Download as pdf or txt
Download as pdf or txt
You are on page 1of 124

Instituto Tecnológico y de Estudios

Superiores de Monterrey

Campus Monterrey
School of Engineering and Sciences

Simulation and modeling of three mechanisms of flow through


porous media.

A thesis presented by

Michel Romero Flores

Submitted to the
School of Engineering and Sciences
in partial fulfillment of the requirements for the degree of

Master of Science
In

Engineering Science in Energy Engineering

Monterrey Nuevo León, December 5th, 2017


Dedication
Esta Tesis está dedicada a mi madre Marisol Flores y mis hermanos Rosman y Miguel Ángel,
quienes me han apoyado en todo momento y ante toda adversidad, y gracias a quienes
siempre he podido dar lo mejor de mí.
A mi Difunto padre Miguel Ángel Romero Sánchez, quien durmió en el eterno descanso en
el transcurso de mis estudios de maestría en el Tecnológico de Monterrey y a quien le debo
todo en esta vida por enseñarme a seguir adelante y no flaquear ante ninguna situación.
A los amigos que hice en la maestría, quienes me dieron su apoyo de forma incondicional
ante toda dificultad, especialmente a Eduardo Salinas y Marco Antonio Rodríguez, a quienes
considero parte de mi familia, a quienes jamás lograre expresar mi eterna gratitud por estar
ahí en los momentos más difíciles y a quienes llevare en mi mente, mi espíritu y mi corazón
por el resto de mi vida.
A mis compañeros y amigos de laboratorio, especialmente a Mariandrea y Diana, quienes
siempre me brindaron su apoyo y amistad durante las incontables horas de estudios.

i
Acknowledgements

Gracias al Dr. José Luis, quien fue un excelente asesor, por su increíblemente extenso
conocimiento del tema, por su apoyo a través de toda mi estadía en la maestría y por ser una
gran persona.
Al Dr. Alejandro Montesinos, quien me brindó su apoyo y consejo en momentos difíciles.
Al Dr. Micheloud por encontrarme y ofrecerme la maravillosa oportunidad de estudiar esta
maestría.
A los Drs. Alejandro Javier García y Carlos Iván Rivera por su apoyo durante la elaboración
de esta tesis.
Al programa de becas del Tecnológico de Monterrey, sin el cual nada de esto habría sido
posible.

ii
Simulation and modeling of three mechanisms of flow through porous media.
by

Michel Romero Flores

Abstract
The mathematical modeling of two-phase flow in saturated porous media, as well as the
modeling of adsorption/retention behavior of surface active materials in a porous medium
composed of a complex network of macro, meso, and micropores in 1-D, 2-D and
axisymmetric cases and the displacement of two phases in capillary conducts, were studied.
Also, a Comparison of two CFD tools is made (COMSOL and ANSYS Fluent). The results
were compared with experimental data from the literature. For the saturated porous media
study and mathematical modeling of adsorption/retention behavior, COMSOL software was
used, while for the capillary conducts displacement, ANSYS Fluent was chosen. In the
saturated porous media analysis, different ways to obtain permeabilities and their effects on
the flow in saturated porous media were compared. For the mathematical modeling of
adsorption/retention behavior of surface active materials in a porous medium different effects
were analyzed: the selection of boundary conditions, the size of the tracer and surfactant
signals, effects of reversible and irreversible adsorption, the difference between local
equilibrium and the rate-limited process. Also, for 2-D and axisymmetric simulations,
heterogeneities, blends of surfactant and dispersion/diffusion effects were studied. The
proposed mathematical model compares favorably with experimental data from literature
when taking macro and mesoporosity into account. This model should be helpful in guiding
the design of dynamic adsorption experiments, and to understand how heterogeneities in the
rock may influence the interpretation of experimental results. Finally, capillary displacement
was analyzed in an axisymmetric system, and the results were compared against a one-
dimensional model and experimental results obtained through experimentation. This was
done to identify the benefits of simulating this type of phenomena, because, sometimes a
simplified model can hide vital information for the experiments. The axisymmetric
simulations were superior when showing the complete information of the phenomena but at
a much higher computational cost than the one-dimensional model.

3
Contents
Declaration of Authorship .................................................... ¡Error! Marcador no definido.

Dedication ................................................................................................................................ i

Acknowledgements ................................................................................................................ ii

Abstract ................................................................................................................................... 3

Contents .................................................................................................................................. 4

List of Figures ......................................................................................................................... 8

List of Tables ........................................................................................................................ 13

Nomenclature........................................................................................................................ 15

Chapter I: Introduction ......................................................................................................... 18

1.1. Problem Statement ................................................................................................. 24

1.2. Objective ................................................................................................................ 25

1.2.1. Specific objectives .......................................................................................... 25

1.3. Justification ............................................................................................................ 25

1.4. Organization of the thesis ...................................................................................... 26

1.5. References .............................................................................................................. 28

Chapter II: Theoretical framework, literature review and applicable tools. ......................... 30

2.1. Theoretical framework ........................................................................................... 30

2.1.1. Adsorption ...................................................................................................... 30

2.1.2. Isotherm of adsorption .................................................................................... 33

2.1.3. Interfacial tension ........................................................................................... 35

2.1.4. Contact angle .................................................................................................. 36

4
2.1.5. Surfactants ...................................................................................................... 37

2.1.6. Porosity and its classification ......................................................................... 38

2.1.7. Wettability and capillarity .............................................................................. 40

2.1.8. Rock saturation and volume fraction .............................................................. 44

2.1.9. Permeability .................................................................................................... 46

2.1.10. Relative permeability .................................................................................. 46

2.2. Literature review .................................................................................................... 47

2.3. Applicable CFD tools ............................................................................................ 49

2.3.1. Finite element method .................................................................................... 50

2.3.2. Finite volume method ..................................................................................... 51

2.3.3. Commercial software available ...................................................................... 52

2.4. References .............................................................................................................. 54

Chapter III: Methodology ..................................................................................................... 58

3.1 Multiphase flow in saturated porous media ........................................................... 58

3.1.1 Model development ........................................................................................ 58

3.1.2 Simplification and dimensionless system of equations .................................. 61

3.1.3 Boundary conditions ....................................................................................... 62

3.1.4 Methodology for model application ............................................................... 63

3.2 Modeling surfactant adsorption/retention and transport through porous media .... 65

3.2.1 Model development ........................................................................................ 65

5
3.2.2 One-dimensional system of equations ............................................................ 69

3.2.3 Boundary conditions ....................................................................................... 70

3.2.4 Methodology for model application ............................................................... 71

3.2.5 Adsorption model analysis ............................................................................. 72

3.2.6 Parameters analysis for calculations in 2-D and axisymmetric porous media 75

3.3 Axisymmetric simulations of displacements in capillary conducts ....................... 78

3.3.1 Experimental setup ......................................................................................... 79

3.3.2 Model used ..................................................................................................... 80

3.3.3 Domain definition and boundary conditions .................................................. 81

3.3.4 Methodology for model application ............................................................... 82

3.4 References .............................................................................................................. 84

Chapter IV: Results and discussion ...................................................................................... 86

4.1. Results of multiphase flow in saturated porous media .......................................... 86

4.1.1. Adjusted parameters ....................................................................................... 86

4.1.2. Sensitivity analysis for alpha coefficient ........................................................ 87

4.2. Results of modeling surfactant adsorption and/or retention .................................. 88

4.2.1. Model element analysis .................................................................................. 91

4.2.2. Adsorption in heterogeneous porous media 2-D ............................................ 96

4.2.3. Axisymmetric 2-D simulations in porous media. ......................................... 101

4.3. Results of axisymmetric simulations for displacements in capillary conducts.... 104

6
4.3.1. Experimental results and comparison ........................................................... 104

4.3.2. Conduct radius effect .................................................................................... 107

4.3.3. Interfacial tension effects.............................................................................. 109

4.3.4. Model comparison ........................................................................................ 110

4.4. References ............................................................................................................ 113

Chapter V: Conclusions and future work ........................................................................... 114

ANNEXES ......................................................................................................................... 119

7
List of Figures
Figure 1.1. World energy consumption by source, 1990-2040 ............................................ 18

Figure 1. 2. Primary energy demand by fuel in Mexico ....................................................... 19

Figure 1.3.Mexican oil production and exports, 1990-2015 ................................................ 20

Figure 1.4. Different EOR Methods .................................................................................... 21

Figure 1. 5. Surfactant injection process .............................................................................. 23

Figure 2.1. Difference between absorption and adsorption .................................................. 30

Figure 2.2. Particles in an aqueous media being adsorbed in a solid surface ....................... 32

Figure 2.3. A wire loop with a slide wire on which a liquid film might be formed and stretched

by an applied force F. ........................................................................................................... 36

Figure 2.4. Profile of a three-phase boundary that defines the contact angle θ.................... 37

Figure 2.5 Molecule of surfactant (Left) and surfactant-stabilized droplet (right). ............. 38

Figure 2.6 Example of a porous media. High porosity – Larger spaces (Left) and low porosity

- Smaller spaces (right) ......................................................................................................... 39

Figure 2.7. Surface tension effects: a) equilibrium position of the fluid-fluid interface near

the solid surface, b) rise of the wetting fluid in a capillary tube, c) pendular ring around the

contact point of two solid grains........................................................................................... 41

Figure 2.8. Spatial configurations of water and air in an unsaturated granular porous medium:

a) adsorbed regime, b) capillary pendular regime, c) capillary funicular regime, d) occluded

air bubble regime, e) fully saturated regime. ........................................................................ 42

Figure 2.9. Pore-scale representative elementary volume .................................................... 44

8
Figure 3.1. Experimental data used by Kulkarni et al for relative permeability of oil (a) and

water (b) vs. water saturation ............................................................................................... 64

Figure 3.2. Experimental data used by Kulkarni et al for capillary pressure vs. water

saturation in a porous media for an oil-water system ........................................................... 65

Figure 3.3. Model showing the heterogeneous porous media where 2-D calculations were

carried out. Kp1 and Kp2 indicates the permeability, p1 and p2 indicates the pressure applied

to the entrance and outlet respectively, being p1> p 2 and the arrows indicates the flow

direction. The dimensions of this system are 1 x 1 m. ......................................................... 76

Figure 3.4. Model showing the heterogeneous porous media where axisymmetric calculations

were carried out. Kp1 and Kp2 indicates the permeability, p1 and p2 indicates the pressure

applied to the entrance and outlet respectively, being p1>p2, the arrows indicates the flow

direction and the red dotted line indicates the axisymmetric condition of the system. The

dimensions of this axisymmetric system are 6 x 1 m. .......................................................... 77

Figure 3.5. Experimental setup used for two-phases capillary displacement experiments .. 79

Figure 3.6. Simulated domain details. a) Represents a slice of the complete simulation while

b) shows the actual simulated system. .................................................................................. 82

Figure 4.1. Simulation results for 20, 70, 100 and 250 minutes, respectively, using the values

of the adjusted coefficients ................................................................................................... 87

Figure 4.2. Water saturation for different alpha values at 20, 70, 100 and 250 minutes,

respectively. .......................................................................................................................... 88

Figure 4.3. Exit concentration (C/C0) as a function of pore volume (PV). Comparison

between computed and experimental results obtained for (a) C1-T (Tracer) and C1-S

(Surfactant) and (b) C2-T (Tracer) and C2-S (Surfactant). .................................................. 89

9
Figure 4.4. Exit concentration (C/C0) as a function of pore volume (PV). Comparison between

computed (our model and Mannhardt´s model) and experimental results obtained for (a) C3-

T (Tracer) and C3-S (Surfactant) and (b) C4-T (Tracer) C4 and C4-S (Surfactant) ................. 90

Figure 4.5. Model prediction of effluent concentration for tracer (IT-T), Langmuir isotherm

(IT-L) and Redlich-Peterson (IT-RP) for a rectangular pulse of 3 PV. ................................ 91

Figure 4.6. Model prediction of (a) solid phase (𝐶(𝑠) ) and aqueous (𝐶(𝑎𝑞)) concentration

for two Langmuir isotherms: isotherm A (IP-A: A=10, k=25.32) and isotherm B (IP-B: A=1.9,

k=4) and (b) effluent concentration for two types of Langmuir isotherm (parameters as in fig

8(a)) and for the tracer (IP-T). .............................................................................................. 92

Figure 4.7. Model results considering different downstream boundary conditions (continuity

and discontinuity) at the location x = 35 cm, and using two different Peclet numbers (5 and

20) (DB: Downstream Boundary condition, Core Length: L1 =35 cm and L5 = 178 cm, Peclet

number: Pe5 = 5 and Pe20 = 20) .......................................................................................... 94

Figure 4.8. Model prediction of (a) outlet concentration using different slug sizes (SS: Slug

Size or injection size) and (b) adsorption obtained using different slug sizes. .................... 94

Figure 4.9. Model prediction of effluent concentration for tracer (NE1-T and NE2-T) and

surfactant (NE2-S and NE2-S) as a function of pore volume (PV). (a) Using the same value

for Nst14 and Nst25 in non-equilibrium and (b) using a constant value of Nst14 = 1 while

varying Nst25 in non-equilibrium. In both legends, the value for the Stanton number that

varies appears after the Nst tag. ............................................................................................ 95

Figure 4.10. Model prediction of outlet surfactant and tracer concentrations as a function of

pore volume for different porous network scenarios. Continuous lines represent the surfactant

(S) concentration and the dashed lines represents tracer (T) concentration. The porous network

considered is described in the legend as: Macro (M), meso (m) and micro (μ) porosity. ....... 96

10
Figure 4.11. Shape of the concentration signal at the effluent of the porous media. ........... 97

Figure 4.12. Time lapse for a 2-D simulation of a heterogeneous porous media. Maximum

normalized concentration (1) is represented with red color, while zero concentration is in

blue. Figure a) is for PV=0.5, b) for PV=1.5, c) for PV=2.5, d) for PV=4, e) for PV=5 and f)

for PV=6 ............................................................................................................................... 98

Figure 4.13. Model showing velocity field during the simulation with lower permeability,

Red zone is the higher pressure, and blue the lowest pressure. a) shows the velocity

magnitude and direction while b) indicates just the velocity direction. ............................... 98

Figure 4.14. Output signal for heterogeneous porous media under the assumption of different

levels of dispersion/diffusion. .............................................................................................. 99

Figure 4.15. Response of the concentration downstream for tracer, and two different

surfactants. The two surfactants present no interaction in this simulation. ........................ 100

Figure 4.16. Tracer and surfactant blend concentration normalized. ................................. 101

Figure 4.17. Surfactant and tracer output normalized concentration for axisymmetric

simulation (Case 1). ............................................................................................................ 101

Figure 4.18. Surfactant and tracer signal for axisymmetric simulations with double peak

(Case 2). .............................................................................................................................. 102

Figure 4.19. Time lapse for an axisymmetric simulation of a heterogeneous porous media

using a rectangular impulse of surfactant in the injection (Case 1). Maximum normalized

concentration (1) is represented with red color, while zero concentration is in blue. Figure a)

is for PV=0.25, b) for PV=0.5, c) for PV=1.2, d) for PV=2, e) for PV=2.5 and f) for PV=3. . 103

Figure 4.20. Experimental results for blue oil for displacement in capillary conducts and the

results obtained using Fluent model and an one-dimensional model. ................................ 104

Figure 4.21. Experimental drop analyzed for the blue oil case. ......................................... 105

11
Figure 4.22. Experimental results for red oil for displacement in capillary conducts and the

results obtained using Fluent model and an one-dimensional model. ................................ 106

Figure 4.23. Experimental drop analyzed for the red oil case. ........................................... 107

Figure 4.24. Drop shape for varying conduct radius. a) drop shape for r=0.1 mm, b) r= 0.2

mm and c) r=0.3mm ........................................................................................................... 107

Figure 4.25. Emptying speed for varying conduct radius .................................................. 108

Figure 4.26. Drop shape for varying interfacial tension. a) IFT = 5e-5, b) IFT=5e-4 and c*)

IFT =5e-3. For the c*) case, injection was used to form the drop, because the oil could not go

out by itself due to the high IFT. ........................................................................................ 109

Figure 4.27. Emptying speed for varying IFT .................................................................... 110

Figure 4.28. Emptying speed models comparison for varying conduct radius .................. 111

Figure 4.29. Emptying speed models comparison for varying IFT .................................... 111

12
List of Tables
Table 1.1. EOR application criteria ...................................................................................... 22

Table 2.1. Classification of pores according to their width. ................................................. 40

Table 3.1. Models for transport in porous media comparison. ............................................. 58

Table 3.2. Values used in Kukarnki et al work .................................................................... 65

Table 3.3. Description of experimental data sets and their tags for model adjustment. ....... 72

Table 3.4. Adsorption key elements for analysis and model parameters used in calculations .. 74

Table 3.5. Porosity (φ), macroporosity fraction (f1) and mesoporosity fraction (f2) used for

analysis of key element......................................................................................................... 75

Table 3.6. Adsorption key elements for analysis and model parameters used in calculations

of 2-D and axisymmetric modeling. ..................................................................................... 78

Table 3.7. Properties for oils used in experiments ............................................................... 80

Table 3.8. Parameters for axisymmetric simulations. .......................................................... 83

Table 4.1. Adjusted model parameters for comparison with Kulkarni’s results .................. 86

Table 4.2. Adjusted model parameters using irreversible isotherm for comparison with

experimental results .............................................................................................................. 89

Table 4.3. Adjusted model parameters using hybrid isotherm for comparison with

experimental results. ............................................................................................................. 90

Table 4.4. Surfactant adsorption (in milligrams of surfactant per gram of rock) for model

prediction when varying isotherm type. ............................................................................... 91

Table 4.5. Surfactant adsorption (in milligrams of surfactant per gram of rock) for model

prediction when varying isotherm parameters. .................................................................... 93

Table 4.6. Correlation coefficients and IFT for the models used in the blue oil experiments ….105

13
Table 4.7. Comparison of IFT’s obtained through two different methods. ........................ 105

Table 4.8. Correlation coefficients and IFT for the models used in the blue oil experiments . 106

Table 4.9. Comparison of IFT’s obtained through two different methods. ........................ 107

14
Nomenclature
𝑑𝑎 Damping or accumulation coefficient

𝑒a Mass or inertial coefficient

a Adsorption coefficient or exchange coefficient for 1-D and 2-D systems

c Diffusion or dispersion coefficient

α Convective flux coefficient

β Advection vector or convection vector coefficient

γ Conservative flux source

f Source coefficient

u Dependent variable to solve for

𝑁𝑆𝑡 Stanton number

𝑁𝑃𝑒 Peclet number

A Adsorption coefficient

C Concentration

K Saturation coefficient

F The function of the adsorption, which is expressed in terms of concentration in the


aqueous phase, in contact with the porous region under examination
ni Saturation index

𝑘𝑐12 Mass transfer coefficient between macro and meso porosity

𝑎𝑣12 Specific area for mass transfer per unit volume [m2/m3]

𝑎𝑣13 Specific area for mass transfer per unit volume [m2/m3], cross section of the necks that
connect macro pores with meso pores

15
𝑎𝑣14 Specific area for mass transfer per unit volume [m2/m3] between the aqueous phase in
the macro porous region and the solid in contact with the macro pores

𝑎𝑣25 Specific area for mass transfer per unit volume [m2/m3] between the aqueous phase in
the meso porous region and the solid in contact with the meso pores

𝑎𝑣36 Specific area for mass transfer per unit volume [m2/m3] between the aqueous phase in
the micro porous region and the solid in contact with the micro pores
l Length of the core sample

v Superficial velocity

𝐷𝑖 Dispersion/diffusion tensor in the porous region, i

𝜑 Porosity

𝜑𝑖 Porosity of the region i, where: 1 = macro pores, 2 = meso pores and 3 = micro pores

𝐶𝐴,𝑖 Concentration in the porous region i, where: 1 = macro pores, 2 = meso pores and 3 =
micro pores

𝐶𝑖𝐿,𝑆 Concentration in the aqueous phase in equilibrium in contact with the solid

𝜒 Dimensionless axial component

𝜏 Pore volume or dimensionless time

𝑝𝑖 Pressure at the boundary i

𝑘𝑝𝑖 Permeability of section i in porous media

𝑓𝑖 Normalized porosity for section i, where: 1 = macro pores, 2 = meso pores and 3 =
micro pores
i Index to identify the region of interest, where 1 = macro pore, 2 = meso pore, 3 =
micro pore, 4 = aqueous phase in equilibrium with the solid, in contact with the macro
pore, 5 = aqueous phase in equilibrium with the solid. in contact with the meso pore,
and 6 = aqueous phase in equilibrium with the solid, in contact with the micro pore
ao Affinity of the medium towards the oil

16
aw Affinity of the medium towards water

Ca Capillary number

Di Hydraulic dispersion coefficient of phase i

g Acceleration of gravity

j Fluid or phase in the system (o = oil, w = water)

J(Sj) Leverett equation for phase j

K Absolute system permeability

krj Fluid permeability j

krj0 Relative fluid permeability j

rj Term of fluid generation j

Sj Saturation of phase j

Srj Residual saturation of fluid j in the system

Uj Flux volumetric phase j

∇pc Capillary pressure gradient

∇pj Pressure gradient of phase j

∇z Distance in the direction of gravity

z Capillary conduct height

𝑧̇ Capillary conduct height

Φ Porosity

σ Interfacial tension

ρj Fluid density j

μj Fluid viscosity j

θ Contact angle

17
Chapter I: Introduction
Energy development links to comfort and prosperity around the world. Reaching the growing
demand for energy in a harmless and responsible way is a crucial challenge.

