You are on page 1of 119

Dynamic Local Grid Refine-

ment
for Incomplete Mixing in Reservoir Sim-
ulation

R.A.M. Gielisse
Technische Universiteit Delft
DYNAMIC L OCAL G RID R EFINEMENT
FOR I NCOMPLETE M IXING IN R ESERVOIR S IMULATION

by

R.A.M. Gielisse

in partial fulfillment of the requirements for the degree of

Master of Science
in Applied Earth Sciences

at the Delft University of Technology,


to be defended publicly on Friday October 30, 2015 at 14:00 PM.

Supervisor: Dr. H. Hajibeygi, TU Delft


Prof. C. P. J. W. van Kruijsdijk, TU Delft/Shell Global Solutions
Dr. J. A. W. M. Groot, Shell Global Solutions

Thesis committee: Dr. J.E. Romate, TU Delft/Shell Global Solutions

This thesis is confidential and cannot be made public until October 30, 2016.

An electronic version of this thesis is available at http://repository.tudelft.nl/.


A BSTRACT
Reservoir simulation is a key tool to get insight in complex fluid flow phenomena occurring in many petroleum
reservoirs, creating the possibility to quantify fluid productions and their uncertainty regarding many un-
known properties. It has been shown by experimental and theoretical research that under certain criteria
immiscible (multiphase) and miscible (single-phase) displacements can develop instabilities. These insta-
bilities cause the displacing fluid to finger though the displaced oil, which is referred to as viscous fingering.
Moreover, reservoir rock heterogeneity has shown to be capable to amply these phenomena.
However, to limit computational costs, reservoir simulations typically use far lower grid resolution than
the scale at which incomplete mixing phenomena occur. Numerical truncation errors imposed by large grid
blocks will conceal small scale physical phenomena that contribute to unstable displacements. These sub-
grid flow features are what is called incomplete mixing, as opposed to the fully mixed scenario inherent to the
use of low resolution grids. To cope with this problem, effective models have been developed that make use
of effective fluid properties in order to capture the sub-grid effect of incomplete mixing. Todd and Longstaff’s
model is the most widely utilized model in the petroleum industry. Like similar models, one major disadvan-
tage is that this effective model depends on an undefined scalar mixing parameter, which must be obtained
from either high resolution simulation or experimental data.
Dynamic local grid refinement (DLGR) is a simulation technique that can adapt the spatial grid resolution
to the scale at which the physical phenomena occur without refining the grid over the entire domain. This can
be achieved by means of local refinement and coarsening of the computational grid dynamically at every time
step where small scale, local phenomena occur using predefined error criteria. This way DLGR could provide
the means to solve for complex flow phenomena arising from instabilities at fluid fronts, while preserving
complex geological features that could amplify this phenomena.
This report investigates the use of state-of-the-art DLGR as an alternative to the effective model developed
by Todd and Longstaff (TL). As DLGR is capable of adapting the spatial grid resolution locally where physical
phenomena of incomplete mixing occur, this technique potentially allows for computational efficient and
accurate results. The first part of this report studies and compares the results obtained using DLGR with
that of the TL model and reference fine scale simulations for miscible transport in homogeneous media.
Elaboration on DLGR’s efficiency in term of CPU time, accuracy and consistency regarding scalability in terms
of reservoir domain size and mobility ratio of the displaced and displacing fluid. Also a comparison is made
for a secondary polymer flooding.
The second part of this report investigates the accuracy of DLGR for miscible flow in heterogeneous reser-
voirs, and comparing the consistency of DLGR and the TL model.

iii
P REFACE
I would like to thank the people who have supported me during my thesis. Firstly, Dr. Hadi Hajibeygi I would
like to thank you for being a great teacher as well as an exceptional supervisor. You have always been able to
motivate me and you made time to discuss any issues I faced regardless of your busy schedule. Your empathy
and input has been essential for my work, I truly felt privileged having you as my university supervisor.

Next, I would like to thank my daily supervisor, Hans Groot. Your guidance during my internship has helped
me a lot. You were always open for questions, and our discussions have made me able to improve my knowl-
edge about reservoir simulation.

Also my gratitude goes to Cor van Kruijsdijk for being my mentor and adviser, our meetings have always
been motivating to me. I appreciate the opportunity you have provided me to do my graduation internship.

I want to thank Johan Romate for being part of my exam committee, but also for the discussions we have
had during my thesis work.

Finally I would like to thank Matteo Cusini, Haiyang Cui, Marco Welling, Ali Fadili, Assaf Mar-Or and the
DARSim group for your input and support.

R.A.M. Gielisse
Delft, October 2015

v
C ONTENTS

List of Figures 3
1 Introduction 9
1.1 Introduction to Incomplete Mixing . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 10
1.1.1 Unstable Flow in Porous Media: One-Dimensional Darcy Analysis. . . . . . . . . . . . . 10
1.1.2 Linear Stability Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 12
1.1.3 Effect of Heterogeneity on Flow Through Porous Media . . . . . . . . . . . . . . . . . . 14
1.2 Research Goals . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 14
2 Governing Equations 17
2.1 Physical models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.1.1 Multi-Phase flow model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 17
2.1.2 Miscible Convection-Dispersion model . . . . . . . . . . . . . . . . . . . . . . . . . . 18
2.2 Effective Models . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2.1 Koval Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 19
2.2.2 Todd and Longstaff Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 20
2.2.3 Calibration of Mixing Parameter of Todd and Longstaff . . . . . . . . . . . . . . . . . . 22
3 Numerical Scheme and Simulation Strategy 23
3.1 Multiphase Model. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2 Convection-Dispersion Model . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 24
3.2.1 Flux Approximation on Locally Refined Grid . . . . . . . . . . . . . . . . . . . . . . . . 24
3.3 Simulation Strategy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 25
4 Dynamic Local Grid Refinement 27
4.1 Local Grid Refinement . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 27
4.2 Data Structure . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
4.3 Property calculation . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 28
4.4 Consistency of the Discretization Scheme on Non-Uniform Grid . . . . . . . . . . . . . . . . . 29
4.4.1 Rotation of flux field . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 30
4.5 DLGR Simulation Strategy . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.6 Refinement Criteria . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 33
4.7 Upscaling Permeability Map . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 35
5 Incomplete Mixing for Homogeneous Porous Media 37
5.1 Performance Assessment of DLGR for Miscible Displacements . . . . . . . . . . . . . . . . . . 37
5.2 DLGR Accuracy for Miscible-Solvent Displacements . . . . . . . . . . . . . . . . . . . . . . . 44
5.2.1 Fine scale reference Solution. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 44
5.2.2 Uncertainty within Unstable Flow Simulation . . . . . . . . . . . . . . . . . . . . . . . 51
5.2.3 Error Analysis . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 55
5.3 Consistency Investigation of DLGR and the TL Model for Fully Miscible Solvent Flooding . . . . 63
5.4 Consistency Investigation of DLGR and the TL Model for Polymer Flooding . . . . . . . . . . . 74
6 Incomplete Mixing for Heterogeneous Porous Media 81
6.1 Consistency of DLGR and the TL Model for Fully Miscible Flooding . . . . . . . . . . . . . . . 82
6.2 Accuracy Investigation of DLGR for Fully Miscible Solvent flooding. . . . . . . . . . . . . . . . 94
7 Conclusions 97
7.1 Incomplete Mixing in Homogeneous Porous Media. . . . . . . . . . . . . . . . . . . . . . . . 97
7.1.1 Performance study DLGR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
7.1.2 Accuracy investigation of DLGR . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 97
7.1.3 Consistency Investigation of DLGR and the TL Model . . . . . . . . . . . . . . . . . . . 98

1
2 C ONTENTS

7.2 Incomplete Mixing for Heterogeneous Porous Media . . . . . . . . . . . . . . . . . . . . . . . 99


8 Recommendations for Future Work 101
A One Dimensional Numerical Diffusion Analysis for First Order Schemes 103
B Accuracy investigation DLGR for Polymer Flooding 105
Nomenclature 108
Bibliography 109
L IST OF F IGURES

1.1 Perturbation (δz) introduced at the interface of displacing fluid 2 and displaced fluid 1. Flow is
in the vertical z-direction aligned with gravitatational force direction. . . . . . . . . . . . . . . . . 10
1.2 Pressure distribution for a scenario with two viscous fingers, one slightly ahead of the other. . . 11
1.3 Dimensionless growth rate (σ∗ ) vs dimensionless wave number (k ∗ ) of instabilities for different
dispersion longitudinal (d l ) and transverse (d t ) dispersivities, showing a lowered growth rate
and cut-off value for an increase in dispersion. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 13

2.1 Frontal dispersion for a stable displacement (upper) and unstable displacement (lower). . . . . 21

3.1 Structured equidistant two-dimensional Cartesian grid, with cell-centered pressure unknowns.
Velocity unknowns are obtained at the interfaces. . . . . . . . . . . . . . . . . . . . . . . . . . . . . 23
3.2 (a) Refined interfaces between grid block i and that of j and k. (b) Auxiliary grid blocks, subdi-
viding grid block i in two parts assuming constant pressure. . . . . . . . . . . . . . . . . . . . . . 25
3.3 Flowchart overview of the simulation strategy used in this work to solve for tracer transport. . . 25

4.1 Local grid refinement of grid block (i + 1, j ) with (r x , r y , r z )=(2,2,1). . . . . . . . . . . . . . . . . . . 27


4.2 Nested-grid DLGR data structure used in this project. . . . . . . . . . . . . . . . . . . . . . . . . . 28
4.3 (a) Base grid (3x3) showing a 2x2 refinement in grid block (x i +1 , y j ). (b) Interpolation of pressure
in order to obtain pressure values perpendicular to the coarse-refined grid block interface. . . . 29
4.4 Two dimensional irregular grid with LGR applied in the x-direction (refinement number of 2) in
parent cell (i+1,j). ∆x i + 1 , j and ∆x i + 1 , j +1 represents the average length ∆x between cell nodes
2 2
(i , j ) and (i + 1, j ), (i , j + 1) and (i + 1, j + 1) respectively. ∆y i , j + 1 and ∆y i +1, j + 1 denotes the
2 2
average grid block length ∆y between cell nodes (i , j ) and (i , j + 1), (i + 1, j ) and (i + 1, j + 1)
respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 31
4.5 Rotation in flux field, a consequence of the constant interpolation for pressure. . . . . . . . . . . 32
4.6 (a) Explicit gridding strategy for DLGR. (b) Semi-implicit gridding strategy of DLGR. . . . . . . . 33
4.7 Multi-level grid permeability obtained by flow-based upscaling. . . . . . . . . . . . . . . . . . . . 36
4.8 Flow-based upscaling; calculating the parent grid block’s upscaled permeability in both x and
y-direction. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 36

5.1 Base case reservoir for stable displacement, with domain size 30 by 120 m. The concentration
profile at 0.3 PVI (top) and the imposed DLGR grid at this corresponding time (bottom) are shown. 38
5.2 Mean CPU time per simulation time step [s]. Scalability of stable displacement for domain sizes:
Nz (base case), 2 × Nz (scaled twofold) and 4 × Nz (scaled fourfold). . . . . . . . . . . . . . . . . . 39
5.3 Mean CPU time per simulation time step [s] of DLGR functionalities. Scalability of stable dis-
placement for domain sizes: Nz (base case), 2 × Nz (scaled twofold) and 4 × Nz (scaled fourfold). 40
5.4 Base case reservoir for an unstable displacement (M =10), with domain size 30 by 120 m. The
concentration profile at 0.3 PVI (top) and the imposed DLGR grid at this corresponding time
(bottom) are shown. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 40
5.5 Mean CPU time per simulation time step [s]. Scalability of unstable displacement for domain
sizes: Nz (base case), 2 × Nz (scaled twofold) and 4 × Nz (scaled fourfold). . . . . . . . . . . . . . . 41
5.6 Mean CPU time per simulation time step [s] of LGR functionalities. Scalability of unstable dis-
placement for domain sizes: Nz (base case), 2 × Nz (scaled twofold) and 4 × Nz (scaled fourfold). 42
5.7 Convergence analysis for zero (left column), medium (middle column) and high (right column)
dispersion (case 1 to 3 of table 5.2) on a static Cartesian grid on a 100 by 100 m domain; grid
refinement of 2 × 2 occurs from top left to bottom. The resolutions are (a-c) 25 × 25; ∆h = 4 [m]
, (d-f ) 50 × 50; ∆h = 2 [m], (g-i) 100 × 100; ∆h = 1 [m], (j-l) 200 × 200; ∆h = 0.5 [m] and (m-o)
400 × 400; ∆h = 0.25 [m]. The colorbar (p) quantifies the concentration profiles to the actual
concentration. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 46

3
4 L IST OF F IGURES

5.8 Cumulative oil production shown for different grid resolutions (25 × 25 to 400 × 400) on a 100 by
100 m domain for zero dispersion (case 1 of table 5.2). . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.9 Solvent rate shown for different grid resolutions (25×25 to 400×400) on a 100 by 100 m domain
for zero dispersion (case 1 of table 5.2). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 47
5.10 Transverse averaged concentration (averaged over z-direction) shown for different grid resolu-
tions (25 × 25 to 400 × 400), at 0.2 pvi and 0.6 pvi for zero dispersion (case 1 of table 5.2). . . . . 48
5.11 Cumulative oil production shown for different grid resolutions (25 × 25 to 400 × 400) on a 100 by
100 m domain for medium dispersion (case 2 of table 5.2). . . . . . . . . . . . . . . . . . . . . . . 48
5.12 Solvent rate shown for different grid resolutions (25×25 to 400×400) on a 100 by 100 m domain
for medium dispersion (case 2 of table 5.2). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
5.13 Transverse averaged concentration (averaged over z-direction) shown for different grid resolu-
tions (25 × 25 to 400 × 400), at 0.2 pvi and 0.6 pvi for medium dispersion (case 2 of table 5.2).
. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 49
5.14 Cumulative oil production shown for different grid resolutions (25 × 25 to 400 × 400) on a 100 by
100 m domain for high dispersion (case 3 of table 5.2). . . . . . . . . . . . . . . . . . . . . . . . . . 50
5.15 Solvent rate shown for different grid resolutions (25×25 to 400×400) on a 100 by 100 m domain
for high dispersion (case 3 of table 5.2). . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 50
5.16 Transverse averaged concentration (averaged over z-direction) shown for different grid resolu-
tions (25 × 25 to 400 × 400), at 0.2 pvi and 0.6 pvi for high dispersion (case 3 of table 5.2). . . . . 51
5.17 Range of possible realizations of cumulative oil production for zero dispersion case obtained
on a high resolution ( 400 × 400) Cartesian grid for a 100 × 100 [m] domain, showing the mean
solution (blue) and 3 standard deviations as range of variation (red error bars). Mean variation
from the mean 1%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5.18 Range of possible realizations of solvent rate for zero dispersion case obtained on a high reso-
lution ( 400 × 400) Cartesian grid for a 100 × 100 [m] domain, showing the mean solution (blue)
and 3 standard deviations as range of variation (red error bars). Mean variation from the mean
11%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 52
5.19 Range of possible realizations of cumulative oil production for medium dispersion case ob-
tained on a high resolution ( 400 × 400) Cartesian grid for a 100 × 100 [m] domain, showing the
mean solution (blue) and 3 standard deviations as range of variation (red error bars). Mean
variation from the mean 1.2%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.20 Range of possible realizations of solvent rate for medium dispersion case obtained on a high
resolution ( 400 × 400) Cartesian grid for a 100 × 100 [m] domain, showing the mean solution
(blue) and 3 standard deviations as range of variation (red error bars). Mean variation from the
mean 17.4%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 53
5.21 Range of possible realizations of cumulative oil production for high dispersion case obtained
on a high resolution ( 400 × 400) Cartesian grid for a 100 × 100 [m] domain, showing the mean
solution (blue) and 3 standard deviations as range of variation (red error bars). Mean variation
from the mean 1.1%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.22 Range of possible realizations of solvent rate for high dispersion case obtained on a high reso-
lution ( 400 × 400) Cartesian grid for a 100 × 100 [m] domain, showing the mean solution (blue)
and 3 standard deviations as range of variation (red error bars). Mean variation from the mean
11.4%. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 54
5.23 Concentration profile and imposed grid at 0.6 PVI using DLGR with space-time derivative of
spatial concentration gradient as criteria. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 56
5.24 Tracer mass flux in the z-direction perpendicular to flow in different stages of refinement after
coarsening. Coarsening causing local rotation of velocity field, causing refinement in successive
repeat time steps. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 57
5.25 DLGR quality evaluation for zero dispersion case. (a) Mean error in cumulative oil production
vs mean active grid blocks (magb), (b) mean error in solvent rate vs magb, (c) Mean normalized
l 2 -norm of concentration vs magb, (d) mean repeat time steps vs magb. Properties used for the
criteria are differentiated by color: Concentration (magenta), viscosity (blue), mobility (green).
The operator used within the criteria are indicated in the legends. . . . . . . . . . . . . . . . . . . 58
5.26 Normalized gradients of phase viscosity (blue), concentration (magenta) and phase mobility
(green) against dimensionless length. The corresponding concentration gradient is shown in red. 59
L IST OF F IGURES 5

5.27 Imposed grid for zero dispersion case at 0.4 PVI. Left depicts the solution obtained with viscosity
gradient criterion and right with mobility gradient criterion. The normalized tolerance was 0.05
as shown in table 5.3. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 60
5.28 DLGR concentration solution with imposed grid at 0.6 PVI. Left shows the concentration differ-
ence criteria, right the concentration gradient criteria at a normalized tolerance of 0.2. . . . . . 61
5.29 Active refinement level grid blocks against PVI, illustrating the grid block size distribution of
concentration difference and gradient based criteria, depicted left (a) and right (b) hand side
respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 61
5.30 DLGR quality evaluation for zero dispersion case. (a) Mean error in cumulative oil production
vs mean active grid blocks (magb), (b) mean error in solvent rate vs magb, (c) Mean normalized
l 2 -norm of concentration vs magb, (d) mean repeat time steps vs magb. Properties used for the
criteria are differentiated by color: Concentration (magenta), viscosity (blue), mobility (green).
The operator used within the criteria are indicated in the legends. . . . . . . . . . . . . . . . . . . 62
5.31 DLGR quality evaluation for zero dispersion case. (a) Mean error in cumulative oil production
vs mean active grid blocks (magb), (b) mean error in solvent rate vs magb, (c) Mean normalized
l 2 -norm of concentration vs magb, (d) mean repeat time steps vs magb. Properties used for the
criteria are differentiated by color: Concentration (magenta), viscosity (blue), mobility (green).
The operator used within the criteria are indicated in the legends. . . . . . . . . . . . . . . . . . . 63
5.32 The effect of the mixing parameter on the cumulative oil production of the TL model using a
resolution of 40x40 on a 100 by 100 m domain. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 64
5.33 The effect of the mixing parameter on the produced liquid rates of the TL model using a resolu-
tion of 40x40 on a 100 by 100 m domain. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 65
5.34 The effect of the mixing parameter on the transverse average saturation profile at 0.4 PVI of the
TL model of the TL model using a resolution of 40x40 on a 100 by 100 m domain. . . . . . . . . . 65
5.35 The effect of grid block size and the mixing parameter on the error [%] made by the TL model
compared to a fine scale simulation in cumulative oil production for medium dispersion (case
2 of table 5.2). Both approaches were simulated linear flow on a 100 by 100 m domain, fine scale
simulation resolution was 400x400. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 66
5.36 The effect of grid block size and the mixing parameter on the error [%] made by the TL (rescaled
color bar) model compared to a fine scale simulation in cumulative oil production for medium
dispersion (case 2 of table 5.2. Both approaches were simulated linear flow on a 100 by 100 m
domain, fine scale simulation resolution was 400x400. . . . . . . . . . . . . . . . . . . . . . . . . . 67
5.37 Scalability evaluation results for cumulative oil production for high resolution simulations (red),
DLGR (green) and TL (blue). The mean active grid block used by DLGR is 28, 29, 33 and 34%
(compared to the fine scale simulations grid size) for 200 (a), 100 (b), 50 (c) and 25 m (c) respec-
tively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 68
5.38 Scalability evaluation results for oil rate for high resolution simulations (red), DLGR (green) and
TL (blue). The mean active grid blocks used by DLGR is 28, 29, 33 and 34% (compared to the
fine scale simulations grid size) for 200 (a), 100 (b), 50 (c) and 25 m (c) respectively. . . . . . . . 69
5.39 Scalability evaluation results for transverse averaged concentration/saturation for high resolu-
tion simulations (red), DLGR (green) and TL (blue). The mean active grid block used by DLGR
is 28, 29, 33 (compared to the fine scale simulations grid size) and 34% for 200 (a), 100 (b), 50 (c)
and 25 m (c) respectively. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 70
5.40 Consistency evaluation of TL and DLGR regarding mobility ratio between displacing and dis-
placed fluids for zero dispersion (case 1). The different colors depict the different mobility
ratios; M=10 (red), M=100 (green) and M=500 (magenta). Two different results for DLGR are
shown, obtained with different average active grid block counts (aagc) over simulation time. . . 71
5.41 Consistency evaluation of TL and DLGR regarding mobility ratio between displacing and dis-
placed fluids for medium dispersion (case 2). The different colors depict the different mobility
ratios; M=10 (red), M=100 (green) and M=500 (magenta). Two different results for DLGR are
shown, obtained with different average active grid block counts (aagc) over simulation time. . . 72
5.42 Consistency evaluation of TL and DLGR regarding mobility ratio between displacing and dis-
placed fluids for high dispersion (case 3). The different colors depict the different mobility
ratios; M=10 (red), M=100 (green) and M=500 (magenta). Two different results for DLGR are
shown, obtained with different average active grid block counts (aagc) over simulation time. . . 73
6 L IST OF F IGURES

5.43 Mixing parameter match against the logarithm of mobility for zero, medium and high dispersed
cases. The results from the mixing parameter relation (Christie 1993 [1]) is shown in yellow
showing good results for zero dispersion cases. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 73
5.44 Solution profiles for the 0.1 PV polymer slug injection at 0.2 PVI liquid injection. From top to
bottom illustrating the solutions of the fine scale simulation and DLGR (without and without
grid), respectively. The left hand side shows the polymer concentration while the right hand
side illustrates the corresponding oil saturation. Resolutions of fine scale simulation and DLGR
are 80x800, 5x50 (base grid; using 2x2 refinements until maximum level of 4). . . . . . . . . . . . 75
5.45 Solution profiles for the 0.1 PV polymer slug injection at 0.4 PVI liquid injection. From top to
bottom illustrating the solutions of the fine scale simulation and DLGR (without and without
grid), respectively. The left hand side shows the polymer concentration while the right hand
side illustrates the corresponding oil saturation. Resolutions of fine scale simulation and DLGR
are 80x800, 5x50 (base grid; using 2x2 refinements until maximum level of 4). . . . . . . . . . . . 76
5.46 Solution profiles for 0.1 PV polymer slug followed by chase water injection at 0.2 PVI (a-d) and
0.4 PVI (e-h) liquid injection, for the high resolution simulation solutions (even rows) and the
solution obtained with the TL model (odd rows). The left hand side shows the fraction of poly-
mer occupied water phase, the right hand side illustrates the corresponding oil saturation. Res-
olutions of high resolution simulation and TL are 80x800 and 8x80, respectively. . . . . . . . . . 77
5.47 Cumulative oil production for the 0.3 PV (red) and 0.1 PV (blue) polymer slug injection. Fine
scale simulations are shown in solid lines, DLGR by dashed lines and TL with solid line with
circled markers. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 78
5.48 Oil rate of the 0.3 PV (red) and 0.1 PV (blue) polymer slug injection. Fine scale simulations are
shown in solid lines, DLGR by dashed lines and TL with solid line with circled markers. . . . . . 78
5.49 Transverse average oil saturation for the 0.3 PV (left) and 0.1 PV (right) polymer slug injection
at two different total PVI (red: 0.2PVI, blue: 0.4PVI). Fine scale simulations are shown in solid
lines, DLGR by dashed lines and TL with solid line with circled markers. . . . . . . . . . . . . . . 79

