of a solute front as it moves along a bond. Therefore, flow velocity
the solute transport. within a bond has no interaction with diffusion; a particle making
Particles move with convective and diffusive steps as described a diffusive jump to another streamline within a bond still travels
previously. The timestep should be small enough to prevent dif- at the same velocity after the jump. Thus, diffusion acts indepen-
fusive jumps from being larger than the bond diameter (Bruderer dently of convective spreading, and the dispersion coefficient is
Bernabe 2001). In the simulations reported here, the timestep was just the sum of the two. Consequently, for small diffusion (large
taken to be one-tenth of this value (evaluated for the average bond Pclet number), we find a linear dependence of dimensionless dis-
diameter) to reduce numerical error. persion coefficient on pore-scale Pclet number. The longitudinal
Fig. 9 shows the effect of diffusion on solute-particle statistics. dispersion coefficient for the full range of Pclet numbers can be
It is evident that diffusion has negligible effect on particle distribu- expressed as
tion for diffusion coefficients ranging from 0 to 109 m2/s (for our
case, with grain radius of 2.1918104 m and average interstitial vD p
DL 1
velocity of 5.12105 m/s, resulting in a dispersion coefficient = + 0.877 . . . . . . . . . . . . . . . . . . . . . . . . . . . . . (4)
of 9.25109 m2/s if stream splitting is accounted for) because Do F Do
diffusion is very small in magnitude compared with convective
spreading. The dispersion coefficient remains constant in this range The first term in Eq. 4 represents the contribution of diffusion,
of diffusion coefficients. For higher diffusion coefficients, diffu- and the second term is the result of convective spreading (stream
sion becomes significant in magnitude compared with convective splitting).
spreading, and, therefore, the variance of particle displacements Thus, Eq. 1, commonly used to describe dispersion in porous
increases and this increases the dispersion coefficient. media, considers mechanical dispersion caused by stream split-
When diffusion is very small, transport is dominated by con- ting only. It neglects the effect of variations in fluid velocity in
vection. However, at very low Pclet numbers when diffusion is the transverse direction and, thus, interaction between convective
large, our flow-rate-weighted rule for deciding an out-flowing spreading and diffusion.
bond captures only part of the particle dynamics at a pore center. With Parabolic Velocity Profile in Bonds. Finally, we examine
Therefore, in this case, the particle statistics deviate from normal the influence on dispersion of all three physical features (stream
distribution at longer times. splitting, velocity variation, and diffusion) simultaneously. We
Fig. 10 shows the a priori prediction of dimensionless disper- repeated the simulations of the preceding subsection but with a
sion coefficient vs. pore-scale Pclet number for this case. The parabolic velocity profile in each network bond, as described.
agreement with the experimental data from the literature is good If a bond has a velocity gradient within it, then the interaction
for small and moderate Pclet number, but it is clear that the scal- between convection and diffusion is no longer trivial. Now, a solute
ing is incorrect for large Pclet numbers. The experimental data front stretches as it travels in a bond, and diffusion becomes a much
suggest a power-law relationship between dispersion coefficient more effective mixing mechanism. (This interaction is the basis of

March 2011 SPE Journal 71


Concentration Profile at 25 seconds Concentration Profile at 50 seconds
0.1 0.06

0.05
0.08

Probability
0.04
Probability

0.06
0.03
0.04
0.02

0.02 0.01

0 0
0 10 20 30 40 50 60 0 20 40 60 80 100
( a) Distance, 0.1 mm (b) Distance, 0.1 mm

Concentration Profile at 100 seconds Concentration Profile at 150 seconds


0.04 0.035

0.03
0
0.03 11013
0.025

Probability
11011
Probability

0.02 11010
0.02 1109
0.015 1108
1107
0.01
0.01
0.005

0 0
0 20 40 60 80 100 120 0 20 40 60 80 100 120
(c) Distance, 0.1 mm (d) Distance, 0.1 mm

Fig. 9Effect of diffusion on spatial statistics of solute particles for plug-flow profile in bonds. Diffusion has negligible effect
unless its magnitude becomes significant compared with that of mechanical dispersion. All diffusion coefficients shown in the
legend are in m2/s.