In this times of global market unpredictability, one thing we do know is that the world needs
energy sources (and in growing amounts) to sustenance economic and social development
and build a better life quality, especially in developing nations. But providing these energy
sources around the globe comes with a responsibility: to establish and use our resources in a
responsible way. We have to protect people and the world while making economic progress.

For developing countries, the necessity of trustworthy and reasonable energy sources is
essential. It can improve and save lives. In these countries, reliable energy sources can
enhance the industry, agriculture, and transportation among other things. These are the bases
that help people to create better lives.

Nowadays, the principal source of energy are the hydrocarbons, being crude oil the dominant
source of energy. As shown in Figure 1.1, international energy demand may increase about
48 percent from 2012 to 2040 according to the Energy Information Association (EIA)
prediction[1] even with the significant advances in energy efficiency made in the past few
years.

Figure 1.1. World energy consumption by source, 1990-2040[1]

18
Global energy production must increase to reach the increasing demand from all the nations.
Renewable energies are not capable of sustaining this growing demand. So, we need all
energy sources available. Thus, an important demand of oil and gas will rise in the next years.
But given the actual depletion of existing fields, these hydrocarbons must be obtained from
non-exploited areas using new or unconventional methods[2] including enhanced oil
recovery methods (EOR).

In our times, oil is the world’s most important source of energy. It provides essential support
to power industry and modern society in general, including transport.

Oil and gas both have key impact on all levels of the economy, since oil and gas prices affect
transportation, health, comfort, among other things. To protect a country economy, oil and
gas industry must proper in a consistent way.[3]

Energy demand in Mexico has increased approximately by 25% since 2000 to this day. This
increment is closely related to the economic expansion of the country. This energy
consumption is predicted to continue growing in the next years[4]. Figure 1. 2 shows the
energy demand by fuel in Mexico. From 2000 to 2014 the natural gas usage has increased by
8% and but oil remains as the predominating source of energy.

Figure 1. 2. Primary energy demand by fuel in Mexico[4]

Nowadays fossil fuels are the primary energy source used in Mexico, principally: oil, natural
gas, and coal. Natural gas usage has increased in the past few years mostly because the
exploitation of shale gas in the USA and the construction of new gas pipelines to Mexico to
increase the gas trade.[4]

19
Mexico is one of the largest oil producers in the world, but in recent years, it production has
declined and the demand has increased. In 2004 crude production was 3.8 million barrels per
day (mb/d) but it decreased to 2.6 mb/d in 2015[4]. This is shown in Figure 1.3.

Figure 1.3.Mexican oil production and exports, 1990-2015[4]

The production of oil fields is declining, and there is an insuficient development of new oil
fields to compensate this decrement. Mexico has a large reserve of oil, but the small
investment in exploration and technologies for extraction, combined with an inadequate
revenues distribution of the Mexican government, is stopping the development of new oil
fields.

Development of new extraction methods is needed to guarantee the economic stability of any
country.

Crude oil expansion and production in reservoirs include three different phases: primary,
secondary, and tertiary (or enhanced) recovery. Through the first recovery, the natural
pressure of the reservoir push oil to the extraction point, together with artificial forces (i.e.,
pumps) they transport the oil to the surface. Around 10 % of a reservoir's oil is produced with
this recovery type. Second recovery technique prolongs a field's productive life usually by
injecting water or gas to move the oil to the production point, obtaining a recovery of
approximately 20 - 40 % of the field’s oil.

However, when the easy oil is extracted, producers must use tertiary or enhanced oil recovery
(EOR) processes. These methods consist in the implementation of various techniques for

20
increasing the amount of crude oil that can be extracted from an oil field[5]. Using EOR
technologies, 30 – 60 %[6], or more, of oil from a reservoir can be extracted, compared with
20 - 40 % for conventional methods[7]. There are many different kinds of EOR methods,
Figure 1.4 shows most of them.

Figure 1.4. Different EOR Methods [5]

Thermal EOR techniques include all processes that introduce heat to the reservoir,
increasing the flow of oil to by reducing its viscosity. These are worldwide the most advanced
EOR methods. In other words, thermal recovery is the use of heat to lower the viscosity of
the oil, increasing the recovery. When heated, oil viscosity decreases, and then it flows more
quickly. This is an important property of oil, so, considerable effort has been made to develop
techniques that include the use of heat in reservoirs to increase recovery of the heavier and
more viscous oils. The viscosity of oils drops as temperature rises. Thus, the purpose of all
thermal oil recovery processes is to heat the oil and make it flow[5].

Chemical EOR methods had the utmost fame in the 1980’s, most of them in sandstone
reservoirs. The projects involving EOR improved in 1986 with polymer flooding as the most
important chemical EOR method. But, since 1990’s, oil production from chemical EOR
methods has been reduced around the world except for China. Still, chemical flooding has
shown to be sensitive to the volatility of oil markets despite recent advances (e.g., low
surfactant concentrations) and lower costs of chemical additives. Polymer flooding is

21
considered as a superior technology and is the most significant EOR chemical method in
sandstone reservoirs. These methods consist in the modification of interfacial properties of
oil trapped in the oil field through the use of chemicals to ease its extraction.[5]

EOR gas flooding is the most applied recovery method for light, condensate, and volatile oil
reservoirs. Nitrogen (N2) injection has been proposed to increase oil recoveries under
miscible conditions favoring the vaporization of small fractions of light oils and condensates.
But, today few N2 floods are ongoing in sandstone reservoirs. In contrast, CO2 flooding has
been the most popular EOR technique for medium and light oil production in oil reservoirs
during last year’s, due to the availability of cheap and accessible CO2 from natural sources.
Gas flooding methods consist of an injection process that introduces miscible gases into the
reservoir. A miscible displacement process maintains reservoir pressure and improves oil
movement reducing the interfacial tension between oil and water.[5]

EOR methods applications vary according to diverse factors, especially, reservoir’s depth
and viscosity of oil. The application criteria for some EOR methods is shown in Table 1.1.

Table 1.1. EOR application criteria[8]

EOR Method Depth, ft Viscosity, cP Permeability, md


Steam 500 – 5,000 50 – 9,000 >=200
In Situ 1,000 – 10,000 10 – 1,000 >=200
Alkaline 0 – 10,000 0 – 500 >=20
Surfactant 0 – 9,000 0 – 50 >=20
Polymer 0 – 9,000 0 – 150 >=20
3,000 – 10,000 and
CO2 0 – 15 Any range
above

The most widespread EOR technique is CO2 extraction. This technology is applied in many
countries around the world. Usually, CO2 used for this method is obtained from natural
reservoirs, but now it’s being produced from industrial processes. This technique has some
critical problems when employed in an oil field with fractures or big pores; it loses efficiency
in this kind of reservoirs. Also, it presents some relevant environmental damages and is very
difficult to achieve a profitable extraction with this method.

22
Another EOR method is the usage of surfactants. This consist on a flooding of the reservoir
using a surfactant. The principal functions of surfactants are to reduce interfacial tension and
wettability of the rock reservoir, allowing the oil to flow. In Figure 1. 5 the injection process
of surfactant is illustrated.

Figure 1. 5. Surfactant injection process[9]

Usually, oil fields are flooded with water after primary and secondary extractions, and an
important portion of the oil will stay trapped in the reservoir after that (called residual oil).
Surfactants must be injected into the reservoir to reduce the quantity of residual oil. Thus,
this is a fundamental EOR method[10]. But, these techniques for recovering residual oil have
been in general not very satisfactory, because chemicals are lost by adsorption on reservoir
rocks, precipitation, and changes in rock properties, increasing the costs of the process. Many
factors affect the variation in the properties of the reservoir after injecting surfactants: the
chemical structure and mix of the surfactants, surface properties of the rock, composition of
the oil and reservoir fluids, nature of the polymers added and solution conditions such as
salinity, pH, temperature and mineralogy of the rock[11]. Therefore, a deep understanding
of these concepts is necessary to predict and mitigate the adverse effects.
Adsorption of surfactants has other applications too. For example, they can change many of
the solid properties where is adsorbed: hydrophobicity, surface charge, and others, which
govern interfacial processes such as flocculation/dispersion, flotation, wetting and
solubilization, detergency, and corrosion inhibition. In general, adsorption is governed by
some forces[12].

23
There is some debate on the effect of parameters affecting the adsorption of surfactants in
porous media. Thus, simplifications are made. But these simplifications may not always be
the best option. Sometimes the simplification of the model can lead us to imprecise results
or misinterpretations of our data. So, to know the relevance of those parameters is of critical
importance.
This work focuses in the analysis of key elements effect in porous media adsorption, to get a
better understanding and prediction of this kind of phenomena, diminishing costs and
operation times. To achieve this goal, the system of equations describing the adsorption of
surfactant in a porous medium for one and two dimensions was solve using a computational
tool. After the system was solved, comparison was made with experimental results from
literature.

1.1. Problem Statement


Crude oil recovery is of the utmost economic importance for producers around the world. As
oil resources become depleted in some countries enhance oil recovery (EOR) techniques start
being considered. Surfactant injection to alter the interfacial tension (IFT) between oil and
the reservoir fluid is an important EOR technique being applied in reservoirs. Modifying
mobility can also enhance the recovery of residual oil. However, the optimal recovery of oil
by surfactant injection requires a comprehensive understanding of the transport of
components and phases in porous media.

A common EOR process based on IFT is the injection of chemical components which are
able to modify either the interfacial tension between oil and the aqueous phase, or the
wettability of oil in the rock formation, or both. To achieve this goal, the injected
formulations may include surfactants, co-surfactants, co-solvents, polymers, sacrificial
agents, chelating agents, hydrotopes, and even ionic species. Therefore, the success of the
process will be strongly related to the appropriate transport of all these species through the
reservoir. The adsorption/retention interaction or selective adsorption of any of these species
over the rocks must be predicted with certainty to guarantee the success of the extraction
process. The complexities of the formulations and the intricacies of the rock structure make
this a challenging process. Thus, a mathematical model able to uncouple all these elements,
will lead to a better understanding of the adsorption mechanism.

24
1.2. Objective
Assess the importance of parameters, computational tools and adequate modeling in the
prediction of transport of components and phases in porous media, and to determinate its
importance for planning and understanding experiments.

1.2.1. Specific objectives


 Solve the System of equations describing the transport of two phases in a porous
medium for one dimension
 Solve the System of equations describing the adsorption of surfactant in a porous
medium for one and two dimensions
 Analyze the effect of the isotherm model and its parameters, the downstream
boundary condition with and without discontinuity, the slug size, the non-local
equilibrium, micro porosity and the heterogeneities in the permeability of the porous
media, in the adsorption of surfactant in the porous media.
 Analyze the displacement of phases in capillary conducts through a CFD tool and
compare it with a one-dimensional model
 Determine the relevance of the analyzed parameters involved in the adsorption of
surfactants in porous media.
 Explore the best suitable computational tools to solve the problems described.
 Determine common “bad practices” of experimentation in porous media, not only
designing the experiments but interpreting them.

1.3. Justification
The study of the flow of phases and components in porous media has acquired a considerable
interest in recent years in the industry of oil extraction. Thus, the search for new and more
efficient extraction methods has increased. Hence, this is essential to get the best profit
margins, not to mention that since 2014 and until the middle of 2015[13] the price of oil has
suffered a drastic decline, something that had not happened since 2009[13]. Thus, it is
necessary to look for cheaper and more efficient extraction methods that allow increasing the
profits of each extraction, if not, the oil reserves of some countries could be depleted. One of

25
the enhanced oil recovery methods (EOR) that have gained interest in recent years is the
extraction through the application of surfactants. Analysis, study, and development of this
technique focuses on the transport of phases and components in the porous media. But
because of the little progress that has been seen before in this field, there is some uncertainty
on the effects of some parameters in the prediction of this type of phenomena, which is of
vital importance for the simulation of the extraction processes that use these methods of
surfactants.

1.4. Organization of the thesis


This Work has the next structure:

I. Introduction

II. Theoretical framework, literature review and applicable tools

III. Methodology.

IV. Analysis of results

V. Conclusions and future work

In chapter 1, the global view of the transport in porous media is presented as well as a broad
view of the energetic situation of Mexico. Also, the different types of Enhanced oil recovery
process are mentioned and their importance in recent times. Finally, the importance and effect
of the adsorption of surfactant are mentioned. The general and specific objectives and
justification of this investigation are also presented.

Chapter 2 contains the fundamental theoretical concepts used in this work. Moreover,
includes historical background about the transport in porous media to get a better
understanding of its origin and its development in recent times. Also, contains an analysis
and discussion of the computational tools available to solve the system of equations
predicting the transport of surfactant in porous media, and the most suitable tool is selected.

Chapter 3 explains all the methodology followed to solve the system of equations describing
the transport two phases in a saturated porous media, the transport of surfactants in porous
media in 1-D, 2-D and axisymmetric cases, the experiments selected for comparison with our

26
model results, key elements selected to analyze (the isotherm model and its parameters, the
downstream boundary condition with and without discontinuity, the slug size, the non-local
equilibrium, micro porosity and the heterogeneities in the permeability of the porous media)
and the displacement of two phases in capillary conducts. The explanation of the
mathematical modeling includes the description of all the equations terms and its variables
as well as the dimensionless form of the equations. Also, includes the description of all the
boundary conditions used for every case analyzed.

Chapter 4 shows the results obtained and a comparison with others model and/or
experimental results. The analysis of the results is aided by the use of graphs of normalized
outlet concentration in function of dimensionless time. Also, different images of the
simulations were included. Description of the behavior observed in all figures is included.

Finally, chapter 5 presents the conclusions of this work and the future work section. The
conclusion made is based on the results obtained using our results. Moreover, the critical
findings of the elements analyzed are discussed in a concretely and accurate way, as well as
their relevance in the transport of surfactant and phases in porous media. Some suggestions
are made for the realization of experimental tests to assure the accuracy of the results
obtained. This section, also includes diverse points for future work, which may contribute to
this investigation. Moreover, these points could be useful to find more critical elements to
consider in the prediction of transport of surfactants in porous media, leading to more precise
results.

27
1.5. References
[1] EIA, “EIA projects 48% increase in world energy consumption by 2040,” Today in
Energy, 2016. [Online]. Available:
https://www.eia.gov/todayinenergy/detail.php?id=26212.

[2] IMPERIAL, “The importance of energy,” 2016. [Online]. Available:


http://www.imperialoil.ca/en-ca/company/about/importance-of-energy/the-
importance-of-energy.

[3] M. Overholt, “The Importance of Oil and Gas In Today’s Economy,” 2016. [Online].
Available: http://www.tigergeneral.com/the-importance-of-oil-and-gas-in-today-s-
economy/.

[4] IEA, “Mexico Energy Outlook,” p. 129, 2016.

[5] M. Masoud, “Comparing Carbon Dioxide Injection in Enhanced Oil Recovery with
other Methods,” vol. 2, no. 2, pp. 1–11, 2015.

[6] U.S. Department of Energy, “Enhanced Oil Recovery,” 2010. [Online]. Available:
https://energy.gov/fe/science-innovation/oil-gas-research/enhanced-oil-recovery.

[7] A. Amarnath, “Enhanced Oil Recovery Scoping Study,” Epri, pp. 1–148, 1999.

[8] C. Udumbasseri, “EOR Methods,” 2015. [Online]. Available:


https://www.slideshare.net/ChandranUdumbasseri1/eor-methods-50680035.

[9] Elnusa, “Enhanced Oil Recovery (EOR),” Drilling and oilfield services, 2016.
[Online]. Available: http://www.elnusa.co.id/idn/product-services/drilling-oilfield-
services/enhanced-oil-recovery-eor/.

[10] J. J. Sheng, “Status of surfactant EOR technology,” Petroleum, vol. 1, no. 2, pp. 97–
105, 2015.

[11] P. Somasundaran and L. Zhang, “Adsorption of surfactants on minerals for wettability


control in improved oil recovery processes,” J. Pet. Sci. Eng., vol. 52, no. 1–4, pp.
198–212, 2006.

28
[12] R. Zhang and P. Somasundaran, “Advances in adsorption of surfactants and their
mixtures at solid/solution interfaces.,” Adv. Colloid Interface Sci., vol. 123–126, no.
Special Issue in Honor of Dr. K. L. Mittal, pp. 213–229, Nov. 2006.

[13] Energy Information Administration, “Annual Energy Outlook 2015 Table: Oil and
Gas Supply,” Annual Energy Outlook 2015, 2015. [Online]. Available:
http://www.eia.gov/beta/aeo/#/?id=14-AEO2015&region=0-
0&cases=ref2015&start=2014&end=2040&f=A&linechart=14-AEO2015.5.

29
Chapter II: Theoretical framework,
literature review and applicable tools.
2.1. Theoretical framework
Many concepts are needed to obtain a better understanding of the model and parameters
analyzed in this work. Thus, the critical of concepts of this work are described in the next
paragraphs.