6.1 Permeability realization with dimensionless correlation length λD =0.01 and Vdp =0.63. . . . . . . 83
6.2 Concentration solution for a heterogeneous reservoir (Vdp =0.63 and λD =0.01, 20 by 100 m) of
static high resolution, DLGR, DLGR with imposed grid and TL consecutively. The mixing pa-
rameter for TL was equal to 0.6 and the mean active grid blocks used by DLGR equaled 35% of
high resolution grid. Resolutions of high resolution simulation,DLGR and TL are 80x400, 5x25
(base grid; using 2x2 refinements until maximum level of 4) and 8x40. . . . . . . . . . . . . . . . 84
6.3 Concentration solution for a heterogeneous reservoir (Vdp =0.63 and λD =0.01, 20 by 100 m) of
static high resolution, DLGR (shown with and without the imposed grid) and TL, respectively.
The mixing parameter for TL was equal to 0.6 and the mean active grid blocks used by DLGR
equaled 17% of high resolution grid. Resolutions of high resolution simulation,DLGR and TL
are 80x400, 5x25 (base grid; using 2x2 refinements until maximum level of 4) and 8x40. . . . . . 85
6.4 Averaged solutions comparing high resolution simulation (80x400) with DLGR and TL (8x40)
(ω=0.6) for heterogeneous case with λD =0.01 of the 20 by 100 m. . . . . . . . . . . . . . . . . . . . 86
6.5 Permeability realization with dimensionless correlation length λ=0.1 and Vd p=0.63 of a domain
of 20 by 100 m. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 87
6.6 Concentration solution for a heterogeneous reservoir (Vdp =0.63 and λD =0.1, 20 by 100 m) of
static high resolution, DLGR, DLGR with imposed grid and TL consecutively. The mixing pa-
rameter for TL was equal to 0.375 and the mean active grid blocks used by DLGR equaled 44%
of high resolution grid. Resolutions of high resolution simulation,DLGR and TL are 80x400, 5x25
(base grid; using 2x2 refinements until maximum level of 4) and 8x40. . . . . . . . . . . . . . . . 88
6.7 Concentration solution for a heterogeneous reservoir (Vdp =0.63 and λD =0.1, 20 by 100 m) of
static high resolution, DLGR, DLGR with imposed grid and TL consecutively. The mixing pa-
rameter for TL was equal to 0.375 and the mean active grid blocks used by DLGR equaled 29%
of high resolution grid. Resolutions of high resolution simulation,DLGR and TL are 80x400, 5x25
(base grid; using 2x2 refinements until maximum level of 4) and 8x40. . . . . . . . . . . . . . . . 89
6.8 Averaged solutions comparing high resolution simulation (80x400) with DLGR and TL (8x40)
(ω=0.375) for heterogeneous case with λD =0.1 of the 20 by 100 m domain as shown in the Figures
above. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 90
L IST OF F IGURES 7

6.9 Averaged solutions comparing high resolution simulation (80x400) with DLGR and TL (8x40)
(ω=0.6) for heterogeneous case with λD =0.01 of the 20 by 100 m domain. . . . . . . . . . . . . . 92
6.10 Averaged solutions comparing high resolution simulation (80x400) with DLGR and TL (8x40)
(ω=0.375) for heterogeneous case with λD =0.1 of the 20 by 100 m domain. . . . . . . . . . . . . . 93
6.11 DLGR accuracy evaluation for heterogeneous medium (Vd p =0.63 and λD =0.01, 20 by 100 m). (a)
Mean error in cumulative oil production vs mean active grid blocks (magb), (b) mean error in
solvent rate vs magb, (c) Mean normalized l 2 -norm of concentration vs magb, (d) mean repeat
time steps vs magb. Properties used for the criteria are differentiated by color: Concentration
(magenta), viscosity (blue), mobility (green). The operator used within the criteria are indicated
in the legends. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 95
6.12 DLGR accuracy evaluation for heterogeneous medium (Vd p =0.63 and λD =0.1, 20 by 100 m). (a)
Mean error in cumulative oil production vs mean active grid blocks (magb), (b) mean error in
solvent rate vs magb, (c) Mean normalized l 2 -norm of concentration vs magb, (d) mean repeat
time steps vs magb. Properties used for the criteria are differentiated by color: Concentration
(magenta), viscosity (blue), mobility (green). The operator used within the criteria are indicated
in the legends. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 96

B.1 DLGR accuracy evaluation for polymer flood of secondary 0.1 PV polymer slug chased by water
injection until 1.5 PVI. (a) Mean error in cumulative oil production vs mean active grid blocks
(magb), (b) mean error in solvent rate vs magb, (c) Mean normalized l 2 -norm of concentration
vs magb, (d) mean repeat time steps vs magb. Properties used for the criteria are differentiated
by color: Concentration (magenta), viscosity (blue), mobility (green). The operator used within
the criteria are indicated in the legends. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 105
1
I NTRODUCTION
Reservoir simulation is a key tool to get insight in complex fluid flow phenomena occurring in many petroleum
reservoirs and, therefore, creating the possibility to quantify fluid productions and their uncertainty regard-
ing many unknown properties.
Much experimental and theoretical research [2–6] has shown that immiscible (multiphase) and miscible
(single-phase) displacements can develop instabilities under certain criteria. Differences in fluid properties
such as viscosity and density across the displaced front lead to instabilities, causing the displacing fluid to
finger through the displaced oil, which is referred to as viscous fingering [7]. This leads to early breakthrough
and therefore decreases the production efficiency. Moreover, reservoir rock heterogeneity has shown to have
a dominating effect on how fluids are displaced through porous reservoir rock. Depending on the variance
and spatial correlation of the permeability field, it can either accentuate or suppress the viscous fingering
phenomenon, depending on the degree of heterogeneity and its correlation length [8–14].
To limit computational costs, reservoir simulations typically use far lower grid resolution than the scale
at which viscous fingering and physical dispersion at the front occurs [15, 16]. Numerical truncation er-
rors imposed by large grid blocks will conceal small scale physical phenomena that contribute to unstable
displacements. These sub-grid flow features are called “incomplete mixing”, as opposed to the fully mixed
scenario inherent to the use of low resolution grids. To cope with this problem, effective models have been
developed that introduce effective upscaled fluid properties, in order to capture the sub-grid effect of incom-
plete mixing [17–19]. Among these models, Todd and Longstaff’s (TL) model [18] is the most widely utilized
model in the petroleum industry. Similar to alternative models, a major disadvantage of TL approach is that
the effective fluid properties are described by an undefined scalar mixing parameter, which must be obtained
from either high resolution simulation or experimental data. The use of full-field high resolution simulations
in order to find the values of these tuning (effective) parameters are computationally expensive. Moreover,
extrapolation of these simulations to different scenarios is questionable, because local mixing is naturally
dependent on changes in fluid properties, scale, geometry and rock heterogeneity.
As mentioned before, computational limitations make classical numerical schemes impractical for ac-
curate simulation of fine-scale displacements in filed-scale geological formations. On the other hand, the
accuracy of effective (upscaled) approaches, in general, is questionable. This challenge motivates the de-
velopment and extension of Dynamic Local Grid Refinement (DLGR) method for processes with incomplete
mixing. DLGR and other members of the rich family of adaptive grid (mesh) refinement approaches are sim-
ulation techniques that adapt the spatial grid resolution to the scale at which the physical phenomena occur
without refining the grid over the entire domain [20–26]. This can be achieved by means of local refinement
and coarsening of the computational grid dynamically at every time step where small scale, local phenomena
(e.g. steep pressure and/or concentration gradients that take place, e.g., near wells and at displacing fronts)
occur, on the basis of predefined error criteria. As such, DLGR provides an accurate and efficient framework
to solve for complex flow phenomena arising from instabilities at fluid fronts, while preserving complex geo-
logical features that could accentuate this phenomena.
Next, an introduction to displacements with incomplete mixing is presented.

9
10 1. I NTRODUCTION

1.1. I NTRODUCTION TO I NCOMPLETE M IXING


Due to the large length scales of natural formations (including petroleum reservoirs), reservoir simulators
employ computational grid block sizes ranging from ten to hundreds of meters [15, 16]. As for a grid block
only a single value for each dedicated property is calculated (e.g., phase saturation), the simulation assumes
complete mixing on the scale of these large grid block sizes. However, it is know that in reality highly complex
flow can evolve at scales much lower than these simulation grid scales. The sub-grid flow features can show
incomplete mixing of fluids, as opposed to the concealed fully mixed assumption of low resolution simulation
grids. An example of incomplete mixing arises when under certain criteria fluid flow through porous media
becomes unstable, during which the displacing fluid penetrates through the displaced fluid. This phenom-
ena is called viscous fingering [7], resulting in early breakthrough [17] and, therefore, inefficient production.
The main reasons behind these instabilities are determined to be the infavorable variation of viscosity and
density across the front [7]. When two fluids are miscible, meaning that that there is no interfacial tension
between them, both of these properties (density and viscosity contrast) depend on the solute concentration.
In such cases, molecular diffusion and mechanical dispersion can affect the stability , thus need to be also
considered. Although fluid properties can render the stability of the displacing front, static disorder caused
by reservoir rock heterogeneity has found to have a dominating effect on fluid flow, having the potential of
creating channeling pathways or strong dispersion of the front[8, 13].
In this section an introduction is given to what is previously described as incomplete mixing phenomena. A
one dimensional Darcy analysis is presented from which criteria are derived for stable and unstable frontal
displacements in porous media. Apart from the criteria that causes instabilities, following the literature, the
linear stability analysis (LSA) is also presented. This analysis allows for describing effects of the different
forces imposing instabilities on the growth rate of viscous fingers. Finally, the effect of static disorder caused
by reservoir rock heterogeneity on these instabilities will be explained.

1.1.1. U NSTABLE F LOW IN P OROUS M EDIA : O NE -D IMENSIONAL D ARCY A NALYSIS


To understand the basic mechanism and the forces that cause a displacing front in porous media to become
unstable, an analysis based on Darcy’s law for fluid flow in one dimension is considered [4, 7]. Two fluids
(1 and 2) fill a vertical porous medium, having a sharp interface with a small perturbation at their contact
location, as depicted in Figure 1.1. Here, µ and ρ stand for fluid viscosity and density, respectively. Also, the
sub-index 1 and 2 denote the corresponding fluid properties.

X
direction of flow

Fluid 2

δz

Fluid 1

Figure 1.1: Perturbation (δz) introduced at the interface of displacing fluid 2 and displaced fluid 1. Flow is in the vertical z-direction
aligned with gravitatational force direction.

Assuming a homogeneous isotropic medium permeability, k, with appropriate continuum assumptions for
incompressible flow, the Darcy’s law for vertical flow (being parallel to the direction of gravity force) can be
expressed as

k
u = − ∇(p + ρg z). (1.1)
µ
1.1. I NTRODUCTION TO I NCOMPLETE M IXING 11

Here, u is the superficial (Darcy) velocity, which is measured in SI unit in [ m ], k the absolute permeability
2
£ kg s
having the dimension of [m ], p the pressure in [Pa], ρ the density in m3 ] and g is the gravitational accel-
eration measured in [ sm2 ] dimension in SI units. Rewriting Eq. (1.1) in terms of the pressure drop across the
perturbed interface resulted from a virtual displacement [7] gives

µ1 − µ2 U ¡
·¡ ¢ ¸
δp = p 2 − p 1 = + ρ 1 − ρ 2 )g δz.
¡ ¢
(1.2)
k

As shown in Eq. (1.2), the viscosity and density ratio across the interface as well as the velocity of flow can
render the front unstable, i.e., causing the perturbation to grow. Assuming a positive downward pressure
gradient, if both µ2 > µ1 and ρ 2 > ρ 1 hold, a virtual displacement of the perturbation is evoked. Scaling the
velocity such that the pressure gradient is zero, i.e., δp/δz = 0, the critical velocity below which flow is always
stable can be found as
ρ 1 − ρ 2 )g k
¡
Uc = ¡ . (1.3)
µ1 − µ2 )
In the absence of gravitational effect, e.g. for horizontal flows, the stability criterion reduces to

µ2 < µ1 , (1.4)

meaning that the perturbation grows if the condition (1.4) holds. This simple analysis is valid if the fluid
flow is incompressible, yet note that it holds valid for both miscible and immiscible displacements, if an
initially sharp interface between the fluid phases exists [7, 27]. Also note that the relative permeability, and
important factor for immiscible displacements, is also neglected in this simple -yet important- study [28].
It is important to be mentioned that specially for miscible displacements, mechanical dispersion and (in
lesser extend) molecular diffusion reduce the sharp contrast of the frontal interface over time, resulting in
the formation of a time-dependent mixing zone. Therefore, criterion (1.4) only holds if the viscosity is a
monotonic function of concentration.
Several experimental studies have been performed to describe different patterns under which viscous fin-
gering occurs. These patterns include shielding, spreading and splitting [5, 7, 29]. As viscous fingers develop,
the leading fingers (ahead of their neighbors) have the tendency to outgrow the fingers lacking behind, which
is called shielding. As shown in Figure 1.2, the pressure gradient inside the displacing fluid is quite smaller
than the displaced one. As such at the tips of the fingers a relatively large pressure gradient exists [7]. The
length-scale of the dominating fingers is affected by the transverse dispersion (for miscible displacements)
and interfacial tensions (IFT, for immiscible displacements). This will be addressed in more detail later in
subsection 1.1.2. In addition, a phenomenon called “ tip splitting ” also affects the length-scale of the domi-
nating fingers. Notice that both transverse dispersion and IFT damp disturbances with length scales below a
certain cut-off value, allowing for instabilities with higher length scales to dominate. On the other hand, these
effects (diffusion and IFT) can also increase viscous fingers spreading in the lateral direction. This increase
can lead fingers exceed the local stability limit, and this, may lead to forming new fingers [7].

Figure 1.2: Pressure distribution for a scenario with two viscous fingers, one slightly ahead of the other.
12 1. I NTRODUCTION

1.1.2. L INEAR S TABILITY A NALYSIS


Characteristics of viscous fingering at early stages have been addressed through several studies in the litera-
ture, on the basis of the onset of instabilities by means of the linear stability analysis (LSA) [3, 4, 27]. To the
best of the author’s knowledge, Hill was perhaps the first who observed instabilities in two-phase flow in
porous media, the phenomena that he reported as "channeling" [4]. However, after the study of instabilities
for immiscible flows by Saffman and Taylor [3], this phenomenon was commonly known as the Saffman-
Taylor instability.
LSA is a powerful mathematical procedure that can be employed to describe the onset of viscous fingers
in homogeneous media, with an initial sharp interface between phases (or components). In this approach,
disturbances are considered to be of the Fourier (exponential complex) functional form of

ξ(σ, k) =e (σt +i k y) , (1.5)

where σ is the growth rate of the perturbation over time t . In addition, i is the imaginary number, as such
describing the oscillatory component of the error, and k stands for the k-th wave number of the perturbation
along y-direction perpendicular to the flow direction (z-direction for the case of Figure 1.1). Next, the solu-
tions of LSA for immiscible and miscible displacements are given [2, 30].

For immiscible displacements it is shown that the dimensionless parameters


µ1 u
C a0 = , (1.6)
γK
µ1 − µ2
A= , (1.7)
µ1 + µ2
and
(ρ 1 − ρ 2 )g K
G= (1.8)
(µ1 + µ2 )u
are important [2, 7]. Here C a 0 is the modified capillary number, A is a normalized viscosity difference quan-
N
tity, and G is the gravity number. Note that µ is the viscosity in [Pa · s], γ the interfacial tension in [ m ], g the
gravitational acceleration in [ sm2 ], K the matrix permeability in [m 2 ] and u the Darcy velocity in [ ms
s ].

Following LSA, the initial growth rate of instabilities , σ, can be descrived as a function of the wave length of
the instability,k, [7] as
(A + 1)k 3
σ = (A +G)k − . (1.9)
2C a 0
If capillary pressure is neglected (i.e. γ = 0), Eq. (1.9) predicts a monotonically increasing function of the
wave number k for growth rate σ, i.e., σ = (A + G)k. Therefore, surface tension has a dampening effect on
short wave numbers. From Eq. (1.9), the wave number k max at which the growth rate is maximum can be
expressed as
¶1
2(A +G)C a 0 2
µ
k max = . (1.10)
3(A + 1)
Likewise, for miscible displacements, i.e. when no interfacial tension between fluids exists, the LSA ap-
proach allows for understanding of instability growth. More about miscible displacements can be found in
chapter 2. For miscible flows dispersion, in contrast to the case of immiscible flows, damps instabilities with
higher than a cut-off value. Also note that, unlike immiscible displacements, there is no steady state solution
for miscible flows, since dispersion will continuously spread the tracer at the front and, therefore, changes
the viscosity over time. The steady-state situation is reached when no concentration gradient in the entire
domain exists. As LSA assumes a sharp front between fluids, the following relationships are only applicable
for short times after the perturbation occurs, such that dispersion has not yet effected the front. Under this
situation, for miscible displacements, the dimensionless parameters

Ll
Pe l = , (1.11)
dl

Lt
Pe t = , (1.12)
dt
1.1. I NTRODUCTION TO I NCOMPLETE M IXING 13

and
µ1 − µ2
A= (1.13)
µ1 + µ2
play role in the flow regime, if gravity is neglected [7].
Here Pe l and Pe t are the longitudinal and transverse non-dimensional Peclet number, respectively, de-
scribing the ratio between convectional effects and longitudinal and transverse dispersion. Similar to immis-
cible flows, the non-dimensional mobility ratio number is also an important factor.
Assuming an initial sharp concentration profile, the dimensionless (denoted as ∗ ) perturbation growth rate,
as described in [30], is given by
1h 1
i
σ∗ = (Ak ∗ − k ∗2 ) − k ∗ (k ∗2 + 2Ak ∗ ) 2 + (1 − ²)k ∗2 , (1.14)
2
where ² is the ratio between the transverse and longitudinal dispersion, i.e.,
Dt
²= .
Dl
For linear flows, being the case of this mathematical description, the dispersion coefficients read
D l = dl u (1.15)
and
D t = d t u. (1.16)
As such, the dimensionless growth rate and wave number are given as
u2
σ = σ∗
Dl
and
u
k = k∗ .
Dl
More discussions about the diffusion coefficients will be presented in chapter 2.
It can be seen from equation (1.14) that apart from the mobility ratio also the degree of longitudinal disper-
sion and the ratio of transverse over longitudinal dispersion play a big role on growth rate of instabilities’
wave lengths. The dependency of growth rate on physical dispersion for the mobility ratio of 10 is illustrated
in Figure 1.3. In this figure, the green curve illustrates the case with isotropic dispersion of d l = d t = 0.1[m]),
the red and blue curves represent the increase of the longitudinal and transverse dispersivities, respectively.
Although both forces reduce the growth rate, and increase dominant wave lengths of viscous fingers, longi-
tudinal dispersion has a greater effect on the growth rate while transverse dispersion acts more strongly as a
cut-off for higher wave numbers (increasing the viscous finger width).

Figure 1.3: Dimensionless growth rate (σ∗ ) vs dimensionless wave number (k ∗ ) of instabilities for different dispersion longitudinal (d l )
and transverse (d t ) dispersivities, showing a lowered growth rate and cut-off value for an increase in dispersion.
14 1. I NTRODUCTION

1.1.3. E FFECT OF H ETEROGENEITY ON F LOW T HROUGH P OROUS M EDIA


As described in the previous sections, fluid properties can determine the stability of the frontal displacement
in porous media. The presented analyses, however, assumes homogeneous porous media which is an unre-
alistic assumption for petroleum reservoirs. Due to the fact that the depositional environments change over
space and geological time, properties such as grain size, porosity and permeability vary with location and
scale. These variations in rock properties in space and scale, called heterogeneity, have a dominating effect
on how fluid flows through petroleum reservoirs [8–14].

Permeability distributions of reservoir formations could be described using log-normal distribution. The
Dykstra-Parson coefficient [31] is a measure of heterogeneity, widely used in the petroleum industry. It char-
acterizes the variation permeability as
k 50 − k 84.1
Vdp = . (1.17)
k 50
Here, k 50 indicates the mean permeability and k 84.1 permeability of the mean plus one standard deviation.
The Dykstra-Parson coefficient ranges from 0 for homogeneous media to 1 for highly heterogeneous media.
Apart from the variation of permeability, also the spatial correlation of the inhomogeneities is of great influ-
ence to fluid flow. The dimensionless correlation length λL reads

l
λL = . (1.18)
L
Here l and L are the correlation and domain length, respectively. The dimensionless correlation length λL can
vary from 0 to 1 for spatially uncorrelated to stratified reservoir formations, respectively. It is the relationship
between both the variance, here denoted as Vd p , and the correlation length λL that influences how fronts are
displaced through the reservoir. Assuming that the stability criterion given by 1.3, or 1.4 for horizontal flow,
is violated it has been observed that, for low variation of permeability, viscous fingering occurs irrespective of
the correlation length of the inhomogeneities [8]. A strong increase of the variation in permeability (higher
Vd p ) causes abrupt changes in the direction of the flow-lines. This effect causes the instabilities created at
the front to disperse and, therefore, has shown to have a stabilizing effect to displacements that surpass the
stability criterion [8]. As the correlation length increases, however, preferred paths are present that dominate
how the fluids flows through the porous medium. This is known as channeling. Although mobility contrast
can amplify this channeling phenomena, it is not similar to viscous fingering. Instead, it results from corre-
lated static disorder of the reservoir rock [8, 13].

1.2. R ESEARCH G OALS


It is evident from the previous sections that incomplete mixing, both occurring in homogeneous and hetero-
geneous systems, can highly alter production forecasts. Due to computational limitations, current methods
of estimating production and reservoir management strategies, in presence of incomplete mixing processes
are based on effective (upscaled) models. This research investigates the applicability of state-of-the-art DLGR
method, as an alternative advanced method for replacement of excessively employed effective model devel-
oped by Todd and Longstaff (TL). As DLGR is capable of adapting the spatial grid resolution locally where
physical phenomena of incomplete mixing occur, this technique potentially allows for computational effi-
cient and, at the same time, accurate results. The research goals are stated as following:

1. Assess the accuracy and efficiency (applicability) of DLGR as a next-generation simulation strategy for
displacement with incomplete mixing.

2. Investigate the accuracy of the TL and DLGR compared with fully resolved reference simulations.

The studies presented in this work are obtained based on a commercial in-house simulator of the sponsoring
company.
The report is structured as following. Governing equations for both immiscible and miscible transport in
porous media are provided in chapter 2. In the same chapter, the TL effective model is also described. In
chapter 3 the finite volume discretization (FVM) used to simulate the governing equations as well as the
solution strategy is presented. chapter 4 provides the theory of DLGR as implemented in the commercial
simulator. The results regarding the research goals are illustrated and discussed for homogeneous porous
1.2. R ESEARCH G OALS 15

media in chapter 5 and for heterogeneous porous media in chapter 6. Finally chapter 7 will conclude the
results and will provide a brief discussion on future work.
2
G OVERNING E QUATIONS
Mathematical formulations describing multiphase miscible flow in porous media entail several complexities
due to nonlinear fluid and heterogeneous rock properties. Such complexities make analytical approaches
applicable only if several assumptions for the fluid physics and reservoir properties are made. As such, com-
putational methods are typically used in order to obtain a more accurate estimate of the complex flow and
transport processes. Although being more accurate in the sense that they impose minimum simplifications
of the physics and geological complexities, they are subject to truncation errors which tend to zero in the
limit of the grid and time step sizes. Before addressing the simulation strategy followed in the work, first,
in this chapter, the governing equations describing immiscible and miscible displacements are provided. In
addition, this chapter addresses effective (upscaled) models for miscible flows.

2.1. P HYSICAL MODELS


2.1.1. M ULTI -P HASE FLOW MODEL
Mass conservation for phase α ∈ {1, ..., N p }, where N p is the total number of phases, reads

∂(φρ α S α )
+ ∇ · ρ α u α ) = qα ρ α ,
¡
(2.1)
∂t
where φ is the porosity, ρ α phase density and S α is the phase saturation. Moreover, u is the Darcy velocity
and q α is the source term. Assuming incompressible fluid and rock, Eq. (2.1) reduces to

∂S α
φ
¡
+ ∇ · u α ) = qα . (2.2)
∂t
Darcy’s law is employed to relate the phase velocity to pressure (potential) gradient, which reads

kk r,α ¡
uα = − ∇ p α + ρ α g h), (2.3)
µα

where k denotes the permeability tensor, µα the phase viscosity, k r,α is the relative permeability, and pα the
phase pressure. Also, g denotes the gravitational acceleration acting in h direction.
Once Eq. (2.3) is substituted in Eq. (2.2), a well-posed system for N p phase saturations and one pressure
(no capillary effect, i.e., p α = p ∀α) unknowns is obtained if the physical constraint

Np
X
S α = 1, (2.4)
α=1

is also used. This constraint states that the phase saturations fill up the entire pore space.
Relative permeability is a function of saturation, which is typically obtained based on experiments. In this
work, the Brooks-Corey model [32] relations are used, i.e.,
2+3n
k r,w = k r,w,e (S w,eff ) n (2.5)

17
18 2. G OVERNING E QUATIONS

2+n
k r,o = k r,o,e (1 − S w,eff )2 (1 − S w,eff
n
). (2.6)
Here, k r,w,e and k r,o,e are the end-point relative permeabilities of water and oil, respectively, k r,o,e the end-
point oil relative permeability, n the sorting coefficient and S w,eff the effective water saturation defined as

S w − S w,c
S w,eff = , (2.7)
1 − S w,c − S o,r

where S w,c is the connate water saturation and S o,r the irreducible water saturation.