the familiar Taylors dispersion.) Because velocity now influences curve fitting, we obtain a power-law coefficient of 1.23. This a
dispersion within a bond, the dispersion coefficient will not be priori quantitative prediction of dispersion coefficient and its scaling
simply the summation of convective spreading and diffusion. behavior indicates that the essential physics of dispersion through the
Fig. 11 shows a comparison of the dimensionless dispersion pore space of a disordered granular material has been captured.
coefficient obtained from simulations with the experimental data. We investigate the interaction of convection and dispersion
The match is excellent for the whole range of Pclet numbers. From more closely. Fig. 12 shows the effect of diffusion on particle

1,000,000
Pfannkuch (1963)
Seymour and Callaghan (1997)
100,000
Kandhai et al. (2002)
Khrapitchev and Callaghan (2003)
10,000 Stohr (2003)
Perkins and Johnston (1963)
1,000 Jha (2005)
DL /Do

Simulation, Plug Flow

100

10

0.1
0.001 0.01 0.1 1 10 100 1,000 10,000 100,000 1,000,000

Dp /Do

Fig. 10Comparison of dimensionless dispersion coefficients simulated with plug flow in bonds in a physically representative
network (beadpack with a bead radius of 2.1918104 m) and experimental data in the literature.

72 March 2011 SPE Journal


1,000,000
Pfannkuch (1963)
Seymour and Callaghan (1997)
100,000
Kandhai et al. (2002)
Khrapitchev and Callaghan (2003)
10,000 Stohr (2003)
Perkins and Johnston (1963)
DL /Do 1,000 Jha (2005)
Simulation, Parabolic Flow
100

10

0.1
0.001 0.01 0.1 1 10 100 1,000 10,000 100,000 1,000,000

Dp /Do

Fig. 11Comparison of dimensionless dispersion coefficient simulated with parabolic flow in bonds and experimental data in
the literature. The simulated data quantitatively match with the experimental data for the whole range of Pclet numbers.

statistics for a physically representative network with a parabolic corresponding to particles in the slow-moving region near the walls
velocity profile in the bonds. Even a vanishingly small amount starts decreasing and almost disappears by 150 seconds. We obtain
of diffusion of 1013 m2/s (for our case, with grain radius of a normal distribution of solute-particle positions.
2.1918104 m and average interstitial velocity of 5.12105 m/s) This is consistent with experimental observations by Lebon et al.
starts moving solute particles from low-velocity regions near the (1997). They studied dispersion at short times using a pulsed-field-
wall to high-velocity regions. After moving out of the low-veloc- gradient/nuclear-magnetic-resonance technique. At short times,
ity zone, solute particles can sample all the regions of pore space the displacement of the molecules is small enough that the local
because of stream splitting. The first peak at small displacement displacement is proportional to the local velocity component along

Concentration Profile at 25 seconds Concentration Profile at 100 seconds


0.06 0.025
Simulated data Fallingdue
Falling duetoto Simulated data
0.05 Normal distribution Normal distribution
0.02 diffusion
diffusion
Probability

0.04
Probability

0.015
0.03
0.01
0.02
0.005
0.01

0 0
0 5 10 15 20 25 30 35 0 20 40 60 80
( a) Distance, 0.1 mm (b) Distance, 0.1 mm

Concentration Profile at 150 seconds -8 Dispersion Coefficient


x 10
0.02 2.2
Simulated data
Dispersion Coefficient, m2/s

2
Normal distribution
0.015
1.8
Probability

1.6
0.01
1.4

1.2
0.005
1 From
From spatial
Spatialstatistics
Statistics
From
From temporal
Temporalstatistics
Statistics
0 0.8
0 20 40 60 80 100 120 0 50 100 150 200

(c) Distance, 0.1 mm (d) Time, seconds

Fig. 12Scanned spatial distribution of solute particles for parabolic flow in the network bonds and Do = 1013 m2/s [(a) through
(c)]. The first peak corresponds to particles in the slow-velocity regions near the pore walls and starts falling because of diffu-
sion. Dispersion is asymptotically Fickian at large times. (d) Dispersion coefficient as evaluated from spatial statistics.