2.1.1. Adsorption
Adsorption may be defined as: The increase in quantity of a component at an interface. In
most usage is positive, but it can be negative. Adsorption may also denote the process of
components accumulating at an interface.[1]

Adsorption may not be confused with absorption. The last is when a component gets inside
a phase and adsorption is when the componenet gets in the interface of two phases. This is
depicted in Figure 2.1.

Figure 2.1. Difference between absorption and adsorption[2]

The surface structure of solids has been an important subject of scientific study for much of
the twentieth century, initially because of its significance in understanding the behavior of
solid catalysts. The early work on the subject was, therefore, heavily influenced by the

30
adsorption behavior of various gases on solids, an issue which was given a dramatic impetus
by the introduction of gas warfare in World War 1. There were, however, many limitations
in these studies, due firstly to the very small numbers of molecules involved in surface
reactions and hence the high sensitivity needed for detection. The other problem was
contamination. It was not until the development of ultra-high vacuum technology that it
became possible to deal with truly clean surfaces on a more or less routine basis. Using the
well-established techniques of the electron microscope and a variety of special spectroscopic
methods, the study of the details of surface atomic structure expanded to become a highly
specialized field from the late 1960s onwards. The scientific, technological, and commercial
driving forces for these developments came from many sources: new catalyst requirements,
harder wearing surfaces and better lubrication, fracture resistant ceramics, chemical sensors,
magnetic tape and other memory devices, dry paper copying, and a host of others, including
the ever increasing demands of the computer microchip industry. The detailed study of these
solid surfaces can involve the application of a barrage of techniques to a carefully prepared
surface under the most stringent purity conditions.[3]

Many colloidal solids would, however, change structure irreversibly under high vacuum and
our primary interest in hydrocolloids also makes much of the high vacuum work of only
marginal relevance, because of the overwhelming influence of water on the structure of the
solid-water interface. Water does not merely orient itself at the interface, it often interacts
chemically with the surface atoms and may even penetrate to some extent into the surface
layers. We will, therefore, concentrate on techniques which can throw some light on those
processes and leave the many other fields for the specialist monographs. Fortunately, there
still are a significant number of spectroscopic and microscopic techniques, as well as
theoretical modeling and computational methods which can throw light on the solid-liquid
interfaces of most interest to us.[3]

Adsorption at solid-liquid interfaces

When adsorption occurs from (aqueous) solution onto a solid, liquid, or gaseous surface, it
is important to recognize that the adsorbate is competing against the solvent (water)
molecules for a place at the interface. Substances which lower the surface tension are said to
be ‘surface active’; they accumulate at the interface because in so doing they reduce the total

31
energy of the system. The result is an adsorption isotherm, an equation of very considerable
importance for the study of the surface properties of liquid- vapor and liquid-liquid
systems[4].

Unfortunately, for the solid-liquid systems with which we are principally concerned, a
particular isotherm is needed, due to the problems related to the measurement of the surface
tension (or surface free energy) of solids. Indeed, in this case, the equation is used oppositely,
to obtain estimates of changes in the surface energy from independent measurements of the
amount of adsorption. Because of the large surface areas which commonly occur in colloidal
systems, the amount of adsorption can often be determined directly from the depletion of the
adjoining liquid phase, especially if the adsorbate is present in fairly dilute solution.
Furthermore, it often turns out that we are most interested in the adsorption properties of
relatively minor components of the liquid phase, especially if they are surface active (i.e.,
strongly adsorbed). Such substances are widely used to modify and control the equilibrium
and kinetic (transport) properties of colloidal dispersions.[3] Figure 2.2 is an illustration on
how molecules are adsorbed in a solid surface, creating a layer on the solid.

Figure 2.2. Particles in an aqueous media being adsorbed in a solid surface[5]

32
2.1.2. Isotherm of adsorption
As mentioned before, the isotherms of adsorption are of great importance when the
adsorption phenomena are analyzed. There are many different types of isotherms, each one
with their particular usage.

As noted above, adsorption of a solute from dilute solution onto a solid surface must be seen
as an exchange process between the solute in the solution and one or more molecules of the
solvent at the surface. That may occur because of a positive affinity of the solute for the
surface or as a result of the rejection of the solute by the solvent. In the case of surface active
agents, it is often a combination of both such effects. The fact that the adsorption occurs from
a dilute solution makes its detection and measurement relatively easy, especially as the area
for adsorption can usually be made large enough to produce a significant change in the solute
concentration.

In setting up possible theoretical isotherms for the adsorption process we need to be aware
of the influence of the surface solvent on the mobility of surface adsorbed solute molecules,
the interactions between the solute and solvent and between solute molecules, both in the
bulk and at the surface and, of course, all the problems associated with surface roughness,
porosity, and heterogeneity, as mentioned in respect of the gas-solid interface. It should be
noted, however, that there are some compensations: water is such a very active molecule both
in its interactions with itself and with other materials that it tends to smooth out some of the
heterogeneities on solid surfaces.

Lyklema6 distinguishes the six most common isotherm types. The linear isotherm is usually
only observed over a very short concentration range since it will inevitably give way to some
curvature as the surface fills with adsorbate. The Langmuir isotherm is commonly observed
and does not always mean that the solute is behaving according to the postulates of
Langmuir’s theory. Freundlich’s isotherm is also often observed and can be explained by a
logarithmic variation in the energy of surface sites. The high-affinity isotherm is usually
regarded as a variation of the Langmuir in which the initial region shows such strong
adsorption that none of the adsorbates remains in solution until the surface nears saturation.
The sigmoid isotherm suggests that some nucleation process must first occur either in

33
solution or, more likely, on the surface, before adsorption can proceed. The step isotherm is
sometimes interpreted regarding bilayer formation but may also be due to the re-orientation
of the surface layer to accommodate more material (e.g., the change from horizontal to
vertical orientation of a long chain molecule). [3]

Isotherms are studied to understand the mechanisms and the forces involved in the adsorption
process and to gain an understanding of the structure of the adsorbed species. Control of the
extent of adsorption and the orientation of the adsorbate can be useful for controlling the
transport and equilibrium properties of a suspension.

Next, the isotherms analyzed in this work are described briefly.

-Langmuir

Langmuir adsorption isotherm was created to describe gas–solid-phase adsorption onto


activated carbon, has traditionally been used to quantify and compare the performance of
different bio-sorbents. In its formulation, this empirical model assumes monolayer adsorption
(the adsorbed layer is one molecule in thickness), with adsorption can only occur at a finite
(fixed) number of definite localized sites, that are identical and equivalent, with no lateral
interaction and steric hindrance between the adsorbed molecules, even on adjacent sites. In
its derivation, Langmuir isotherm refers to homogeneous adsorption, which each molecule
possess constant enthalpies and sorption activation energy (all locations possess an equal
affinity for the adsorbate), with no transmigration of the adsorbate in the plane of the
surface[6]. The Langmuir isotherm has the next form:

𝐴 𝐶𝐴
𝐹(𝐶𝐴 ) = (2.1)
1 + (𝑘 𝐶𝐴 )

-Redlich-Peterson

Redlich–Peterson isotherm is a hybrid isotherm featuring both Langmuir and Freundlich


isotherms, which incorporate three parameters into an empirical equation. This model
presents a linear dependence on concentration in the numerator and an exponential function
in the denominator to represent adsorption equilibria over a wide concentration range, that
can be applied either in homogeneous or heterogeneous systems due to its versatility.

34
Typically, a minimization procedure is adopted in solving the equations by maximizing the
correlation coefficient between the experimental data points and theoretical model
predictions with solver add-in function of the Microsoft Excel. In the limit, it approaches
Freundlich isotherm model at high concentration (as the exponent β tends to zero) and is in
accordance with the low concentration limit of the ideal Langmuir condition (as the β values
are all close to one)[6]. The Redlich-Peterson isotherm has the next form:

𝐴1 𝐶𝐴 𝐴2 𝐶𝐴
𝐹(𝐶𝐴 ) = 𝑛
+ (2.2)
1 + (𝑘1 𝐶𝐴 ) 1 1 + (𝑘2 𝐶𝐴 )𝑛2

2.1.3. Interfacial tension


As the name implies, surface tension (denoted by γ) is a force that operates on a surface and
acts perpendicular and inward from the boundaries of the surface, tending to decrease the
area of the interface. To illustrate this, a simple apparatus can be used. In fact, other methods
are used, but the arrangement of Figure 2.3 has the advantage of simplicity and serves to
illustrate how the tension in the liquid surface is indeed measured by γ. The figure represents
a loop of wire with one movable side on which a film could be formed by dipping the frame
into a liquid. The surface tension of a stretched film in the loop will cause the slide wire to
move in the direction of decreasing film area unless an opposing force F is applied. In an
actual apparatus, the friction of the slide wire might be sufficient for this. In an idealized,
frictionless apparatus like that in Figure 2.3, the force opposing γ could be measured. The
force evidently operates along the entire edge of the film and will vary with the length P of
the slide wire. Therefore it is the force per unit length of edge that is the intrinsic property of
the liquid surface. Since the film in Figure 2.3 has two sides, the surface tension as measured
by this apparatus equals[7]:

𝐹
𝛾= (2.3)
2𝑙

35
Equation (2.3) defines de units defines the units of surface tension to be those of force per

length or Nm-1 in the International System of Units (SI) or dynes cm-1 in the Centimetre–

Gram–Second System (CGS).

Figure 2.3. A wire loop with a slide wire on which a liquid film might be formed and stretched
by an applied force F[7].

2.1.4. Contact angle


A quantity that is closely related to surface tension is the contact angle. The contact angle θ
is defined as the angle (measured in the liquid) that is formed at the junction of three phases,
for example, at the solid-liquid-gas junction as shown in Figure 2.4. Although the surface
tension is a property of the two phases that form the interface, θ requires that three phases be
specified for its characterization, as mentioned above. The above definition of contact angle
is, however, highly simplified, but is enough to get a good understanding of the concept.[7]

36
Figure 2.4. Profile of a three-phase (solid, liquid, gas) boundary that defines the contact angle
θ.[7]

2.1.5. Surfactants
Surfactants are among the most useful products of the chemical industry, appearing in such
diverse products as the motor oils we use in our automobiles, the pharmaceuticals we take
when we are ill, the detergents we use in cleaning our laundry and our homes, the drilling
muds used in prospecting for petroleum, and the flotation agents used in benefication of ores.
The last decades have seen the extension of surfactant applications to such high-technology
areas as electronic printing, magnetic recording, biotechnology, micro-electronics, and viral
research.

A surfactant (a contraction of the term surface-active agent ) is a substance that, when present
at low concentration in a system, can be adsorbed onto the surfaces or interfaces of the system
and alter to a marked degree the surface or interfacial free energies of those surfaces (or
interfaces). The term interface indicates a boundary between any two immiscible phases; the
term surface denotes an interface where one phase is a gas, usually air.

The surfactant usually has a head and a tail, where the tail as a strong affinity to a substance
and the head to another substance. The most common case is a surfactant being hydrophobic

37
and hydrophilic. So, is because of this property that these molecules are adsorbed in the
interface of two phases (usually liquid-liquid), as shown in Figure 2.5. Although, there are
many types of surfactants with different configurations.

Figure 2.5 Molecule of surfactant (Left) and surfactant-stabilized droplet (right)[8].

A surfactant is, therefore, a substance that at low concentrations is adsorbed at some or all of
the interfaces in the system and significantly changes the amount of work required to expand
those interfaces. Surfactants usually reduce interfacial free energy rather than to increase it,
although there are occasions when they are used to increase it.[9]

2.1.6. Porosity and its classification


Porosity may be defined as the percentage of pore volume or void space or that volume within
rock that can contain fluids. Porosity can be a leftover of deposition (space between grains
that were not compacted together completely) or can originate through alteration of the rock
(such as when grains are dissolved from sandstones). Porosity can be created by the
development of fissures, in which case it is called fracture porosity. Effective porosity is the
interconnected pore volume that contributes to fluid flow in a reservoir. Isolated pores are
excluded. Total porosity is the sum of void space in the rock included isolated pores. Thus,
effective porosity is usually less than total porosity. Shale gas reservoirs tend to have high
porosity, but the alignment of platy grains such as clays makes their permeability very
low.[10] In Figure 2.6 a graphical representation of high and low porosity is shown.

38
Figure 2.6 Example of a porous media. High porosity – Larger spaces (Left) and low porosity
- Smaller spaces (right) [11]

The pore systems of solids are of many different types. The pores may vary greatly in size
and shape within a given solid, and between solids. A point of special interest for many
purposes is the pores width, e.g., the diameter of a cylindrical pore, or the distance between
pores. A convenient classification of pores according to their average width originally
proposed by Dubining and now officially adopted by the International Union of Pure and
Applied Chemistry[12] is summarized in Table 2.1. The classification is based on that each
of the size ranges corresponds to characteristic adsorption effects as manifested in the
isotherm. In micropores, the interaction potential is higher than in wider pores because the
proximity of the walls and the amount adsorbed is correspondingly enhanced. In mesopores,
capillary condensation takes place. In the macropores region, the pores are so wide that it is
virtually impossible to map out the isotherm in detail because the relative pressures are very
close to the unity. The borderlines between the different classes are not hard and fast,
depending as they do both in the shape of the pores and on the nature of the adsorptive
molecule. Thus, the highest value of w at which the enhancement of adsorption occurs, i.e.,
the upper limit of the micropore range, will vary from one adsorptive to another. It frequently
happens that the micropores effect, the enhancement of interaction potential and the resultant
adsorption, ceases to appear when the value of w is still below the beginning of the hysteresis
loop.

39
Table 2.1. Classification of pores according to their width.[12]

Porosity type Width


Micropores Less than 2 nm
Mesopores Between 2 and 50 nm
Macropores More than 50 nm
Within recent years, the micropores range has therefore been subdivided into the very narrow
pores or ultra-micropores, where the enhancement effect is found, and supermicropores
which fill the gap between the ultramicropore and the mesopore ranges. It is perhaps worth
emphasizing that, amongst solids as a whole, a broad and continuous range of pore sizes is
to be found, from macropores through mesopores and micropores, to sub-atomic "pores".
There is an important reminder of the danger of unconsciously assuming that nature arranges
the pore sizes in solids to suit the particular scientific instruments and methods that happen
to have been devised so far.[12]

2.1.7. Wettability and capillarity


When two fluids are present in a pore space, usually one of them is preferentially attracted to
the surface of the solid structure. This fluid is called the “wetting fluid” (or phase), while the
other fluid phase is called “non-wetting fluid”. Usually, hydrophilic porous media is more
widespread in nature. The term wetting phase will be used as a synonym for water and the
term non-wetting phase as a synonym for air. Immiscible fluids are separated by a well-
defined interface, which, if observed at a scale much larger than the molecule size, can be
considered infinitely thin. Since the cohesion between fluid molecules at one side of the
interface is different from that on the other side, the interface is characterized by some surface
energy (or surface tension), which is a measure of the forces necessary to change its shape.
The existence of the surface tension causes a difference in the equilibrium pressures of air
and water, divided by a curved interface, due to unbalanced tangential forces at the surface.
The pressure drop between the air and water phases can be calculated from the Laplace
equation[13]:
1 1
∆𝑝 = 𝑝𝑎 − 𝑝𝑤 = 𝜎𝑎𝑤 ( + ) (2.4)
𝑟𝑐1 𝑟𝑐2

40
Where the subscripts a and w denote the air and water phases, respectively, σaw is the air-
water surface tension of the interface, and rc1 and rc2 are the curvature radii of the interface.
The value of air-water surface tension is 0.0726 Nm−1 for a given temperature of 20 ◦C, and
it may decrease with increasing temperature[13].

Figure 2.7. Surface tension effects: a) equilibrium position of the fluid-fluid interface near the
solid surface, b) rise of the wetting fluid in a capillary tube, c) pendular ring around the
contact point of two solid grains.[13]

The pressure will always be smaller in the fluid occupying the concave side of the interface.
When no external forces are present, the interface of a droplet of a fluid contained in another
fluid tends to create a spherical shape, minimizing the surface energy. In the presence of a
solid surface the shape of the interface is determined by the relation of the surface tensions
between all three phases (Figure 2.7 a):
𝜎𝑎𝑤 𝑐𝑜𝑠𝜓 = 𝜎𝑠𝑎 − 𝜎𝑠𝑤 (2.5)

Where σsa and σsw are the surface tension between the solid phase and air and water,
respectively, and ψ is called the wetting angle. When a fluid can wet perfectly a surface, ψ =
0°, i.e., the fluid tends to spread evenly over the whole solid surface. In the other hand, when
a fluid can't wet a surface, ψ = 180°, resulting in the formation of spherical droplets on the
surface. The surface tension effect can be observed in the origin of small tubes (Figure 2.7
b). The wetting fluid is attracted by the tube wall, and a meniscus is formed between water
and air. The pressure drop across the interface is called capillary pressure and can be
calculated for a cylindrical tube as:

41
2 𝜎𝑎𝑤 𝑐𝑜𝑠𝜓
∆𝑝 = 𝑝𝑐 = 𝑝𝑎 − 𝑝𝑤 = (2.6)
𝑟𝑐

Where rc is the tube radius. Considering that the surface tension value between air and water
corresponds to the temperature of 20°C, the tube material is perfectly wettable (ψ = 0°), air
is at constant atmospheric pressure everywhere, and water is incompressible, one obtains the
well-known formula for the height of the capillary rise:
1.5𝑥10−5
ℎ𝑐 = (2.7)
𝑟𝑐

Where both hc and rc are in meters. Equation (2.7) is usually applied to approximate the height
of capillary rise in a porous media characterized by small wettability angles.
There exists a relationship between Darcy-scale capillary pressure and water saturation, it is
known under some names, such as the capillary function, suction function, retention function,
or soil water characteristic function, and it is of great relevance in the prediction of saturation
in porous media.
The geometry of pores in natural porous media is much more complex. Thus, the
representation of pore system as a bundle of capillary tubes does not hold in many situations,
and more complex configurations of air and water in the pore space are encountered, which
will be discussed on the example of a granular porous medium[13].

Figure 2.8. Spatial configurations of water and air in an unsaturated granular porous
medium: a) adsorbed regime, b) capillary pendular regime, c) capillary funicular regime, d)
occluded air bubble regime, e) fully saturated regime.[13]

42
Considering that water molecules are preferentially attracted to the surface of the solid phase,
they can be adsorbed from the vapor present in the pore air. Thus, small amounts of water
are always present in the form of thin films covering the surface of the solid skeleton (Figure
2.8 a). This layer thickness depends on the strength of molecular interactions between solid
and water and on the relative air humidity. Due to the bonding of the adsorbed water to the
solid surface, in many practical problems, it is considered as immobile. As the amount of
water in a porous medium increase, the cohesion forces attract the water to the water adsorbed
at the solid surface, but on the other hand, tends to minimize the area of the interface. This
kind of water is known as capillary water and occurs initially in sharp corners of the pores.
If the porous structure is made of grains or something similar, water forms pendular rings
around the contact points, (Figure 2.8 b and c). The corresponding water-air configuration is
known as the pendular stage. Pendular water is less tightly bound to the solid phase, but it
occurs in isolated regions and does not form continuous flow paths so that it can be
considered as macroscopically immobile. The pressure drop at the interface between
pendular water and pore air can be theoretically calculated from Eq. (2.6), assuming a
negative value for rc1 and a positive value for rc2 (Figure 2.8 c). If more water is added to the
system, the regions occupied by pendular water coalesce and continuous thicker films are
formed along the pore walls. At this stage, known as the funicular stage, the flow of liquid
water is possible (Figure 2.8 c). For all three water configurations mentioned above, air
occupies continuously the central part of the pores. As the amount of water in the system
increases further, the water films become thicker, and pores can be filled with water at the
points where their cross-section is smaller. The air phase loses its continuity, and no
macroscopic air flow is possible. This is called the occluded air bubbles stage (Figure 2.8 d).
With time the air can dissolve in water, and full water saturation is reached (Figure 2.8 e)
[13].