2.1.2. M ISCIBLE C ONVECTION -D ISPERSION MODEL


Peaceman and Rachford developed a system of partial differential equations (PDE) to describe transport of
solvent [33], with properties being averaged over scales much larger that of the pore size. In their formula-
tions, they considered buoyancy forces, viscosity and density of the fluids, spatial distribution of permeabil-
ity, molecular diffusion and dispersion. Assuming that the two fluids (e.g., solvent and oil) are first-contact
miscible, the Peaceman model reads
∂φw s
+ ∇ · w s u − ∇ · φD∇w s = q s ,
¡ ¢ ¡ ¢
(2.8)
∂t
kg 2
where w s is the weight concentration of the solvent [ m3 ] and D is the dispersion tensor [ ms ] [34]. Moreover,
kg
q s denotes the solvent source term [ m3 s ]. Assuming that the fluids and rock are incompressible and that no
volume change occurs upon mixing (ideal mixing), the mass conservation can be rewritten as

∂c
φ + ∇ · cu) − ∇ · φD∇c) = q̃ s ,
¡ ¡
(2.9)
∂t
where the solvent volume fraction, c, is defined as
ws
c= . (2.10)
ρs

Darcy’s velocity, u, reads


k
∇ p + ρ mi x g h),
¡
u=− (2.11)
µmi x
which is a single phase flow formulation, since both solvent and hydrocarbons are assumed to mix instanta-
neously on molecular level. In Eq. (2.11), the mixture viscosity, µmi x , is calculated based on a quarter-power
mixing rule [17], i.e.,
1 − c) −4
¡
c
µ ¶
µmi x c) =
¡
1
+ 1
. (2.12)
µs4 µo4
Here, the mixture (phase) density ρ mi x is found based on ideal mixing, i.e.,

ρ mi x c) = cρ s + 1 − c)ρ o .
¡ ¡
(2.13)

Molecular diffusion and mechanical dispersion [34] are described in the tensorial coefficient D as
¡
D x, u) = d m I + |u|d l û ⊗ û + |u|d t [I − û ⊗ û]. (2.14)
q
u
Here û = |u| is the unit velocity, where |u| = u x2 + u 2y . This results in the dispersion tensor for two-dimensional
displacements as
· ¸
D xx Dxy
, (2.15)
Dyx Dyy
where
u x2 u 2y
D xx = d m + d l + dt , (2.16)
|u| |u|
u 2y u x2
D y y = dm + dl + dt , (2.17)
|u| |u|
2.2. E FFECTIVE M ODELS 19

and
ux u y
D x y = D y x = (d l − d t ) . (2.18)
|u|
2
Here d m is the molecular diffusion constant [ ms ], and d l , and d t are the longitudinal and transverse disper-
sion constants [m], respectively.

Defining the dimensionless parameters [35], denoted by D subindex, as

φL x A q inj
t= tD , x = L x xD , y = L y y D , u = uD
qi n j A
q inj L x q inj L y
D xx = D xx,D , D y y = D y y,D ,
A A

allows for representation of Eq. (2.9) for non-dimensional concentration c D as

∂c D ∂ ∂ 1 ∂c D ∂ 1 ∂c D
+ (c D u D,x ) − ( )− ( ) = qD , (2.19)
∂t D ∂x D ∂x D Pel ∂x D ∂y D Pet ∂y D

where molecular diffusion is neglected and

1 Ll
Pel = = , (2.20)
Dxx,D dl

and

1 Lt
Pet = = . (2.21)
Dyy,D dt

Here, Pe l and Pe t are the longitudinal and transverse Peclet number, respectively.

2.2. E FFECTIVE M ODELS


Effective models have been developed in order to solve for incomplete mixing on scales larger than that of the
physics describing it [17–19]. These empirical models make use of effective fluid parameters, such that their
effective results fit the data obtained from experimental work or high resolution simulations. This way the
tuned models are used in field development plans to predict the consequences of incomplete mixing using
low resolution (computational efficient) grids.

2.2.1. KOVAL M ODEL


As the Peaceman model fails to describe the effect of viscous fingering on production forecasts in one dimen-
sion, Koval suggested the use of a modified Buckley-Leverett model for immiscible displacements in 1963
[17]. This modified Buckley-Leverett model, describing miscible displacements, is based on the multi-phase
flow model assuming incompressible fluids, and neglecting capillary pressure. Koval’s model for the solvent
"phase" reads
∂S s u t ∂ f (S s )
+ = q̃ s . (2.22)
∂t φ ∂x
Here, S s denotes the solvent phase saturation and u t is the total velocity, i.e.,
³S So ´
s
u t = −k + ∇p. (2.23)
µs µo

Moreover, f (S s ) is the fractional flow function defined as

Ss
µs λs (S s )
f (S s ) = = . (2.24)
So
+ Ss λo (S s ) + λs (S s )
µo µs
20 2. G OVERNING E QUATIONS

Finally, q˜s the source term of the solvent phase [ 1s ]. The Koval model assumes first contact miscible dis-
placements, and then introduces a modification to the fractional flow function by incorporating the effective
mobility ratio of the solvent-oil interface for homogeneous media, i.e.,

Ss
f (S s ) = , (2.25)
S s + (1 − S s ) M1
eff

where
µo
M eff = . (2.26)
µs,eff
Here, µo is the oil viscosity, and the µs,eff is the effective solvent viscosity defined in Eq. (2.12). Koval found
that one single effective solvent volume fraction of 0.22 was sufficient in determining the effective mobility
that fitted the experimental data of miscible floods in homogeneous cores.He then extended the fractional
flow formulation to include the effect of heterogeneity by multiplying it with the heterogeneity parameter H ,
i.e.,
µo
M eff = H . (2.27)
µs,eff
The factor H was found to be a function of the variation of permeability expressed using the Dykstra-Parson
coefficient, i.e.,
VDP
log(H) = . (2.28)
(1 − VDP )0.2

2.2.2. T ODD AND L ONGSTAFF M ODEL


Todd and Longstaff (TL) approach was based on the black oil model describing multiphase flow through
reservoir rock [18]. The motivation behind their work was to capture the sub-grid effect of mixing in the typ-
ical reservoir simulator grid block sizes. This model, however, introduces a tuning parameter, which requires
a fine-scale fully resolved simulation to be found, and, as such, its accuracy and applicability for field-scale
problems is questionable.
Mass balance equations for a system consisting of three phases of water (w), oil (o) and solvent (s) can be
described as
∂S w ¡ kk r,w ¡
φ ∇ p + ρ w g z = qw ,
¢¢
+∇· − (2.29)
∂t µw

∂S o ¡ kk r,o ¡
φ ∇ p + ρ o g z = qo ,
¢¢
+∇· − (2.30)
∂t µo
and
∂S s ¡ kk r,s ¡
φ ∇ p + ρ s g z = qs ,
¢¢
+∇· − (2.31)
∂t µs
where, capillary effects are neglected. Moreover, the relative permeability of oil and solvent read

So So
k r,o = k r,n = k r,n (2.32)
Sn So + Ss

and
Ss Ss
k r,s = k r,n = k r,n , (2.33)
Sn So + Ss
respectively, since they are assumed to be miscible. Note that k r,n is the nonwetting phase relative perme-
ability from, e.g., Brooks-Corey correlations [32] in the absence of solvent. Figure 2.1 shows two physical
representation of incomplete mixing of miscible displacements. The top shows a stable dispersed front and
the bottom shows an unstable displacement with both frontal dispersion and viscous fingering. Depending
on the grid size used with respect to the degree of dispersion that occurs at the front, two extreme cases can
be considered. One where the solvent and oil fluid parameters can be seen as completely mixed, if grid block
sizes are small with respect to the dispersed zone, and therefore can be described by the corresponding mix-
ing rules [18] (i.e., Eqs. (2.12) and (2.13)). If frontal dispersion is negligibly small with respect the the grid
block size, the viscosity and density can be described by that of their pure components [18]. However, it is
reasonable to expect that the reality would lie somewhere in between, for which the properties can be de-
scribed by their effective parameters that lie in between that of the mixed and pure properties. Doing so it is
2.2. E FFECTIVE M ODELS 21

believed that this can effectively solve for the small scale flow features, as shown in Figure 2.1, upscaled on
the coarse grid blocks overlaying it.

Solvent Dispersed zone Oil

Figure 2.1: Frontal dispersion for a stable displacement (upper) and unstable displacement (lower).

To determine the effective parameters (µo,eff and µs,eff ) Todd and Longstaff suggested the use of the fol-
lowing mathematical relation, i.e.,
ω
µo,eff = µ1−ω
o µmix (2.34)
and
µs,eff = µs1−ω µω
mix . (2.35)
The mixture viscosity µmix are calculated from Eq. (2.12). It can be observed that ω = 1 corresponds to com-
plete mixing and ω = 0 corresponds to a situation without effective mixing. Rearranging this viscosity mixing
rule gives
1 ¡ µo ¢ 1
S o M 4 − µmix
4

= 1
. (2.36)
Sn M 4 −1
This equation is valid for any mixture viscosity and can therefore be used to determine the effective fractional
solvent saturations, i.e.,
1 ¡ µ ¢1
¡ So ¢ M4 − µ o 4
s,eff
= , (2.37)
S n s,eff 1
M −1
4

and
1 ¡ µ ¢1
¡ So ¢ M4 − µ o 4
o,eff
= . (2.38)
S n o,eff 1
M 4 −1
Here SSns s,eff denotes the effective volume fraction of solvent in the solvent phase, and SSns o,eff the effective
¡ ¢ ¡ ¢

volume fraction of solvent in oil phase. These ratios are then used to determine the effective densities, i.e.,
22 2. G OVERNING E QUATIONS

¡ So ¢ So ¢
ρ o,eff = ρ o o,eff + ρ s
¡
1− (2.39)
Sn S n o,eff
and
¡ So ¢ So ¢
ρ s,eff = ρ o s,eff + ρ s
¡
1− . (2.40)
Sn S n s,eff
As can be seen from the equation above the effective fractional saturation cannot be determined for dis-
µ
placements with unit viscosity ratio (ie. M = µos = 1). In this situation Todd and Longstaff advised the use
of:

ρ o,eff = 1 − ω)ρ o + ωρ mix


¡
(2.41)

ρ s,eff = 1 − ω)ρ s + ωρ mix


¡
(2.42)
The final unknown in this model is the fitting parameter ω that determines the amount of mixing between
solvent and oil. Todd and Longstaff reported that the use of ω = 23 gave reasonable results for the simulation
of miscible displacements. However, it should be noted that this empirical mixing parameter can be highly
dependent on changes in properties such as grid block size, permeability, geometry and fluid properties and,
therefore, cannot be easily generalized for a model in two or three dimensions.

2.2.3. C ALIBRATION OF M IXING PARAMETER OF T ODD AND L ONGSTAFF


Literature has described calibration of the mixing parameter for miscible flow. These calibrations are found
by equating Koval’s fractional flow formulation [17] , assuming an effective concentration for the mixing zone
(see Eq. (2.25)), with the fractional flow formulation obtained from the Todd and Longstaff model. This
procedure leads to
1 1
f o,TL (c) = = f o,Koval (c) = . (2.43)
S s µo 1−ω Ss 1
1+ S µ 1 + So M
o s eff

Employing an effective solvent concentration of 0.22 (based on experimental data) in Koval’s model [17], the
effective mobility can be denoted as

µs,eff ³ ¡ µo ¢ 1 ´−4
M eff = = 0.22 4 + 0.78 . (2.44)
µo µs

Finally, the mixing parameter ω reads


¡ 1 ¢
4log 0.22M 4 + 0.78
ω = 1− . (2.45)
log(M)

Similarly, for heterogeneous cases, the heterogeneity factor H in the Koval model can be employed to recal-
ibrate the mixing parameter [36]. For heterogeneous media, Koval altered the effective mobility parameter
as ³µ ´
s,eff
M eff = H . (2.46)
µo
The heterogeneity factor is said to be a function of the Dykstra-Parson coefficient (VDP ), i.e.,

VDP
log(H) = . (2.47)
(1 − VDP )0.2

Again by equating the fractional flow functions from Koval’s and TL’s models, the mixing parameter can be
obtained as:
log(H)
ω∗ = ω − µ . (2.48)
log( µos )
3
N UMERICAL S CHEME AND S IMULATION
S TRATEGY
An overview of the Finite-Volume (FV) numerical approach followed in this project is presented in this chap-
ter. A MATLAB code was developed from scratch as a prototype and, for the published results, an in-house
simulator of the sponsoring company was also used.

3.1. M ULTIPHASE M ODEL


As and example, mass conservation for water phase, i.e.,

∂S w
φ
¡
+ ∇ · u w ) = qw , (3.1)
∂t
is considered. In order to find a FV-based discrete form of this equation, it is integrated over the finite volumes
Ω, which after applying using Gauss’ law reads

∂S w
Z Z Z
¡
dV + uw ) · ndS = qw dV, (3.2)
Ω ∂t ∂Ω Ω

where n denotes a unite vector pointing out of surface ∂Ω. For the following explanations a 2D Cartesian grid
with N x × N y grid cells, as shown in Fig. 3.1 for a case of N x = N y = 3, is considered.

i-1, j+1 i, j+1 i+1, j+1

i-1, j i, j i+1, j

i-1, j-1 i, j-1 i+1, j-1

y
x

Figure 3.1: Structured equidistant two-dimensional Cartesian grid, with cell-centered pressure unknowns. Velocity unknowns are ob-
tained at the interfaces.

Using Darcy’s law to relate velocities and pressure gradients, the pressure unknowns are then obtained at

23
24 3. N UMERICAL S CHEME AND S IMULATION S TRATEGY

the cell centers, while the velocity values are obtained at the interfaces. After some mathematical manipula-
tions, using Euler backward scheme for time, one obtains

1 ¡ n+1 n
− T n+1 1 (p i +1, j − p i , j )n+1 + Tw,i−1,j
n+1
(p i , j − p i −1, j )n+1
¢
S w,i,j − S w,i,j
∆t w,i+ 2 ,j

− T n+1 1 (p i , j +1 − p i , j )n+1 + Tw,i,j−1


n+1
(p i , j − p i , j −1 )n+1 = q in+1
,j , (3.3)
w,i,j+ 2

where Twn+1 denotes the implicit transmissibility, e.g., for cells (i ± 12 , j ) it reads

CTRi± 1 ,j
T n+1 1 = MOBn+1 1 2
, (3.4)
w,i± 2 ,j w,i± 2 ,j Vi,j

where,
¡ kr,w ¢n+1
MOBn+1
w = , (3.5)
µw
and
∆A int k H
CT R = . (3.6)
∆x
While the rock property k H is obtained using harmonic averaging, the phase properties, i.e., mobilities, are
calculated using an upwind scheme. This would ensure the consistency and stability of the discrete equa-
tions. Note that to determine the upwind direction of a phase, its phase velocities is used, and that the upwind
direction of phase α can be different than that of the phase β.

3.2. C ONVECTION -D ISPERSION M ODEL


Conservation of concentration c considering convection-dispersion effects reads

∂c ¡ ¢ ¡ ¢
+ ∇ · cu − ∇ · D∇c = q̃. (3.7)
∂t
Following FV scheme, after integration and transformation of volume integrals with surface flux integrals,
one obtains
∂c
Z Z Z Z
¡ ¡
dV + cu) · ndS − D∇c) · ndS = q̃dV. (3.8)
Ω ∂t ∂Ω ∂Ω Ω
Assuming linear flow in the x-direction the dispersion tensor (2.15) results in
· ¸ · ¸
D xx 0 dl u x 0
D= = . (3.9)
0 Dyy 0 dt u x
It is important to note that here the concentration transport equation is solved once the pressure-saturation
system is fully converged, and that the velocities are calculated from pressure solutions. As such, in Eq. (3.8),
u is known at the interfaces, and therefore cell-centered concentrations can be obtained by solving this equa-
tion. To maintain the stability, an upwind strategy is followed to calculate concentration at the interface for
the convective term. The rest of the discrete formulation derivations becomes a classical approach, and thus
will be skipped here.

3.2.1. F LUX A PPROXIMATION ON L OCALLY R EFINED G RID


When discrediting the reservoir transport equations on irregular grids, as a result of local grid refinement (see
chapter 4), one should consider new types of grid block interfaces. Figure 3.2(a) illustrates an example of a
grid block (i ) neighbouring refined grid blocks ( j and k). Assuming that pressure is constant within a grid
block, grid block i can be subdivided into auxiliary smaller conforming grid blocks with pressures equal to i
(See Figure 3.2(b)). As such, we can approximate the the total flux from grid block i to blocks j and k as

k ∂p ´ p j − pi ´ pk − pi ´
Z Z ³ ³ ³
u i , j ,k n · d S = − n · d S = −A i , j MOB i k iH, j 3 − A i ,k MOB i k iH,k 3 , (3.10)
∂Ωi , j ,k ∂Ωi , j ,k µmi x ∂x 4 ∆x i 4 ∆x i

where A i , j and A i ,k are the refined interfacial areas equal to half of the interfacial area of grid block i (A i , j =
A i ,k = 12 A i ). Also note that the distance between the center of grid blocks i and j is 34 ∆x i .
3.3. S IMULATION S TRATEGY 25

j j

k k

(a) (b)

Figure 3.2: (a) Refined interfaces between grid block i and that of j and k. (b) Auxiliary grid blocks, subdividing grid block i in two parts
assuming constant pressure.

3.3. S IMULATION S TRATEGY


In order to solve for flow and transport, the in-house simulator provides the flexibility to do this using differ-
ent strategies for both multi-phase flow (section 3.1) and miscible displacements (section 3.2).
Multi-phase flow was solved using a fully-implicit time integration scheme. In order to simulate for miscible
displacements, a passive tracer model was used. The tracer model (2.9) was solved using a different solver
than that is used for the normal phases. Succeeding the phase solver time step, tracer transport was solved
along the phase flow lines. The tracer transport equation is solved using a sequential technique, called oper-
ator splitting. The linear advection equations are solved using an explicit discretization scheme. The diffu-
sion/dispersion term was solved using a fully implicit strategy, due to stringent stability considerations. The
convection-dispersion equation used in this study included the second order physical dispersion term. Con-
ventional reservoir simulators use a second order accurate spatial discretization scheme for pressure equa-
tion. However, for phase transport a single-point upwind scheme, which is first order accurate in space, is
employed. As such, the truncation error introduces a second order diffusion error to the solution, capable
of concealing the physical dispersion added to the system. Therefore, for the studies of this work, the tracer
model was set to use a second-order (two-point upstream) explicit spatial discretization scheme. This higher
order scheme used a flux limiter in order to avoid unstable solutions in combination with high-resolution
time stepping in order to reduce the first order temporal discretization error. Figure 3.3 shows a flowchart
overview representing the solution strategy used in this study.

Figure 3.3: Flowchart overview of the simulation strategy used in this work to solve for tracer transport.
4
DYNAMIC L OCAL G RID R EFINEMENT
Today’s state-of-the-art reservoir simulators employ grid resolutions much lower than the complex physical
processes involved in the displacement [15, 16], causing severe numerical dispersion, which conceals the
physical phenomena occurring at a small scale. Increasing the grid resolution on the entire domain is not
possible, due to the computational limitations.
Dynamic local grid refinement (DLGR) is a method that can adapt the spatial grid resolution to the scale at
which the physical phenomena occur without refining the grid everywhere. This can be achieved by means
of local refinement and coarsening of the computational grid dynamically at every time step where small
scale, local phenomena (e.g. steep pressure and/or concentration gradients) occur, using predefined refine-
ment criteria. In this chapter, an overview of the DLGR method, as an advanced method for next-generation
simulators, is presented.

4.1. L OCAL G RID R EFINEMENT


Dynamic local grid refinement (DLGR) introduces a dynamic strategy on a given local grid refinement (LGR)
method. As such, the LGR is first introduced in this sub-section. There exist two approaches in the LGR
community in terms of grid coarsening and refinement, namely, nested and non-nested approaches. In this
project, we focus only on the nested approaches. An example is provided in Fig.4.1, where a 2D domain
with 3 × 3 grid cells is considered. The refinement number (r x , r y , r z ) is assigned at each time-step for a grid
cell, e.g., (i + 1, j ). The refinement number for the shown example in Fig.4.1 is (2,2,1) for cell (i + 1, j ), and
remains (1,1,1) for the rest of the cells. Although the LGR method is general in the sense that a cell can receive
any refinement number, due to numerical accuracy, a cell is not allowed to be refined more than one level
compared to the neighboring cells [21, 37]. Throughout the report, the smaller, i.e., refined grid cells, will be
called child cells, while the grid blocks from which they are created will be referred to as the parent cells.

Figure 4.1: Local grid refinement of grid block (i + 1, j ) with (r x , r y , r z )=(2,2,1).

27
28 4. DYNAMIC L OCAL G RID R EFINEMENT

4.2. D ATA S TRUCTURE


Conventional structured equidistance grids have the advantage that each grid block can be addressed using
three indices (X , Y , Z ) from which one can easily determine the neighbouring grid blocks. However, when
LGR is applied, this clear structure becomes more complex, therefore, a more general data structure (i.e., close
to unstructured) is needed in order to keep track of each parent and child grid blocks and their interfaces.
As mentioned, we focus on nested grid LGR approach. As an example, consider a hierarchical locally
refined grid shown in Fig. 4.2. Here the two-dimensional base grid, consisting of 3 × 2 cells, is refined at three
different locations. Firstly, the upper left grid block is refined in both dimensions, then the top left child cell
is again refined in 2x2 grid cells. One base grid block to the right is refined in the vertical direction, and the
bottom left base grid block is refined in the horizontal direction. All refinement are represented as dashed
lines.

Grid

Refinement 1:
Refinement table Refinement table
Refinement 1.1 No refinements

Refinement 2:
Refinement table
No refinements

Refinement 3: Refinement table


No refinements

Figure 4.2: Nested-grid DLGR data structure used in this project.

The hierarchical data structure uses refinement tables in order to to keep track of the nested refinements,
as shown in Fig. 4.2. Here, refinement 1 is shown to have a 2 × 2 sub-grid. Looking at the appropriate refine-
ment table, it can be seen that this sub-grid’s top left grid block is refined by 2×2. Similarly, both refinements
of the neighbouring grid blocks can be addressed, showing only a refinement in the horizontal and vertical
direction. This conceptional representation of the hierarchical data structure also applies to all grid block
properties (i.g. grid block volumes, saturation, mobilities etc.).

4.3. P ROPERTY CALCULATION


On a hierarchical LGR grid, the properties of the newly refined grid blocks need to be calculated from the
parent cell and stored in the data structure. The properties can be categorized into 4 groups:

1. Basic geometric properties, such as nodal heights, are calculated from the grid block coordinates.

2. Intensive properties, i.e. properties that are independent of the grid block size, such as pressure, satu-
ration and mobility, are copied to child cells or averaged to the parent cell if coarsening is applied.
4.4. C ONSISTENCY OF THE D ISCRETIZATION S CHEME ON N ON -U NIFORM G RID 29

3. Extensive properties, properties that depend on the grid block size, e.g. grid block volume, are calcu-
lated such that upn refinement, the sum of the child blocks’ linear properties are equal the original
parent block value.

4. Interface properties, such as interface area, should add up on a sub-grid to the parent grid block value
at the corresponding interface.

4.4. C ONSISTENCY OF THE D ISCRETIZATION S CHEME ON N ON -U NIFORM G RID


For uniform grids, the single-point upwind scheme is consistent and first-order accurate in spatial coordi-
nate. For irregular LGR grids, however, the upwind scheme is not consistent with the underlying conserva-
tion law [38]. Moreover, the assumption that pressure is constant within grid blocks (through refinement,
or averaging for coarsening) causes an additional zero-order truncation error. Figure 4.3(a) illustrates a two-
dimensional grid consisting of n x × n y grid blocks. In this example cell (x i +1 , y j ) is locally refined into four
child grid blocks. As can be observed, the path between the nodes of cell (x i , y j ) and the child cell (x i + 3 , y j + 1 )
4 4
are not parallel to the outward normal of the shared interface. The flux calculation procedure at the interface
of refined and non-refined grid cells, assuming constant pressure through the entire cells, can lead to non-
consistency [20], which then negatively impacts the convergence rate. To avoid this, from the literature, a
consistent flux-based approximation, where the pressure is linearly interpolated between cells, needs to be
followed (see Fig. 4.3(b)). The inconsistent truncation error as a result of the lack of interpolation causes a
local rotation of the flux field at the interface of the coarse-refined grid blocks as will be explained below.

i-1, j+1 i, j+1 i+1, j+1 i-1, j+1 i, j+1 i+1, j+1

i+3/4, j+1/4 i+5/4, j+1/4 ycf i+3/4, j+1/4 i+5/4, j+1/4

i-1, j i, j i-1, j i, j
i+3/4, j-1/4 i+5/4, j-1/4 i+3/4, j-1/4 i+5/4, j-1/4

i-1, j-1 i, j-1 i+1, j-1 i-1, j-1 i, j-1 i+1, j-1

y xcf y
x x

(a) (b)

Figure 4.3: (a) Base grid (3x3) showing a 2x2 refinement in grid block (x i +1 , y j ). (b) Interpolation of pressure in order to obtain pressure
values perpendicular to the coarse-refined grid block interface.