March 2011 SPE Journal 73


11107

D L, m2/s
11108

11109
11013 11012 11011 11010 1109 1108 1107 1106

DO , m2/s
Fig. 13Change in dispersion coefficient with diffusion coefficient for parabolic velocity profile in bonds. Dispersion coefficient
decreases with increased diffusion unless diffusion becomes significant in magnitude compared with mechanical dispersion.

the magnetic-field gradient. At mean displacements larger than the main flow stream. Subsequently, solute particles can encounter
five bead diameters, the displacement distribution was found to be a wide range of velocities because of convective spreading, and this
Gaussian. At intermediate displacements, the measured distribu- gives rise to the Fickian behavior of dispersion (Koch and Brady
tion displayed two peaks. With increasing diffusion coefficient, the 1985; Duplay and Sen 2004). Therefore, diffusion, even though
peak at small displacements disappears quickly (Jha 2008). small in magnitude, is essential for Fickian behavior of dispersion.
Fig. 13 shows the effect of increasing diffusion on dispersion. Now, dispersion is a result of an interaction between convective
For zero diffusion, there is a wide range of particle positions because spreading and diffusion, and this interaction leads to nonmechani-
of velocity gradient in the bonds. As diffusion is increased, solute cal dispersion, as termed by Koch and Brady (1985).
particles move in the transverse direction in the bonds and the effect The plug-flow velocity profile does not have regions of zero
of velocity gradient in bonds is reduced. This reduces the spread in velocity. Therefore, nonmechanical dispersion is absent there
solute-particle positions, and, therefore, the dispersion coefficient is and dispersion scales linearly with NPe. However, in the case of a
reduced. This trend continues with increasing diffusion and reverses parabolic velocity profile in the network bonds, there are regions
when the magnitude of the diffusion coefficient becomes large of zero velocity near pore walls. Therefore, we find a nonmechani-
compared with convective spreading. The dispersion-coefficient-vs.- cal dispersion that becomes important at high Pclet numbers and
diffusion-coefficient plot goes through a minimum. results in a mild nonlinear dependence of dispersion coefficient
on Pclet number.
Discussion Taylor (1953) showed that, in a single capillary tube, the inter-
Dispersion in porous media results from an interaction between action of diffusion and the parabolic velocity profile yields a much
convective spreading and diffusion. Convective spreading occurs stronger dependence of dispersion coefficient on Pclet number
because of variations in path lengths and velocities of solute ( = 2). Our model invokes tubes, though of short length, in which
particles traveling along different streamlines. There are two com- a parabolic velocity profile exists. Why then does transport in the
ponents of convective spreading: (1) stream splitting caused by physically representative networkand in experimentsexhibit a
splitting and joining of flow streams at every pore and (2) velocity much smaller value of than Taylors analysis? The reason is the
gradient in pore throats in the direction transverse to flow. In this asymmetric splitting and joining of flow paths around grains in
study, we decouple all three key phenomena giving rise to disper- the porous medium. The splitting/joining process in a disordered
sion in porous media. medium causes convective spreading to dominate the dispersive
For dispersion to be Fickian, all solute particles must undergo a behavior at large values of Pclet number.
range of independent velocities. If we consider the effect of stream Interaction of convective spreading and diffusion is also the
splitting only (by having plug flow in bonds where there is no reason of irreversibility of dispersion. Purely convective spreading
variation in particle velocities in the transverse direction), different is reversible (i.e., all solute particles retrace their path back to the
bonds carry different flow rates and there is no long-range correla- inlet when the flow direction is reversed). Thus, convective spread-
tion of these flow rates. Thus, particle positions eventually become ing is canceled and echo dispersion is zero. However, an interaction
uncorrelated as flow continues, even in the absence of diffusion. of convective spreading with diffusion causes the fluids to mix (as
Consequently, as per the central-limit theorem, convective spread- opposed to pure spreading). This makes dispersion irreversible and
ing caused by stream splitting is Fickian and grows like NPe. makes echo dispersion as large as forward (or transmission) disper-
However, with a parabolic velocity profile in the bonds, when the sion. Mixing and irreversibility of dispersion in terms of interaction
effect of variation in velocity in the transverse direction is considered between convective spreading and diffusion have been explained in
along with stream splitting, solute particles near the pore walls are not detail by Garmeh et al. (2007) and Jha et al. (2009).
free to move. Thus, those particles do not have independent velocities.
Consequently, the central-limit theorem is not applicable and disper- Summary and Conclusions
sion because of convective spreading becomes non-Fickian. We study single-phase solute transport through porous media and
Dispersion caused by convective spreading is also called investigate the three key features giving rise to dispersion: (1)
mechanical dispersion. Steady-state single-phase flow in the net- splitting and joining of flow streams at every pore, (2) velocity
work is a linear process. If the overall pressure drop across the gradient in pore throats in the direction transverse to flow, and (3)
network is changed, the flow rate in each bond changes propor- diffusion. The first two of these features collectively give rise to
tionately. Consequently, mechanical dispersion is proportional to convective spreading.
the mean velocity. We track movement of solute particles through a physically
When diffusion is superimposed on convection, solute particles representative network model. We infer dispersion from particle-
in the low-velocity layer near solid surfaces can move out and enter displacement statistics. We introduce deterministic (nonstochastic)