43
2.1.8. Rock saturation and volume fraction

Figure 2.9. Pore-scale representative elementary volume[13]

Description of a multiphase porous medium at the Darcy scale indicates that the relevant
physical quantities defined at a point x represent averages over a pore-scale representative
elementary volume (REV) associated with that point (Figure 2.9). At the Darcy scale, a point
can be occupied by all three phases at the same time; this is represented by the concepts of
volume fractions and saturation. The volume fraction of phase α is the ratio between the
volume of the part of the REV occupied by phase α and the total volume of the REV:
|𝑈𝛼 |
𝜃𝛼 = (2.8)
|𝑈|

Porosity can be calculated as the sum of the volume fractions of the two pore fluids:
|𝑈𝑤 | + |𝑈𝛼 |
𝜑 = = 𝜃𝛼 + 𝜃𝑤 (2.9)
|𝑈|

Also, the saturation of each phase, can be defined as the fraction of the pore space occupied
by a given fluid:
𝜃𝛼
𝑆𝛼 = (2.10)
𝜑

44
The sum of the air and water saturations must be equal to one:
𝑆𝛼 + 𝑆𝑤 = 1 (2.11)

Each saturation vary from 0 to 1. But, the range of variability can be smaller in some
applications. For instance, when a fully water-saturated system is drained, at some point the
mobile water (water not adsorbed) loses its continuity, and the liquid flow will not be
possible. The value of the water saturation is called residual water and is denoted by Srw.
However, it should be noted that the value of water saturation can be further decreased by
natural evaporation or oven drying. Similarly, during imbibition, it is usually not possible to
achieve a complete water saturation, as a part of the pores will always be occupied by some
isolated air bubbles. The corresponding residual air saturation is denoted as Sra. However, the
water saturation can increase above the value of 1-Sra, for example, if the air is compressed
or dissolves in water. Therefore, the residual saturations must be considered as problem-
specific parameters and not as material parameters.
For practical purposes, the fluid saturations are often normalized with respect to the range of
values occurring in the problem under consideration. The resulting normalized (effective)
saturations are defined as:
𝑆𝛼 − 𝑆 𝑚𝑖𝑛
𝑆𝑒𝛼 = (2.12)
𝑆𝛼𝑚𝑎𝑥 − 𝑆𝛼𝑚𝑖𝑛

Where 𝑆𝑎𝑚𝑎𝑥 and 𝑆𝑎𝑚𝑖𝑛 are the maximum and minimum saturation values occurring for a
given problem.
In soil physics and hydrology it is more common to quantify the relative amount of fluid
phases in soil using the volumetric fractions θα. If the compressibility of the solid skeleton is
neglected (φ=const), the volumetric phase contents are uniquely defined by phase saturations.
In field conditions the volumetric fraction of water varies between the residual water content:
𝜃𝑟𝑤 = 𝜑 𝑆𝑟𝑤 (2.13)

and the so-called saturated water content:


𝜃𝑠𝑤 = 𝜑(1 − 𝑆𝑟𝑎 ) (2.14)

The latter value refers to the state of maximum attainable water saturation. Equivalent limit
values can also be defined for the volumetric air content.[13]

45
2.1.9. Permeability
The permeability may be defined as the measurement of rock's ability, to transmit fluids,
typically measured in darcies or millidarcies. The term was described by Henry Darcy, who
showed that the general mathematics of heat transfer could be modified to describe fluid flow
in porous media adequately. Formations that transmit fluids easily, such as sandstones, are
described as permeable and tend to have large and well-connected pores. Impermeable
structures, such as shales and siltstones, tend to have finer grains or to have few
interconnected pores. Absolute permeability is the permeability when a single fluid, or phase,
is present in the porous structure (rock). Effective permeability is when the rock exhibits a
preference for a particular fluid, increasing its flow when other immiscible fluids exist in the
reservoir. The saturation of the fluids and the reservoir nature, influence the effective
permeability. Relative permeability is defined as the ratio of effective permeability of a fluid
at an absolute permeability. If a single fluid is present in a rock, its relative permeability is 1.
Calculation of relative permeability permits the comparison of the different fluids abilities to
flow in the presence of each other because the presence of more than one fluid inhibits
flow[14].

2.1.10. Relative permeability


Since the relative permeability of fluid phase α varies from krα = 0 for Sα = Srα to krα = 1 for
Sα = 1, it can be conveniently represented by a function of the normalized saturation of each
phase, as given by Eq. (2.12), assuming 𝑆𝛼𝑚𝑖𝑛 = Srα and 𝑆𝛼𝑚𝑎𝑥 =1. However, to simplify the
model formulation, the capillary function and permeability functions for both fluids are
typically defined with respect to the water saturation normalized in the range between 𝑆𝑤𝑚𝑖𝑛 =
Srw and 𝑆𝑤𝑚𝑎𝑥 = 1− Sra. In such a case, the relative permeability equals unity for the actual fluid
saturation smaller than one. To keep the physical consistency of the model, ks is interpreted
as the maximum permeability attainable for the considered problem, which can be different
for each phase, and is smaller than the permeability tensor for single-phase flow. In soil
hydrology and soil physics the maximum attainable value of the water conductivity:

46
(𝜌𝑤 𝑔)
𝑘𝑠𝑤 = 𝑘𝑠 (2.15)
𝜇𝑤

Equation (2.15) is often referred to as the saturated hydraulic conductivity. However, it


corresponds to the state of apparent saturation, with the corresponding volumetric water
content θsw smaller than the porosity φ. The formulae for relative permeability presented
below can be used in conjunction with various definitions of the normalized water saturation
Sew. Simple power-type relationships between the normalized saturation and the relative
permeabilities are often postulated, e.g.:
𝑘𝑟𝑤 = (𝑆𝑒𝑤 )𝜂𝑤 (2.16)

𝑘𝑟𝑎 = (1 − 𝑆𝑒𝑤 )𝜂𝑎 (2.17)

Where the exponents ηw and ηa are fitting parameters. Relationships of this type can be
obtained on the theoretical basis, for simple models of laminar flow in bundles of capillary
conducts. As it is the wetting fluid, water tends to move along the solid phase surface and
preferentially fills smaller pores. Therefore, at the same saturation of each fluid, the
resistance of the medium to the flow of water is larger than the resistance to the flow of air.
Thus, the exponent ηw is typically larger than ηa. [13]

2.2. Literature review


The transport of solute in porous media may be analyzed using different approaches, all of
these take into consideration that pore space or porosity is divided into different classes,
according to their size. For instance, Bai and Roegiers[15] have indicated that different
transport processes take place within the macro, meso, and micro pores.

The porosity distribution and the kind of structures inside the porous medium have been
identified by a number of authors who have made efforts to explain and model these
structures. A number of approaches with different degrees of simplification have been used
in different fields to model flow and the transport of components inside porous media. These
include: Studies of flow in naturally-fractured reservoirs [16]; the transport of inert species
in porous media [17]; analysis of the transport of solutes that undergo sorption [18]; the
transport of fluids in miscible displacement[19], [20]; the measurement and modeling of the
transport of different species during immiscible displacement [21], [22], and the effect of

47
mineralogy has been studied by Liu [23]. However, these studies and models were limited
by the computational capabilities of the time, and by the simplifications used to reproduce
and explain the experimental results. For instance, Bai and Roegiers [15] coupled the
boundary condition with the mass balance in the meso porous region and assumed the same
concentration for both the meso porous region and the solid surface. As a result, their model
neglects diffusional flow in the micro porous region.

The transport of solute in a porous medium is also related to solute adsorption on the solid
surface. The adsorption of compounds over the surface has been widely studied. This
mechanism can be simplified using an isotherm equation. Although adsorption is a time-
dependent process, the usage of the instantaneous equilibrium condition can offer a helpful
simplification. The concept of instantaneous equilibrium and the distribution throughout the
porous medium have been studied by Schwartz et al. [24], who assumed a linear relationship
between adsorbed species concentration and bulk concentration. However, they neglected
adsorption and diffusion in the micro pores, which may become important at these small
scales. Van Genuchten and Wierenga[25] adopted the same principle of instantaneous
adsorption, using the Freundlich isotherm in a porous medium. They also took into account
boundary conditions applicable to semi-infinite media. Lapidus and Amudson [26] analyzed
the effect of instantaneous adsorption comparing it with another approach, in which they
incorporated the rate of adsorption for a porosity model. The effect of boundary conditions
plays a fundamental role as well as discussed by Balhoff et al. [27]. However, their model is
simplified in order to be solved analytically. Differences at low flow rates were observed.
They also discussed the problem of semi-infinite boundary conditions assuming a single type
of porosity. Sardin et al. [28] presented a broader spectrum of different mathematical models
and simulated the effect of having different geometries representing the micro pores. They
also compared the effect of instantaneous equilibrium or adsorption/desorption kinetics for
the adsorbed species, but no experiments were included in their study.

In recent years, increasing computational power has enabled the usage of more complex
models. As a result, models in several fields have included non-equilibrium approaches or
have combined nano, micro, and macro pores [29], [30]. However, to get a better
understanding of the transport of solutes in porous media, complex models are needed. These

48
models should include both, the pore distribution of real rock formations and the different
mechanisms of adsorption. Diverse types of rock and geological formations are found in the
recovery of crude oil. A group of surfactants with different composition have potential use
in modifying the IFT, creating different adsorption mechanisms. A robust model is needed
to cover the full spectrum of potential formulations.

In this work, we use earlier ideas about pore distribution to build a model that involves three
types of pores (macro, meso, and micro pores). Modeling results were compared with
experimental data before additional simulations were carried out. Moreover, the effects of
diverse parameters were analyzed using additional results.

2.3. Applicable CFD tools


Computational Fluid Dynamics (CFD) is a fluid mechanic's divisions that uses numerical
methods and algorithms to solve and analyze problems about the flow of fluids. Computers
are used to perform millions of calculations required to simulate the interaction of liquids
and gases with complex surfaces. Even using simplified equations and with high-
performance supercomputers, only approximate results can be achieved in many cases.
Continuous research, however, allows the incorporation of software that increases the speed
of calculation as well as increasing accuracy, while allowing to analyze increasingly complex
situations such as transonic fluids and turbulent flows. The verification of the data obtained
by CFD is usually done in wind tunnels or other physical models at scale.
One of the founders of this method of analysis was Suhas Patankar who is currently an
emeritus professor at the University of Minnesota.
The method consists of discretizing a region of space by creating what is known as spatial
mesh, dividing a region of space into small volumes of control. Then the discretized
conservation equations are solved in each of them so that an algebraic matrix in each cell is
solved iteratively until the residue is sufficiently small.[31]
There are several methods for solving equation systems in CFD, but the two methods mainly
used are the finite element method and the finite volume method. The first is widely used in
structural physics, while the latter method has better applications in fluid mechanics.

49
2.3.1. Finite element method
The finite element method (FEM) is a numerical method for the approximation of solutions
for very complex partial differential equations used in various engineering and physics
problems.
The FEM is intended for use in computers. It allows solving differential equations associated
with a physical problem on complicated geometries. The FEM is employed in the design and
improvement of industrial products and applications, as well as in the simulation of complex
physical and biological systems. The variety of problems where it can be applied has grown
enormously, being the basic requirement to know boundary and initial conditions in advance.
[32]
Four stages compose the development of a finite element algorithm to solve a problem
defined by differential equations and boundary conditions:
1. The problem must be reformulated in a variational way.
2. The domain of independent variables (usually a spatial domain) must be divided into
subdomains, called finite elements. Then, we construct a vector space of finite
dimension, called finite element space. The approximate numerical solution obtained
by finite elements is a linear combination in said vector space.
3. We get the projection of the original variational problem on the finite element space
obtained. With this, we obtain a system with a finite number of equations, but with
a large number of unknown variables. The number of unknown variables will be
proportional to the dimension of the finite element space obtained. With a smaller
dimension of the finite element space, a better numerical approximation is obtained.
4. The last step is the numerical calculation of the solution for the equation’s system.
The previous steps allow the creation of a differential calculus problem in a linear algebra
problem. This issue arises on a vector space of non-finite dimension, but it can be solved
approximately by finding a projection on a subspace of finite dimension, and therefore, with
a finite number of equations (although the number of equations will be elevated, typically
thousands or even hundreds of thousands). The finite element discretization helps to construct
a simple projection algorithm, also making the solution by the finite element method accurate
in a finite set of points. These points usually coincide with the vertices of the finite elements.

50
For the resolution of the enormous system of algebraic equations, the conventional methods
of linear algebra can be used in spaces of finite dimension.
Advantage
 Model an irregular shape fairly quickly
 Models bodies composed of several different materials because the same elements
are evaluated individually.
 To vary the size of the elements to make possible the use of small elements where
necessary.
 Modifies finite elements relatively easily and cheaply.

Disadvantages
 Requires the application of weighted residues
 Not suitable for infinite problems
 Requires mesh refining
 Domain and contour must be modeled separately

2.3.2. Finite volume method


The method of finite control volumes allows to discretize and numerically solve differential
equations. First, consider a mesh of discretization for the fluid space. Around each point of
this mesh, a control volume is constructed. Thus, the total volume of fluid becomes equal to
the sum of the control volumes considered. The differential equation to be solved is integrated
on each control volume, which results in a discretized version of the equation. To perform
the integration, it is necessary to specify variation profiles of the dependent variable between
the mesh points, to be able to evaluate the resulting integrals. The main property of the
resulting system of discretized equations is that the solution obtained satisfies exactly the
conservation equations considered, regardless of the size of the mesh.
By putting the equation to be solved at each point of the mesh, a system of algebraic equations
is obtained. If the coefficients of this system of equations are functions of the state variables,
then the resulting algebraic equation system is nonlinear. In that case, the way to solve the
system is by iterations. It starts by making an initial assumption of the values, with which the

51
values of the coefficients are determined. Then, solving the linear problem with constants
and known coefficients. Thus, the assumed value of the coefficients can be revised to solve
a new linear problem with corrected coefficients. After successive iterations, the solutions of
the linear problem converge, and the solution of the original nonlinear problem is obtained.
The four basic rules established in previous sections makes possible to ensure that this
iterative procedure effectively converges.[33]
Advantage
 It is robust and efficient in solving conservation equations
 Can handle complex geometries
 It can give adequate results with few elements in the mesh.
Disadvantages
 It presents a relatively low precision of the results
 Difficulties may arise in trying to solve structural physics systems
 May have meshing difficulties in complex geometries

2.3.3. Commercial software available


There are many usefully CFD software in the present as COMSOL, Open Foam, Star CCM,
ANSYS Fluent, etc. For this thesis we will be comparing COMSOL and ANSYS Fluent,
which are two of the most used CFD software in the ITESM of Monterrey. COMSOL is a
CFD software based on the finite element method (FEM), while ANSYS is based on finite
volume method.

As mentioned above, FEM is especially useful when dealing with structural phenomena,
while the finite volume method is a more convenient technique when solving fluids flow
phenomena. Thus, COMSOL is more suitable for structural physics problems while ANSYS
has a better performance in fluids mechanics problems (higher calculation speed).
Nevertheless, COMSOL has an important advantage when compared with ANSYS. It can
solve any equation or system of equations introduced by the user using a mathematical
module included in the software, in any number of dimensions (1D, 2D, 3D and
axisymmetric). Though, ANSYS can only solve the equations predefined by the software.
Thus, ANSYS may not always be the most suitable software for a problem that involves

52
fluids mechanics; it may depend on the equations included in the software, which can be an
important limitation when dealing with complex phenomena.

53
2.4. References
[1] J. D. Shosa and L. L. Schramm, “Surfactants: Fundamentals and Applications in the

Petroleum Industry,” Palaios, vol. 16, no. 6, p. 614, Dec. 2001.

[2] Marc, “What’s the Difference Between Absorption and Adsorption?,” 2013. [Online].

Available: http://chemistrytwig.com/2013/09/18/whats-the-difference-between-

absorption-and-adsorption/.

[3] R. J. Hunter, Foundations of colloid science, Second. USA: OXFORD UNIVERSITY

PRESS, 2007.

[4] G. Limousin, J.-P. Gaudet, L. Charlet, S. Szenknect, V. Barthès, and M. Krimissa,

“Sorption isotherms: A review on physical bases, modeling and measurement,” Appl.

Geochemistry, vol. 22, no. 2, pp. 249–275, Feb. 2007.

[5] A. Santosh, “Adsorption,” 2012. [Online]. Available: http://chemistry-

desk.blogspot.mx/2012/10/adsorption.html.

[6] K. Y. Foo and B. H. Hameed, “Insights into the modeling of adsorption isotherm

systems,” Chem. Eng. J., vol. 156, no. 1, pp. 2–10, Jan. 2010.

[7] P. C. Hiemenz and R. Rajagopalan, Principles of colloid and surface chemistry.,

Third. New York: Marcel Dekker, Inc., 1997.

[8] University of Waikato, “Surfactants,” 2012. [Online]. Available:

https://www.sciencelearn.org.nz/images/1291-surfactants.

[9] M. J. Rosen and J. T. Kunjappu, Surfactants and interfacial phenomena. John Wiley

& Sons, 2012.

54
[10] Schlumberger, “porosity,” Oilfield Glossary, 2001. [Online]. Available:

http://www.glossary.oilfield.slb.com/Terms/p/porosity.aspx.

[11] P. R. Davila Farias, “PROCESSAMENTO DE IMAGENS DE MICRO - CT NA

CARACTERIZAÇÃO DE BIOFILME BACTERIANO,” UFRJ, 2017.

[12] S. J. Gregg, K. S. W. Sing, and H. W. Salzberg, “Adsorption surface area and

porosity,” J. Electrochem. Soc., vol. 114, no. 11, p. 279C–279C, 1967.

[13] A. Szymkiewicz, Modelling Water Flow in Unsaturated Porous Media: Accounting

for Nonlinear Permeability and Material Heterogeneity, vol. 9. 2013.

[14] Schlumberger, “permeability,” Oilfield Glossary, 2002. [Online]. Available:

http://www.glossary.oilfield.slb.com/Terms/p/permeability.aspx.

[15] M. Bai, D. Elsworth, H. I. Inyang, and J. C. Roegiers, “Modeling contaminant

migration with linear sorption in strongly heterogeneous media,” J. Environ. Eng., vol.

123, no. 11, pp. 1116–1125, 1997.