To describe the inconsistency regarding the lack of interpolation of pressure within coarse grid block we first
define the truncation error for the volume flux approximation. From the Taylor expansion of pressure in grid
block node (x i + ∆x c f , y j + ∆y c f ) around location (x i , y j + ∆y c f ), as depicted in Figure 4.3, we see that:

2
∂p ∆x c f ∂2 p
p(x i + ∆x c f , y i + ∆y c f ) = p(x i , y j + ∆y c f ) + ∆x c f (x i , y j + ∆y c f ) + (x i , y j + ∆y c f ) + O(∆y c3f ) (4.1)
∂x 2 ∂x 2
30 4. DYNAMIC L OCAL G RID R EFINEMENT

Expanding each term around point (x i , y j ) gives:

2
∂p ∆y c f ∂2 p
p(x i , y j + ∆y c f ) = p(x i , y j ) + ∆y c f (x i , y j ) + (x i , y j ) + O(∆y c3f ) (4.2)
∂y 2 ∂y 2
∂p ³ ∂p ∂2 p ´
∆x (x i , y j + ∆y c f ) = ∆x c f (x i , y j ) + ∆y c f (x i , y j ) + O(∆x c2 f ) (4.3)
∂x ∂x ∂y∂x
∆x c2 f ∂2 p ∆x c2 f ³ ∂2 p ´
(x i , y j + ∆y c f ) = (x i , y j ) + O(∆x c f ) (4.4)
2 ∂x 2 2 ∂x 2

Substitution of (4.2) to (4.4) into (4.1) we obtain:


2
∂ ∂ ∆x c f ∂2
p(x i + ∆x c f , y j + ∆y c f ) = p(x i , y j ) + ∆x c f p(x i , y j ) + ∆y c f p(x i , y j ) + p(x i , y j )
∂x ∂y 2 ∂x 2
2
∂2 ∆y c f ∂2
+ ∆x c f ∆y c f p(x i , y j ) + p(x i , y j ) + O(∆x c f , ∆y c f )3 (4.5)
∂x∂y 2 ∂y 2

Resulting that the truncation error approximation of the pressure gradient used to calculate fluxes at the
interface between grid blocks (x i , y j ) and (x i + ∆x c f , j + ∆y c f ) is equal to:

p(x i + ∆x c f , j + ∆y c f ) − p(x i , y j )
=
∆x c f
2
∂ ∆y c f ∂ xc f ∂2 ∂2 ∆y c f ∂2
p(x i , y j ) − p(x i , y j ) + ∆ 2 p(x i , y j ) − ∆y c f p(x i , y j ) + p(x i , y j ) + O(∆x c f , ∆y c f )2
∂x ∆x c f ∂y 2 ∂x ∂x∂y 2x c f ∂y 2
(4.6)

As the fluxes are calculated using the pressure gradient as stated above, it can readily be seen that the flux
approximation (used in the industry-grade simulator of this project) is order zero if ∆y c f 6= 0. In Eq. (4.6),
∆y c f ∂
th truncation error term ( ∆x p(x i , y j )) is dependent only on the refinement ratio. This means that the
c f ∂y
local error is not reduced with the reduction of grid block size while using the same refinement ratio, and
therefore is not consistent. As mentioned before, the reason behind this inconsistency is the lack of proper
interpolation of pressure.

4.4.1. R OTATION OF FLUX FIELD


The low-order pressure calculation cases artificial rotation in the flux field. By definition, rotation is the curl
of the pressure gradient, i.e.,
¡ ∂ ∂p ∂ ∂p ¢
∇ × ∇p = − · ez , (4.7)
∂x ∂y ∂y ∂x
which should be zero, for any potential gradient vector. Here e z is the unit basis vector for the z-direction
perpendicular to the two-dimensional x-y plane. Assuming an equidistant grid, as shown in Figure 4.1, the
finite-difference discretization of the derivatives in Eq. (4.7) gives

∂ ∂p p i +1, j +1 − p i +1, j p i , j +1 − p i , j
µ ¶Á ³
1 ´
= 1 −1 ∆x i , j + 1 + ∆x i +1, j + 1 (4.8)
∂x ∂y 2 (∆y i +1, j +1 + ∆y i , j +1 ) 2 (∆y i +1, j + ∆y i , j )
2 2 2

∂ ∂p p i +1, j +1 − p i , j +1 p i +1, j − p i , j
µ ¶Á ³
1 ´
= 1 −1 ∆y i + 1 , j + ∆y i + 1 , j +1 (4.9)
∂y ∂x 2 (∆x i +1, j +1 + ∆x i , j +1 ) 2 (∆x i +1, j + ∆x i , j )
2 2 2

For the equidistant equidistance grid, ∆x i , j = ∆y i , j holds. Omitting indices of ∆x i j = ∆x, ∆y i j = ∆y, we
observe that for a structured equidistant grid the finite difference approximation of the curl of the velocity
field is zero, i.e.,

∂ ∂p 1 ³ ´ ∂ ∂p
= p i +1, j +1 − p i +1, j − p i , j +1 + p i , j = . (4.10)
∂x ∂y ∆x∆y ∂y ∂x
4.4. C ONSISTENCY OF THE D ISCRETIZATION S CHEME ON N ON -U NIFORM G RID 31

Now consider a non-equidistant grid, as depicted in Figure 4.4, that could be generated in LGR method.

xi+1/2,j+1
i-1, j+1 i, j+1 i+1, j+1
yi+1,j+1/2
yi,j+1/2
xi+1/2,j
i-1, j i, j i+1, j

i-1, j-1 i, j-1 i+1, j-1

y
x

Figure 4.4: Two dimensional irregular grid with LGR applied in the x-direction (refinement number of 2) in parent cell (i+1,j). ∆x i + 1 , j
2
and ∆x i + 1 , j +1 represents the average length ∆x between cell nodes (i , j ) and (i + 1, j ), (i , j + 1) and (i + 1, j + 1) respectively. ∆y i , j + 1
2 2
and ∆y i +1, j + 1 denotes the average grid block length ∆y between cell nodes (i , j ) and (i , j + 1), (i + 1, j ) and (i + 1, j + 1) respectively.
2

In this example the first child cell on the left is denoted as cell (i +1, j ) (instead of (i + 43 , j )) to keep a consistent
formulation as stated for the derivation of a regular grid. Subtracting the formulations of the derivatives ((4.8)
and (4.9)) and sorting it into parts with equivalent pressure terms gives:

∂ ∂ ∂ ∂
µ ¶
1 1
p− p= − p i +1, j +1
∂x ∂y ∂y ∂x ∆y i +1, j + 1 ∆x i + 1 , j ∆x i + 1 , j +1 ∆y i , j + 1
2 2 2 2
µ ¶
1 1
− − p i +1, j
∆y i +1, j + 1 ∆x i + 1 , j ∆x i + 1 , j ∆y i , j + 1
2 2 2 2
µ ¶
1 1
− − p i , j +1
∆y i , j + 1 ∆x i + 1 , j ∆x i + 1 , j +1 ∆y i , j + 1
2 2 2 2
µ ¶
1 1
+ − pi , j
∆y i , j + 1 ∆x i + 1 , j ∆x i + 1 , j ∆y i , j + 1
2 2 2 2

6= 0 (4.11)

Here ∆x i + 1 , j and ∆x i + 1 , j +1 represent the average length between cells (i , j ) and (i +1, j ), (i , j +1) and (i +1, j +
2 2
1), respectively. Likewise ∆y i , j + 1 and ∆y i +1, j + 1 denote the average length between cells (i , j ) and (i , j + 1),
2 2
(i + 1, j ) and (i + 1, j + 1), respectively. From Figure 4.4 we can conclude that the pressure gradient field is not
rotation free at the interface between grid cells of non-equal size, because

∆x i + 1 , j 6= ∆x i + 1 , j +1
2 2

∆y i , j + 1 6= ∆y i +1, j + 1 .
2 2

Figure 4.7 shows the effect of the zeroth order truncation error causing rotation of the flux field. A 3x3 base
grid is shown in which only the center grid block is not refined. As the flow direction is from left to right, the
error is made at the interfaces between the different grid sizes parallel to the pressure gradient. The net flow
direction (not the magnitude) is indicated by the arrows.
32 4. DYNAMIC L OCAL G RID R EFINEMENT

Figure 4.5: Rotation in flux field, a consequence of the constant interpolation for pressure.
4.5. DLGR S IMULATION S TRATEGY 33

4.5. DLGR S IMULATION S TRATEGY


Commercial reservoir simulators are typically developed on the basis of fully implicit approaches. This is
motivated by the fact that fully implicit methods are found to be the most stable approach, accounting for the
coupling between the unknowns. As it is generally cumbersome to evaluate where the front will be located
at the end of a time step when using implicit discretization, this reservoir simulator makes use of a semi-
implicit evaluation of grid refinement. Figure 4.6(a) shows the conventional, explicit way of grid refinement
compared to the newly developed semi-implicit evaluation.

(a) (b)

Figure 4.6: (a) Explicit gridding strategy for DLGR. (b) Semi-implicit gridding strategy of DLGR.

In order to simulate hyperbolic (for transport) and parabolic (for flow) equations implicitly, the equations
are linearized using the Newton-Raphson (NR)-scheme. When the system is converged, the grid is evaluated
and refined and/or coarsened dependent on the preset refinement criteria. For the explicit method, shown
in Figure 4.6(a), at a new time step the fluid front could have traveled to the unrefined cells, depending on
the time step size. As such, the explicit refinement approach can smear out the front. In order to avoid this,
the semi-implicit method (shown in Figure 4.6(b)) introduces a second loop that evaluates the location of
refinements before the end of each simulation time step. If the fluid front has traveled ahead of the refined
area, the grid is refined at these locations and the solution is recalculated on this newly refined grid. This
loop is repeated until convergence is reached. Then, coarsening is applied only after the refinement loop
is converged successfully. Previous studies [22] showed that the overhead inherent to the inclusion of the
additional loop is minimal. This is due to the fact that the initial guess used to solve the system of linear
equations in the repeat time step(s) uses the solution of the previous converged solution. Therefore, this
reduces the amount of iterations needed for the additional repeat time steps.

4.6. R EFINEMENT C RITERIA


To the best of the author’s knowledge, the application of DLGR in reservoir simulation was introduced in
34 4. DYNAMIC L OCAL G RID R EFINEMENT

1979 [39]. Using the method of characteristics to calculate the shock front saturation velocity, refinement
at the front was successfully applied to a one-dimensional Buckeley-Leverett model resulting in accurate
simulations. However, one can imagine that using DLGR for transport of fluids within multi-dimensional,
complex reservoir models introduces the need of different refinement criteria. Literature describes the use of
many different refinement criteria, most of which can be directly approximated by lower-order discretization
methods, which are computationally inexpensive. Most of the spatial inaccuracy evolves around fluid fronts,
which can be easily located using simple criteria. Saturation, concentration and temperature threshold val-
ues [23, 40], saturation differences across neighboring grid cells [41] and saturation gradients [21, 24, 40, 42]
have been found effective to accurately predict fluid interfaces. Also mass or volume flux, in combination
with different criteria, have been shown to be able to indicate where front locations reside, resulting in accu-
rate simulation of frontal displacements [16, 24, 25].
The DLGR method, implemented in the industry-grade simulator used in this project, allows for the flex-
ibility to use built-in error criteria, scripted criteria or a combination of the two. During DLGR simulation at
every (repeat) time step, the error criteria are calculated locally and evaluated to estimate the locations that
need increased spatial resolution. Depending on the prescribed tolerances, refinement and/or coarsening of
the grid is applied. The tolerances for refinement and coarsening can be set independently. The only require-
ment is that the tolerances are not contradictory, i.e. δu ≤ δr , where u stands for unrefinement (coarsening)
and r for refinement. The calculation and use of normalized criteria is available in the simulator, which al-
lows more intuitive way of selecting tolerances for particular criteria. It is clear that a selected error threshold
may not be achieved, as such, a maximum refinement level is set so that grid is not refined any further than
that.
For this research different criteria on the basis of cell-centered and interface differences are used. All param-
eters of concentration, mobility and viscosity have been considered. In addition, the time derivative of the
spatial gradient of concentration and the third order spatial derivative of concentration were used. Below the
criteria summarized:

1. ∇C 5. ∇µmix

2. ∆C 6. ∆µmix

3. ∂t ∇C 7. ∇mobmix

4. ∇3C 8. ∆mobmix

The description of the criteria are similar for all properties, therefore, here it is illustrated for the concentra-
tion property only.

C ONCENTRATION D IFFERENCE
The difference of concentration (or saturation) for miscible (or multi-phase) flow is an intuitive error criterion
because of its normalized and dimensionless property. During DLGR simulation on every (repeat) time step
the difference of concentration was calculated and evaluated if it locally violates the preset conditions:

∆C > δu (4.12)
∆C < δr (4.13)

C ONCENTRATION G RADIENT
The gradient criterion is defined as:

∇C > δu (4.14)
∇C < δr (4.15)

During this research the calculated local gradients of concentration were normalized by dividing the obtained
gradient field by its maximum value on every time step. This way a more intuitive range scaled between 0 to
1 is generated, avoiding the use of non-trivial gradients tolerances. The normalized concentration gradient is
defined as:
4.7. U PSCALING P ERMEABILITY M AP 35

∇C
> δu (4.16)
||∇C ||∞
∇C
< δr (4.17)
||∇C ||∞

T HIRD O RDER C ONCENTRATION D ERIVATIVE


A the transport solver used to model incomplete mixing for miscible displacements is second order accu-
rate, as described in section 3.3, the local truncation error includes a third order spatial derivative term. The
approximation of this third order spatial derivative was used as criterion defined as:

∇3C > δu (4.18)


∇3C < δr (4.19)

As with first order spatial gradients the third order derivatives of concentration may not be known before-
hand. Therefore, during this research the normalized third order spatial derivative was used, defined as:

∇3C
> δu (4.20)
||∇3C ||∞
∇3C
< δr (4.21)
||∇3C ||∞

T IME D ERIVATIVE OF C ONCENTRATION G RADIENT


The time derivative of the local gradient is reported to have good front detection properties for stable dis-
placements [22]. During the performed simulations the local concentration gradients of the previous time
step can be saved in order to calculate the local time derivative of the spatial gradient. The calculation is
performed by subtracting the the locally obtained gradients in the two different time steps, and dividing it by
the time step size. This criteria concentrates strongly around the front as the change of gradients in time are
greatest at the start and end of the front. The criteria is given by:


∇C > δu (4.22)
∂t

∇C < δr (4.23)
∂t
During this research the normalized criteria were used:


∂t ∇C

> δu (4.24)
|| ∂t ∇C ||∞

∂t ∇C

< δr (4.25)
|| ∂t ∇C ||∞

4.7. U PSCALING P ERMEABILITY M AP


As discussed in subsection 1.1.3 the permeability field of reservoir can have a determining effect on fluid
flow in porous media. As DLGR uses multi-level hierarchical refinements in order to capture complex flow
features, it is clear that the underlying permeability field of heterogeneous reservoirs equally needs to be re-
vealed. As illustrated in Figure 4.8, starting from an imported fine scale permeability map (level 2) it shows
the possible levels of grid resolution that DLGR can attain using 2 levels of refinement. If DLGR’s imposed
grid includes grid blocks corresponding to a level other than the imported fine scale permeability field (level
1 or 0), local flow-based upscaling was performed in order to calculate the local permeability.
36 4. DYNAMIC L OCAL G RID R EFINEMENT

Figure 4.7: Multi-level grid permeability obtained by flow-based upscaling.

The principle of flow based upscaling originates from the assumption of Darcy’s law that flow relates linearly
to the applied pressure gradient.

Figure 4.8: Flow-based upscaling; calculating the parent grid block’s upscaled permeability in both x and y-direction.

For single-phase flow, Darcy’s law states u = − µk ∇P . Once the grid coarsening is initiated, the particular
group of children grid blocks that contain the parent grid blocks are isolated (see Figure 4.7). Across this local
domain a pressure drop is applied, while considering no-flow boundary conditions at the interface parallel to
the applied pressure drop. Doing so the mean flow velocity across the the coarse interface can be calculated,
and therefore, the upscaled permeability can be obtained using Darcy’s law:

û x · µ · ∆x
k eff,x = − (4.26)
∆P
û y · µ · ∆y
k eff,y = − (4.27)
∆P
5
I NCOMPLETE M IXING FOR H OMOGENEOUS
P OROUS M EDIA
In this chapter the results regarding the simulation of incomplete mixing in homogeneous media are shown
an discussed. section 5.1 investigates the efficiency of DLGR compared to corresponding high resolution
simulations. The error investigation of the solutions obtained with DLGR are discussed in section 5.2. This
section describes the high resolution reference cases used to compare DLGR, the inherent uncertainty of
unstable viscous flow as well as the error analysis of DLGR.
The consistency of DLGR as well as the TL model regarding scalability and mobility ratio between the
displaced and displacing fluids are shown in section 5.3. Finally section 5.4 compares results obtained with
DLGR and the TL model for polymer flooding in homogeneous media.

5.1. P ERFORMANCE A SSESSMENT OF DLGR FOR M ISCIBLE D ISPLACEMENTS


A thorough assessment of performance of any method, including DLGR, needs to be done on the basis of
CPU time measurement. This is due to the fact that several complexities are involved in a method, making
analytical estimate of its performance, once applied to a complex nonlinear multidimensional problem, quite
challenging. As such, in this section a detailed investigation of DLGR once applied to miscible displacements
is presented. It is important to note that the performance is split into setup and solution phases, to allow for
more detailed investigation of its advantages and opportunities for further developments. As DLGR dynami-
cally adjusts the local grid resolution, by refining and coarsening the grid, its overall CPU time is expected to
be much smaller than that of the fine-scale fully resolved case. However, one should note that dynamic grid
comes with additional overhead that requires an efficient implementation, which otherwise can become the
bottleneck.
Table 5.1 shows three reservoir cases which were chosen to evaluate the scalability in terms of CPU time
with respect to the grid size for both high resolution simulations and DLGR. The base case reservoir has di-
mensions of 120 × 10 × 30 [m]. The static fine scale simulation employed a grid consisting of 384 × 1 × 96 cells.
On the other hand, DLGR used a base grid of 12 × 1 × 3, making use of the norm of the solvent concentration
gradient as a criteria for refinement and coarsening (tolerance was set to 0.1). Each grid block was locally
refined in 2 × 1 × 2, with the maximum refinement level of 5, corresponding to the fine scale resolution. The
domain was scaled in the z-direction, perpendicular to the mean flow direction, doubling and quadrupling
the size. This would seem a fair comparison as DLGR refines along the fluid front perpendicular to the flow
direction (z-direction). This scalability test was performed for both a stable (i.e. unit mobility ratio) and un-
stable (i.e. mobility ratio of 10) displacement of solvent injection in a fully saturated oil reservoir assuming
incompressible fluids. The injection well perforates all grid blocks in the vertical direction on the left bound-
ary of the domain, while the production occurs through the entire right boundary cells. Statically refined
grids were imposed for the regions close to the wells, as to minimize the effect of capturing the well infectivity
in different simulations. Simulations were run until first break through was observed.

S TABLE D ISPLACEMENT
Shown in the top part of Figure 5.1 is physical problem for the base case reservoir model at 0.3 pore volume
injected (PVI). The lower part illustrates the imposed grid for DLGR, clearly illustrating the static LGR at the

37
38 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

injector and producer wells as well as DLGR along the fluid front. Figure 5.2 illustrates the scalability of
the average CPU-time per simulation time step for both fine scale and DLGR simulations. As Nz denotes
the number of grid blocks in the z-direction (i.e. perpendicular to flow), 2 × Nz and 4 × Nz correspond to
the twofold and fourfold scaled system, respectively. The static high resolution base case model consisted
of approximately 37.000 active grid blocks. The scaled systems therefore, consisted of 74.000 and 147.000
active grid blocks respectively. The initialized grid size for DLGR was 36, 72 and 144 for the three different
cases without static refinment along the wells. The total amount of active grid cells were adjusted in order to
capture the displacing front. On average 8% active grid blocks were used compared to the high resolutions
reference cases.

Figure 5.1: Base case reservoir for stable displacement, with domain size 30 by 120 m. The concentration profile at 0.3 PVI (top) and the
imposed DLGR grid at this corresponding time (bottom) are shown.

As can be observed from Figure 5.2, the increased complexity of the data structure for DLGR is reflected by an
increased amount of operations indicated by the different colored segments. Grid initialization, depicted in
bright yellow, consists out of the calculation time spent on an interfacial property used for the discretization
of physical dispersion. For static grid simulations this procedure is only performed once as a preprocessing
step, resulting in an invisible fraction of the CPU time. However, for DLGR these calculations have to be per-
formed every time that the grid structure changes.
Shown in red and blue are the phase and tracer solver average CPU times per time step, respectively. Solver
times are significantly reduced using the DLGR implementation. DLGR makes use of a multi-grid precon-
ditioner in order to solve for the hierarchical nested LGR grid. The reduced active grid size used by DLGR
resulted in lower average CPU time compared to the solver time of the high resolution static grid.
The grid adaptation operations are shown in purple for local grid refinement (LGR) and green for local grid
coarsening (LGC), naturally absent in the static grid simulation column bars. The DLGR functionalities con-
sists of error criteria calculation and the refinement or coarsening operation requiring recalculation of the
adjusted (interfacial) grid block properties. It is observed in Figure 5.2 that on average a fraction of 0.68 of the
mean CPU time per time step is spent on the actual adaptive restructuring and recalculation of the locally
refined/coarsened grid.
The overall efficiency gain of using DLGR in these cases, while maintaining similar accuracy, was found to be
approximately 35% compared to the reference high resolution simulation. This shows the potential of DLGR
as well as the overhead associated with the loss of a conventional structured grid. Both fine scale and DLGR
5.1. P ERFORMANCE A SSESSMENT OF DLGR FOR M ISCIBLE D ISPLACEMENTS 39

CPU time were found to be scalable within the limit of the tested grid sizes, allowing for reliable efficiency
behavior upon resizing of the system.

Figure 5.2: Mean CPU time per simulation time step [s]. Scalability of stable displacement for domain sizes: Nz (base case), 2×Nz (scaled
twofold) and 4 × Nz (scaled fourfold).

In this example the LGR and LGC functionalities have shown to give most of the overhead during the simula-
tions in this study, combined, taking almost 70% of the total simulation time.

Figure 5.3 shows the LGR/LGC CPU time for the three scaled systems. Observed is that the LGR/LGC
operations consist of the calculation of absolute gradients, locating the maximum gradient (i.e. calculating
the norm of the gradients), acquiring neighboring interfaces and to adaptively update the grid properties
after refinement or coarsening. Locating the maximum gradient within the system takes almost 45% of the
CPU time required by the LGR/LGC monitors. This can be improved by using dimensional gradients as an
error criterion rather than the norm of the gradient. The fact that on average LGC uses less CPU time can be
explained by the fact that LGC is only enforced at the end of a simulation time step while LGR is applied in
all potential repeat time steps within a simulation time step. The difference observed is small as the average
amount of repeat time steps needed per simulation time step was found to be 0.35. It is observed that the
LGC monitor calculates for the error criterion, similar to that of the LGR monitor. This is required when the
error criterion for coarsening is different from that of the refinement. However, during these simulations only
one criterion was used, such that the value of this criterion would have already been calculated during the
last check for refinement. A potential efficiency gain could therefore be made when the values of the error
criteria are saved after refinement.
40 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

Figure 5.3: Mean CPU time per simulation time step [s] of DLGR functionalities. Scalability of stable displacement for domain sizes: Nz
(base case), 2 × Nz (scaled twofold) and 4 × Nz (scaled fourfold).

U NSTABLE D ISPLACEMENT

As many incomplete mixing phenomena evolve from unfavourable mobility ratio’s between displacing and
displaced fluids, a CPU time analysis for unstable displacement was performed. Discussed below are the CPU
timings of geometrically similar domains as described above but for an unfavorable mobility ratio (M =10).
Permeability perturbations (variation of 5% of the reservoir permeability) at the injector well location were
used to trigger instabilities. Shown below, in Figures 5.4 and 5.5, are the physical problem for the base case
reservoir as well as the scalability in terms of CPU time.

Figure 5.4: Base case reservoir for an unstable displacement (M =10), with domain size 30 by 120 m. The concentration profile at 0.3 PVI
(top) and the imposed DLGR grid at this corresponding time (bottom) are shown.
5.1. P ERFORMANCE A SSESSMENT OF DLGR FOR M ISCIBLE D ISPLACEMENTS 41

Figure 5.5: Mean CPU time per simulation time step [s]. Scalability of unstable displacement for domain sizes: Nz (base case), 2 × Nz
(scaled twofold) and 4 × Nz (scaled fourfold).