74 March 2011 SPE Journal


rules to map particle position from an in-flowing bond to a point Bijeljic for helping us with a particle-tracking algorithm in a tube.
on an out-flowing bond. The rules are essentially geometric and This work was conducted with the support of the National Petroleum
depend on the network flow field (rates in individual bonds and the Technology Office of the US Department of Energy through contract
local configuration of inlet and outlet bonds at each pore body). DE-PS26-04NT15450-3F. Larry W. Lake holds the W.A. (Monty)
These rules provide a simple way to obtain particle paths through Moncrief Centennial Chair at The University of Texas at Austin.
the network equivalent to those obtained from Stokes law. Thus,
the particle path is completely known in the absence of diffu-
sion. These rules also yield the exact cancellation of convective References
spreading if the flow direction is reversed. Thus, they enable us Acharya, R.C., van Dijke, M.I.J., Sorbie, K.S., van der Zee, S.E.A.T.M.,
to attribute correctly the contribution of convective spreading to and Leijnse, A. 2007. Quantification of longitudinal dispersion by
core-scale dispersion without requiring a detailed (subpore) solu- upscaling Brownian motion of tracer displacement in a 3D pore-scale
tion of the flow field through the porous medium. Subsequently, network model. Advances in Water Resources 30 (2): 199213. doi:
we superimpose diffusion and study the effect of the interaction 10.1016/j.advwatres.2005.04.017.
between diffusion and convective spreading. Bear, J. 1972. Dynamics of Fluids in Porous Media. New York: Elsevier.
Because our methodology invokes no arbitrary assumptions either Bijeljic, B., Muggeridge, A.H., and Blunt, M.J. 2004. Pore-scale modeling
about the geometry of the porous medium or about the solute-trans- of longitudinal dispersion. Water Resources Research 40 (11): 19. doi:
port mechanism, the simulations provide a priori (no adjustable 10.1029/2004WR003567.
parameters) predictions of dispersion coefficient as a function of Bruderer, C. and Bernabe, Y. 2001. Network modeling of dispersion: Tran-
pore-scale Pclet number. The predicted trends quantitatively match sition from Taylor dispersion in homogeneous networks to mechanical
the experimental data found in the literature for a wide range of Pclet dispersion in very heterogeneous ones. Water Resources Research 37
numbers, including the well-known empirical observation that the (4): 897908. doi: 10.1029/2000WR900362.
scaling exponent has a value of approximately 1.2. We also obtain near Bryant, S.L., King, P.R., and Mellor, D.W. 1993. Network Model Evalu-
independence of dispersion coefficient for small Pclet numbers. The ation of permeability and spatial correlation in a real random sphere
agreement indicates that the key features of the model correspond to packing. Transport in Porous Media 11 (1): 5370. doi: 10.1007/
the key physical phenomena causing dispersion in porous media. BF00614635.
The model permits rigorous attribution of the contribution Cenedese, A. and Viotti, P. 1996. Lagrangian Analysis of Nonreactive
of the phenomena individually and of the interaction between Pollutant Dispersion in Porous Media by Means of the Particle Image
combinations of phenomena. Fickian behavior of solute transport Velocimetry Technique. Water Resour. Res. 32 (8): 23292343. doi:
is asymptotically observed when solute particles displacements 10.1029/96WR00605.
are independent, identically distributed, and random. If we ignore Chandrasekhar, S. 1943. Stochastic Problems in Physics and Astronomy.
variation in fluid velocity in pore throats in the transverse direction Rev. Mod. Phys. 15 (1): 189. doi: 10.1103/RevModPhys.15.1.