[16] J. E. Warren and P. J. Root, “The Behavior of Naturally Fractured Reservoirs,” Soc.

Pet. Eng. J., vol. 3, no. 3, pp. 245–255, 1963.

[17] K. H. Coats and B. D. Smith, “Dead-End Pore Volume and Dispersion in Porous

Media,” Soc. Pet. Eng. J., vol. 4, no. 1, pp. 73–84, 1964.

[18] M. L. Brusseau, R. E. Jessup, and P. S. C. Rao, “Modeling the transport of solutes

influenced by multiprocess nonequilibrium,” Water Resour. Res., vol. 25, no. 9, pp.

1971–1988, Sep. 1989.

[19] W. E. Brigham, P. W. Reed, and J. N. Dew, “Experiments on Mixing During Miscible

55
Displacement in Porous Media,” Soc. Pet. Eng. J., vol. 1, no. 1, pp. 1–8, 1961.

[20] M. Bai and D. Elsworth, “On the modeling of miscible flow in multi-component

porous media,” Transp. Porous Media, vol. 21, no. 1, pp. 19–46, Oct. 1995.

[21] K. K. Mohanty, “Multiphase Flow in Porous Media: III. Oil Mobilization, Transverse

Dispersion, and Wettability,” in SPE Annual Technical Conference and Exhibition,

1983.

[22] H. Valiollahi, Z. Ziabakhsh, and P. L. J. Zitha, “Computers & Geosciences

Mathematical modeling of chemical oil-soluble transport for water control in porous

media,” Comput. Geosci., vol. 45, pp. 240–249, 2012.

[23] M. Liu, M. Shabaninejad, and P. Mostaghimi, “Computers & Geosciences Impact of

mineralogical heterogeneity on reactive transport modelling,” Comput. Geosci., vol.

104, no. March, pp. 12–19, 2017.

[24] R. C. Schwartz, A. S. R. Juo, and K. J. McInnes, “Estimating parameters for a dual-

porosity model to describe non-equilibrium, reactive transport in a fine-textured soil,”

J. Hydrol., vol. 229, no. 3–4, pp. 149–167, Apr. 2000.

[25] M. T. Van-Genuchten and P. J. Wierenga, “Mass Transfer Studies in Sorbing Porous

Media I. Analytical Solutions,” Soil Sci. Soc. Am. J., vol. 40, no. AUGUST, pp. 473–

480, 1976.

[26] L. Lapidus and N. R. Amundson, “Mathematics of adsorption in beds. VI. The effect

of longitudinal diffusion in ion exchange and chromatographic columns,” J. Phys.

Chem., vol. 56, no. 8, pp. 984–988, 1952.

56
[27] M. T. Balhoff, K. E. T. Ã, and M. Hjortsø, “Coupling pore-scale networks to

continuum-scale models of porous media,” Comput. Geosci., vol. 33, pp. 393–410,

2007.

[28] M. Sardin, D. Schweich, F. J. Leij, and M. T. van Genuchten, “Modeling the

Nonequilibrium Transport of Linearly Interacting Solutes in Porous Media: A

Review,” Water Resour. Res., vol. 27, no. 9, pp. 2287–2307, Sep. 1991.

[29] T. D. Le, C. Moyne, and M. A. Murad, “A three-scale model for ionic solute transport

in swelling clays incorporating ion–ion correlation effects,” Adv. Water Resour., vol.

75, pp. 31–52, Jan. 2015.

[30] M. A. Maraqa and S. A. Khashan, “Modeling solute transport affected by

heterogeneous sorption kinetics using single-rate nonequilibrium approaches.,” J.

Contam. Hydrol., vol. 157, pp. 73–86, Feb. 2014.

[31] J. D. Anderson, Computational fluid dynamics: The basics with applications, 1st ed.

McGraw-Hill, 1999.

[32] A. Cardona and V. Fachinotti, “Introduccion al metodo de los elementos finitos,”

2014. [Online]. Available:

http://www.cimec.org.ar/twiki/pub/Cimec/CursoFEM/cursofem_0.pdf.

[33] Y. Niño, “Metodo de los volumenes finitos.” Chile University, p. 25, 2002.

57
Chapter III: Methodology
In this chapter, the methodology used for the solution of each case is shown. A comparison
was done using the model proposed by different authors in their works and the models
proposed here. In Table 3.1 the different models are shown.

Table 3.1. Models for transport in porous media comparison.

Transport mechanism Author Equation


𝑁𝑐

𝑘𝑟𝑖 (𝑠𝑖 ) = ∑ 𝐶𝑗𝑖 𝐵𝑗𝑚 (𝑠𝑖 , ⃗⃗⃗


𝑦 𝑖 ),
𝑗=1
Kulkarni et al.[1] 𝑁𝑐

Two phase flow in 𝑝𝑐 (𝑠𝑖 ) = ∑ 𝐶𝑗𝑐 𝐵𝑗𝑚 (𝑠𝑖 , ⃗⃗⃗⃗


𝑦𝑐)
porous media 𝑗=1
𝑘𝑟𝑤 = 𝑘𝑟𝑤0 (𝑆𝑤 )𝑎𝑤
𝑘𝑟𝑜 = 𝑘𝑟𝑜0 (1 − 𝑆𝑤 )𝑎𝑜
This work 1
𝜙 2
𝑝𝑐 = 𝑗(𝑆𝑤 ) 𝜎 𝑐𝑜𝑠(𝜃) ( )
𝑘
𝜕𝑥𝑖 𝜕 2 𝑥𝑖 𝜕𝑥𝑖 1 − 𝜑 𝜌𝑟 𝜕𝑛𝑖𝑒𝑎
Species transport in Mannhardt et al.[2], [3] −𝑣 +𝐷 = +
𝜕𝑦 𝜕𝑦 2 𝜕𝑡 𝜑 𝜌𝑖 𝜕𝑡
porous media 𝜕𝐶A,1 (1 − 𝜑)
3
𝑘𝑐 1,𝑚 𝑎𝑣 1,𝑚
This work −𝑣1 ∙∇ (𝐶A,1 ) + ∇ ∙ (𝐷1 ∙∇ 𝐶A,1 ) =
𝜕𝑡
(1 + 𝐹1′
𝜑
)+ ∑
𝑓1 φ
(𝐶A,1 − 𝐶𝐴,𝑚 )
𝑚=1

𝜕(𝜌𝑜 𝑧̇ 𝑧+𝜌𝑤 𝑧̇ (𝐿−𝑧)) 2𝑏2 𝜎𝐹 𝑐𝑜𝑠𝜃𝐹 2𝑏2 𝜎𝐹 𝑐𝑜𝑠𝜃𝐹 32𝜇𝑜𝑧̇ 𝑧 32𝜇𝑤 𝑧̇ (𝐿−𝑧)
- =− − + ∆𝜌 𝑔 𝑧 + + + (𝜌𝑤 𝑧̇ 𝑧̇ − 𝜌𝑜 𝑧̇ 𝑧̇)
𝜕𝑡 𝑅𝐹 𝑅𝐹 𝐷2 𝐷2
J. L. Lopez Salinas[4]
𝜕(𝜌𝑜 𝑧 + 𝜌𝑤 (𝐿 − 𝑧))
= −(𝜌𝑤 𝑧̇ − 𝜌𝑜 𝑧̇ )
𝜕𝑡
Two phase flow in 𝜕(𝜌𝑣 )
capillary conducts + ∇ ∙ (𝜌𝑣 𝑣) = −∇𝑝 + ∇ ∙ (𝜏̿) + 𝜌𝑔 + 𝐹
𝜕𝑡
This work 𝑛
1 ∂
[ (𝛼 𝜌 ) + ∇ ∙ (𝛼𝑞 𝜌𝑞 𝑣
⃗⃗⃗⃗𝑞 ) = 𝑆𝛼𝑞 + ∑(𝑚̇𝑝𝑞 − 𝑚̇𝑞𝑝 )]
𝜌𝑞 ∂t 𝑞 𝑞
𝑝=1

3.1 Multiphase flow in saturated porous media


3.1.1 Model development

The methodology for the resolution of this case consists, in the solution of the continuity
equation for two-phase flow in a single dimension. Incompressible and immiscible phases of
water and oil are considered. The system is saturated initially with water, and it will be

58
displaced by an oil flow. Gravity forces are also neglected to make comparisons with the
results obtained by Kulkarni et al. [1]

First, a mass balance for a single phase (water) is considered.

𝜕(𝜙𝑆𝑤 𝜌𝑤 ) (3.1)
+ 𝛻(𝜌𝑤 𝒖𝒘 ) − 𝛻(𝜙𝑆𝑤 𝑫𝒘 𝛻𝜌𝑤 ) + 𝑟𝑤 = 0
𝜕𝑡
Moreover, volumetric flows for oil and water are defined as follows

𝑘𝑟𝑤
𝒖𝒘 = − 𝐾(𝛻𝑝𝑤 + 𝜌𝑤 𝑔𝛻𝑧) (3.2)
𝜇𝑤
𝑘𝑟𝑜
𝒖𝒐 = − 𝐾(𝛻𝑝𝑜 + 𝜌𝑜 𝑔𝛻𝑧) (3.3)
𝜇𝑜
Capillary pressure can be defined as

𝛻𝑝𝑐 = 𝛻𝑝𝑜 − 𝛻𝑝𝑤 (3.4)

Isolating the oil pressure of equation (3.4) and substituting in equation (3.3) we have

𝛻𝑝𝑜 = 𝛻𝑝𝑤 + 𝛻𝑝𝑐 (3.5)

𝑘𝑟𝑜 𝑘𝑟𝑜
𝒖𝒐 = − 𝐾(𝛻𝑝𝑜 + 𝜌𝑜 𝑔𝛻𝑧) = − 𝐾(𝛻𝑝𝑤 + 𝛻𝑝𝑐 + 𝜌𝑜 𝑔𝛻𝑧) (3.6)
𝜇𝑜 𝜇𝑜
The total volumetric flux is defined by

𝑢 = 𝑢𝑤 + 𝑢𝑜 (3.7)

Substituting equations (3.2) and (3.6) in equation (3.7) and arranging variables we obtain

𝑘𝑟𝑤 𝑘𝑟𝑜 𝑘𝑟𝑜 𝑑𝑝𝑐 𝑘𝑟𝑤 𝑘𝑟𝑜


𝑢 = −( + ) 𝐾 𝛻𝑝𝑤 − 𝐾 𝛻𝑆𝑤 − ( 𝜌𝑤 + 𝜌 ) 𝑔 𝐾 𝛻𝑧 (3.8)
𝜇𝑤 𝜇𝑜 𝜇𝑜 𝑑𝑆𝑤 𝜇𝑤 𝜇𝑜 𝑜
Where

𝑑𝑝𝑐
𝛻𝑝𝑐 = 𝛻𝑆 (3.9)
𝑑𝑆𝑤 𝑤
Isolating ∇pw from equation (3.8) and substituting in equation (3.2) we get

𝑘𝑟𝑜 𝑑𝑝𝑐 𝑘𝑟𝑜


𝑢𝑤 = 𝐹𝑤 (𝑢 + 𝐾 𝛻𝑆𝑤 − ∆𝜌 𝑔 𝐾 𝛻𝑧) (3.10)
𝜇𝑜 𝑑𝑆𝑤 𝜇𝑜
Where

59
1 𝑘𝑟𝑜 𝜇𝑤 −1
𝐹𝑤 = = (1 + ) (3.11)
𝑘 𝜇 𝑘𝑟𝑤 𝜇𝑜
1 + 𝑟𝑜 𝑤
𝑘𝑟𝑤 𝜇𝑜
∆𝜌 = 𝜌𝑤 − 𝜌𝑜 (3.12)

Thus, we combined the equations of the fluxes of both phases to obtain the flux of water
phase as a function of the total flux, oil relative permeability and viscosity.

To obtain water saturation by coupling the equations of both phases, we consider


incompressible flow in equation (3.1) to obtain

𝜕(𝜙𝑆𝑤 ) (3.13)
𝜌𝑤 + 𝜌𝑤 𝛻(𝒖𝒘 ) + 𝑟𝑤 = 0
𝜕𝑡
Dividing by density we obtain

𝜕(𝜙𝑆𝑤 ) (3.14)
+ 𝛻(𝒖𝒘 ) + 𝑞𝑤 = 0
𝜕𝑡
Where

𝑟𝑤
𝑞𝑤 = (3.15)
𝜌𝑤
Now, substituting equation (3.10) in equation (3.14), we obtain the water saturation in the
system using a model that couples the equations of two phases, water and oil.

𝜕(𝜙𝑆𝑤 ) 𝑘𝑟𝑜 𝑑𝑝𝑐 𝑘𝑟𝑜


+ 𝛻 (𝐹𝑤 (𝑢 + 𝐾 𝛻𝑆𝑤 − ∆𝜌 𝑔 𝐾 𝛻𝑧)) + 𝑞𝑤 = 0 (3.16)
𝜕𝑡 𝜇𝑜 𝑑𝑆𝑤 𝜇𝑜

The total volumetric balance equations can be expressed as

𝜕𝜙
+ 𝛻𝒖 + 𝑞 = 0 (3.17)
𝜕𝑡
Isolating ∇u we obtain

𝜕𝜙 (3.18)
𝛻𝒖 = − −𝑞
𝜕𝑡
Substituting eq (3.18) in eq (3.16) and simplifying we obtain

𝜕(𝜙𝑆𝑤 ) 𝜕𝜙 𝜕𝐹𝑤 𝑘𝑟𝑜 𝑑𝑝𝑐 𝑘𝑟𝑜


+ (𝑆𝑤 − 𝐹𝑤 ) + 𝑢𝛻𝑆𝑤 + 𝛻(𝐹𝑤 𝐾 𝛻𝑆𝑤 ) − 𝛻(𝐹𝑤 ∆𝜌 𝑔 𝐾 𝛻𝑧) − 𝐹𝑤 𝑞 + 𝑞𝑤 = 0 (3.19)
𝜕𝑡 𝜕𝑡 𝜕𝑆𝑤 𝜇𝑜 𝑑𝑆𝑤 𝜇𝑜

Considering q = qw = 0 and a homogeneous and isotropic porous media we get

60
𝜕𝑆𝑤 𝜕𝐹𝑤 𝑘𝑟𝑜 𝑑𝑝𝑐 𝑘𝑟𝑜
𝜙 + 𝑢𝛻𝑆𝑤 + 𝛻 (𝐹𝑤 𝐾 𝛻𝑆𝑤 ) − 𝛻 (𝐹𝑤 ∆𝜌 𝑔 𝐾 𝛻𝑧) = 0 (3.20)
𝜕𝑡 𝜕𝑆𝑤 𝜇𝑜 𝑑𝑆𝑤 𝜇𝑜
Finally, considering gravity as negligible because out system is in a horizontal position, and
a one dimensional system, the next expression is obtained

𝜕𝑆𝑤 𝜕𝐹𝑤 𝜕𝑆𝑤 𝜕 𝑘𝑟𝑜 𝑑𝑝𝑐 𝜕𝑆𝑤


𝜙 + 𝑢 + (𝐹𝑤 𝐾 )=0 (3.21)
𝜕𝑡 𝜕𝑆𝑤 𝜕𝑥 𝜕𝑥 𝜇𝑜 𝑑𝑆𝑤 𝜕𝑥
With this equation is possible to determinate the water saturation in a porous media with a
constant oil injection.

3.1.2 Simplification and dimensionless system of


equations
An easier way to solve the established model is to use the dimensionless form of the equations
as follows

𝐿 𝜕𝑆𝑤 𝜕𝐹𝑤 𝜕𝑆𝑤 𝜕 𝑘𝑟𝑜 𝑑𝑝𝑐 𝜕𝑆𝑤


[𝜙 + 𝑢 + (𝐹𝑤 𝐾 ) = 0] (3.22)
𝑢 𝜕𝑡 𝜕𝑆𝑤 𝜕𝑥 𝜕𝑥 𝜇𝑜 𝑑𝑆𝑤 𝜕𝑥
𝜕𝑆𝑤 𝜕𝐹𝑤 𝜕𝑆𝑤 𝐾 𝜕 𝑑𝑝𝑐 𝜕𝑆𝑤
𝜙 + + (𝐹𝑤 𝑘𝑟𝑜 )=0 (3.23)
𝜕𝜏 𝜕𝑆𝑤 𝜕𝑧 𝑢 𝜇𝑜 𝐿 𝜕𝑧 𝑑𝑆𝑤 𝜕𝑧
Where

𝑡𝑢
𝜏= (3.24)
𝐿
𝑥
𝑧= (3.25)
𝐿
Capillary pressure may be represented as water saturation function through Leverett
equation[5].