Comparison of Figure 5.2 and 5.5 shows that the efficiency of DLGR is highly dependent on the physical
problem at hand. For unstable displacements, such as depicted in Figure 5.4, the viscous fingers will increase
the surface area of the front dramatically. This increases the amount of required refinements during DLGR
simulation, resulting in an increasing overhead. As mentioned for stable displacements the speed up of us-
ing DLGR was 35%, whereas for these unstable displacement realizations the efficiency was calculated to be
more than 3 times lower than that of the static grid simulations.
As mentioned previously for the stable displacement, grid initialization (depicted in yellow) accounts for the
calculation of an interfacial grid property used within the simulator, in order to calculate transmissibilities as
well as for the discretization of physical diffusion or dispersion. If the simulator does not save these values,
which was the case of our simulator, these parameters need to be recalculated over the entire grid whenever
grid adaptation is applied. Such an unnecessary frequent recalculations take more than 30% of the simula-
tion CPU time. This could be mitigated by disabling physical dispersion, or by saving this grid property.
The fact that on average 24% active grid blocks were used with DLGR compared to the fine scale simulations
resulted that the multi-grid preconditioned solver was capable of solving for both phase and tracer transport
more efficiently.
Finally, the remaining segment depicted in orange represent the viscosity multiplier used in order to modify
the viscosity in the tracer model. This procedure is more efficient for DLGR due to the reduced active grid
blocks used during simulation.
As shown in Figure 5.5, both the CPU time of the high resolution reference simulations and DLGR were found
to be scalable within the limit of the tested grid sizes, allowing for reliable efficiency behavior upon resizing
of the system for unstable flow.

The CPU break-down of the LGR and LGC functionality is shown in Figure 5.6. Due to the increased frontal
surface area of the physical problem, the amount of active grid blocks increases as well. As a result the amount
of time used to update the grid numbering and neighbouring connections as well as recalculation of grid
properties takes much of the LGR and LGC CPU time, illustrated by blue and purple. It can be seen that the
refinements take up more time than the coarsening procedures. As explained before this is due to the fact
that only refinement is applied during repeat time steps. On average 1.8 repeat time steps per simulation
time step were needed, clearly illustrating the difference in CPU time observed.
42 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

Figure 5.6: Mean CPU time per simulation time step [s] of LGR functionalities. Scalability of unstable displacement for domain sizes: Nz
(base case), 2 × Nz (scaled twofold) and 4 × Nz (scaled fourfold).
5.1. P ERFORMANCE A SSESSMENT OF DLGR FOR M ISCIBLE D ISPLACEMENTS 43

Geometric and fluid/reservoir properties


Base case Fine Scale DLGR
x length [m] 120 120
y length [m] 10 10
z length [m] 30 30
x base grid size 384 12
y base grid size 1 1
z base grid size 96 3
2 x base case Fine Scale DLGR
x length [m] 120 120
y length [m] 10 10
z length [m] 60 60
x base grid size 384 12
y base grid size 1 1
z base grid size 192 6
4 x base case Fine Scale DLGR
x length [m] 120 120
y length [m] 10 10
z length [m] 120 120
x base grid size 384 12
y base grid size 1 1
z base grid size 384 12
Properties Fine Scale DLGR
ft
Injection rate [ day ] 1 1
∆t [s] 1 CFL 1 CFL
(P w,inj )max [bar] 450 450
(P w,prod )min [bar] 100 100
µo [Cp] 10 10
µs [Cp] (stable) 10 10
µs [Cp] (unstable) 1 1
S w,c 0 0
S o,ini 1 1
d l [m] 0.2 0.2
d tr [m] 0.02 0.02
Permeability [mD] 400 400
Porosity 0.3 0.3

Table 5.1: Geometric and fluid/reservoir properties.


44 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

5.2. DLGR A CCURACY FOR M ISCIBLE -S OLVENT D ISPLACEMENTS


Apart from the efficiency, a thorough assessment needs to also include the quality of the results obtained from
DLGR. As described in section section 4.6, different refinement criteria have been used in order to evaluate
their effectiveness in reducing the local and global errors. It can be readily understood that the set tolerances
for the criteria have a determining effect on the sensitivity, and therefore, accuracy of DLGR results.
In this section, the quality of DLGR solutions for incomplete mixing arising from unstable displacements in
homogeneous porous media is studied. This evaluation is performed by comparing the results with corre-
sponding fine scale solutions in terms of cumulative production, production rates and local concentration
profile. As dispersion has a strong effect on the physical behaviour of viscous fingering (see section 1.1),
three different dispersion cases were selected in order to evaluate the accuracy of DLGR for different mixing
scenarios.

5.2.1. F INE SCALE REFERENCE S OLUTION


In order to compare the quality of the DLGR results, it is necessary to obtain a reference solution. As it is not
yet possible to construct an analytical solutions for unstable flow, these reference solutions were obtained
on a high resolution static Cartesian grid. The model included longitudinal and transversal hydrodynamic
dispersion, but neglected molecular diffusion. Ideally, one would like to have a unique reference solution
in order to compare the results obtained with DLGR, however the required grid resolution can be extremely
high.
Conventional reservoir simulators use first order accurate schemes. Using first order schemes one would
obtain a converging reference solution if the numerical diffusion introduced by the dominant truncation
error terms are reduced to a value far below that of the physical dispersion. As both the numerical diffusion
and physical dispersion terms are second order spatial derivatives, estimates can be derived for the spatial
and temporal resolution for which the numerical diffusion term is below that of the physical dispersion term
(see Appendix A). However, in this study a second order accurate scheme (see section 3.3) was used in order to
limit numerical diffusion concealing the physical dispersion. Therefore, one would only be possible to obtain
a converging reference solution if the numerical error in the solution is sufficiently reduced below that of the
physical dispersion included in the model. At this point the physical dispersion would be the determining
factor altering the solution.
In order to test the limits of DLGR, three different reference cases with varying degrees of longitudinal and
transversal dispersion have been used to evaluate the accuracy of DLGR results. The first case did not include
any physical dispersion, the second case includes both longitudinal and transverse dispersion and for the
third case the amount of transverse dispersion was increased fivefold. This is denoted in Table 5.2, which
includes the dimensionless Peclet numbers describing the ratio between convection and dispersion. All high
resolution reference solutions that were used to investigate the accuracy of DLGR were obtained using a grid
spacing of ∆h = 0.25m, corresponding to a 400×400 grid resolution imposed on a 100×100 [m] domain. This
subsection describes the grid size sensitivity of the reference solutions.

Dispersion cases
Dispersion cases Longitudinal Transverse Length Length Longitudinal Transverse
dispersion dispersion x z Peclet number Peclet number
Case 1: zero 0 [m] 0 [m] 100 [m] 100 [m] - -
dispersion
Case 2: medium 0.3 [m] 0.01 [m] 100 [m] 100 [m] 333 10.000
dispersion
Case 3: high 0.3 [m] 0.05 [m] 100 [m] 100 [m] 333 2000
dispersion

Table 5.2: Three different simulation scenario’s using different physical dispersion, with corresponding dimensionless Peclet numbers.

Figure 5.7 shows the concentration profile at 0.2 pore volume injected (PVI) upon grid refinement for the
three different cases denoted in Table 5.2, starting at the top with (25 × 25 resolution; ∆h = 4 [m]) to the
bottom (400 × 400 resolution; ∆h = 0.25 [m]) on a 100 × 100 [m] domain. From left to right the columns in
Figure 5.7 illustrate the cases 1 to 3, respectively. The left column of Figure 5.7 correspond to the solutions
obtained without modeling of physical dispersion, therefore, numerical truncation errors are the only cause
of smearing out the solution. As discussed in subsection 1.1.2, the lack of physical dampening results that
5.2. DLGR A CCURACY FOR M ISCIBLE -S OLVENT D ISPLACEMENTS 45

smaller and smaller instabilities become increasingly unstable. Consequently, the solution will not converge.
It is observed that the system becomes increasingly unstable as the grid is refined more and more, illus-
trated by the appearance of narrower viscous fingers. As stated above, this can be explained by the fact that
no physical dispersion is present that in reality would act as a cut-off to smaller perturbations. In this case
only numerical dispersion plays a role on the solution, being a function of grid resolution. Therefore, when
the grid is refined, numerical dispersion, a stabilizing force, is reduced. As such, smaller fingers grow.
Figure 5.8 to 5.10 show the corresponding grid sensitivity results for the dimensionless cumulative oil produc-
tion, solvent rate and transverse average solvent concentration. Similar to the local concentration solution,
the averaged solutions do not show a converging trend.

The middle column of Figure 5.7 and Figures 5.11 to 5.13 illustrate solutions with respect to grid refinement
for the second case denoted in Table 5.2. For this case the amount of added physical dispersion seems to be
greater than the numerical error introduced by the O(∆x 2 ) discretization scheme for the higher resolution
(400 × 400; ∆h = 0.25m).
Observed is that the degree of viscous fingering is reduced compared to the case without physical disper-
sion, explained by the added physical dispersion that acts on instabilities in the longitudinal and, in lesser
extend, the transverse direction. As a results, smaller perturbation are smeared out, allowing only for fewer
larger dominant fingers to grow.
Figure 5.12 shows the solvent rate against simulation time. Similar to the concentration profile, the break-
through time seems to show a converging trend, however, the solution in time still seems to be sensitive to
the increased grid resolution. The transverse average concentration solution at 0.2 and 0.6 PVI is illustrated
in Figure 5.13. It can be observed from the resolution of 200 × 200 (∆h = 0.5m) that the magnitude of the im-
posed truncation error becomes lower than the added physical dispersion, so that the two highest resolution
simulations show a converging trend.

The third column of Figure 5.7 and Figure 5.14 to 5.16 illustrate the solutions with respect to grid refinement
for case 3 (shown in Table 5.2). In this case the longitudinal dispersion is kept the same as case 2. How-
ever, the ratio between longitudinal and transversal dispersion is lowered by a factor of 5. The transversal
dispersion further spreads out instabilities in the direction perpendicular to the flow, resulting in reduction
of viscous fingers. Similar to the second case described previously, the solutions seem to show a converging
trend. From the solutions of the third column of Figure 5.7 it can be seen that there almost no change be-
tween 5.7(k) (200 × 200 resolution; ∆h = 0.5m) and 5.7(n) (400 × 400 resolution; ∆h = 0.25m). The same can
be said about the averaged solutions.
46 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

(a) (b) (c)

(d) (e) (f)

(g) (h) (i)

(j) (k) (l)

(m) (n) (o)

(p)

Figure 5.7: Convergence analysis for zero (left column), medium (middle column) and high (right column) dispersion (case 1 to 3 of table
5.2) on a static Cartesian grid on a 100 by 100 m domain; grid refinement of 2×2 occurs from top left to bottom. The resolutions are (a-c)
25 × 25; ∆h = 4 [m] , (d-f) 50 × 50; ∆h = 2 [m], (g-i) 100 × 100; ∆h = 1 [m], (j-l) 200 × 200; ∆h = 0.5 [m] and (m-o) 400 × 400; ∆h = 0.25 [m].
The colorbar (p) quantifies the concentration profiles to the actual concentration.
5.2. DLGR A CCURACY FOR M ISCIBLE -S OLVENT D ISPLACEMENTS 47

Figure 5.8: Cumulative oil production shown for different grid resolutions (25 × 25 to 400 × 400) on a 100 by 100 m domain for zero
dispersion (case 1 of table 5.2).

Figure 5.9: Solvent rate shown for different grid resolutions (25 × 25 to 400 × 400) on a 100 by 100 m domain for zero dispersion (case 1 of
table 5.2).
48 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

Figure 5.10: Transverse averaged concentration (averaged over z-direction) shown for different grid resolutions (25 × 25 to 400 × 400), at
0.2 pvi and 0.6 pvi for zero dispersion (case 1 of table 5.2).

Figure 5.11: Cumulative oil production shown for different grid resolutions (25 × 25 to 400 × 400) on a 100 by 100 m domain for medium
dispersion (case 2 of table 5.2).
5.2. DLGR A CCURACY FOR M ISCIBLE -S OLVENT D ISPLACEMENTS 49

Figure 5.12: Solvent rate shown for different grid resolutions (25×25 to 400×400) on a 100 by 100 m domain for medium dispersion (case
2 of table 5.2).

Figure 5.13: Transverse averaged concentration (averaged over z-direction) shown for different grid resolutions (25 × 25 to 400 × 400), at
0.2 pvi and 0.6 pvi for medium dispersion (case 2 of table 5.2).
50 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

Figure 5.14: Cumulative oil production shown for different grid resolutions (25 × 25 to 400 × 400) on a 100 by 100 m domain for high
dispersion (case 3 of table 5.2).

Figure 5.15: Solvent rate shown for different grid resolutions (25 × 25 to 400 × 400) on a 100 by 100 m domain for high dispersion (case 3
of table 5.2).
5.2. DLGR A CCURACY FOR M ISCIBLE -S OLVENT D ISPLACEMENTS 51

Figure 5.16: Transverse averaged concentration (averaged over z-direction) shown for different grid resolutions (25 × 25 to 400 × 400), at
0.2 pvi and 0.6 pvi for high dispersion (case 3 of table 5.2).

5.2.2. U NCERTAINTY WITHIN U NSTABLE F LOW S IMULATION


Reservoir simulations are mostly deterministic, determined by the initial and boundary conditions. In reality
however no porous media are the same, leading that experimental observations of unstable displacements
will never be identical but rather give a cloud of different solutions inherent to the variations in the different
porous media. Therefore, we expect that the solution of unstable flow would have a statistical nature rather
than a deterministic one. As DLGR refines locally at the location of the front, the numerical error in pressure
(which is calculated globally) or concentration due to the grid adaptation could possibly alter the solutions.
In order to obtain an allowable tolerance for the error caused by DLGR, ten different realizations were sim-
ulated on a fine scale Cartesian grid, each of which has a different but statistically equal mean distribution
of tracer concentration at the location at the injector well at initialization. Each perturbed realization was
triggered with the concentration variation between 0 and 1. Analyzing the different results one can obtain
an average solution. In order to determine the uncertainty of possible physical realizations, the range of 3
standard deviations from the mean was used. From this the average variation was calculated and used as an
indicator of the allowable deviation of DLGR’s results for the different dispersion cases.

Figures 5.17 to 5.22 show the mean cumulative oil production and solvent rate of all realizations for the dif-
ferent dispersion cases. The variations, expressed as 3 standard deviations, are illustrated with red error bars.
For the zero, medium and high dispersion cases the average variation from the mean, expressed as the mean
error in percentage, was calculated to be 1, 1.2 and 1.1 % for the cumulative production of oil and 11, 17.4 and
11.4% for the solvent rate, respectively. It is observed that the variation in the zero dispersion case is the low-
est. This could be explained by the fact that almost every perturbation can grow strongly resulting in many
competing fingers, that are in average less subject to the variation in perturbation. The addition of dispersion
in the medium case results in fewer viscous fingers. Therefore, the trigger used to perturb the front has a more
determining effect resulting in a slightly higher variation. The third case with a higher transversal dispersion
causes the variation to reduces again, as the increased transverse dispersion strongly reduces instabilities.
The variation in cumulative production is observed to be about one order of magnitude lower than for the
solvent rate. This is because the effect of variations in oil production rate on the cumulative oil production
reduces strongly over time due to the accumulation of oil production.
52 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

Figure 5.17: Range of possible realizations of cumulative oil production for zero dispersion case obtained on a high resolution ( 400×400)
Cartesian grid for a 100 × 100 [m] domain, showing the mean solution (blue) and 3 standard deviations as range of variation (red error
bars). Mean variation from the mean 1%.

Figure 5.18: Range of possible realizations of solvent rate for zero dispersion case obtained on a high resolution ( 400 × 400) Cartesian
grid for a 100 × 100 [m] domain, showing the mean solution (blue) and 3 standard deviations as range of variation (red error bars). Mean
variation from the mean 11%.
5.2. DLGR A CCURACY FOR M ISCIBLE -S OLVENT D ISPLACEMENTS 53

Figure 5.19: Range of possible realizations of cumulative oil production for medium dispersion case obtained on a high resolution (
400 × 400) Cartesian grid for a 100 × 100 [m] domain, showing the mean solution (blue) and 3 standard deviations as range of variation
(red error bars). Mean variation from the mean 1.2%.

Figure 5.20: Range of possible realizations of solvent rate for medium dispersion case obtained on a high resolution ( 400×400) Cartesian
grid for a 100 × 100 [m] domain, showing the mean solution (blue) and 3 standard deviations as range of variation (red error bars). Mean
variation from the mean 17.4%.
54 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

Figure 5.21: Range of possible realizations of cumulative oil production for high dispersion case obtained on a high resolution ( 400×400)
Cartesian grid for a 100 × 100 [m] domain, showing the mean solution (blue) and 3 standard deviations as range of variation (red error
bars). Mean variation from the mean 1.1%.

Figure 5.22: Range of possible realizations of solvent rate for high dispersion case obtained on a high resolution ( 400 × 400) Cartesian
grid for a 100 × 100 [m] domain, showing the mean solution (blue) and 3 standard deviations as range of variation (red error bars). Mean
variation from the mean 11.4%.
5.2. DLGR A CCURACY FOR M ISCIBLE -S OLVENT D ISPLACEMENTS 55

5.2.3. E RROR A NALYSIS


DLGR employs the fine grid when and where needed, typically around saturation fronts, wells, and geologi-
cally complex features (fractures and faults). As such it employs a fraction of the fine-scale refined grid cells,
making it attractive to cast the next-generation reservoir simulation framework. As discussed before, the
quality and efficiency of the DLGR results depend on the error criterion and the specified threshold values.
As such, similar to many advanced methods, accuracy-efficiency tradeoff needs to be done for practical pur-
poses. In this chapter, quality of the DLGR solutions compared with the fine-scale reference ones (obtained
in the previous chapter) is investigated.

S TUDIED ERROR CRITERIA


As described in section 4.6, several error criteria have been considered in this work, i.e.,

1. ∇C 5. ∇µmix

2. ∆C 6. ∆µmix

3. ∂t ∇C 7. ∇mobmix

4. ∇3C 8. ∆mobmix

Concentration criteria are being frequently used, because of their easy operations. As with all other proper-
ties, the spatial gradient and the difference across grid block interfaces have been evaluated. Moreover, the
time derivative of the spatial gradient of concentration has been studied as it has shown great front tracking
potential in the literature [22]. The third order spatial derivative of concentration has also been tested, justi-
fied by the fact that the truncation error of the tracer solver is of O(∆h 2 ).
Because this study evaluates displacements in homogeneous media, instabilities and local mixing of compo-
nents are solely driven by viscous forces. Therefore, both the gradient and difference across grid interfaces
of phase viscosity µmix have been tested. Similar to the viscosity, the solvent phase mobility mob mix was se-
lected as this parameter is determining for flow as stated by Darcy’s law.
The volume flux, was predicted to be a good indicator for the local error. As stated by Darcy’s law the volume
flux contains both the mobility and the pressure gradient. Although both mobility and pressure are in essence
a function of concentration, the pressure gradient is the driving force for transport. As pressure is solved glob-
ally, the use of fluxes could give good results for high mobility ratios between displacing and displaced fluids
and/or highly heterogeneous systems where high variations in pressure exist. Unfortunately, the use of fluxes
as criteria could not be evaluated as will be explained in the final remarks of this section.
For all criteria used to assess the quality of DLGR the obtained error indicators were normalized. In total
seven simulations per criterion per dispersion coefficient have been performed, each having an increased
normalized tolerance as given in Table 5.3.

Normalized tolerances
Simulation Normalized
run tolerance
Run 1 0.005
Run 2 0.01
Run 3 0.05
Run 4 0.1
Run 5 0.15
Run 6 0.20
Run 7 0.25

Table 5.3: The set tolerances for the different simulations per criteria in order to perform quality study.

Although the space-time derivative shows sensitive responds to fronts, one should take care in solely using
this criteria for incomplete mixing caused by viscous fingering. The reason is that coarsening of static gra-
dients will occur at the interfaces of viscous fingers parallel to mean flow direction. Although the flux in the
56 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

direction perpendicular to the mean flow direction is small (resulting in low numerical error), grid coarsening
at the interfaces of fingers will occur at early stages. Therefore, it will highly reduce the instabilities observed
in DLGR simulations illustrated in Figure 5.23.

Figure 5.23: Concentration profile and imposed grid at 0.6 PVI using DLGR with space-time derivative of spatial concentration gradient
as criteria.

Therefore, in order to evaluate this criterion it was modified such that it was used for refinement while for
coarsening the gradient of concentration was added to the unrefined statement. This way strong gradients
with low time derivative were preserved.

A concluding remark is that the use of flux gradients as criteria for DLGR in reservoir simulators that
assume constant pressure within grid blocks can cause problems. This is especially true when using repeat
time steps to ensure adequate refinement. The reason has been found to be the lack of interpolation of
pressure inside coarse grid blocks neighboring refined grid blocks to the value perpendicular to the interfaces
for the flux approximations. This results in a zero order truncation error (as discussed in section 4.4). This
zeroth order error causes local rotation of the velocity field. Therefore, this local rotation of velocity field
results in steep local volume flux variations at the coarse grid block interfaces, forcing DLGR to refine. This
is illustrated in Figure 5.24 showing the tracer mass flux in the z-direction perpendicular to flow. First grid
coarsening behind the front is initiated, resulting in local inconsistent truncation errors causing the grid to
refine again in the next repeat time step.
5.2. DLGR A CCURACY FOR M ISCIBLE -S OLVENT D ISPLACEMENTS 57

Figure 5.24: Tracer mass flux in the z-direction perpendicular to flow in different stages of refinement after coarsening. Coarsening
causing local rotation of velocity field, causing refinement in successive repeat time steps.

E RROR I NVESTIGATION DLGR


In order to calculate the error made by DLGR, as a result of the trade off between active grid blocks and numer-
ical accuracy, the high resolution simulations discussed in subsection 5.2.1 were compared with the results
obtained from DLGR. The error was quantified using different error indicators based on the absolute error in
cumulative oil production, solvent rate and the concentration profile. The absolute error in cumulative oil
production is given as:
|CPdlgr (pvi) − CPfs (pvi)|
²cp (pvi) = (5.1)
CPfs (pvi)
Here CP denotes the dimensionless cumulative oi production as a function of PVI.The absolute error in sol-
vent rate:
|SRdlgr (pvi) − SRfs (pvi)|
²sr (pvi) = (5.2)
SRfs (pvi)
m 3
Here SR describes the solvent rate [ day ] as a function of PVI. The normalized l 2 -norm of the concentration
solution:

if c > 0.01
s
NOE
P ¡ ¢2
cdlgr (i) − cfs (i)
i=1
l2 norm(pvi) = (5.3)
NOE

Here NOE denotes the number of grid blocks that contain a concentration (c) higher than 0.01.

For all criteria used in this report, seven simulations were performed each with different tolerances according
to Table 5.3. As the active grid blocks per simulation time step were saved, one could calculate the average
amount of active grid blocks used for a particular simulation. The average grid block count is a direct indi-
cator for the efficiency of DLGR. In this report the average active grid size is given as a percentage of the grid
blocks used with the high resolution reference simulations.
The mean value of the error indicators stated above as well as the mean repeat time steps needed per simu-
lation time step for the different error criteria are depicted in Figures 5.25(a) to 5.25(d) for the zero dispersion
case. The property of each criterion is differentiated by color. Magenta illustrates criteria using solvent con-
centration, blue using mixture viscosity and green using mixture mobility. The red line indicates the allowable
mean deviation from the reference solution as explained in subsection 5.2.2. For each criteria the ∆ marker
symbolizes the difference across interfaces and the ∇ marker the spatial derivative.
58 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

(a) (b)

(c) (d)

Figure 5.25: DLGR quality evaluation for zero dispersion case. (a) Mean error in cumulative oil production vs mean active grid blocks
(magb), (b) mean error in solvent rate vs magb, (c) Mean normalized l 2 -norm of concentration vs magb, (d) mean repeat time steps
vs magb. Properties used for the criteria are differentiated by color: Concentration (magenta), viscosity (blue), mobility (green). The
operator used within the criteria are indicated in the legends.

Looking at the mean error of cumulative oil production and solvent rate, Figure 5.25(a) and 5.25(b), it is
observed that the criteria’s behaviour are similar. Down to a mean average active grid block count of around
40% the error resides below the allowable deviation, after which the error made linearly increases with the
relaxation of the tolerances and, thus, the reduction of active grid blocks. When looking at the colors one can
observe that in general the error made by the viscosity based criteria (shown in blue) are the lowest followed
by the concentration based criteria (magenta) and the mobility based (green) criteria is positioned highest.
The latter was not able to reach the accuracy for cumulative production as shown in Figure 5.25(a).

A possible explanation for the observed difference in quality can be deduced when analyzing the normal-
ized gradients illustrated in Figure 5.26.
5.2. DLGR A CCURACY FOR M ISCIBLE -S OLVENT D ISPLACEMENTS 59

Figure 5.26: Normalized gradients of phase viscosity (blue), concentration (magenta) and phase mobility (green) against dimensionless
length. The corresponding concentration gradient is shown in red.