(i.e., plug flow occurs in network bonds), Fickian behavior can Coats, K.H., Whitson, C.H., and Thomas, L.K. 2004. Modeling confor-
occur without diffusion. Convective spreading and diffusion act mance as dispersion. Paper SPE 90390 presented at the SPE Annual
independently of each other, and dispersion coefficient is the sum Technical Conference and Exhibition, Houston, 2629 September. doi:
of the two. In the more-realistic case of parabolic velocity profile 10.2118/90390-MS.
in the bonds, purely convective (i.e., no diffusion) spreading is Duplay, R. and Sen, P.N. 2004. Influence of local geometry and transition
not asymptotically Fickian. Diffusion is required to move solute to dispersive regime by mechanical mixing in porous media. Phys. Rev.
particles from low-velocity regions near pore walls. Subsequently, E 70 (6): 066309. doi: 10.1103/PhysRevE.70.066309.
stream splitting causes independent, random movement of solute Garmeh, G., Johns, R.T., and Lake, L.W. 2007. Pore Scale Simulation of
particles and gives rise to Fickian behavior of dispersion. Dispersion in Porous Media. Paper SPE 110228 presented at the SPE
In the absence of diffusion, convective spreading in porous Annual Technical Conference and Exhibition, Anaheim, California,
media results in a linear dependence of DL on NPe. Interaction USA, 1114 November. doi: 10.2118/110228-MS.
between convective spreading and diffusion results in a weak Hiby, J.W. 1962. Longitudinal and transverse mixing during single-phase
nonlinear dependence of DL on NPe, in agreement with the experi- flow through granular beds. In Proceedings of the Symposium on the
mental observations. Interaction Between Fluids and Particles, London, 2022 June, 1962,
ed. P.A. Rottenburg. London: Institution of Chemical Engineers.
Nomenclature Jha, R.K. 2008. Investigation of local mixing and its influence on core scale mix-
DL = longitudinal-dispersion coefficient, m2/s ing (dispersion). PhD dissertation, University of Texas, Austin, Texas.
Do = molecular-diffusion coefficient, m2/s Jha, R.K., John, A., Bryant, S.L., and Lake, L.W. 2009. Flow Reversal and Mix-
ing. SPE J. 14 (1): 4149. SPE-103054-PA. doi: 10.2118/103054-PA.
Dp = particle diameter, m
John, A.K., Lake, L.W., Bryant, S.L., and Jennings, J.W. 2008. Investiga-
F = formation resistivity factor tion of Field Scale Dispersion. Paper SPE 113429 presented at the
g = hydraulic conductivity of a bond, m3/s/Pa SPE/DOE Symposium on Improved Oil Recovery, Tulsa, 1923 April.
l = length of a bond connecting two neighbors, m doi: 10.2118/113429-MS.
NPe = pore scale Pclet number, NPe = vDp /Do Kandhai, D., Hlushkou, D., Hoekstra, A.G., Sloot, P.M.A., Van As, H.,
q = flow rate through a bond, m3/s and Tallarek, U. 2002. Influence of Stagnant Zones on Transient
rc = radius of the largest circle to fit in a pore throat, m and Asymptotic Dispersion in Macroscopically Homogeneous Porous
rdiff = magnitude of diffusive jump, m Media. Phys. Rev. Lett. 88 (23): 234501. doi: 10.1103/PhysRev-
re = radius of a circle having same area as a pore throat, m Lett.88.234501.
reff = effective radius of a bond, m Khrapitchev, A.A. and Callaghan, P.T. 2003. Reversible and irreversible
dispersion in a porous medium. Phys. Fluids 15 (9): 26492660. doi:
= power-law coefficient characterizing dependence of dis-
10.1063/1.1596914.
persion coefficient on Pclet number
Koch, D.L. and Barady, D.F. 1985. Dispersion in Fixed Beds. Journal of
 z2 = variance of solute particles in z direction, m2 Fluid Mechanics 154: 399427. doi: 10.1017/S0022112085001598.
 = fluid viscosity, Pas Lake, L.W. 1989. Enhanced Oil Recovery. Englewood Cliffs, New Jersey:
v = interstitial fluid velocity, m/s Prentice Hall.