1
𝜙 2 (3.26)
𝑝𝑐 = 𝑗(𝑆𝑤 ) 𝜎 𝑐𝑜𝑠(𝜃) ( )
𝑘
In this case a polynomial will be used to estimate the 𝑗(𝑆𝑤 ) coefficient, the expression
suggested by Amir[6] will be used

𝑗(𝑆𝑤 ) = 𝑎1 (1 − 𝑆𝑤 ) + 𝑎2 (1 − 𝑆𝑤 )2 + 𝑎3 (1 − 𝑆𝑤 )3 + ⋯ (3.27)

61
Where 𝑎1 , 𝑎3 and 𝑎3 are coefficients, Amir[6] uses the next values for a third order
polynomial.

𝑎1 = 1.417

𝑎2 = -2.12

𝑎3 = 1.263

The derivative with respect to water saturation of this polynomial is

𝑑𝑗(𝑆𝑤 )
= −𝑎1 − 2𝑎2 (1 − 𝑆𝑤 ) − 3𝑎3 (1 − 𝑆𝑤 )2 (3.28)
𝑑𝑆𝑤
Substituting equation (3.26) in equation (3.23) we get

𝜕𝑆𝑤 𝜕𝐹𝑤 𝜕𝑆𝑤 𝜕 𝑑𝑗(𝑆𝑤 ) 𝜕𝑆𝑤


𝜙 + + 𝛼 (𝐹𝑤 𝑘𝑟𝑜 )=0 (3.29)
𝜕𝜏 𝜕𝑆𝑤 𝜕𝑧 𝜕𝑧 𝑑𝑆𝑤 𝜕𝑧
Where

𝐾 𝜙 √𝜙 𝑘 (3.30)
𝛼= 𝜎 𝑐𝑜𝑠(𝜃) √ =
𝑢 𝜇𝑜 𝐿 𝑘 𝐿 𝐶𝑎
𝑢 𝜇𝑜
𝐶𝑎 = (3.31)
𝜎 𝑐𝑜𝑠(𝜃)
𝑘𝑟𝑜 𝜇𝑤 −1
𝐹𝑤 = (1 + ) (3.32)
𝑘𝑟𝑤 𝜇𝑜
𝑘𝑟𝑜 = 𝑘𝑟𝑜0 (1 − 𝑆𝑤 )𝑎𝑜 (3.33)

𝑘𝑟𝑤 = 𝑘𝑟𝑤0 (𝑆𝑤 )𝑎𝑤 (3.34)

𝑘𝑟𝑜 𝜇𝑤 −1
𝜕𝐹𝑤 𝜕 (1 + ) 𝑘𝑟𝑜 𝜇𝑤 −2 𝜇𝑤 𝑘𝑟𝑤 𝑘𝑟𝑜
′ ′
− 𝑘𝑟𝑜 𝑘𝑟𝑤
𝑘𝑟𝑤 𝜇𝑜 (3.35)
= = − (1 + ) ( )( 2
)
𝜕𝑆𝑤 𝜕𝑆𝑤 𝑘𝑟𝑤 𝜇𝑜 𝜇𝑜 𝑘𝑟𝑤

𝑘𝑟𝑤 = 𝑎𝑤 𝑘𝑟𝑤0 (𝑆𝑤 )𝑎𝑤−1 (3.36)

𝑘𝑟𝑜 = −𝑎𝑜 𝑘𝑟𝑜0 (1 − 𝑆𝑤 )𝑎𝑜−1 (3.37)

With equation (3.29) we will obtain the water and oil saturation in the porous media

3.1.3 Boundary conditions


For the system analyzed here, we have the next initial conditions for water saturation

62
𝑆𝑤 = 𝑆𝑤∗ (𝑥), 0 ≤ 𝑥 ≤ 𝐿, 𝑡 = 0 (3.38)

Where, for this analysis 𝑆𝑤∗ =1

For the oil injection boundary condition in the inlet the next condition is used

𝜕𝑆𝑤 𝜇𝑜 (𝑢𝑜 )𝑖𝑛𝑖𝑐𝑖𝑎𝑙 𝑑𝑝𝑐 −1


=− [ ] , 𝑥 = 0, 𝑡 > 0 (3.39)
𝜕𝑥 𝐾 𝑘𝑟𝑜 𝑑𝑆𝑤
Besides, for a incompressible flow

𝑢𝑜 = (𝑢𝑜 )𝑖𝑛𝑖𝑐𝑖𝑎𝑙 , 0 ≤ 𝑥 ≤ 𝐿, 𝑡 > 0 (3.40)

For the outlet boundary condition

𝑆𝑤 = 𝑆𝑤 |𝑃𝑐=0 , 𝑥 = 𝐿, 𝑡 > 0 (3.41)

Where, for this analysis 𝑆𝑤 |𝑃𝑐=0 = 1

Dimensionless boundary condition

We just need to obtain the dimensionless form for the injection boundary condition, for is the
only one with dimensions. We just need to substitute equation (3.26) in equation (3.39) and
multiplying both sided by L to get the dimensionless form of this boundary condition

1 −1
𝜕𝑆𝑤 𝜇𝑜 (𝑢𝑜 )𝑖𝑛𝑖𝑐𝑖𝑎𝑙 𝜕 𝜙 2
𝐿[ =− [ 𝑗(𝑆𝑤 ) 𝜎 𝑐𝑜𝑠(𝜃) ( ) ] ] (3.42)
𝜕𝑥 𝐾 𝑘𝑟𝑜 𝜕𝑆𝑤 𝑘

−1 −1
𝜕𝑆𝑤 𝐿 𝜇𝑜 (𝑢𝑜 )𝑖𝑛𝑖𝑐𝑖𝑎𝑙 𝜕 1 𝜕
=− 1 [ 𝑗(𝑆𝑤 ) ] = − [ 𝑗(𝑆𝑤 ) ] (3.43)
𝜕𝑧 𝜕𝑆𝑤 𝑘𝑟𝑜 𝛼 𝜕𝑆𝑤
𝜙 2
𝐾 𝑘𝑟𝑜 𝜎 𝑐𝑜𝑠(𝜃) ( )
𝑘

3.1.4 Methodology for model application


For this model resolution, COMSOL software was used, using its differential equations
module. To solve the model we need to know the parameter’s values. For the relative
permeability coefficients in the cited work[1], an adjustment was made using the
experimental data by least squares method, performing interpolations with the B-spline
technique. It is a precise and useful method in large databases with nonlinear behaviors, but
the determination of these splines requires a computational algorithm of some complexity,
which can difficult its manipulation in certain software, such is the case of COMSOL. In the

63
same way, an adjustment of the capillary pressure was made using the same method. In
Kulkarni’s work, the values of the coefficients obtained and the number of terms used were
not mentioned. For this reason, the coefficients were determined using the data presented in
the graphs of the experimental data (Figure 3.1 and Figure 3.2) by the least squares technique.
Relative permeabilities were adjusted using the exponential models of equations (3.33) and
(3.34), this allows us to know also the coefficients of affinity of the fluids towards the porous
material ao and aw. Whereas for the capillary pressure the Leverett equation was used with
the polynomial of 3rd order suggested by Amir including the terms of interfacial tension,
permeability, porosity and contact angle. The adjusted parameters are shown in the results
section. Also, a sensitivity analysis was done with α coefficient, to observe its effect in the
displacement of water in the saturated porous media.
(b)

(a)
Figure 3.1. Experimental data used by Kulkarni et al for relative permeability of oil (a) and
water (b) vs. water saturation[1]

64
Figure 3.2. Experimental data used by Kulkarni et al for capillary pressure vs. water
saturation in a porous media for an oil-water system[1]

The results were compared against the low injection speed experiment of the cited work. The
values used in the cited work are shown in Table 3.2.

Table 3.2. Values used in Kukarnki et al work

Water viscosity (cp) 1


Oil viscosity (cp) 1.5
Core lenght (cm) 10
Core diameter (cm) 3.5
Porosity (%) 30.0
Absolute permeability (mD) 50.0
Initial water saturation 1.0
Injection velocity (cm3/ min)
Water 0.0
Oil 0.1

3.2 Modeling surfactant adsorption/retention and


transport through porous media
3.2.1 Model development
The porosity of a solid may be broadly classified in two: Connected pores or macroporosity
and the not-connected pores or meso porosity. These two different types of porosity coexist
and interact with each other. Two common ways to model a medium that possesses both
macroporosity and mesoporosity are by specifying dual porosity with single permeability,
and dual porosity with dual permeability. Another approach is to consider the two sections
as a single heterogeneous section allowing the porosity and permeability to vary rapidly and
discontinuously over the entire domain, but this approach is not considered here and the
computational cost is high.

Using concepts developed by Salter and Mohanty [7], a mathematical model is proposed to
examine the transport of components in porous media. This model can be used to understand
and predict the adsorption of surfactants in consolidated porous media. In addition, the model
can be tuned or tested by analyzing experimental results. The model can be generalized for

65
the liquid phase in the three porous regions in all directions: macro, meso and microporous
region:
3
𝜕𝐶A,1 𝑘𝑐 1,𝑚 𝑎𝑣 1,𝑚 𝑘𝑐 1,4 𝑎𝑣 1,4
−𝑣1 ∙∇ (𝐶A,1 ) + ∇ ∙ (𝐷1 ∙∇ 𝐶A,1 ) = +∑ (𝐶A,1 − 𝐶𝐴,𝑚 ) + (𝐶A,1 − 𝐶A,4 ) (3.44)
𝜕𝑡 𝑓1 φ 𝑓1 φ
𝑚=1
3
𝜕𝐶A,2 𝑘𝑐 2,𝑚 𝑎𝑣 2,𝑚 𝑘𝑐 2,5 𝑎𝑣 2,5
−𝑣2 ∙∇ (𝐶A,2 ) + ∇ ∙ (𝐷2 ∙∇ 𝐶A,2 ) = +∑ (𝐶A,2 − 𝐶𝐴,𝑚 ) + (𝐶A,2 − 𝐶A,5 ) (3.45)
𝜕𝑡 𝑓2 φ 𝑓2 φ
𝑚=1

𝜕𝐶A,3 𝑘𝑐 3,6 𝑎𝑣 3,6


∇ ∙ (𝐷3 ∙∇ 𝐶A,3 ) = + (𝐶A,3 − 𝐶A,6 ) (3.46)
𝜕𝑡 𝑓3 φ

Velocity v1 represents the velocity field existing in the macro porous region and velocity v2
the velocity field existing in the meso porous region, 𝐷1 , 𝐷2 and 𝐷3 are the

dispersion/diffusion coefficients in the macroporous, mesoporous and microporous regions,


respectively. 𝐶𝐴,𝑙 is the concentration of component A in region l, t is time, 𝑘𝑐 𝑙,𝑚 is the mass
transfer coefficient between regions l and m, 𝑎𝑣 𝑙,𝑚 is the specific area for mass transfer per
unit volume, 𝜑 is the total porosity and f1, f2 and f3 are the normalized macro, meso and micro
porosity, respectively. Regions l and m may be: 1 for the macroporous region, 2 for the
mesoporous region, 3 for the microporous region, 4 for the solid in contact with the
macropores, 5 for the solid in contact with the mesopores and 6 for the solid in contact with
the micropores.

The two terms of the LHS of Eq. (3.44) from left to right account for convection and
diffusion, respectively. Similarly for the RHS of Eq. (3.44), the terms account for
accumulation, mass exchange between porous regions and the exchange of mass between
phases, respectively. Each equation includes mass transfer in the porous region between the
fluid and its respective solid interface (solid-liquid interface). This is needed if the solute
(surfactant) has the potential of being adsorbed. In the case of a solute with no adsorption
potential (tracer), the equilibrium phase term is discarded.

When the double-permeability model is used, two different velocities v1 and v2, arise due to
the two values of permeability that are considered. These two velocities may be determined
by Darcy´s law for each region. This approach is similar to the one used in transport in
fractured reservoirs and has been used to model horizontal wells by Jia and Niu [8].When
homogeneous media is considered, v2 is discarded, and the velocity field 1 is assumed to be
uniform in axial direction.

66
In the mass exchange between phases term, the concentration difference (𝐶A,𝑙 − 𝐶A,𝑚 ) is the
solute concentration in the liquid phase minus the solute concentration in the solid phase (l=
1, 2, 3 and m= 4, 5, 6).

Some simplification can be made according to the system selected, for instance, by assuming
that the solute concentration equilibrates instantaneously during adsorption with the solid
phase (instant local equilibrium). Taking into account that exchange of mass sites (designated
sites where adsorption of solute may occur) are distributed randomly throughout the porous
media, the equilibrium phase term can be replaced with the solute isotherm, and equations
(3.44) to (3.46) can be simplified as follows:
3
𝜕𝐶A,1 (1 − 𝜑) 𝑘𝑐 1,𝑚 𝑎𝑣 1,𝑚
−𝑣1 ∙∇ (𝐶A,1 ) + ∇ ∙ (𝐷1 ∙∇ 𝐶A,1 ) = (1 + 𝐹1′ )+ ∑ (𝐶A,1 − 𝐶𝐴,𝑚 ) (3.47)
𝜕𝑡 𝜑 𝑓1 φ
𝑚=1
3
𝜕𝐶A,2 (1 − 𝜑) 𝑘𝑐 2,𝑚 𝑎𝑣 2,𝑚
−𝑣2 ∙∇ (𝐶A,2 ) + ∇ ∙ (𝐷2 ∙∇ 𝐶A,2 ) = (1 + 𝐹2′ )+ ∑ (𝐶A,2 − 𝐶𝐴,𝑚 ) (3.48)
𝜕𝑡 𝜑 𝑓2 φ
𝑚=1

𝜕𝐶A,3 (1 − 𝜑)
∇ ∙ (𝐷 ∙∇ 𝐶A,3 ) = (1 + 𝐹′3 ) (3.49)
3 𝜕𝑡 𝜑

This simplification reduces the number of differential equations to three for each solute in
the porous media. The model is then easier to handle, and the number of parameters to be
estimated is reduced. In these equations 𝐹1′ , 𝐹2′ and 𝐹3′ represent the adsorption isotherm
derivative with respect to the solute concentration in the aqueous phase for the macroporous,
mesoporous and microporous regions, respectively.

Simplification of the previous equations can lead to different models used by researchers in
the past [2]. The limitations of a model that results from simplifying equations (3.44) to
(3.46), arise as complete terms are neglected (e.g. not considering instant local equilibrium,
or eliminating meso and micro porosity).

One advantage of the model presented here is that it includes all the different pore networks
previously discussed, hence, it can be used to adequately model surfactant flow through
intricate porous media.

Since the solute adsorption process is complex, dealing with surface active materials or
polymers cannot be described adequately using a single isotherm. Hence, for most models,

67
the Langmuir isotherm and its derivative are used. This isotherm is the most commonly used
for systems involving large molecules (Molecular weight >200 g/mol) [9]

𝐴 𝐶𝐴
𝑎 = 𝐹(𝐶𝐴 ) = (3.50)
1 + (𝑘 𝐶𝐴 )
𝐴
𝐹′(𝐶𝐴 ) = (3.51)
(1 + 𝑘 𝐶𝐴 )2
In equations (3.50) and (3.51), a is adsorption in mg/g of rock, A is the adsorption coefficient,
k is the desorption coefficient and 𝐶𝐴 is the concentration of component A.

The Redlich-Peterson isotherm [10] was used for comparison. It takes the form:

𝐴1 𝐶𝐴 𝐴2 𝐶𝐴
𝑎 = 𝐹(𝐶𝐴 ) = + (3.52)
1 + (𝑘1 𝐶𝐴 ) 1 1 + (𝑘2 𝐶𝐴 )𝑛2
𝑛

𝐴1 [1 − (𝑛1 − 1)(𝑘1 𝐶𝐴 )𝑛1 ] 𝐴2 [1 − (𝑛2 − 1)(𝑘2 𝐶𝐴 )𝑛2 ]


𝐹 ′ (𝐶𝐴 ) = + (3.53)
[1 + (𝑘1 𝐶𝐴 )𝑛1 ]2 [1 + (𝑘2 𝐶𝐴 )𝑛2 ]2
Where 𝐴1 and 𝐴2 are the adsorption coefficients for the first and second terms, 𝑘1 and 𝑘2 are
the adsorption coefficients for the first and second terms and 𝑛1 and 𝑛2 are the exponential
coefficients for the first and second terms respectively

The irreversibility conditions for the isotherms are specified by the equations:

𝐴 𝐶𝐴 𝑑𝐶𝐴
𝑎= , >0 (3.54)
1 + (𝑘𝐶𝐴 )𝑛 𝑑𝜏
𝑑𝐶𝐴
𝑎 = 0, <0 (3.55)
𝑑𝜏
Also, a hybrid Langmuir isotherm may be used. For this isotherm in macro and meso
porosity, the conditions are given by:

𝐴𝑖,I 𝐶𝐴,𝑖 𝑑𝐶
𝑎𝑖 = 𝑛 , >0 (3.56)
1 + (𝑘𝑖,I 𝐶𝐴,𝑖 ) 𝑑𝜏

𝐴𝑖,II 𝐶𝐴,𝑖 𝑑𝐶
𝑎𝑖 = 𝑛 , <0 (3.57)
1 + (𝑘𝑖,II 𝐶𝐴,𝑖 ) 𝑑𝜏

Where 𝑎𝑖 is adsorption for region i, Ai,I and Ai,II are the adsorption coefficients and ki,I and
ki,II are the desorption coefficients, for the macro (i=1) and meso (i=2) porous regions

68
Finally, to evaluate the appropriateness of the assumption of non-local equilibrium, a network
with only macro and meso porosity was selected. The equations describing the process are
(3.44), (3.45), (3.58) and (3.59):

𝜕𝐶A,4 ′
𝑘𝑐 1,4 𝑎𝑣 1,4 (𝐶𝐴,1 − 𝐶𝐴,4 ) = 𝑓1 (𝐹 4 (1 − 𝜑)) (3.58)
𝜕𝑡
𝜕𝐶A,5 ′
𝑘𝑐 2,5 𝑎𝑣 2,5 (𝐶𝐴,2 − 𝐶𝐴,5 ) = 𝑓2 (𝐹 5 (1 − 𝜑)) (3.59)
𝜕𝑡
In the resulting model, concentration 𝐶𝐴,4 is the solute concentration in the solid in contact
with the liquid phase in the macro porous region. Concentration 𝐶𝐴,5 is the solute
concentration in the solid in contact with the liquid phase in the mesoporous region.

3.2.2 One-dimensional system of equations


The equations describing the one-dimensional transport of solute in the porous medium are
recast in dimensionless form. Equations (3.47) to (3.49) take the following form:

𝜕𝐶A,1 𝑓1 𝜕 2 𝐶A,1 𝜕𝐶A,1 (1 − 𝜑)


−𝑏1 + 2
− 𝑁𝑆𝑡 1,2 (𝐶𝐴,1 − 𝐶𝐴,2 ) = 𝑓1 (1 + 𝐹 ′1 ) (3.60)
𝜕𝜒 𝑁𝑃𝑒,1 𝜕𝜒 𝜕𝜏 𝜑
𝜕𝐶A,2 𝜕𝐶A,2 (1 − 𝜑)
−𝑏2 + 𝑁𝑆𝑡 1,2 (𝐶𝐴,1 − 𝐶𝐴,2 ) + 𝑁𝑆𝑡 2,3 (𝐶𝐴,3 − 𝐶𝐴,2 ) = 𝑓2 (1 + 𝐹 ′ 2 ) (3.61)
𝜕𝜒 𝜕𝜏 𝜑

𝑎 1 𝜕 𝜕𝐶A,3 𝜕𝐶A,3 (1 − 𝜑)
( 2 (𝜁 2 )) = (1 + 𝐹 ′ 3 ) (3.62)
𝑁𝑃𝑒 2,3 𝜁 𝜕𝜁 𝜕𝜁 𝜕𝜏 𝜑

Where 𝑏1 and 𝑏2 are the dimensionless velocity terms for the macro and meso porous regions,
respectively (for the cases studied here 𝑏1 = 1 and 𝑏2 = 0), 𝜒 and τ are the dimensionless
distance and time, respectively (this dimensionless time is also known as pore volumes or
PV), and 𝜁 is the dimensionless radius of the spheres that represent the grains, 𝑁𝑆𝑡 is the
Stanton number and 𝑁𝑃𝑒 is the Peclet number, where:

𝜁 = 𝑟/𝑅 (3.63)

𝜒 = 𝑥/𝐿 (3.64)
𝑡𝑣
𝜏= (3.65)
𝐿
(𝑘𝑐 𝑙,𝑚 𝑎𝑣 𝑙,𝑚 𝐿)
𝑁𝑆𝑡 𝑙,𝑚 = (3.66)
𝑣𝑙

69
𝑣𝑙 𝐿
𝑁𝑃𝑒 𝑙 = (3.67)
𝜑 𝐷𝑙
Where x is the coordinate of the system, t is time, 𝑣 is the velocity in the system (𝑣1 ), and L
is the actual core length (L is the characteristic length used to perform the dimensionless
analysis for the geometry). R is the radius of the sphere representing the grain and r is the
position within the sphere.