The viscosity is assumed to be a strong function of concentration, formulated by (2.12) in chapter 2. This
results that the viscosity gradient is steepest at the start of solvent-oil interface (where solvent concentration
is low), concentrating the normalized gradient curve to the right of the normalized concentration gradient
curve. The phase mobility is essentially the inverse of the phase viscosity as given by:

1
λmix (c) = (5.4)
µmix (c)

Therefore, the normalized gradient of the phase mobility can be rewritten in terms of viscosity by taking the
derivative in terms of viscosity and space:

1
|∇λmix | = | ∇µmix | (5.5)
µ2mix

Due to the adverse mobility ratio we know that:

µmix,l << µmix,r

Such that the change in phase mobility is larger to the left of the front:

1 1
|∇µmix | >> |∇µmix | (5.6)
µ2mix,l µ2mix,r

This is explains the shift between the use of the norm of the gradient of phase viscosity and phase mobility
observed in Figure 5.26. As a consequence it can then be seen that the viscosity criteria will keep most of
the right hand side of the front intact. However, for the mobility criterion the set tolerances will more easily
lose refinement at the right hand side of the front and thus will introduce numerical error sooner. Figure 5.27
illustrates this concept by showing the concentration solution and its imposed gird of DLGR at 0.2 PVI for the
normalized viscosity and mobility gradient criteria using a tolerance of 0.05. The left and right column show
the viscosity and mobility gradient criteria respectively. It is observed that as motivated above, the viscosity
refines more at the front while the mobility criteria refines more along the fingers that are fully occupied by
solvent. As the tolerances are relaxed the mobility criterion loses resolution at the front, observed by dark
60 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

blue patches at the edges of some of the fingers, whereas the viscosity criteria still has complete refinement
along the edges. As a result the number of repeat time steps needed upon relaxing of the tolerances for the
viscosity criteria reduces while all other criteria have an increasing trend, as shown in Figure 5.25(d). How-
ever, it should be stressed that these results obtained for the phase viscosity and mobility criteria are based
under the assumption that the mixture viscosity can be described using the quarter-power rule as mentioned
in subsection 2.1.2. Therefore, the results obtained using phase viscosity for miscible displacements could be
inconsistent when different viscosity models and mobility ratios are assumed.
When observing the mean l 2 norm of the concentration solution, both increasing and decreasing trends are
recognized. This is caused by the fact that the errors imposed by the adaptive grid strongly affects the so-
lution, if no physical dispersion is present. Therefore, when tolerances are relaxed, non-overlapping fingers
could be coarsened and consequently show lower error than the DLGR simulations using tight tolerances.
Because of this no consistent trends can be distinguished for the l 2 norm as shown in Figure 5.25(c).
Both the space-time derivative and the third order spatial derivative of concentration were able to performed
within the allowable range of cumulative oil production and solvent rate, 1% and 11% respectively (as de-
scribed in subsection 5.2.2. However, the quality was not notably better than the first order spatial gradient
or difference criteria. Moreover, the mean repeat time steps needed per simulation time step strongly in-
creases for both error estimators with the relaxation of the tolerances (as high as 5.3 repeat time steps per
simulation time step shown in Figure 5.25(d)). In addition to computational expenses of the evaluation of the
criteria, the increased repeat time steps needed reduces the potential efficiency of DLGR.

(a) (b)

Figure 5.27: Imposed grid for zero dispersion case at 0.4 PVI. Left depicts the solution obtained with viscosity gradient criterion and right
with mobility gradient criterion. The normalized tolerance was 0.05 as shown in table 5.3.

Figure 5.30 and 5.31 show the accuracy evaluation for the medium and high dispersion case, respectively. For
the medium dispersion case the error made in cumulative oil production suggest an allowable lower limit of
10-15% mean active grid blocks, while the error made by DLGR simulations for all criteria for the high disper-
sion fell below the allowable variation. From the mean l 2 norm as well as the repeat time steps, depicted in
Figure 5.30 and 5.31, a clear separation is observed between the gradient and difference based criteria, show-
ing that for fewer mean active grid blocks the additional error made by DLGR is lower for difference based
criteria. This observed difference in local error can be explained by the different asymptotic behaviour upon
refinement, as shown by (5.7) and (5.8).

∆f ∂f
lim = (5.7)
∆x→0 ∆x ∂x
lim ∆ f = 0 (5.8)
∆x→0

Here f is an arbitrary function of location x. If we take the limit of x towards zero, the gradient converges to a
constant (C) while the difference between values of f reduces to zero. This means that gradient criteria need
a preset maximum refinement level in order to limit the amount of refinements. In addition, they can cause a
critical situation if during simulation the local gradients are smeared out by dispersion and suddenly surpass
5.2. DLGR A CCURACY FOR M ISCIBLE -S OLVENT D ISPLACEMENTS 61

the preset tolerance. Therefore, if tolerances are relaxed too much, gradient based criteria could coarsen up
finger tips, spread out by dispersion, completely (as the gradient on the coarsened grid will always stay below
the preset tolerance). This is illustrated in Figure 5.28 showing the solution of concentration based difference
(left) and gradient (right) criterion at 0.6 PVI using a normalized tolerance of 0.2.

(a) (b)

Figure 5.28: DLGR concentration solution with imposed grid at 0.6 PVI. Left shows the concentration difference criteria, right the con-
centration gradient criteria at a normalized tolerance of 0.2.

As shown in the Figure above the asymptotic nature of the difference based criteria allows for a more evenly
spread distribution of refinement levels according to the local variation in space, therefore, beneficially influ-
ences DLGR’s potential efficiency. Figure 5.29 shows the amount of grid blocks per refinement level during
the simulation time for both the concentration difference and gradient criteria, confirming that more inter-
mediate refinement levels are used with the difference criterion.

(a) (b)

Figure 5.29: Active refinement level grid blocks against PVI, illustrating the grid block size distribution of concentration difference and
gradient based criteria, depicted left (a) and right (b) hand side respectively.

The distinct difference between gradient based and difference based criteria in the mean repeat time
steps (Figure 6.12(d) and 5.31(d)), again would suggest the added benefit of the use of differences across grid
interfaces as pre-set criteria.
Observed is that the error criteria using the space-time derivative of the spatial gradient of the solvent con-
centration performed well for simulations including dispersion. This is reflected in the error in production
62 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

data and the concentration solution, as can be seen in Figure 5.31. However, at the relaxation of the tolerances
the error in cumulative production strongly increases, suggesting that the space-time derivative criteria are
smaller than the tolerances and thus no local grid refinements are introduced. Additionally, the amount of
extra repeat time steps needed for this criteria (as high as an average of 6.5 repeat time steps per simulation
time step, as opposed to around 2 for the difference based criteria) increases the computational load.

(a) (b)

(c) (d)

Figure 5.30: DLGR quality evaluation for zero dispersion case. (a) Mean error in cumulative oil production vs mean active grid blocks
(magb), (b) mean error in solvent rate vs magb, (c) Mean normalized l 2 -norm of concentration vs magb, (d) mean repeat time steps
vs magb. Properties used for the criteria are differentiated by color: Concentration (magenta), viscosity (blue), mobility (green). The
operator used within the criteria are indicated in the legends.
5.3. C ONSISTENCY I NVESTIGATION OF DLGR AND THE TL M ODEL FOR F ULLY M ISCIBLE S OLVENT F LOODING63

(a) (b)

(c) (d)

Figure 5.31: DLGR quality evaluation for zero dispersion case. (a) Mean error in cumulative oil production vs mean active grid blocks
(magb), (b) mean error in solvent rate vs magb, (c) Mean normalized l 2 -norm of concentration vs magb, (d) mean repeat time steps
vs magb. Properties used for the criteria are differentiated by color: Concentration (magenta), viscosity (blue), mobility (green). The
operator used within the criteria are indicated in the legends.

5.3. C ONSISTENCY I NVESTIGATION OF DLGR AND THE TL M ODEL FOR F ULLY


M ISCIBLE S OLVENT F LOODING
The Todd and Longstaff (TL) model, developed in 1972, is an effective model that makes use of effective fluid
properties in up-scaled grid blocks in order to capture the sub-grid effect of incomplete mixing [18]. As de-
scribed in chapter 2 one has to fit a dimensionless mixing parameter, ranging from 0 to 1, to experimental
or high resolution simulation data. Although literature describes the mixing parameter to be of physical na-
ture, one could question if the physical effect of instabilities observed in incomplete mixing processes can be
upscaled. However, due to the ease of use of TL’s model and its low computational costs to run simulations,
this approach became industrial standard ever since its appearance. As illustrated in subsection 5.2.3, DLGR
shows credible potential to overcome the limitation of effective models as it has been shown to be capable of
simulating real physics of instable flow in homogeneous media with an adaptive grid that uses in average only
a fraction of the grid blocks used in high resolution simulations. Therefore, this approach allows for accurate
description of complex processes that occur in enhanced oil recoveries.
This section evaluates the quality of TL’s model to simulate incomplete mixing as well as its consistency com-
pared to the DLGR for homogeneous solvent flooding regarding scalability in terms of reservoir size and mo-
bility.
64 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

E FFECT OF THE M IXING PARAMETER ON THE TL M ODEL

The model of TL is based on the black-oil (multi-phase) model, which it uses to effectively simulate incom-
plete mixing between miscible phases (single phase) on a sub-grid level. The mixing parameter (ω), is a
parameter that needs to be adjusted in order to fit the effective result to the data. As described in chapter 2, ω
scales the phase viscosity and density (gravity effects due to density variations are not evaluated in this work)
between that of the pure phase properties (ω=0) and the fully miscible properties (ω=1).
Figure 5.32 to 5.34 show the effect of the mixing parameter on the cumulative oil production, liquid rates and
transverse average saturation (concentration) of solvent. All figures illustrate that low values of ω result in
early break-through. This is observed by a reduction of cumulative oil production, early break-through in
solvent rate (and drop in oil rate) and the spread out transverse average saturation curve. If ω is set to zero,
one effectively solves a miscible displacement without saturation dependent viscosity and physical disper-
sion. Increasing the value of ω the viscosity (and density) increasingly behaves like a fully miscible property,
thus, reducing the mobility contrast at the front. This reduced mobility contrast stabilizes the front resulting
in an increasing cumulative oil production, later break-through time and a more stable front.

Figure 5.32: The effect of the mixing parameter on the cumulative oil production of the TL model using a resolution of 40x40 on a 100 by
100 m domain.
5.3. C ONSISTENCY I NVESTIGATION OF DLGR AND THE TL M ODEL FOR F ULLY M ISCIBLE S OLVENT F LOODING65

Figure 5.33: The effect of the mixing parameter on the produced liquid rates of the TL model using a resolution of 40x40 on a 100 by 100
m domain.

Figure 5.34: The effect of the mixing parameter on the transverse average saturation profile at 0.4 PVI of the TL model of the TL model
using a resolution of 40x40 on a 100 by 100 m domain.

E RROR I NVESTIGATION OF THE TL M ODEL


Although the TL effective model ought to be fitted to high resolution data, it is of interest to know to what
degree of accuracy the up scaled results can mimic the physical problem of incomplete mixing. To this end,
the high resolution reference simulation with dispersion (case 2 of table 5.2 in subsection 5.2.1) was used to
compare the effective results of TL for a range of different mixing parameter values and grid resolutions. As
discussed in subsection 5.2.2, the average allowable deviation from the mean cumulative oil production was
calculated to be 1.2% for the medium dispersion case. Figure 5.35 shows the surface plot of the error made
by the TL model in cumulative oil production plotted against grid block size and mixing parameter. For three
66 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

different locations on the surface plot the actual difference in cumulative production is visualized in order to
relate the error.

Cumul
ati
veoi
lpr
oduct
ion
Cumul
ati
veoi
lpr
oduct
ion

Cumul
ati
veoi
lpr
oduct
ion

Figure 5.35: The effect of grid block size and the mixing parameter on the error [%] made by the TL model compared to a fine scale
simulation in cumulative oil production for medium dispersion (case 2 of table 5.2). Both approaches were simulated linear flow on a
100 by 100 m domain, fine scale simulation resolution was 400x400.

From Figure 5.35 it can be observed that the range of accurate results obtained for given grid resolution is
narrow. In fact we know that the physical unstable displacement are allowed to vary around 1.2% from the
mean solution. Rescaling the color bar as shown in Figure 5.36 shows that the TL model was able to solve
the problem effectively around a 1% error. It can be seen that as the grid resolution reduces the optimal
mixing parameter increases. From a physical point of view, as grid resolution reduces the mixing zone of the
physical front would become relatively less important with respect to the grid block size. This would suggest
the use of a lower valued mixing parameter instead of the increased value observed. This indicates that the
mixing parameter used is a tuning parameter. As grid resolution decreases, the numerical diffusion increases
therefore the tuning parameter needs to increase (stabilizing effect on flow) to counter balance this effect.
5.3. C ONSISTENCY I NVESTIGATION OF DLGR AND THE TL M ODEL FOR F ULLY M ISCIBLE S OLVENT F LOODING67

Figure 5.36: The effect of grid block size and the mixing parameter on the error [%] made by the TL (rescaled color bar) model compared
to a fine scale simulation in cumulative oil production for medium dispersion (case 2 of table 5.2. Both approaches were simulated linear
flow on a 100 by 100 m domain, fine scale simulation resolution was 400x400.

S CALABILITY A NALYSIS OF DLGR AND THE TL M ODEL

Scalability of results of a model or simulation technique is an important aspect for consistent results. In this
section the scalability of TL and DLGR is discussed, comparing it to high resolution simulations. In prac-
tise effective models are fitted to high resolution simulations. It is therefore of great interest to know if the
matched solutions can be extended to different scales without the need of time consuming retuning.
The simulation model to evaluate the scalability is based on 4 reservoir reservoir layers, each of equal height
(25 m) and width (10 m) but ranging in length from 200, 100, 50, to 25 m. Starting from a reservoir of 200 m in
length, TL’s mixing parameter was fitted to match the fine scale simulation after which all reservoir and fluid
properties as well as the injection rate (1 ft/day) were kept constant when scaling down the reservoir length.
This way it could be evaluated if TL and DLGR are capable of capturing the development of viscous fingering
at all stages of their development.
The results for a zero dispersion case are illustrated in Figure 5.37 to 5.39. In each figure the high resolution
simulation curves are depicted in red, DLGR in green and TL’s effective model in blue. The top left (a) cor-
responds to a reservoir length of 200 m, which was used to fit the mixing parameter of TL. DLGR uses the
concentration difference across grid block interfaces as error criterion, using a threshold of 0.05. On average
this resulted in a mean active grid block count of 31% of the fine scale simulation grid size. The reservoir
length was reduced from the top to bottom and from left to right. As expected it is clearly observed that
DLGR is capable of capturing the physics independent of the scale of the system. It is also observed that the
mean active grid blocks used by DLGR reduces for longer length scales. This is caused by the fact that fingers
merge over time, therefore reducing the surface area of the instable front. TL was tuned to fit the production
data of the 200 m case, however, after reducing the reservoir length it is clearly observable that TL’s model is
not able to scale with the physical solution. The mixing parameter used shows an underprediction of local
mixing, resulting in incrementally earlier solvent break-through for the downscaled reservoir sizes. This is
illustrated by the relatively lowered cumulative oil production, a reduction in solvent break-through time and
an increasingly stretched averaged solvent profile as depicted in Figure 5.37 to 5.39.
68 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

(a) (b)

(c) (d)

Figure 5.37: Scalability evaluation results for cumulative oil production for high resolution simulations (red), DLGR (green) and TL (blue).
The mean active grid block used by DLGR is 28, 29, 33 and 34% (compared to the fine scale simulations grid size) for 200 (a), 100 (b), 50
(c) and 25 m (c) respectively.
5.3. C ONSISTENCY I NVESTIGATION OF DLGR AND THE TL M ODEL FOR F ULLY M ISCIBLE S OLVENT F LOODING69

(a) (b)

(c) (d)

Figure 5.38: Scalability evaluation results for oil rate for high resolution simulations (red), DLGR (green) and TL (blue). The mean active
grid blocks used by DLGR is 28, 29, 33 and 34% (compared to the fine scale simulations grid size) for 200 (a), 100 (b), 50 (c) and 25 m (c)
respectively.
70 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

(a) (b)

(c) (d)

Figure 5.39: Scalability evaluation results for transverse averaged concentration/saturation for high resolution simulations (red), DLGR
(green) and TL (blue). The mean active grid block used by DLGR is 28, 29, 33 (compared to the fine scale simulations grid size) and 34%
for 200 (a), 100 (b), 50 (c) and 25 m (c) respectively.

C ONSISTENCY R EGARDING M OBILITY OF DLGR AND THE TL M ODEL


Analogously to the scalability, consistency regarding the mobility ratio, M , between the displacing and dis-
placed fluid was evaluated. Three different mobility ratio’s have been used: M =10, M =100 and M =500. The
mobility ratio corresponds to the mobility ratio of oil over solvent. The DLGR simulations used the concen-
tration difference criterion. The mixing parameter of TL was matched with the high resolution simulation for
M=10. The consistency of DLGR and TL are compared for the three different dispersion cases, as have been
shown in Table 5.2 of subsection 5.2.1.
Figures 5.40 to 5.42 show the results for the consistency study regarding mobility ratio. Red depicts M=10,
green shows the results for M=100 and purple indicates an mobility ratio of 500. The high resolution simula-
tion are shown by a solid curve, TL with circles and two realization of DLGR are depicted with squares (empty
and yellow faced). The curve with yellow faced squares represent a similar DLGR simulation but with relaxed
tolerances. The average active grid count (aagc) is denoted in the legend.
From all three dispersion cases, DLGR is observed to be able to accurately capture the instable front resulting
in similar cumulative oil production curves as the high resolution simulations. In general the mean active grid
blocks used increased with the increased mobility ratio independent of the amount of physical dispersion in-
cluded in the model. This is due to the increasing interfacial surface area caused by the increased instability.
Reading from M=10 to M=500 the average grid count for DLGR ranges from 17-25% to 39-60% for the zero
dispersion case and as low as 4-5% to 20-32% for the highly dispersed case. The physical dispersion causes
the amount of viscous fingering to decrease as well as for smooth gradient, resulting that the concentration
5.3. C ONSISTENCY I NVESTIGATION OF DLGR AND THE TL M ODEL FOR F ULLY M ISCIBLE S OLVENT F LOODING71

difference criteria required fewer mean active grid blocks.


The results using TL’s model matched for the M=10 cases, as this were the realizations for which the mixing
parameter was fitted. However, when this fitting parameter value was used for different mobility ratios the
TL model was not able to obtain accurate results. For reasonable results the mixing parameter needs to be
re-tuned, hence, the effective model is not consistent with respect to mobility change. Figure 5.43 illustrates
the matched mixing parameter values against mobility ratio for each dispersion case. Clearly a linear rela-
tion exist on a semi-log scale, allowing for potential rescaling of the mixing parameter in order to fit the data
consistently. Literature reported a relation of the mixing parameter as a function of the mobility ratio (see
subsection 2.2.3), obtained by equating the fractional flow formulation derived from TL effective model with
the fractional flow obtained from Koval’s effective model assuming an average solvent concentration at the
front (see chapter 2). This relationship, depicted with yellow markers, shows a good fit with the zero disper-
sion case, however, incapable of predicting an optimal mixing parameter for cases with added dispersion.

Figure 5.40: Consistency evaluation of TL and DLGR regarding mobility ratio between displacing and displaced fluids for zero dispersion
(case 1). The different colors depict the different mobility ratios; M=10 (red), M=100 (green) and M=500 (magenta). Two different results
for DLGR are shown, obtained with different average active grid block counts (aagc) over simulation time.
72 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

Figure 5.41: Consistency evaluation of TL and DLGR regarding mobility ratio between displacing and displaced fluids for medium dis-
persion (case 2). The different colors depict the different mobility ratios; M=10 (red), M=100 (green) and M=500 (magenta). Two different
results for DLGR are shown, obtained with different average active grid block counts (aagc) over simulation time.
5.3. C ONSISTENCY I NVESTIGATION OF DLGR AND THE TL M ODEL FOR F ULLY M ISCIBLE S OLVENT F LOODING73

Figure 5.42: Consistency evaluation of TL and DLGR regarding mobility ratio between displacing and displaced fluids for high dispersion
(case 3). The different colors depict the different mobility ratios; M=10 (red), M=100 (green) and M=500 (magenta). Two different results
for DLGR are shown, obtained with different average active grid block counts (aagc) over simulation time.

Figure 5.43: Mixing parameter match against the logarithm of mobility for zero, medium and high dispersed cases. The results from the
mixing parameter relation (Christie 1993 [1]) is shown in yellow showing good results for zero dispersion cases.
74 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

5.4. C ONSISTENCY I NVESTIGATION OF DLGR AND THE TL M ODEL FOR P OLY-


MER F LOODING
During water floods of viscous oil reservoirs, instable displacements (due to an adverse mobility ratio as de-
scribed in subsection 1.1.1) as well as the appearance of thief zones in strongly correlated heterogeneous
reservoirs can be observed. The addition of water-soluble polymer to the injected water reduces the viscos-
ity ratio between oil and water, thereby increasing the volumetric sweep efficiency and lowering the swept
oil saturation [43]. As polymer is expensive, polymer slugs are usually chased by water injection in order to
drive the polymer through the reservoir. As the polymer slug is significantly more viscous than water, an
unfavourable mobility ratio exists between the chase water and the polymer slug. This could allows for insta-
bilities at the water-polymer slug front to grow and potentially decrease the efficiency of a polymer flood.
In these two-dimensional simulations a heavy oil (200 cP) reservoir of 200 m long, 20 m high and 10 m wide is
modeled. The viscosity of water (1 cP) is assumed to increase linearly with the polymer concentration (15 cP
at c polymer =1). No viscosity dependency is assumed regarding, water composition, temperature, pressure and
share rate. Also, no adsorption reduced permeability or hydrodynamic acceleration of the polymer is added
to the model. Refer to Table 5.4 and 5.5 for the geometric and other properties used during simulation.

Geometrical properties
High resolution static grid
Length X 200 [m] Properties
Length Y 10 [m] Rock permeability 400 mD
Length Z 20 [m] Water viscosity 1 [cP]
Resolution (X,Y,Z) (800,1,80) Polymer viscosity 1·c polymer +14 [cP]
DLGR Oil viscosity 100 [cP]
Length X 200 [m] S w,c 0.2
Length Y 10 [m] S o,r 0.2
Length Z 20 [m] k o,e 0.8
Base grid resolution (X,Y,Z) (50,1,5) k w,e 0.8
Refinement (X,Y,Z) (2,1,2) no 3.5
Max refinement level 4 (corresponding to nw 2.5
400x1x80 resolution) porosity 0.3
TL Injection rate 0.5 [ft/day]
Length X 200 [m] Time step size DLGR/FS: 1 day, TL: 10 days
Length Y 10 [m] Total PVI 1.5
Length Z 20 [m]
Table 5.5: Simulation properties.
Base grid resolution (X,Y,Z) (80,1,8)

Table 5.4: Geometrical properties and grid resolutions.

Figure 5.44 shows simulation results of a 0.1 PV polymer slug injection followed by 0.1 PVI chase water (total
of 0.2 PVI). Similarly, Figure 5.45 shows the same simulations at a later time after 0.3 PVI chase water (0.4 PVI).
For both Figure 5.44 and 5.45 the top row illustrates the high resolution simulation results, while the middle
and bottom row illustrate the fluid solutions obtained with DLGR, shown without and with adaptive grid,
respectively. The left column represents the polymer concentration and the right column the corresponding
oil saturation.
DLGR used a combination of both the normalized gradient of polymer concentration as well as that of the
water phase using a tolerance of 0.05, resulting in an average of around 30% grid blocks were used during
simulation. Looking at Figure 5.44 it can be clearly observed that DLGR is capable of simulating viscous fin-
gering at the chase water-polymer slug interface. Figure 5.44(e) shows refinement in front of the polymer
bank caused by a second water shock front as can be seen in the oil saturation profile. This water shock front
is caused by the fact that the connate water is assumed to be miscibly displaced by the polymer slug, thus
causing an increased saturation in front of the polymer slug bank.
Differences with the high resolution simulation can be observed at later times such as depicted in Figure 5.45.
This is most probably caused by the cumulative error caused by the dynamic gridding. However, the prop-
agation of the instabilities is still captured. Refer to Appendix B for a quantitative error analysis regarding
different error criteria.
5.4. C ONSISTENCY I NVESTIGATION OF DLGR AND THE TL M ODEL FOR P OLYMER F LOODING 75

Figure 5.46 shows the comparison of the high resolution simulation with the effective solution obtained
with TL model (ω=0.875). As the model of TL solves for phases, the polymer slug phase of TL is compared
with the water saturation multiplied with the polymer concentration of the high resolution simulation. The
first two rows depict the simulation result at 0.2 PVI, the last two rows at 0.4 PVI. The even rows (Figure
5.46(a),5.46(b),5.46(e),5.46(f )) depict the solution obtained by the high resolution simulation, while the odd
rows (Figure 5.46(b),5.46(c),5.46(g),5.46(h)) show the solution obtained from the TL model. In this case the
left column represents the polymer phase, while the right column represents the corresponding oil saturation
solution.

When looking at the solution profiles of Figure 5.46(e) to 5.46(h) it can be observed that the solution of TL
seems to be able to fit the average solution of the water penetrated polymer slug. Also, the numerical diffusion
added to the TL solution at the polymer slug front, caused by the first order accurate discretization scheme
on a relatively coarse grid, also seems to account for the viscous fingers of water that slightly penetrate the oil
bank.pHowever, as this numerical effect scales differently with time (mixing zones caused by diffusion scale
with t while that of viscous fingering scales linearly with time [8]), it can be seen that at earlier stages of
the displacement (shown in Figure 5.46(a) to 5.46(d)) the numerical diffusion does not match the physical
process.