= average fluid velocity through a bond, m/s Lebon, L., Leblond, J., and Hulin, J.P. 1997. Experimental measurement
 = porosity of dispersion processes at short times using a pulsed gradient NMR
technique. Phys. Fluids 9 (3): 481490. doi: 10.1063/1.869208.
Acknowledgments Manz, B., Alexander, P., and Gladden, L.F. 1999. Correlations between
We would like to thank Cynthia Thane for providing coordinates dispersion and structure in porous media probed by nuclear magnetic
of random packing of spheres. We would also like to thank Branco resonance. Phys. Fluids 11 (2): 259267. doi: 10.1063/1.869876.

March 2011 SPE Journal 75


Mellor, D.W. 1989. Random close packing (RCP) of equal spheres: For example, in Fig. A-1a, the fourth bond is closest to the
structure and implications for use as a model porous medium. PhD in-flowing bond and the third bond is the farthest. First, we find
thesis, Department of Earth Sciences, The Open University, Milton (numerically) the pair of points on the first and fourth bond faces
Keynes, UK. that are closest to each other. These are reference points (Reference
Moroni, M. and Cushman, J.H. 2001. Three dimensional particle tracking Point 1 on Bond 1 and Reference Point 6 on Bond 4). A solute
velocimetry studies of the transition from pore dispersion to Fickian particle exiting Bond 1 at Reference Point 1 on Bond 1 will enter
dispersion for homogeneous porous media. Water Resour. Res. 37 (4): Bond 4 at Reference Point 6. These reference points also guide
873884. doi: 10.1029/2000WR900364. mapping of other points.
Perkins, T.K., and Johnston, O.C. 1963. A Review of Diffusion and Dis- Starting with Reference Point 1, we mark a segment of in-
persion in Porous Media. SPE J. 3 (1): 7084; Trans., AIME, 228. flowing bond that carries the same flow rate as the out-flowing
SPE-480-PA. doi: 10.2118/480-PA. Bond 4. A solute particle exiting Bond 1 from this segment will
Pfannkuch, H.O. 1963. Contribution a letude des deplacements de fluie- enter Bond 4.
des miscibles dans un milieu poreux. Revue Francaise de linstitut du Similarly, we mark a middle segment of the in-flowing bond that
Ptrole 18: 215270. carries the same flow rate as the farthest out-flowing bond. A solute
Saffman, P.G. 1959. A theory of dispersion in a porous medium. Journal of particle exiting the in-flowing bond on the middle segment will enter
Fluid Mechanics 6 (3): 321349. doi: 10.1017/S0022112059000672. the farthest bond. Solute particles exiting Bond 1 on the remaining
Sahimi, M., Hughes, B.D., Scriven, L.E., and Davis, H.T. 1986. Dispersion third segment of the in-flowing bond will enter the second bond.
in flow through porous mediaI. One-phase flow. Chemical Engineer- Reference Point 2 on the in-flowing face is taken as the point
ing Science 41 (8): 21032122. doi: 10.1016/0009-2509(86)87128-7. where the line joining Reference Point 1 to the center of the face
Seymour, J.D. and Callaghan, P.T. 1997. Generalized approach to NMR intersects the boundary of the first segment. Reference Point 3 is
analysis of flow and dispersion in porous media. AIChE Journal 43 (8): diametrically opposite of Reference Point 1. Reference Point 4 is
20962111. doi: 10.1002/aic.690430817. the point on Tube 2 that is closest to Reference Point 3. Similarly,
Sthr, M. 2003. Analysis of flow and transport in refractive index matched Reference Point 5 is the point on Tube 3 that is closest to Refer-
porous media. PhD thesis, University of Heidelberg, Heidelberg, Ger- ence Point 2.
many (July 2003).
Taylor, G.I. 1953. Dispersion of soluble matter in solvent flowing slowly Configuration 2: Three In-Flowing Bonds and One Out-Flow-
through a tube. Proceedings of the Royal Society of London: Series ing Bond. This configuration is exactly the opposite of Configura-
A, Mathematical and Physical Sciences 219 (1137): 186203. doi: tion 1 (Fig. A-1b). Three in-flowing bonds feed to one out-flowing
10.1098/rspa.1953.0139. bond. Therefore, the out-flowing stream consists of three segments,
Taylor, G.I. 1972. Low Reynolds number flows. VHS produced by Educa- each segment receiving flow from one of the in-flowing bonds. The
tional Services Incorporated under the direction of the National Com- procedure described for connecting segments and calculating refer-
mittee for Fluid Mechanics Films. Chicago, Illinois: Encyclopaedia ence points in Configuration 1 is applicable in this case, also.