For the non-equilibrium system, equations (3.44), (3.45), (3.58) and (3.59) take the form:

𝜕𝐶A,1 𝑓1 𝜕 2 𝐶A,1 𝜕𝐶A,1


−𝑏1 + − 𝑁𝑆𝑡 1,2 (𝐶𝐴,1 − 𝐶𝐴,2 ) − 𝑁𝑆𝑡 1,4 (𝐶𝐴,1 − 𝐶𝐴,4 ) = 𝑓1 (3.68)
𝜕𝜒 𝑁𝑃𝑒 1 𝜕𝜒 2 𝜕𝜏
𝜕𝐶A,2 𝜕𝐶A,2
−𝑏2 + 𝑁𝑆𝑡 1,2 (𝐶𝐴,1 − 𝐶𝐴,2 ) − 𝑁𝑆𝑡 2,5 (𝐶𝐴,2 − 𝐶𝐴,5 ) = 𝑓2 (3.69)
𝜕𝜒 𝜕𝜏
𝜕𝐶A,4 ′ (1 − 𝜑)
𝑁𝑆𝑡 1,4 (𝐶𝐴,1 − 𝐶𝐴,4 ) = 𝑓1 (𝐹 4 ) (3.70)
𝜕𝜏 𝜑
𝜕𝐶A,5 ′ (1 − 𝜑)
𝑁𝑆𝑡 2,5 (𝐶𝐴,2 − 𝐶𝐴,5 ) = 𝑓2 (𝐹 5 ) (3.71)
𝜕𝜏 𝜑
Equations (3.70) and (3.71), when used to model the transport of tracer or inert solutes, have
values of F’4 and F’5 equal to 0.

3.2.3 Boundary conditions


The boundary conditions considered when solving the model are Robin or third-type for the
injection, and Newmann or second-type for the production (outlet). Upstream perturbations
include rectangular steps or rectangular impulses with normalized concentrations going from
zero to one, injected directly to the flowing fraction (i.e. into the macro porosity). Walls are
considered impermeable.

The Robin and Newmann boundary conditions (BCs) are known as Danckwerts BCs [11] in
the context of the mass balance. The dimensionless boundary conditions are:

1 𝜕𝐶𝐴,1
𝐶1∗ = 𝐶𝐴,1 − ( ) for 𝜒 = 0 (3.72)
𝑁𝑃𝑒 1 𝜕𝜒
𝜕𝐶𝐴,1
( )=0 for 𝜒 = 1 (3.73)
𝜕𝜒

70
Many publications regarding appropriate boundary conditions are available, including
Hulburt [12] and Danckwerts [11]. More recent discussions have been reported by Mott and
Green [13].

In this study, Danckwerts BCs were used because they are appropriate for finite systems,
which are typically found in laboratories working with core material for reservoir
applications, and also because they satisfy the whole spectrum of flow conditions (0<NPe<∞).

For the initial concentrations of the solute, we have the following initial conditions:

𝐶𝐴,1 = 𝐶𝐴,2 = 𝐶𝐴,3 = 0, 𝜏=0 (3.74)

Some additional boundary conditions for the non-equilibrium system are:

𝑁𝑆𝑡 1,4 = 0 for 𝐶𝐴,4 > 𝐶𝐴,1 (3.75)

𝑁𝑆𝑡 2,5 = 0 for 𝐶𝐴,5 > 𝐶𝐴,2 (3.76)

In the micro porosity domain, there are some extra boundary conditions for the micro-meso
porosity interface. The convective mass transfer between meso and micro porosity equals the
diffusion at the surface of the grain where the micro porosity resides. Therefore, the boundary
conditions for the interface are:

𝑎 𝜕𝐶A,3
𝑁𝑆𝑡 1,2 (𝐶𝐴,3 − 𝐶𝐴,2 ) = 𝑛 𝑓3 (− ) (3.77)
𝑁𝑃𝑒 2,3 𝜕𝜁 𝜁=1

𝑎 1 𝜕 𝜕𝐶A,3 𝜕𝐶A,3 (1 − 𝜑)
( (𝜁 𝑛−1 )) = (1 + 𝐹 ′ 3 ) (3.78)
𝑁𝑃𝑒 2,3 𝜁 𝑛−1 𝜕𝜁 𝜕𝜁 𝜕𝜏 𝜑

The micro porosity grain structure can be modeled as a solid cylinder using a coefficient of
2 or 3 for a spherical grain (n= 2 or 3). For a rectangular slab-type, a coefficient of 1 should
be used (n = 1). In this work, a reversible Langmuir isotherm was used.

3.2.4 Methodology for model application


Model parameter adjustment

Various experimental data sets from the literature [2]–[4] were studied to test if
simplifications in the model would reproduce the observed behavior. In this section, we
assumed macro and meso porosity only, one-dimensional (1-D) transport, and local

71
equilibrium (equations (3.60) and (3.61) with boundary and initial conditions as established
in equations (3.74)‒(3.76)). Simplifications imposed for comparison purposes are those used
in each reference.

Model adjustment was made through comparison of experimental data sets as reported in
three references (See Table 3.3) with model calculations that resulted from the best fit of the
adsorption and desorption coefficients. The rest of the parameters were obtained directly
from the corresponding references. The parameters used for each comparison (Nst, NPe, f1, f2
and φ) are shown in the Results and discussion section.

Table 3.3. Description of experimental data sets and their tags for model adjustment.

Experimental data set Model results tag and its description


C1-T Tracer in high mesoporosity
High meso porosity (C1) [4].
C1-S Surfactant in high mesoporosity
C2-T Tracer in high macroporosity
High macro porosity (C2) [4].
C2-S Surfactant in high macroporosity
C3-T Long injection of tracer
Long injection (C3) [3].
C3-S Long injection of surfactant
C4-T Short injection of tracer
Short injection (C4) [2].
C4-S Short injection of surfactant

3.2.5 Adsorption model analysis


After model adjustment, six key elements that affect adsorption were analyzed to determine
their importance in the prediction of transport in porous media. Table 3.4 and Table 3.5
present the parameter values used for each case. In the following subsections, the importance
of each selected element is presented. These elements are: Isotherm model (IT), Isotherm
parameters (IP), Downstream boundary condition (DB), Slug size (SS), Non-local
equilibrium (NE) and Porosity (P)

Isotherm model (IT)

Many researchers consider only the plateau region of the isotherm to be important for
adsorption, neglecting the shape of the isotherm used. To test this assumption, two different

72
irreversible isotherms will be compared: Langmuir (equation (3.51)) and Redlich-Peterson
(equation (3.53)), solving equations (3.60) and (3.61) for macro and meso porosity.

Isotherm parameters (IP)

The shape of the isotherm is often ignored in this type of calculations, although it may have
an impact on the adsorption of the solute.

To examine the effect of the isotherm parameters on the adsorption process, results will be
obtained with two different irreversible Langmuir isotherms that show the same surfactant
concentration in solid phase at equilibrium with an aqueous concentration equal to one (this
is a normalized concentration).

Downstream boundary condition (DB)

There is some debate on how to consider the downstream boundary condition. The system to
be modeled is generally considered as a solid porous medium of infinite length rather than a
finite solid, for better agreement with experimental results. In this case, the adsorption of
solute at x = 35 cm in a system of length L=35 cm will be compared with the solute adsorption
at x = 35 cm in a system of length 5 L=178 cm, which is considered the closest to infinite
boundary condition. Tests will be performed both with high dispersion (NPe = 5) and low
dispersion (NPe = 20). An irreversible Langmuir isotherm will be used.

Slug size (SS)

The effect of slug size (rectangular signal) on the adsorption model calculations in the
unsaturated case will be tested using irreversible Langmuir isotherms.

Non-local equilibrium (NE)

Both local and non-local equilibrium assumptions will be applied to the adsorption model
using irreversible Langmuir isotherms. Results will be compared to determine when a non-
equilibrium assumption may be important. Only macro and meso porosity will be considered.

Porosity (P)

The effect of micro porosity in adsorption processes is often neglected under the assumption
that it has a weak impact. This may not be completely true and may depend on the proportion
of the micro porosity conforming the porous network, and on the rock’s properties. To test
this effect, equations (3.60) to (3.62) will be solved using irreversible Langmuir isotherms.

73
Table 3.4. Adsorption key elements for analysis and model parameters used in calculations.
Adsorption model element Tag Nst 1,2 Nst 2,3 Nst 1,4 Nst 2,5 NPe,1 NPe 2,3 A1 k1 n1 A2 k2 n2 PV
Langmuir IT-L 1 - - - 10 - 1.9 4 1 - - - 3
Isotherm
Type Redlich-Peterson IT-RP 1 - - - 10 - 6 5.17 2 0.21 0.28 1 3
Langmuir IT-T 1 - - - 10 - 0 0 1 - - - 3
IP-A 1 - - - 20 - 10 25.33 1 - - - 3
Isotherm
Parameters IP-B 1 - - - 20 - 1.9 4 1 - - - 3

DB-L1-Pe5 1 - - - 5 - 1.9 4 1 - - - 3

Downstream DB-L1-Pe20 1 - - - 20 - 1.9 4 1 - - - 3


Boundary
Condition DB-L5-Pe5 1 - - - 5 - 1.9 4 1 - - - 3

DB-L5-Pe20 1 - - - 20 - 1.9 4 1 - - - 3

Slug Size SS-0.5,3a 1 - - - 20 - 1.9 4 1 - - - 0.5-3

NE1-T 1 - - - 20 - 0 0 1 - - - 1.46

NE1-S-Nst0.1-10b 1 - 0.1-10 0.1-10 20 - 1.9 4 1 - - - 1.46


Non-Local
Equilibrium
NE2-T 1 - - - 20 - 0 0 1 - - - 1.46

NE2-S-Nst0.1-10c 1 - 1 0.1,10 20 - 1.9 4 1 - - - 1.46

P-M,m-T 1 - - - 69.43 - 0 0 1 - - - 5.48

P-M,m-S 1 - - - 69.43 - 1.9 4 1 - - - 5.48


Porosity
P-M,m, 1 1 - - 69.43 69.43 0 0 1 - - - 5.48

P-M,m,S 1 1 - - 69.43 69.43 1.9 4 1 - - - 5.48


a The values go from 0.5 to 3, with a step of 0.5.
b The values for Nst14 = Nst25 were 0.1, 0.5, 1 and 10.
c The values for Nst were 0.1 and 10.
25
- Denotes a parameter that does not apply for that case

74
Table 3.5. Porosity (φ), macroporosity fraction (f1) and mesoporosity fraction (f2) used for
analysis of key element.

Adsorption model element φ f1 f2


Isotherm type
Isotherm parameters
0.45 0.9 0.1
Downstream boundary
Slug size
Non-Equilibrium 0.5 0.8 0.2
P-M,m-T
0.5 0.5 0.5
P-M,m-S
Porosity
P-M,m,
0.5 0.5 0.25
P-M,m,S

3.2.6 Parameters analysis for calculations in 2-D and


axisymmetric porous media
As said in the introduction, experimentally, the analysis of parameters in the adsorption of
surfactant, is carried out with rectangular cores (for consolidated porous media) and
cylindrical cores (for non-consolidated porous media), hence, 2-D calculations were used to
model a rectangular core, and axisymmetric calculations to model a cylindrical core.

2-D calculations are more similar to the real cases of flow through porous media in soils and
rocks. Also, these calculations can be useful to easier address the effects of a heterogeneous
permeability, variable diffusion/dispersion, among others effects, and can provide a shorter
solution time. Axisymmetric calculations are more similar to laboratory experiments, but
takes longer computational time to be solved.

2-D modeling

Tests in CFD solver were done, in order to address the effect of heterogeneities in the porous
media, which may be present in natural rock formations. These calculations were done in 2-
D, using a porosity of 0.45 for all cases and considering only macroporosity, the rest of the

75
properties were evaluated in agreement with Table 3.6. Figure 3.3 shows the system where
calculations were carried out.

Figure 3.3. Model showing the heterogeneous porous media where 2-D calculations were
carried out. Kp1 and Kp2 indicates the permeability, p1 and p2 indicates the pressure applied to
the entrance and outlet respectively, being p1> p 2 and the arrows indicates the flow direction.
The dimensions of this system are 1 x 1 m.

The inspiration for the 2D calculations were experiments in the composed heterogeneous
rock used where results after continuous injection, presented a change in the slope for both
tracer, and surfactant (in an outlet-concentration vs time chart). In this sense, in the present
work different calculations were carried out, addressing a list of parameters for the 2-D
calculations which are is shown next.

Dispersion/ Diffusion effect in a heterogeneous porous media.

In the calculation to explore the effect of dispersion in a heterogeneous porous media, a


pressure difference was applied between the left and right boundaries, the upper and lower
boundaries were impermeable. After injecting a rectangular pulse the effluent was analyzed,
three diffusion/dispersion values were used. The parameters used for this case are shown in
Table 3.6.

Permeability contrast.

For a system of low dispersion calculations using different contrast in permeability were
conducted. The parameters used for this case are shown in Table 3.6.

76
Effect of having a surfactant blend where each component interacts with the rock
independently.

The same heterogeneous system was analyzed continuously injecting two surfactants that
interact with the rock but not between them, the parameters used for this case are shown in
Table 3.6.

2-D Axisymmetric calculations.

An axisymmetric porous media was modeled, to analyze the heterogeneities effect.

Heterogeneities

To see if heterogeneities in the system may cause an outlet concentration signal for the
surfactant with two peaks under a finite impulse signal. The mechanisms of adsorption in
colloidal solutions is complex, and can be misinterpreted if multiple peaks are found which
are result of complex pore structure more than a complex adsorption mechanism. Two cases
were studied, one with higher inlet pressure (Case 2).

The system used is similar to the one from Figure 3.3 and is shown in Figure 3.4. Parameters
used for these calculations are shown in Table 3.6.

Figure 3.4. Model showing the heterogeneous porous media where axisymmetric calculations
were carried out. Kp1 and Kp2 indicates the permeability, p1 and p2 indicates the pressure
applied to the entrance and outlet respectively, being p1>p2, the arrows indicates the flow
direction and the red dotted line indicates the axisymmetric condition of the system. The
dimensions of this axisymmetric system are 6 x 1 m.

77
Table 3.6. Adsorption key elements for analysis and model parameters used in calculations of
2-D and axisymmetric modeling.

Disp.
p1 p2 kp1 kp2 𝒄𝒎𝟐
Study Tag A k PV
(Pa) (Pa) ( 𝒎𝟐 ) ( 𝒎𝟐 ) ( )
𝒔
DDE-Low 3100 0 1.9 4 50e-12 50e-13 4.8228e-7 2.04
Dispersion/
DDE-Middle 3100 0 1.9 4 50e-12 50e-13 12.8228e-7 2.04
Diffusion effect
DDE-High 3100 0 1.9 4 50e-12 50e-13 48.228e-7 2.04
PC-Low 3100 0 1.9 4 50e-12 50e-16 4.8228e-7 2.04
Permeability
PC-Middle 3100 0 1.9 4 50e-12 50e-14 4.8228e-7 2.04
contrast
PC-High 3100 0 1.9 4 50e-12 50e-13 4.8228e-7 2.04

SBI-T 3100 0 0 0 50e-12 50e-13 4.8228e-7 Continuous


surfactant blend
interaction SBI-S1 3100 0 1.9 4 50e-12 50e-13 4.8228e-7 Continuous
SBI-S2 3100 0 3.8 4 50e-12 50e-13 4.8228e-7 Continuous

Axisymmetric AC1-T 3100 0 0 0 50e-12 50e-13 4.8228e-7 1.58


Case 1 AC1-S 3100 0 1.9 4 50e-12 50e-13 4.8228e-7 1.58

Axisymmetric AC2-T 6200 0 0 0 50e-12 50e-13 1.6076e-7 1.58


Case 2 AC2-S 6200 0 1.9 4 50e-12 50e-13 1.6076e-7 1.58

3.3 Axisymmetric simulations of displacements in


capillary conducts
Two-phase flow in capillary conducts is a phenomena not very wide studied, but is of critical
importance to measure the IFT from experimental dynamic data. Knowing the time needed
for a certain fluid to flow through a capillary conduct, with gravity as the only driving force,
can lead to the determination of the IFT for that system. But the equations needed to do so
may be of complex solution, so Jose Lopez[4] suggest a one-dimension model to determinate
this. But this models assumes a homogeneous flow through the capillary conduct, which may
not always be true. Hence, a comparison with an axisymmetric model in fluent was done.
Moreover, an axisymmetric model could let us see the drop formation in the outlet of the
capillary conduct, with which another technique of determination of the IFT can be applied.
This is of special importance for cases were the IFT is too low (IFT<1e-6 N/m), for the
displacement speed of the fluid remains constant in these cases, so it becomes impossible to

78
determinate the IFT through the fluid displacement speed. However, the drop formed in the
outlet is always affected by the IFT, even at very lows IFT’s, so the importance of knowing
the drop shape.

3.3.1 Experimental setup


For this case, experimentation was needed, due to the scarce information available about this
topic. For the experimentation, the setup shown in Figure 3.5 was used.

Figure 3.5. Experimental setup used for two-phases capillary displacement experiments

As observed in Figure 3.5 a capillary conduct filled with oil was held into a vessel filled with
water. Then, time was counted as water began to replace the oil, while the oil escaped by the
upper part of the conduct. Oil height was measured for different times, obtaining data of
height vs time. Once this data was obtained, a comparison with Jose Lopez model and Fluent
model was done.
Two different oils were used for comparison. Blue oil (more viscous) and red oil (less
viscous). The properties of these oils are shown in Table 3.7. This values were used to obtain
the results with the one-dimensional model and with fluent model. The dimensions of the
capillary conduct used in the experiments were L: 54 mm, ID= 0.59436 mm and OD= 0.8636
mm.

79
Table 3.7. Properties for oils used in experiments

Parameter Red oil Blue oil


Density (Kg/m3) 860.91 816.66
Viscosity (Kg/m-s) 0.006856 0.1315

Also, surface tension was obtained from the drop formed in the outlet of the capillary
conducts through an open source code by Berry et al[17]. This was done for the experimental
results and then compared with the models results.

3.3.2 Model used


These capillary displacement simulations were done using the Volume Of Fraction (VOF)
model included in Fluent (CFD software). The tracking of the interface(s) between the phases
is accomplished by the solution of a continuity equation for the volume fraction of one (or
more) of the phases. The VOF model is described as follows according to ANSYS Inc[14].

𝑛
1 ∂
[ (𝛼 𝜌 ) + ∇ ∙ (𝛼𝑞 𝜌𝑞 ⃗⃗⃗⃗
𝑣𝑞 ) = 𝑆𝛼𝑞 + ∑(𝑚̇𝑝𝑞 − 𝑚̇𝑞𝑝 )] (3.79)
𝜌𝑞 ∂t 𝑞 𝑞
𝑝=1

Where 𝑚̇𝑞𝑝 is the mass transfer from phase q to phase p and 𝑚̇𝑝𝑞 is the mass transfer from
phase p to phase q. By default, the source term on the right-hand side of Equation (3.79) 𝑆𝛼𝑞
is zero, but you can specify a constant or user-defined mass source for each phase.

Surface tension

The VOF model can also include the effects of surface tension along the interface between
each pair of phases. The model can be augmented by the additional specification of the
contact angles between the phases and the walls. You can specify a surface tension coefficient
as a constant, as a function of temperature through a polynomial, or as a function of any
variable through a user defined function (UDF). Variable surface tension coefficient effects
are usually important only in zero/near-zero gravity conditions.[15]

In ANSYS Fluent, two surface tension models exist: the continuum surface force (CSF) and
the continuum surface stress (CSS). For these simulations, CSF model was used (which is
the default model).