(a) (b)

(c) (d)

(e) (f)

(g)

Figure 5.44: Solution profiles for the 0.1 PV polymer slug injection at 0.2 PVI liquid injection. From top to bottom illustrating the solutions
of the fine scale simulation and DLGR (without and without grid), respectively. The left hand side shows the polymer concentration while
the right hand side illustrates the corresponding oil saturation. Resolutions of fine scale simulation and DLGR are 80x800, 5x50 (base
grid; using 2x2 refinements until maximum level of 4).
76 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

(a) (b)

(c) (d)

(e) (f)

(g)

Figure 5.45: Solution profiles for the 0.1 PV polymer slug injection at 0.4 PVI liquid injection. From top to bottom illustrating the solutions
of the fine scale simulation and DLGR (without and without grid), respectively. The left hand side shows the polymer concentration while
the right hand side illustrates the corresponding oil saturation. Resolutions of fine scale simulation and DLGR are 80x800, 5x50 (base
grid; using 2x2 refinements until maximum level of 4).
5.4. C ONSISTENCY I NVESTIGATION OF DLGR AND THE TL M ODEL FOR P OLYMER F LOODING 77

(a) (b)

(c) (d)

(e) (f)

(g) (h)

(i)

Figure 5.46: Solution profiles for 0.1 PV polymer slug followed by chase water injection at 0.2 PVI (a-d) and 0.4 PVI (e-h) liquid injection,
for the high resolution simulation solutions (even rows) and the solution obtained with the TL model (odd rows). The left hand side
shows the fraction of polymer occupied water phase, the right hand side illustrates the corresponding oil saturation. Resolutions of high
resolution simulation and TL are 80x800 and 8x80, respectively.

A comparison of DLGR and the TL model regarding the effect of incomplete mixing on different sized polymer
slugs is shown in Figure 5.47 to 5.49, showing the cumulative oil production, oil rate and the transverse aver-
age oil saturation at two different time steps during simulation, respectively. The TL model was fitted to the
cumulative oil production solution of the high resolution simulation for the case of 0.3 PV secondary polymer
slug injection followed by chase water until 1.5 total PVI, resulting in a mixing parameter of 0.825. This value
was consequently used for the simulation of a 0.1 PV secondary polymer slug chased by water injection until
a total of 1.5 PVI. Looking at the oil production in Figure 5.47 and 5.48 it can be observed that both TL and
DLGR (using on average 8% active grid blocks, error criteria and tolerances are stated in Table 5.6) were ca-
pable of capturing the effective production correctly. As it was observed that viscous fingers were suppressed
when entering the oil phase (potentially caused by relative permeability effects). The effect of mixing mostly
were restricted to the polymer slug. This could explain the fact why the TL model showed consistent simula-
tion results for different polymer slug sizes. However, when looking at the transverse average oil saturations
in Figure 5.49, one can see that, although the average solution matches the fine scale solution reasonably well
at the moment of break through (around 0.4 PVI), the same simulation at earlier stages does not fit the satu-
ration profiles obtained with that of DLGR and the high resolution simulations. This inconsistency regarding
scalability was discussed in Figure 5.3.
78 5. I NCOMPLETE M IXING FOR H OMOGENEOUS P OROUS M EDIA

Figure 5.47: Cumulative oil production for the 0.3 PV (red) and 0.1 PV (blue) polymer slug injection. Fine scale simulations are shown in
solid lines, DLGR by dashed lines and TL with solid line with circled markers.

(a) (b)

Figure 5.48: Oil rate of the 0.3 PV (red) and 0.1 PV (blue) polymer slug injection. Fine scale simulations are shown in solid lines, DLGR by
dashed lines and TL with solid line with circled markers.
5.4. C ONSISTENCY I NVESTIGATION OF DLGR AND THE TL M ODEL FOR P OLYMER F LOODING 79

(a) (b)

Figure 5.49: Transverse average oil saturation for the 0.3 PV (left) and 0.1 PV (right) polymer slug injection at two different total PVI (red:
0.2PVI, blue: 0.4PVI). Fine scale simulations are shown in solid lines, DLGR by dashed lines and TL with solid line with circled markers.

DLGR Error Criteria Polymer Slug Size Consistency


Error criterion Tolerances
∆c polymer 0.1 [-]
1
∇S w 0.5 [ m ]

Table 5.6: DLGR error criteria used for the consistency analysis for different polymer slug sizes.
6
I NCOMPLETE M IXING FOR
H ETEROGENEOUS P OROUS M EDIA

As discussed in subsection 1.1.3, static disorder can have a determining effect on fluid flow in porous me-
dia depending on the variation of permeability and its correlation length [8–14]. In this chapter a quality
investigation of DLGR for heterogeneous cases will be shown. The permeability maps, assumed to be locally
isotropic, were statistically generated using a log-normal distribution with a mean permeability of 90 mD
and variation corresponding to a Dyskstra-Parson (Vdp ) coefficient of 0.63. Two different spatial correlations
lengths for permeability were used (λD =0.01 and λD =0.1). As DLGR uses multilevel nested refinements, the
permeability was upscaled during simulation run time in order to obtain the effective permeability corre-
sponding to the adaptive grid. For the highest refinement level (equal to the resolution of the static fine scale
reference simulations), the fine scale permeability map was used, however, for each lower level grid block
flow-based upscaling was performed to approximate the coarsened grid block’s effective permeability. The
grid resolution used for the TL model was ten times lower than for the fine scale simulations. Table 6.1 and
6.2 illustrate the dimensions, grid resolutions and properties used for the simulations.
First section 6.1 will compare DLGR to the TL model and the high resolution reference solutions. section 6.2
illustrates the error made by DLGR using different error criteria with respect to high resolution reference sim-
ulations.

81
82 6. I NCOMPLETE M IXING FOR H ETEROGENEOUS P OROUS M EDIA

Dimensions and grid resolutions


High resolution simulation
Length X 100 [m]
Length Y 10 [m]
Length Z 20 [m]
Resolution (X,Y,Z) (400,1,80)
DLGR Properties
Length X 100 [m] Solvent viscosity 1 cP
Length Y 10 [m] Oil viscosity 10 cP
Length Z 20 [m] Mean rock permeability 90 mD
Base grid resolution (X,Y,Z) (25,1,5) Dykstra-Parson coefficient 0.63
Refinement (X,Y,Z) (2,1,2) Injection rate 1 ft/day
Max refinement level 4(corresponding to Time step size FS/DLGR: 1 day, TL: 10 days
400,1,80 resolution)
TL Table 6.2: Simulation properties.
Length X 100 [m]
Length Y 10 [m]
Length Z 20 [m]
Resolution (X,Y,Z) (40,1,8)

Table 6.1: Geometrical properties and grid resolutions for all


heterogeneous reservoir simulations.

6.1. C ONSISTENCY OF DLGR AND THE TL M ODEL FOR F ULLY M ISCIBLE F LOOD -
ING
Figure 6.1 illustrates one statistical realization of a reservoir with a mean permeability of 90 mD, correlation
length of 0.01 and a Dykstra-Parson coefficient 0.63. This fine scale permeability map was used as the foun-
dation of the simulations from which flow based upscaling was performed for TL and DLGR.
The concentration solutions for all three simulations techniques as well as the grid generated by DLGR for
this realization at 0.4 PVI are shown in Figures 6.2 and 6.3. The difference between the two figures are the
tolerance set for DLGR, the first figure shows the result for DLGR using a concentration difference tolerances
of 0.05 while for the second the tolerance was relaxed to 0.15.
The dynamic adaptive grid of DLGR using the concentration difference criteria was capable of capturing
all small scale flow features using in average 35% active grid blocks compared to the fine scale simulation
(Figure 6.2(b) and 6.2(c)). Relaxing the tolerance caused the grid to concentrate only on the dominant flow
features resulting in a mean grid block count to 17% (Figure 6.3(b) and 6.3(c)).
The calibration of the mixing parameter of TL’s model for heterogeneous systems by equating the fractional
flow formulation of Koval and TL ((2.45) and (2.48)), is the only method described in literature [36]. How-
ever, as stated in literature this method can give unsatisfactory results for high Dykstra-Parson coefficients
(high variation in heterogeneity). Apart from the fact that this equation for recalibration does not take into
account the correlation length of permeability, a property that can have high influence on local and global
mixing, it was found in this study that the estimation of the mixing parameter resulted in an erroneous neg-
ative value (-0.043). By tuning the mixing parameter to the data of the high resolution reference simulation
it was found that a mixing parameter of 0.6 gave the best fit, simulating the effective flow correctly (Figure
6.2(d) and 6.3(d)).
Figure 6.4 illustrates the cumulative oil production 6.4(a), oil rate 6.4(b) and the transverse average concen-
tration/saturation of solvent 6.4(c). The high resolution simulation is shown in red, DLGR in solid green
(tolerance of 0.05) and dashed green (tolerance of 0.15), TL is depicted in blue with circle markers. One can
observe that DLGR was able to accurately simulate for the complex flow features illustrated by 6.4(b) and
6.4(c). The underlying permeability field with small correlation length caused abrupt changes in local flow
directions, resulting in increased mixing and therefore relatively small frontal surface area. This in turn al-
lowed DLGR to operate with a reduced grid of on average 17% active grid blocks, introducing only marginal
error to the solution. It is observed that the TL method, after calibration of the mixing parameter, gave sat-
isfactory results for the effective flow. However, unlike DLGR the TL model was incapable of solving for the
variations in oil production and local solvent distribution.
6.1. C ONSISTENCY OF DLGR AND THE TL M ODEL FOR F ULLY M ISCIBLE F LOODING 83

Figure 6.1: Permeability realization with dimensionless correlation length λD =0.01 and Vdp =0.63.
84 6. I NCOMPLETE M IXING FOR H ETEROGENEOUS P OROUS M EDIA

(a)

(b)

(c)

(d)

(e)

Figure 6.2: Concentration solution for a heterogeneous reservoir (Vdp =0.63 and λD =0.01, 20 by 100 m) of static high resolution, DLGR,
DLGR with imposed grid and TL consecutively. The mixing parameter for TL was equal to 0.6 and the mean active grid blocks used by
DLGR equaled 35% of high resolution grid. Resolutions of high resolution simulation,DLGR and TL are 80x400, 5x25 (base grid; using
2x2 refinements until maximum level of 4) and 8x40.
6.1. C ONSISTENCY OF DLGR AND THE TL M ODEL FOR F ULLY M ISCIBLE F LOODING 85

(a)

(b)

(c)

(d)

(e)

Figure 6.3: Concentration solution for a heterogeneous reservoir (Vdp =0.63 and λD =0.01, 20 by 100 m) of static high resolution, DLGR
(shown with and without the imposed grid) and TL, respectively. The mixing parameter for TL was equal to 0.6 and the mean active grid
blocks used by DLGR equaled 17% of high resolution grid. Resolutions of high resolution simulation,DLGR and TL are 80x400, 5x25 (base
grid; using 2x2 refinements until maximum level of 4) and 8x40.
86 6. I NCOMPLETE M IXING FOR H ETEROGENEOUS P OROUS M EDIA

(a) (b)

(c)

Figure 6.4: Averaged solutions comparing high resolution simulation (80x400) with DLGR and TL (8x40) (ω=0.6) for heterogeneous case
with λD =0.01 of the 20 by 100 m.
6.1. C ONSISTENCY OF DLGR AND THE TL M ODEL FOR F ULLY M ISCIBLE F LOODING 87

Figure 6.5 shows a realization with a permeability field with mean permeability of 90 mD, a correlation length
of 0.1 and a Dykstra-Parson coefficient of 0.63. As in the previous case with small correlation length, this fine
scale permeability map was flow-based upscaled (locally) for TL and the different grid levels of DLGR during
simulation run time.

Figure 6.5: Permeability realization with dimensionless correlation length λ=0.1 and Vd p=0.63 of a domain of 20 by 100 m.

Figure 6.6 illustrates the concentration profile at 0.4 PVI for the static high resolution simulation, DLGR with
a mean average grid block count of 44% (shown with and without imposed adaptive grid, resulting from a
concentration difference criterion using a tolerance of 0.05) and TL’s effective model. Figure 6.7 shows a sim-
ilar representation but here DLGR’s tolerances were relaxed to 0.2, resulting in 29% mean active grid blocks
with respect to the high resolution reference simulation.
By tuning the mixing parameter to the data of the high resolution reference solution it was found that a mix-
ing parameter of 0.375 gave the best fit, simulating the effective flow correctly (Figure 6.6(d) and 6.7(d)). The
increased permeability correlation length caused channelization of solvent flow, thus reducing local mixing
reflected by the low mixing parameter used for the TL model.
Again DLGR was capable of capturing the physics of fluid flow altered by the underlying permeability field.
However, due to the increased correlation length of permeability, the frontal surface area is higher than in
the previously described case resulting in a higher amount of refinements during simulation. Figure 6.8(a) to
6.8(c) show the cumulative oil production, solvent rate and transverse solvent concentration/saturation re-
spectively. The high resolution simulation is shown in red, DLGR in solid green (tight tolerances) and dashed
green (relaxed tolerances), TL is depicted in blue with circle markers. The figures indicate clearly that TL’s
model can forecast the oil production as its effective results are fit to the fine scale simulation data, but it
does not accurately predict the solvent distribution through time.
88 6. I NCOMPLETE M IXING FOR H ETEROGENEOUS P OROUS M EDIA

(a)

(b)

(c)

(d)

(e)

Figure 6.6: Concentration solution for a heterogeneous reservoir (Vdp =0.63 and λD =0.1, 20 by 100 m) of static high resolution, DLGR,
DLGR with imposed grid and TL consecutively. The mixing parameter for TL was equal to 0.375 and the mean active grid blocks used
by DLGR equaled 44% of high resolution grid. Resolutions of high resolution simulation,DLGR and TL are 80x400, 5x25 (base grid; using
2x2 refinements until maximum level of 4) and 8x40.
6.1. C ONSISTENCY OF DLGR AND THE TL M ODEL FOR F ULLY M ISCIBLE F LOODING 89

(a)

(b)

(c)

(d)

(e)

Figure 6.7: Concentration solution for a heterogeneous reservoir (Vdp =0.63 and λD =0.1, 20 by 100 m) of static high resolution, DLGR,
DLGR with imposed grid and TL consecutively. The mixing parameter for TL was equal to 0.375 and the mean active grid blocks used
by DLGR equaled 29% of high resolution grid. Resolutions of high resolution simulation,DLGR and TL are 80x400, 5x25 (base grid; using
2x2 refinements until maximum level of 4) and 8x40.
90 6. I NCOMPLETE M IXING FOR H ETEROGENEOUS P OROUS M EDIA

(a) (b)

(c)

Figure 6.8: Averaged solutions comparing high resolution simulation (80x400) with DLGR and TL (8x40) (ω=0.375) for heterogeneous
case with λD =0.1 of the 20 by 100 m domain as shown in the Figures above.
6.1. C ONSISTENCY OF DLGR AND THE TL M ODEL FOR F ULLY M ISCIBLE F LOODING 91

The consistency between DLGR and TL for different, but statistical equal (mean permeability of 90 mD
and Vdp =0.63), heterogeneous realizations were tested for the permeability dimensionless correlation length
of 0.01 and 0.1. The TL model was fitted to the first realizations described above (ω=0.6 for λD =0.01 and
ω=0.375 for λD =0.1). These mixing parameter were used for the different statistically equal realizations. Re-
sults obtained from DLGR are shown using both tight and relaxed concentration difference tolerance (0.05
and 0.15, respectively).
Figure 6.9 and 6.10 show the consistency of both approaches compared to high resolution simulations for
the correlations lengths of 0.01 and 0.1, respectively. From top to bottom the cumulative oil production, oil
rate and transverse average solvent concentration/saturation are shown. The left column corresponds to
DLGR with tight tolerance (0.05) resulting in on average 37% and 47% active grid blocks for λD =0.01 and 0.1,
respectively. The right column shows results using the relaxed tolerance (0.15) resulting in on average 19%
and 30% active grid blocks for λD =0.01 and 0.1, respectively. The three different heterogeneous permeability
realizations are depicted in the colors red, blue and green. High resolution results are shown with a solid
curve, DLGR with a dashed curve and the solid curve with circled markers show the results obtained with
the TL model. It can be observed that the TL model shows low sensitivity to the variation in production and
local solvent concentration distributions observed in the high resolution and DLGR simulations. This is re-
flected by the production data and transverse average solvent concentration profiles of both Figure 6.9 and
6.10. One could argue that the low sensitivity of the TL’s model is due to the fact that low grid resolution were
used to solve for incomplete mixing effectively. However, this is as yet reality for reservoir simulation due to
computational limitations. The fact that DLGR can locally refine the grid allows for local preservation of the
underlying geological features, clearly indicating the potential of DLGR to solve for incomplete mixing in het-
erogeneous media. This is illustrated by the fact that the production results of DLGR (shown in both Figure
6.9 and 6.10) clearly show the variations similarly observed in the high resolution simulations. Furthermore,
one could control the level of accuracy by either tightening or relaxing of the pre-set tolerances of the error
criteria, depending on the demand of accuracy needed. It should be noted that this preservation of the un-
derlying heterogeneity is at a cost, as local flow-based upscaling is computational expensive. However, this is
not further studied in this work.
92 6. I NCOMPLETE M IXING FOR H ETEROGENEOUS P OROUS M EDIA

(a) (b)

(c) (d)

(e) (f)

Figure 6.9: Averaged solutions comparing high resolution simulation (80x400) with DLGR and TL (8x40) (ω=0.6) for heterogeneous case
with λD =0.01 of the 20 by 100 m domain.
6.1. C ONSISTENCY OF DLGR AND THE TL M ODEL FOR F ULLY M ISCIBLE F LOODING 93

(a) (b)

(c) (d)

(e) (f)

Figure 6.10: Averaged solutions comparing high resolution simulation (80x400) with DLGR and TL (8x40) (ω=0.375) for heterogeneous
case with λD =0.1 of the 20 by 100 m domain.
94 6. I NCOMPLETE M IXING FOR H ETEROGENEOUS P OROUS M EDIA

6.2. A CCURACY I NVESTIGATION OF DLGR FOR F ULLY M ISCIBLE S OLVENT FLOOD -


ING

Figure 6.11 illustrates the quantitatively behaviour of DLGR for a heterogeneous reservoir of Vdp =0.63 and
λD =0.01, as previously described in section 6.1. As with the previous analyses the used properties for the
error criteria are differentiated by color. Magenta illustrates criteria using solvent concentration, blue using
phase viscosity and green using phase mobility. For each criterion the ∆ marker symbolizes the difference
across interfaces and the ∇ marker the spatial gradient. In the top left and right, Figure 6.11(a) and 6.11(b),
the mean error in cumulative oil production and solvent rate are depicted with respect to the mean active
grid blocks used during simulation as a result of the pre-set tolerances. The red solid curve illustrates the
uncertainty with respect to fluid flow assuming that the deterministic permeability distribution is unknown.
From this uncertainty quantification, obtained from five different but statistically similar realizations, it can
be said that the mean maximum deviation inherent to the uncertainty of the local distribution of permeability
corresponds to 4.5% with respect to cumulative oil production. Using this observation as an allowable error,
DLGR was able to simulate down to an average of 10-20% active grid blocks compared to the high resolution
reference simulation while residing within the inherent uncertainty of the defined problem.
Figure 6.11(c) shows the average normalized l 2 norm of the concentration solution obtained during simula-
tion with DLGR with respect to the high resolution simulation. From this analysis one can observe that the
solvent concentration criteria performed marginally better than the criteria using viscosity or mobility. This
could be explained by the fact that fluid transport is dominated by variations in heterogeneity. These domi-
nant flow paths created by strong variations in permeability are best described by the solvent concentration
criteria, because the use of phase viscosity or phase mobility shifts the focus of refinement to the start or
end of the front, respectively (as discussed for fully homogeneous systems in subsection 5.2.3). Therefore,
the refinements made by the concentration criteria would capture the location of strong variation of perme-
ability and consequently pressure the best, contributing to a lower mean error. A possible explanation phase
viscosity based criteria outperformed the phase mobility criteria in terms of accuracy (similar to what was
observed for homogeneous systems) could be due this reduced sensitivity of phase mobility based criteria to
variations at lower solvent concentration. This causes parts of beginning of the front to be progressively less
refined upon relaxation of the tolerances.

When looking at the repeat time steps needed per simulation time steps, viscosity based criteria clearly on
average need 30% less adaptations of the grid as a consequence the higher sensitivity at the start of the front.
Due to the existence of strong gradients, there is no strong difference observed between the use of gradient
and difference based criteria.
Figure 6.12 illustrates the same analysis for λD =0.1. Observations are similar to the previously described
case, however, the incremental error made by DLGR upon the relaxation of the tolerances is larger. This
can be explained by the fact that the increase permeability correlation length causes for more channelized
transport, increasing the frontal interfacial area of the physical problem during simulation time. Inherent
to the necessity of refinements along the front this resulted in increased mean active grid blocks needed to
accurately capture transport.
6.2. A CCURACY I NVESTIGATION OF DLGR FOR F ULLY M ISCIBLE S OLVENT FLOODING 95

(a) (b)

(c) (d)

Figure 6.11: DLGR accuracy evaluation for heterogeneous medium (Vd p =0.63 and λD =0.01, 20 by 100 m). (a) Mean error in cumulative
oil production vs mean active grid blocks (magb), (b) mean error in solvent rate vs magb, (c) Mean normalized l 2 -norm of concentration
vs magb, (d) mean repeat time steps vs magb. Properties used for the criteria are differentiated by color: Concentration (magenta),
viscosity (blue), mobility (green). The operator used within the criteria are indicated in the legends.
96 6. I NCOMPLETE M IXING FOR H ETEROGENEOUS P OROUS M EDIA

(a) (b)

(c) (d)

Figure 6.12: DLGR accuracy evaluation for heterogeneous medium (Vd p =0.63 and λD =0.1, 20 by 100 m). (a) Mean error in cumulative oil
production vs mean active grid blocks (magb), (b) mean error in solvent rate vs magb, (c) Mean normalized l 2 -norm of concentration vs
magb, (d) mean repeat time steps vs magb. Properties used for the criteria are differentiated by color: Concentration (magenta), viscosity
(blue), mobility (green). The operator used within the criteria are indicated in the legends.
7
C ONCLUSIONS

7.1. I NCOMPLETE M IXING IN H OMOGENEOUS P OROUS M EDIA


7.1.1. P ERFORMANCE STUDY DLGR
It can be concluded that the efficiency of the state-of-the-art DLGR as currently implemented is highly de-
pendent on the physical behavior of fluid flow. For unstable flow, the interfacial area of the front increases
due to the growth of viscous fingers which in turn increases the requirement of local refinements. It was
found that the increase of active grid blocks for DLGR to simulate unstable displacements highly reduced the
efficiency in terms of CPU time.
As discussed in section 5.1 it was found that:

• The simulations of stable miscible displacements using DLGR were able to simulate with only an av-
erage of 8% active grid blocks in comparison to the reference fine scale simulations. This reduction of
active grid blocks significantly lowered the solver time, however, the overhead caused by DLGR for the
calculation of error criteria and the recalculation of (interfacial) grid properties was significant. This
overhead dedicated to DLGR’s functionalities summed up to an average of 70% of the total CPU time
per time step. In the tested case of stable displacements, an performance gain of 35% compared to
the high resolution simulation was observed. Both the high resolution and the DLGR simulations were
found to scale approximately linear with the increased grid size in the direction orthogonal to flow.

• DLGR simulation of unstable transport of miscible transport was performed using on average 24% ac-
tive grid blocks compared to the high resolution reference simulation. Due to the non-trivial data struc-
ture needed to account for the refinements made by DLGR, and the amount of repeat time steps (on
average 1.8 repeat time steps per simulation time step) needed to properly refine the grid resulted that
the performance was on average three times slower than that of the static high resolution reference
simulations. Again also for unstable displacements both DLGR and the high resolution static grid sim-
ulation’s CPU time were found to approximately increase linearly with the increased grid size, allowing
for reliable efficiency behaviour upon rescaling of the system.

• As the in-house reservoir simulator originally was not constructed for adaptive gridding techniques, the
grid property that accounts for the ratio of grid sizes of neighbouring grid blocks (DDTR) is not saved
during simulation. Therefore, when simulating tracer transport including diffusion and/or dispersion
using DLGR, this property needs to be recalculated over the entire grid whenever grid adaptation is
applied. This highly reduced the efficiency of DLGR (holding for around 30% of the total simulation
CPU time).

7.1.2. A CCURACY INVESTIGATION OF DLGR


The accuracy investigation of DLGR to solve for incomplete mixing through instable displacements of mis-
cible transport was performed. The error of solution obtained by DLGR was compared with high resolution
reference simulations. The effect of the accuracy on the use of different error criteria for grid adaptation
based on solvent concentration, viscosity and mobility have been studied .