Britannica Educational Corporation. There is more than one possible flow configuration when we have
Thane, C.G. 2006. Geometry and Topology of Model Sediments and Their two in-flowing and two out-flowing bonds at a pore junction.
Influence on Sediment Properties. MS thesis, University of Texas at
Austin, Austin, Texas. Configuration 3: Third Bond Is the Farthest, and Flow in First
Bond Is Smaller Than That in the Fourth Bond. Because flow
Appendix ADeterministic Rules for rate in the first bond is smaller than that in the fourth bond, the
Mapping an Entrance Point of a Particle fourth bond must receive flow from both the in-flowing bonds.
Because some streamlines move from Bond 2 to Bond 4 and
on an Out-Flowing Bond
streamlines cannot cross, no streamline can move from Bond 1 to
A particle enters the pore through an in-flowing bond, and it leaves Bond 3. Therefore, Bond 3 will get all its flow from Bond 2 and
through an out-flowing bond. The task is to develop rules for calculat- all flow from Bond 1 will go to Bond 4.
ing which out-flowing bond will be the exit and to map the entrance Marking of segments and calculation of reference points are
point of the solute particle on the out-flowing bond. Here, we describe similar to that described for Configuration 1.
a simple and computationally tractable approach for this purpose.
We identify six flow configurations feasible at a pore body. In Configuration 4: Third Bond Is the Farthest, and Flow in First
each case, we determine splitting of streams on the basis of the Bond is Greater Than That in the Fourth Bond. This configu-
flow configuration and flow rates. This tells us which segment of ration is similar to Configuration 3, except that, in this case, the
which out-flowing bond a solute particle exiting an in-flowing flow in the first bond is larger than that in the fourth bond. In this
bond at a particular point will enter. In each case, we calculate case, the fourth bond receives all its flow from Bond 1 and all of
some reference points that serve as guiding points to map the exact the second bonds flow enters the third bond. The first bond feeds
entrance position of solute particle on the out-flowing bond. to both the out-flowing bonds.
For ease of illustration and for calculating reference points, we
displace all the bonds along their axes by equal distance. The dis- Configuration 5: Fourth Bond is the Farthest. Because flow rate
placement should be enough to remove overlap between the bonds. in the fourth bond is greater than that in the second bond, the fourth
After identifying flow configuration and calculating reference bond receives flow from both the inflowing bonds. Because stream-
points, all the bonds are moved back to their original location. lines cannot cross, flow from the second bond cannot enter the third
The bonds are numbered according to their flow rates. Inflow bond. Hence, all the flow from the second bond enters the fourth
is assigned positive sign and outflow negative. Then, flow rates bond. Also, the third bond receives all its flow from the first bond. As
are sorted in descending order along with their sign. Thus, the first evident from Fig. A-1e, Bond 1 feeds to both the out-flowing bonds
bond is the one carrying maximum inflow and the fourth one is and Bond 4 receives flow from both the in-flowing bonds.
carrying maximum outflow. The first bond is taken as the reference
bond. The distances of the face centers of all the bonds from the Configuration 6: The Second Bond is the Farthest. In this configu-
face center of the first bond are calculated. The closest bond will ration, the farthest bond is in-flowing. Or, in other words, the second
have the biggest influence on flow. bond is the farthest. In this case, both the inflowing bonds will feed to
both the out-flowing ones, as shown in Fig. A-1f. Flow segments and
Configuration 1: One In-Flowing Bond and Three Out-Flowing reference points are calculated as previously described.
Bonds. In this case, one in-flowing bond feeds all the out-flowing The reader is referred to Jha (2008) for calculation details.
bonds. Therefore, the incoming stream splits into three segments
on the basis of the flow rates of the three out-flowing bonds Mapping an Incoming Point to an Out-Flowing Bond. After decid-
(Fig. A-1a). ing the incoming and outgoing segments of the bonds, we must map