97
98 7. C ONCLUSIONS

O BSERVED E RRORS
• DLGR has showed to be capable of simulating incomplete mixing to the desired numerical accuracy,
similar to that obtained from static high resolution simulations.
For the case without physical dispersion, it was found that the error made in cumulative production
resided below that of the physical variation of the problem using DLGR with a mean active grid block
count down to 40-45%. This significant refinement needed can by explained by the fact that the lack
of physical dispersion allows for many small instabilities to grow, thus resulting in a high interfacial
area of the front. On average around 2 repeat time steps per simulation time step were needed by the
refinement criteria to operate within the error range of the physical problem.

• Accuracy evaluation of DLGR for the cases of instable displacement with added physical dispersion
showed a lower overall error. This can be explained by the fact that the solution is less sensitive to nu-
merical error imposed by the adaptive grid due to the physical dampening introduced in the model.
When analysing the error made in cumulative oil production, the medium dispersion case shows that
DLGR can operate within the physical variation down to the use 10-20% mean active grid blocks com-
pared to the high resolution reference simulation. For the highest dispersed case the error resided
below the physical range of solutions at a mean active grid size as low as 5%.

E RROR C RITERIA
• In general it was shown that viscosity based criteria imposed the lowest incremental error upon re-
duction of average active grid blocks (relaxing of the predefined tolerances) during simulation among
the studied criteria. This can be explained by the fact that viscosity of miscible flow is assumed to
be a strong non-linear function of component concentration, so that for unfavourable mobility ratio
the change of this property, and thus its sensitivity, was the largest at beginning of front (low solvent
concentration). Therefore, criteria using the mixture viscosity kept most of the start of the fluid front
refined upon relaxation of the tolerances. This resulted in a lower incremental error compared to the
concentration and mobility based criteria.
However, it should be noted that this conclusion is based under the assumption that the mixture vis-
cosity can be described using the quarter-power rule as mentioned in subsection 2.1.2. Therefore,
the results obtained using phase viscosity for miscible displacements could differ when other viscosity
models or viscosity ratio’s are assumed.

• For dispersed instable fronts a distinction separation was observed between the incremental error
made by gradient criteria in comparison to difference based criteria upon relaxation of the tolerances.
This is observed for the cumulative oil production, solvent rate as well as the normalized average l 2
norm of the concentration solution.
The reason why gradient based criteria introduce a larger incremental error compared to difference
based criteria can be explained by the fact that gradient based criteria are not able to converge to a
refinement level according to spatial change of a property, whereas difference based criteria do. There-
fore, gradient based criteria could coarsen up viscous finger tips, that are spread out by physical disper-
sion, completely when the gradient falls below the predefined tolerance. In contrary, the use difference
based criteria resulted in an more evenly distribution of different refinement levels.

• Both the time derivative of the spatial concentration gradient and the third order spatial derivative of
concentration were able to reach errors within the allowable margin denoted by the physical variation.
However, the quality was not notably better than the first order spatial gradient or difference criteria.
Moreover, the mean repeat time steps needed per simulation time step strongly increases for both er-
ror estimators upon relaxation of the tolerances. As these additional derivatives are computationally
expensive, the increased number of repeat time steps reduces the potential efficiency of DLGR.

7.1.3. C ONSISTENCY I NVESTIGATION OF DLGR AND THE TL M ODEL


E RROR I NVESTIGATION OF THE TL M ODEL
In order to determine the accuracy of the TL model, an error analysis was performed for a mobility ratio of 10,
quantifying the mean error made in cumulative oil production with respect to mixing parameter (ω) and the
grid resolution. It was observed that the TL model is able to generate results around the physical allowable
variation. However, from this analysis showed that when grid resolution reduces, the optimal mixing param-
eter increases. From physical point of view, as grid resolution reduces the mixing of the physical front would
7.2. I NCOMPLETE M IXING FOR H ETEROGENEOUS P OROUS M EDIA 99

become relatively less important with respect to the grid block size. This indicates that the mixing parameter
used is a tuning parameter, counteracting the amount of physical diffusion upon grid coarsening, rather than
a physical scaling parameter.

S CALABILITY A NALYSIS OF DLGR AND THE TL M ODEL


When simulating miscible unstable transport, DLGR showed to be capable of capturing the physics of instable
displacement independent of the scale of the system. It was also observed that the mean active grid blocks
used by DLGR reduces for longer length scales, compared to the high resolution simulations. This is caused
by the fact that miscible fingers merge over time, therefore reducing the surface area of the instable front.
The solution obtained with TL was not able to scale with the physical solution. As the mixing parameter was
tuned to the production data of the largest reservoir test case, this mixing parameter showed an increasing
underprediction of local mixing upon downscaling of the reservoir domain, resulting in incrementally earlier
solvent break-through.

C ONSISTENCY R EGARDING M OBILITY OF DLGR AND THE TL M ODEL


Similar as for the scalability, consistency regarding the mobility ratio between the displacing and displaced
fluid for DLGR was observed. The solutions were able to accurately capture the propagation of the unstable
front resulting in an accurate prediction of the cumulative oil production. In general the mean active grid
count increased with the increased mobility ratio independently of the amount of physical dispersion due to
the increased instability at the front.
Fitting the effective solution of the TL model to the production data of one particular mobility ratio did not
scale like the reference solution upon the change of mobility. However, equating the fractional flow formula-
tion of the Koval model [17] with that of the TL model resulted in an accurate mixing parameter estimation
for unstable miscible displacements for different mobility ratios, but only when physical dispersion was ne-
glected.

QUALITY I NVESTIGATION OF DLGR AND THE TL M ODEL FOR P OLYMER F LOODING


DLGR was found to be capable of simulating viscous fingering behind a secondary polymer slug injection,
caused by the adverse mobility ratio between the chase water and the polymer slug. By analyzing the error
made by DLGR it was found that the use of an adaptive grid down to 20% active grid blocks compared to the
high resolution simulation gave adequate results within the uncertainty range caused by the unstable flow.
Both DLGR and TL showed consistent simulation results for different polymer slug sizes. However, the TL
model was not able to match concentration profiles at early stages before break-through of polymer, again
illustrating the inconsistency of the TL model regarding scalability.

7.2. I NCOMPLETE M IXING FOR H ETEROGENEOUS P OROUS M EDIA


DLGR was found to be capable of reproducing complex flow behavior of unstable flow dominated by static
disorder of heterogeneous reservoirs. The amount of grid refinement needed to accurately solve the trans-
port equations for a constant variation in permeability was found to be dependent on the correlation length
of the heterogeneities.

• For small correlation length (λD =0.01) of permeability, local mixing is accelerated by the abruptly chang-
ing flow lines. This mixing allowed DLGR to accurately solve for transport down to a mean active grid
block count of 10%.
Increasing the permeability correlation length caused channeled transport of solvent. This phenomenon
of channeling reduced the local mixing and increased the interfacial area of the front. Consequently,
this resulted in a requirement of at least 30% of mean active grid blocks when compared to the high
resolution reference solutions.

• The TL model showed reasonably good results effectively, however, it showing low sensitivity to varia-
tions observed in production solutions and local solvent concentration distributions compared to the
high resolution and DLGR simulations. On could argue that the low sensitivity of the TL’s model is due
to the fact of the upscaled, low resolution grid used to solve for incomplete mixing effectively, however,
this is as yet reality for reservoir simulation due to computational limitations.
As for every DLGR simulation the fine scale permeability data was stored, local preservation of the
underlying geological features was obtained by the use of local flow-based upscaling of permeability
100 7. C ONCLUSIONS

occording to the adaptively refined grid. This clearly indicates the potential of DLGR to solve for in-
complete mixing in heterogeneous media.
8
R ECOMMENDATIONS FOR F UTURE W ORK
R ECOMMENDATION F LUX A PPROXIMATION
As the in-house reservoir simulator assumes pressure to be constant within grid blocks, the lack of pressure
interpolation for the flux approximation at the interface of grid blocks of different sizes results in a rotation in
the flux field, see section 4.4. Not only could this give rise to unnecessary numerical errors when using static
LGR, but these rotations of the flux field at the location of coarsened grid blocks or at the border of refine-
ments during DLGR simulations triggers refinement and therefore makes the use of fluxes as error criteria less
practical. As fluxes are properties that contain both a concentration dependent property (mobility) as well as
the potential gradients, the use of this property as error criteria for DLGR could yield positive results when
solving for highly heterogeneous and/or fluids with high mobility contrast. The fact that fluxes are already
calculated when solving for transport, means that they could simply be stored during the solution procedure
allowing for low computational costs.
It is recommended to implement a simple first order pressure interpolation for the approximating of fluxes
when using DLGR. By interpolation to the corresponding pressure perpendicular to the interface of grid
blocks of different sizes, the zero order error as described in section 4.4 should be mitigated. This could
result in an elimination of the local rotation of the flux field. Apart from an increased local spatial accuracy
of the flux approximation for irregular grids, this implementation would allow for further investigation of the
use of fluxes as error criteria for DLGR using the semi-implicit gridding strategy.

R ESEARCH IN I MPROVING E FFICIENCY OF DLGR


From this study it was found that both the error criteria calculation and the adaptive recalculation of grid
properties upon grid adaption take up most of the simulation CPU time (up to 70%) when simulating for
incomplete mixing phenomena. Investigation of possibilities that could increase the efficiency is crucial in
order for DLGR to become a practical and fast alternative simulation technique in field scale simulations.
The calculation of error criteria can be very time consuming depending on the amount of active grid blocks
used during simulation. The pressure and saturation/concentration residuals of an implicitly solved lin-
earized system of equations could give a direct indication of the local error and, therefore, could liberate
the need of additional error criteria calculations. However, it is not possible to access the residuals of the
converged solutions. It is recommended to investigate the potential of the use of the residuals as error crite-
ria.
To overcome limitations of the increased complexity of data structures and the adaptive calculation of com-
putational expensive grid properties (e.g. traditional upscaling of permeability and the calculation of con-
vection transmissibilities), one could investigate the use of DLGR in conjunction with algebraic multiscale
methods for finite volume discretizations [44–46]. This approach could potentially result in efficiency gains
as one would keep the structure of the underlying static fine scale grid.
Finally, the use of parallel computing could be one way of speeding up current bottle necks of DLGR. As the
proper use of GPU processing power with the DLGR data structure is not straightforward, investigation of the
potential efficiency gain that could be accomplished when using parallel computations is recommended.

101
102 8. R ECOMMENDATIONS FOR F UTURE W ORK

I NVESTIGATION OF M ORE A CCURATE E FFECTIVE M IXING M ODELS


From this study it can be concluded that although the empirical TL model can be tuned to fit the effective re-
sults that are obtained from high resolution simulations, it is inconsistent to variations in scale, Peclet num-
ber and mobility. Although the use of DLGR looks promising, the development of mixing models that can
reproduce fine scale simulation results of unstable flow is also highly desired. It is therefore recommended to
further investigate possible mixing models that upscale incomplete mixing phenomena.
Recent studies [47–50] have proposed a mixing model that quantifies the decay of concentration variance (an
indication of mixing) and the evolution of the mean scalar dissipation rate, assuming statistical homogeneity
of viscous fingering. It is reported to be capable of capturing the characteristic stretching of the unstable in-
terfaces over which diffusive mixing takes place. Investigating the practicality of these models for the use of
reservoir simulation, and studying its potential to simulate for heterogeneous systems is a topic that would
be of high interest.
A
O NE D IMENSIONAL N UMERICAL
D IFFUSION A NALYSIS FOR F IRST O RDER
S CHEMES
Considering a one dimensional domain with an injector well on the left hand side and a producer well at the
right hand side. A truncation error analysis is performed in order to quantitatively estimate the amount of
numerical diffusion a first order accurate scheme imposes to the solution. This numerical diffusion can be
equated to the physical dispersion (both terms consist of second order differential terms assuming constant
velocity), leading to an estimation of grid block sizes needed for which numerical dispersion is lower than
that of the physical dispersion.

The partial differential equation of the convection-diffusion equation in one dimension (here omitting the
diffusion term) is given by:

∂c ∂c
φ +u =q (A.1)
∂t ∂x
Using the Taylor expansion technique we can construct a numerical approximation of the temporal and spa-
tial derivatives for a typical upwind scheme.

∂c ∆x 2 ∂2 c
c(x i −1 , t ) = c(x i , t ) − ∆x (x i , t ) + (x i , t ) − O(∆x 3 ) (A.2)
∂x 2 ∂x 2

c(x i , t ) − c ( x i −1 , t ) ∂c ∆x ∂2 c
= (x i , t ) − (x i , t ) + O(∆x 2 ) (A.3)
∆x ∂x 2 ∂x 2
The same can be done to approximate the time derivative.

∂c ∆t 2 ∂2 c
c(x i , t n+1 ) = c(x i , t n )∆t (x i , t n ) + (x i , t n ) − O(∆t 3 ) (A.4)
∂t 2 ∂t 2

c(x i , t n+1 ) − c(x i , t n ) ∂c ∆t ∂2 c


= (x i , t n ) − (x i , t n ) + O(∆t 2 ) (A.5)
∆t ∂t 2 ∂2 t
Substituting the derivatives of the exact solutions with their numerical approximations gives.

c(x i , t n+1 ) − c(x i , t n ) c(x i , t n ) − c(x i −1 , t n )


φ +u =q (A.6)
∆t ∆x
The numerical approximation is subtracted from the analytical solution, resulting in the truncation error.

∆t ∂2 c ∆x ∂2 c
φ + u (A.7)
2 ∂t 2 2 ∂x 2

103
104 A. O NE D IMENSIONAL N UMERICAL D IFFUSION A NALYSIS FOR F IRST O RDER S CHEMES

As we want to compare the the numerical diffusion and second order physical dispersion, the time derivative
term is rewritten in terms of x.

∂c u ∂c ∂2 c u ∂ u ∂c u2 ∂2 c
=− thus, =− (− )= 2 2 (A.8)
∂t φ ∂x ∂t2 φ ∂x φ ∂x φ ∂x
This results in:

¡ ∆t u 2 ∆x ¢ ∂2 c
+u (A.9)
2 φ 2 ∂x 2
Using the Courant–Friedrichs–Lewy (CFL) condition we can rewrite ∆t in terms of ∆x. Considering a CFL
number of 1 the above stated equation is rewritten.

u ∆t
1 = cfl = (A.10)
φ ∆x

∆x
∆t = φ (A.11)
u

1 ∂2 c ∂2 c
(∆xu + ∆xu) 2 ≤ D 2 with, D = u · dl (A.12)
2 ∂x ∂x
This results in a criterion for which the numerical diffusion is less than the physical dispersion:

∆x ≤ d l (A.13)

It should be noted that this is a simplistic estimate of a grid resolution at which the numerical diffusion is
in the same order or lower than that of the physical dispersion. For multi-dimensional problems one must
consider the fact that flow velocity changes over time and space.
B
A CCURACY INVESTIGATION DLGR FOR
P OLYMER F LOODING

(a) (b)

(c) (d)

Figure B.1: DLGR accuracy evaluation for polymer flood of secondary 0.1 PV polymer slug chased by water injection until 1.5 PVI. (a)
Mean error in cumulative oil production vs mean active grid blocks (magb), (b) mean error in solvent rate vs magb, (c) Mean normalized
l 2 -norm of concentration vs magb, (d) mean repeat time steps vs magb. Properties used for the criteria are differentiated by color:
Concentration (magenta), viscosity (blue), mobility (green). The operator used within the criteria are indicated in the legends.

105
N OMENCLATURE
S YMBOLS
Symbols Description

λD Dimensionless correlation length permeability


ρ Density
µ Viscosity
Ω Control volume
∂Ω Surface area of control volume
A Area
c concentration
Ca’ Modified capillary number
CTR Convective transmissibility
DLGR Dynamic local grid refinement
D Dispersion tensor
f Fractional flow
g Gravitational acceleration
G Gravity number
h Height
H Heterogeneity factor
k Permeability
kr Relative permeability
k r,e Relative permeability end-point
LGR Local grid refinement
MOB Mobility
n Normal vector
p Pressure
Pe Peclet number
q Source term
S Saturation
t Time
T Transmissibility
TL Todd and Longstaff
u Velocity
V Volume
Vdp Dykstra-Parson coefficient
w Weight concentration

S UBSCRIPTS /S UPERSCRIPTS
D Dimensionless
eff Effective
i Index for x-direction
j Index for y-direction
l Longitudinal
n Current time step
mix Phase mixture
m Molecular
t Transverse
H Harmonic average
ω Mixing parameter of the Todd and Longstaff model

107
108 B. A CCURACY INVESTIGATION DLGR FOR P OLYMER F LOODING

P HASES
n Non-wetting
w Water
o Oil
s Solvent
B IBLIOGRAPHY
[1] M. Christie, A. Muggeridge, J. Barley, et al., 3d simulation of viscous fingering and wag schemes, SPE
reservoir engineering 8, 19 (1993).

[2] R. Chuoke et al., The instability of slow, immiscible, viscous liquid-liquid displacements in permeable
media, (1959).

[3] P. G. Saffman and G. Taylor, The penetration of a fluid into a porous medium or hele-shaw cell containing
a more viscous liquid, Proceedings of the Royal Society of London. Series A. Mathematical and Physical
Sciences 245, 312 (1958).

[4] S. Hill, Channeling in packed columns, Chemical Engineering Science 1, 247 (1952).

[5] C. Park and G. Homsy, The instability of long fingers in hele-shaw flows, Physics of Fluids (1958-1988) 28,
1583 (1985).

[6] R. Slobod, R. Thomas, et al., Effect of transverse diffusion on fingering in miscible-phase displacement,
Society of Petroleum Engineers Journal 3, 9 (1963).

[7] G. M. Homsy, Viscous fingering in porous media, Annual review of fluid mechanics 19, 271 (1987).

[8] J. R. Waggoner, J. L. Castillo, and L. W. Lake, Simulation of EOR (enhanced oil recovery) processes in
stochastically generated permeable media, Tech. Rep. (Sandia National Labs., Albuquerque, NM (USA);
Texas Univ., Austin, TX (USA). Dept. of Petroleum Engineering, 1990).

[9] D. Moissis, M. Wheeler, C. Miller, et al., Simulation of miscible viscous fingering using a modified method
of characteristics: effects of gravity and heterogeneity, SPE Advanced Technology Series 1, 62 (1993).

[10] H. Tchelepi, F. Orr Jr, et al., Interaction of viscous fingering, permeability heterogeneity, and gravity segre-
gation in three dimensions, SPE Reservoir Engineering 9, 266 (1994).

[11] R. Solano, S.-T. Lee, P. R. Ballin, T. P. Moulds, et al., Evaluation of the effects of heterogeneity, grid refine-
ment, and capillary pressure on recovery for miscible-gas injection processes, in SPE Annual Technical
Conference and Exhibition (Society of Petroleum Engineers, 2001).

[12] J. R. Waggoner, V. J. Zapata, L. W. Lake, et al., Viscous mixing in unstable miscible displacements, preprint
(1991).

[13] U. G. Araktingi, F. Orr Jr, et al., Viscous fingering in heterogeneous porous media, Ph.D. thesis, Stanford
University (1988).

[14] D. Brock, F. Orr Jr, et al., Flow visualization of viscous fingering in heterogeneous porous media, in SPE
Annual Technical Conference and Exhibition (Society of Petroleum Engineers, 1991).

[15] P. Audigane, M. J. Blunt, et al., Dual mesh method in upscaling, in SPE Reservoir Simulation Symposium
(Society of Petroleum Engineers, 2003).

[16] J. V. Lambers, M. G. Gerritsen, D. Fragola, et al., Multiphase, 3-d flow simulation with integrated upscal-
ing, mpfa discretization, and adaptivity, in SPE Reservoir Simulation Symposium (Society of Petroleum
Engineers, 2009).

[17] E. Koval et al., A method for predicting the performance of unstable miscible displacement in heteroge-
neous media, Society of Petroleum Engineers Journal 3, 145 (1963).

[18] M. Todd, W. Longstaff, et al., The development, testing and application of a numerical simulator for pre-
dicting miscible flood performance, Journal of Petroleum Technology 24, 874 (1972).

109
110 B IBLIOGRAPHY

[19] F. J. Fayers et al., An approximate model with physically interpretable parameters for representing miscible
viscous fingering, SPE reservoir engineering 3, 551 (1988).

[20] T. Hermitte, D. Guerillot, et al., A more accurate numerical scheme for locally refined meshes in heteroge-
neous reservoirs, in SPE Symposium on Reservoir Simulation (Society of Petroleum Engineers, 1993).

[21] Z. Heinemann, G. Gerken, G. von Hantelmann, et al., Using local grid refinement in a multiple-
application reservoir simulator, Society of Petroleum Engineers (1983).

[22] D. W. Van Batenburg, A. De Zwart, P. M. Boerrigter, M. Bosch, and J. C. Vink, Application of dynamic
gridding techniques to ior/eor processes, in IOR 2011 (2011).

[23] D. Han, D. Han, C. Yan, L. Peng, et al., Spe 16014 a more flexible approach of dynamic local grid reservoir
modeling, (1987).

[24] W. Mulder, R. Meyling, et al., Numerical simulation of two-phase flow using locally refined grids in three
space dimensions, SPE advanced technology series 1, 36 (1993).

[25] Y. Ding, P. Lemonnier, et al., Development of dynamic local grid refinement in reservoir simulation, in SPE
Symposium on Reservoir Simulation (Society of Petroleum Engineers, 1993).

[26] H. Hoteit, A. Chawathe, et al., Making field-scale chemical eor simulations a practical reality using dy-
namic gridding, in SPE EOR Conference at Oil and Gas West Asia (Society of Petroleum Engineers, 2014).

[27] G. Rousseaux, A. De Wit, and M. Martin, Viscous fingering in packed chromatographic columns: Linear
stability analysis, Journal of Chromatography A 1149, 254 (2007).

[28] J. Hagoort, Measurement of relative permeability for computer modeling reservoir simulation, Oil & Gas
Journal 82, 62 (1984).

[29] P. Tabeling, G. Zocchi, and A. Libchaber, An experimental study of the saffman-taylor instability, Journal
of Fluid Mechanics 177, 67 (1987).

[30] C. T. Tan and G. M. Homsy, Stability of miscible displacements in porous media: Rectilinear flow, Physics
of Fluids 29 (1986).

[31] H. Dykstra and R. Parsons, The prediction of oil recovery by waterflood, Secondary Recovery of Oil in the
United States 160 (1950).

[32] A. T. Corey, The interrelation between gas and oil relative permeabilities, Producers monthly 19, 38 (1954).

[33] D. Peaceman, H. Rachford Jr, et al., Numerical calculation of multidimensional miscible displacement,
Society of Petroleum Engineers Journal 2, 327 (1962).

[34] J. Bear, Hydrodynamic dispersion, Flow through porous media , 109 (1969).

[35] D. Moissis, Simulation of viscous fingering during miscible displacement in nonuniform porous media,
(1988).

[36] F. Fayers, M. Blunt, M. Christie, et al., Comparisons of empirical viscous-fingering models and their cali-
bration for heterogeneous problems, SPE reservoir engineering 7, 195 (1992).

[37] M. Edwards, M. Christie, et al., Dynamically adaptive godunov schemes with renormalization in reservoir
simulation, in SPE Symposium on Reservoir Simulation (Society of Petroleum Engineers, 1993).

[38] C. Simeoni, Remarks on the consistency of upwind source at interface schemes on nonuniform grids, Jour-
nal of Scientific Computing 48, 333 (2011).

[39] J. Douglas Jr, B. L. Darlow, M. Wheeler, R. P. Kendall, et al., Self-adaptive galerkin methods for one-
dimensional, two-phase immiscible flow, in SPE Reservoir Simulation Symposium (Society of Petroleum
Engineers, 1979).

[40] J. Manik and T. Ertekin, Development and application of dynamic and static local grid refinement algo-
rithms for water coning studies, in SPE Eastern regional meeting (1997) pp. 153–160.
B IBLIOGRAPHY 111

[41] M. C. Michaud, F. Bisshopp, et al., Accurate waterflood simulation using biased differencing and selective
grid refinement, (1981).

[42] W. D. Gropp, A test of moving mesh refinement for 2-d scalar hyperbolic problems, SIAM Journal on Sci-
entific and Statistical Computing 1, 191 (1980).

[43] L. W. Lake, Enhanced oil recovery, (1989).

[44] H. Hajibeygi, G. Bonfigli, M. A. Hesse, and P. Jenny, Iterative multiscale finite-volume method, Journal of
Computational Physics 227, 8604 (2008).

[45] H. Hajibeygi and P. Jenny, Adaptive iterative multiscale finite volume method, Journal of Computational
Physics 230, 628 (2011).

[46] Y. Wang, H. Hajibeygi, and H. A. Tchelepi, Algebraic multiscale solver for flow in heterogeneous porous
media, Journal of Computational Physics 259, 284 (2014).

[47] B. Jha, L. Cueto-Felgueroso, and R. Juanes, Fluid mixing from viscous fingering, Physical review letters
106, 194502 (2011).

[48] B. Jha, L. Cueto-Felgueroso, and R. Juanes, Quantifying mixing in viscously unstable porous media flows,
Physical Review E 84, 066312 (2011).

[49] .

[50] P. Jenny, J. S. Lee, D. W. Meyer, and H. A. Tchelepi, Scale analysis of miscible density-driven convection in
porous media, Journal of Fluid Mechanics 749, 519 (2014).

You might also like