76 March 2011 SPE Journal


(a) (b) (c)

q4 q4 q4
Reference Point 6 Reference Point 4 Reference Point 4
Reference Point 6

Reference Point 5 Reference Point 1 Reference Point 3


Reference Point 1 Reference Point 1

Reference Point 5
Reference Point 2

q1 q3 q1 q3 q1 q3

Reference Point 3 Reference Point 3

Reference Point 4 Reference Point 2


q2 q2 q2 Reference Point 2

(d) (e) (f)

q4 q3 q4
Reference Point 4 Reference Point 4 Reference Point 4

Reference Point 1 Reference Point 1


Reference Point 1

q1 q3 q1 q4 q1 q2

Reference Point 3 Reference Point 3 Reference Point 2

Reference Point 2 Reference Point 2


q2 q2 q3 Reference Point 3

Fig. A-1Flow configurations at a pore body.

the incoming point to a corresponding point on the outgoing section. point. Therefore, we evaluate the angle , the angle that the posi-
Incoming and outgoing sections are arbitrary sections of a circle. We tion vector of the incoming point makes with the reference vector
impose geometric rules that are physically reasonable: (1) Reference on the incoming segment; and relative radial distance of the point
point on the in-flowing segment will connect to the corresponding from the boundary, r/R. Here, r is the distance of incoming point
reference point on the out-flowing segment, and (2) the center of the from the center and R is distance from the center to the boundary of
incoming segment connects to the center of the outgoing segment. the segment. We place the outgoing point at the same  and r/R on
We take the center of the circular segment as its geometric center (or the outgoing section (Fig. A-2). This rule provides a deterministic,
center of gravity). We call the vector joining the reference point to the one-to-one mapping between exit and entrance points.
center of the segment the reference vector. We mark reference vectors
on incoming as well as outgoing segments (Fig. A-2). Raman K. Jha is a research engineer with Chevron Corporation
in Houston. He holds MS and PhD degrees from The University of
In a polar coordinate system, two parameters are sufficient to Texas at Austin and a BTech degree from the Indian School of
describe the outgoing point: (1) the angle the position vector of Mines, Dhanbad, India, all in petroleum engineering. His research
the outgoing point (here, defined with reference to the center of interests include heavy oil, enhanced oil recovery, reservoir simula-
the segment) makes with the reference vector and (2) its relative tion, and hydrocarbon-phase behavior. Steven L. Bryant is an asso-
radial distance from the center (defined as distance of the point ciate professor in the Department of Petroleum and Geosystems
from the center normalized by distance to boundary in that direc- Engineering at The University of Texas at Austin, where he holds the
tion). We impose another geometric rule: These parameters for J.H. Herring Professorship and the George H. Fancher Centennial
the outgoing point have the same value as those for the incoming Teaching Fellowship in petroleum engineering. He worked in
research laboratories of BP and Eni before joining The University of
Texas faculty. Bryant holds a BE degree from Vanderbilt University
and a PhD degree from The University of Texas at Austin, both in
Outgoing Point
Reference

chemical engineering. He is the director for the Geological CO2


Storage Joint Industry Project. His current research efforts include
Vector

Flow Out
grain-scale models for unconventional gas reservoirs and the role
of methane hydrates in the Earths carbon cycle. Bryant has pub-
lished more than 70 papers and one textbook with applications
in production engineering, reservoir engineering, and formation
evaluation. He served as SPE Distinguished Lecturer in 20012002.
Incoming Point
Larry W. Lake holds the W.A. (Monty) Moncrief Centennial Chair
in the Department of Petroleum and Geosystems Engineering at
The University of Texas at Austin. Lake has published more than
100 peer-reviewed journal articles and has taught industrial and
Reference Vector professional short courses in enhanced oil recovery and reservoir
characterization around the world. He is the author or coauthor
of four textbooks and the editor of three bound volumes. Lake
Flow In has been teaching at The University of Texas for 30 years, before
which he worked for Shell Development Company in Houston. He
holds BS and PhD degrees in chemical engineering from Arizona
Fig. A-2Marking the outgoing point on the basis of reference State University and Rice University, respectively. Lake is a member
angle and relative radial distance of the incoming point. of the US National Academy of Engineering.

March 2011 SPE Journal 77

You might also like