80
The continuum surface force (CSF) model proposed by Brackbill et al[16] has been
implemented such that the addition of surface tension to the VOF calculation results in a
source term in the momentum equation. For two phases, this source term can be expressed
as follows.

(𝜌𝑘𝑖 𝑛)
𝐹𝑣𝑜𝑙 = 𝜎𝑖𝑗 (3.80)
1
2 (𝜌𝑖 + 𝜌𝑗 )
Where 𝜌 is the volume average density, 𝜌𝑖 and 𝜌𝑗 are the phases density, 𝑘𝑖 is the surface
curvature for phase i and 𝑛 is the normal vector for surface i.

Contact angle

An option to specify a wall contact angle in conjunction with the surface tension model is
also available in the VOF model. The model is taken from work done by Brackbill et al.[16]
Rather than impose this boundary condition at the wall itself, the contact angle that the fluid
is assumed to make with the wall is used to adjust the surface normal in cells near the wall.
This so-called dynamic boundary condition results in the adjustment of the curvature of the
surface near the wall.[15]

If 𝜃𝑤 is the contact angle at the wall, then the surface normal at the live cell next to the wall
is.
𝑛 = 𝑛𝑤 cos 𝜃𝑤 + 𝑡𝑤 sin 𝜃𝑤 (3.81)

Where 𝑛𝑤 and 𝑡𝑤 are the unit vectors normal and tangential to the wall, respectively. The
combination of this contact angle with the normally calculated surface normal one cell away
from the wall determine the local curvature of the surface, and this curvature is used to adjust
the body force term in the surface tension calculation.

3.3.3 Domain definition and boundary conditions


For these simulations an axisymmetric model was used. The model consists in a circular
capillary conduct filled with oil and submerged in water. Oil goes out from the upper side of
the capillary conduct as water enters by the lower part, this motion is caused by the buoyancy
forces. The model and its specifications are shown in Figure 3.6.

81
a) b)
Figure 3.6. Simulated domain details. a) Represents a slice of the complete simulation while b)
shows the actual simulated system.

In accordance with Figure 3.6, the red zone is oil and the blue one is water (immiscible phases
were considered) and the empty rectangle defines the conduct thickness (glass). A length of
50 mm was used for the conduct and different radius were used. For the left border (Figure
3.6 b), an axisymmetric condition was selected. All the other borders were defined as walls
with adhesion. A contact angle was specified for the water phase on these walls. Also, an
interfacial tension was specified for the oil-water interface. An atmospheric pressure of 1
Atm and a temperature of 25 C° were specified for the system. Constant temperature was
assumed. A gravitational force going downwards Y axis was assumed. For the initial time,
the conduct was filled with oil, as shown in Figure 3.6.

3.3.4 Methodology for model application


As mentioned before, for this system resolution, ANSYS Fluent software was used, using its
VOF module. To solve the model we need to know the system and fluids properties values.
All of these are shown in Table 3.8. Interfacial tension and conduct radius were varied to
address its effects on the displacement speed and emptying time. The results were compared
with those obtained using the one-dimensional model proposed by Jose Luis Lopez [4].

82
Table 3.8. Parameters for axisymmetric simulations.

Parameter Water Oil


Density (Kg/m3) 1000 900
Viscosity (Kg/m-s) 0.001 0.005
Contact angle (°) 30 150
Interfacial tension (N/m) 5e-5, 5e-4, 5e-3
Conduct radius (mm) 0.1, 0.2, 0.3
Conduct length (mm) 50

83
3.4 References
[1] R. Kulkarni, a. T. Watson, J.-E. Nordtvedt, and A. Sylte, “Two-phase flow in porous

media: Property identification and model validation,” AIChE J., vol. 44, no. 11, pp.

2337–2350, 1998.

[2] K. Mannhardt, L. L. Schramm, and J. J. Novosad, “Adsorption of anionic and

amphoteric foam-forming surfactants on different rock types,” Colloids and Surfaces,

vol. 68, no. 1–2, pp. 37–53, Nov. 1992.

[3] K. Mannhardt and J. J. Novosad, “Chromatographic effects in flow of a surfactant

mixture in a porous sandstone,” J. Pet. Sci. Eng., vol. 5, no. 2, pp. 89–103, Feb. 1991.

[4] J. L. Lopez-Salinas, “Transport of Components and Phases in a Surfactant/Foam,”

Rice University, 2013.

[5] M. C. Leverett, “Capillary Behavior in Porous Solids,” Trans. AIME, vol. 142, no. 1,

pp. 152–169, Dec. 1941.

[6] A. Faghri and Y. Zhang, Transport phenomena in multiphase systems. USA:

Academic Press, 2006.

[7] K. K. Mohanty, “Multiphase Flow in Porous Media: III. Oil Mobilization, Transverse

Dispersion, and Wettability,” in SPE Annual Technical Conference and Exhibition,

1983.

[8] R. N. Y. M. Y. Jia and F. Z. X. Y. X. Niu, “Dual Porosity and Dual Permeability

Modeling of Horizontal Well in Naturally Fractured Reservoir,” Transp. Porous

Media, vol. 92, pp. 213–235, 2012.

[9] V. K. C. Lee, J. F. Porter, G. McKay, and A. P. Mathews, “Application of solid-phase

84
concentration-dependent HSDM to the acid dye adsorption system,” AIChE J., vol.

51, no. 1, pp. 323–332, Jan. 2005.

[10] N. S. Maurya and A. K. Mittal, “ Applicability of Equilibrium Isotherm Models for

the Biosorptive Uptakes in Comparison to Activated Carbon-Based Adsorption,” J.

Environ. Eng., vol. 132, no. 12, pp. 1589–1599, Dec. 2006.

[11] P. V. Danckwerts, “Continuous flow systems: Distribution of residence times,” Chem.

Eng. Sci., vol. 2, no. 1, pp. 1–13, Feb. 1953.

[12] H. M. Hulburt, “Chemical Processes in Continuous-Flow Systems,” Ind. Eng. Chem.,

vol. 36, no. 11, pp. 1012–1017, Nov. 1944.

[13] H. V Mott and Z. A. Green, “On Danckwerts’ Boundary Conditions for the Plug-Flow

with Dispersion/Reaction Model,” Chem. Eng. Commun., vol. 202, no. 6, pp. 739–

745, Jun. 2015.

[14] ANSYS, “Volume Fraction Equation,” 2001. [Online]. Available:

https://www.sharcnet.ca/Software/Ansys/17.0/en-

us/help/flu_th/flu_th_sec_vof_eq.html.

[15] ANSYS, “Surface Tension and Adhesion,” 2001. [Online]. Available:

https://www.sharcnet.ca/Software/Ansys/17.0/en-

us/help/flu_th/flu_th_vof_surf_tens.html#flu_th_wall_adhesion.

[16] J. U. Brackbill, D. B. Kothe, and C. Zemach, “A continuum method for modeling

surface tension,” J. Comput. Phys., vol. 100, no. 2, pp. 335–354, 1992.

[17] J. D. Berry, M. J. Neeson, R. R. Dagastine, D. Y. C. Chan, and R. F. Tabor,

“Measurement of surface and interfacial tension using pendant drop tensiometry,” J.

Colloid Interface Sci., vol. 454, pp. 226–237, 2015.

85
Chapter IV: Results and discussion
In this chapter, results from all the simulations that were carried out throughout all the work
are shown. A detailed description is presented for each result obtained. The outcomes of the
simulations of two phase flow in porous media is presented. As well as the results for the
modeling of surfactant adsorption/retention and transport through porous media.

In this chapter, also the results for the axisymmetric simulations of displacements in capillary
conducts are shown. A detailed analysis is made for all the results presented.

4.1. Results of multiphase flow in saturated porous


media
In this section, the results for the model of multiphase flow in saturated porous media is
presented. The solution using the coefficients suggested in the literature are compared with
the results obtained by Kulkarni et al. Moreover, the results obtained using adjusted
coefficients are shown, as well as its comparison with Kulkarni et al work.

Finally, a sensitivity analysis was made to determinate the impact of the variable a in the
flow of both phases in the porous media. Additionally, an analysis of the numerical solver
problems is done.

4.1.1. Adjusted parameters


Experimental and theoretical results were used to adjust model parameters, which are shown
in Table 4.1.

Table 4.1. Adjusted model parameters for comparison with Kulkarni’s results

Parameter Value
aw 7.836
ao 5.29
Kro0 0.8557
Krw0 1.019
σ [mN/m] 2.56
θ [°] 0
α 1.2

86
In Figure 4.1saturation of water through time is compared with those results obtained by
Kulkarni et al. The model predicts, as expected, that the oil slowly displaces the water from
the porous media, leaving always a residual water concentration. For the model used in this
work, there exists an overestimation for the water displaced in the initial time, but this
difference starts to decrease as time increase leading, however, to a notorious difference in
the response concerning Kulkarni’s results. The difference may be caused by the capillary
pressure model used, it may not work properly under this type of permeability curves. The
correlations values obtained for Kulkarni for each time are: 0.9643, 0.985, 0.981 and 0.9683
respectively. For this work the correlation values obtained are: 0.9438, 0.9823, 0.9355 and
0.9572.

1.0
0.9
0.8
0.7
Water Saturation

0.6
0.5
0.4
0.3
0.2 Experimental results (Kulkarni)
Model results (Kulkarni)
0.1 Model results (This work)

0.0
0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9 1.0
Dimensionless Lenght

Figure 4.1. Simulation results for 20, 70, 100 and 250 minutes, respectively, using the values of
the adjusted coefficients

4.1.2. Sensitivity analysis for alpha coefficient


In Figure 4.2, the different colored curves are the quantification of the water saturation. Using
a smaller alpha value leads to a speed-up water saturation signal reaching the outlet. This
results in a difference in the saturation time of the porous media. Moreover, using a larger
value of alpha leads to a delay in the saturation speed of the porous media. Alpha value have
such important impact in water saturation due to the properties it includes (Interfacial tension,
contact angle, porosity, length and permeability).

87
Figure 4.2. Water saturation for different alpha values at 20, 70, 100 and 250 minutes,
respectively.

The properties of the phases are very difficult to obtain. Thus, Leverett's equation for
capillary pressure can be useful when determining some of these properties, such as
interfacial tension and contact angle.

4.2. Results of modeling surfactant adsorption and/or


retention
Experimental results [1] were used to adjust model parameters (A and k) which are shown in
Table 4.2.

88
Table 4.2. Adjusted model parameters using irreversible isotherm for comparison with
experimental results

Tag NSt NPe φ f1 A k PV R2


C1-T 0.7 10 0.14 0.275 0 0 Continuous 0.947
C1-S 1 4.125 0.14 0.275 1 0.65 Continuous 0.987
C2-T 1 18 0.14 0.9 0 0 Continuous 0.996
C2-S 1 6.75 0.14 0.9 0.25 2 Continuous 0.971

Figure 4.3 shows the comparison of the model results to experimental data of normalized (C0
= 1% wt) concentration as a function of pore volume (PV). The model predicts, as expected,
that the tracer exits first followed by the surfactant. For C1-T model results, there exists a
delay in the response with respect to experimental results. The high macro porosity in C2 (90
%) resulted in rapid and total recovery of both tracer and surfactant. For C1-S model results,
agreement with experiment is very good. The difference between model and experimental
results for C1-T may be caused by an erroneous determination of the rock properties such as
porosity or macroporosity fraction.

1.0 1.0

0.8 0.8

0.6 0.6
C/C0

C2-T
C/C0

0.4 C1-T 0.4 C2-S


C1-S Tracer (Experimental)
0.2 Tracer (Experimental) 0.2
Surfactant (Experimental)
Surfactant (Experimental)
0.0 0.0

0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5 6 7 8 9 10
PV PV

(a) (b)
Figure 4.3. Exit concentration (C/C0) as a function of pore volume (PV). Comparison between
computed and experimental results obtained for (a) C1-T (Tracer) and C1-S (Surfactant) and
(b) C2-T (Tracer) and C2-S (Surfactant).

Experimental results [2], [3] were used to adjust hybrid isotherm parameters which are shown
in Table 4.3.

89
Table 4.3. Adjusted model parameters using hybrid isotherm for comparison with
experimental results.

Macro porosity Meso porosity


Tag Nst Npe φ f1 region (M) region (m) PV R2
A1 k1 A2 k2 A1 k1 A2 k2
C3-T 1 50 0.2 0.9 0 0 0 0 0 0 0 0 3 0.9953
C3-S 1 15 0.2 0.9 1.58 2 0.02 0.02 0.95 2 1.9 2 3 0.9605
C4-T 1 50 0.182 0.9 0 0 0 0 0 0 0 0 1.46 0.9887
C4-S 2 15 0.182 0.9 0.4 18 0.6 18 2 18 0 0 1.46 0.9898

Figure 4.4 shows how the model results compare to the experimental data of normalized
concentration as a function of pore volume (PV). For C3-S there are differences after 5 PV.
It is also important to note that after this value of PV, a constant experimental outlet surfactant
concentration is observed. The results from C3-T seem to have a better adjustment with
experiment than C3-S. For C4-S, after adjusting our model parameters to the experimental
data, a better agreement is obtained than with the model proposed by Mannhardt for PV >2.5.
Considering meso porosity in our model may result in this improved agreement.

1.0 C3-T 1.0 C4-T


C3-S C4-S
0.8 0.8
Mannhardt C3 Model Mannhardt
0.6 Tracer 0.6 C4 Model
(Experimental) Tracer
C/C0

C/C0

0.4 Surfactant 0.4 (Experimental)


(Experimental) Surfactant
0.2 0.2
(Experimental)
0.0 0.0

0 1 2 3 4 5 6 7 8 9 10 0 1 2 3 4 5
PV PV

(a) (b)
Figure 4.4. Exit concentration (C/C0) as a function of pore volume (PV). Comparison between
computed (our model and Mannhardt´s model) and experimental results obtained for (a) C3-
T (Tracer) and C3-S (Surfactant) and (b) C4-T (Tracer) C4 and C4-S (Surfactant)

90
4.2.1. Model element analysis
Isotherm model

In Figure 4.5, the difference in the area under the curve between the tracer response and any
of the two responses corresponding to the two types of isotherms, is the quantification of the
surfactant adsorption. Using the Redlich-Peterson isotherm leads to a delay in the surfactant
signal reaching the outlet. There is also a steeper slope than when using the Langmuir
isotherm. This results in a difference in the adsorption between the two isotherms. This is
important because, as mentioned before, many researchers see only the plateau region of the
isotherm as being an important value for the adsorption.

Table 4.4 shows the adsorption for both isotherms. We can see that the Redlich-Peterson
isotherm has higher adsorption than the Langmuir isotherm according to the parameters used.

1.0
IT-T
0.8 IT-L
IT-RP
0.6
C/C0

0.4

0.2

0.0

0 1 2 3 4 5 6
PV

Figure 4.5. Model prediction of effluent concentration for tracer (IT-T), Langmuir isotherm
(IT-L) and Redlich-Peterson (IT-RP) for a rectangular pulse of 3 PV.

Table 4.4. Surfactant adsorption (in milligrams of surfactant per gram of rock) for model
prediction when varying isotherm type.

Tag Adsorption
mg/g
IT-L 0.41
IT-RP 0.52
IT-T 0

91
Isotherm Parameters

In Figure 4.6 (a) model results for two isotherms of identical analytical form, but with
parameter values as presented for IT-A and IT-B in Table 3.3, are plotted together. Both
isotherms reach the same solid phase concentration at an aqueous phase concentration equal
to 1, but presenting a different shape. Isotherm A has a steeper slope than isotherm B,
producing a larger area under the curve before reaching the concentration equal to 1 in the
aqueous phase. Usually, it would be considered that both isotherms cause the same
adsorption, but in Table 4.5 the difference can be seen. In Figure 4.6 (b), we can see that the
response corresponding to isotherm A, with the larger area under the curve, has a larger
adsorption than the response corresponding to isotherm B. It is therefore possible to infer that
the shape of the curve has a significant influence on the adsorption analysis, although the
isotherms are of the same analytical form. In addition, the effluent concentration resulting
from the use of isotherm A shows a steeper slope than isotherm B. These patterns may change
the interpretation of results and the formulation of conclusions by researchers when analyzing
experimental results.

1.0

0.4 0.8

0.6
C(s) (mg/g)

C/C0

0.2 0.4 IP-T


Isotherm "A"
IP-A
Isotherm "B"
0.2 IP-B

0.0 0.0

0 1 2 0 1 2 3
C(aq) (% wt) PV

(a) (b)
Figure 4.6. Model prediction of (a) solid phase (𝑪(𝒔) ) and aqueous (𝑪(𝒂𝒒)) concentration for
two Langmuir isotherms: isotherm A (IP-A: A=10, k=25.32) and isotherm B (IP-B: A=1.9,
k=4) and (b) effluent concentration for two types of Langmuir isotherm (parameters as in fig
8(a)) and for the tracer (IP-T).

92
Table 4.5. Surfactant adsorption (in milligrams of surfactant per gram of rock) for model
prediction when varying isotherm parameters.

Tag Adsorption mg/g


IP-A 0.568
IP-B 0.488
IP-T 0

Downstream Boundary Condition

The effect of considering different boundary conditions at the outlet on surfactant


concentration as a function of pore volume is presented in Figure 4.7. Two different core
lengths were considered, L = 35 cm (DB-L1) and L = 178 cm (DB-L5). For the same core
position, results for DB-L5 show a displaced output signal toward the right with respect to
results from DB-L1, and this is more evident for a Peclet number of 5. This may be caused
by the effect of diffusion on the transport of surfactant in the porous media. Whereas in the
DB-L1 calculations the χ = 1 position represents the end of the core, for DB-L5 the same
position represents a continuity in the core.

Diffusion and dispersion in the outlet and in the inlet may modify the results, which in this
case resembles a zero type boundary condition at the outlet. As a result, an increasing
variation in the output signals (surfactant concentration) is related to an increase in diffusion
and/or dispersion. This is caused by small Peclet numbers.

For Peclet numbers greater than 20, the effect of the outlet boundary condition seems
negligible. On the contrary, for systems with low Peclet numbers (high dispersion and/or
diffusion, e.g., NPe <5), the boundary condition at the outlet has a great impact on the
downstream response. Choosing the appropriate boundary condition (i.e., dC/dχ=0 at the
outlet of a relatively small core instead of C=0 at the outlet of a very long core) is a decisive
factor for the analysis when relatively short cores are used in experimental tests.

93
1.0
DB-L1-Pe20
0.8 DB-L1-Pe5
DB-L5-Pe20
0.6 DB-L5-Pe5
C/C0
0.4

0.2

0.0

0 1 2 3 4 5 6
PV
Figure 4.7. Model results considering different downstream boundary conditions (continuity
and discontinuity) at the location x = 35 cm, and using two different Peclet numbers (5 and
20) (DB: Downstream Boundary condition, Core Length: L1 =35 cm and L5 = 178 cm, Peclet
number: Pe5 = 5 and Pe20 = 20)

Slug Size

The effect of varying slug size on effluent concentration and on surfactant adsorption is
shown in Figure 4.8 (a) and (b), respectively. The legend of Figure 4.8 (a) shows the slug
size in pore volume following the SS identifying tag. In Figure 4.8 (a), we can see that for
slug size less than 2, an outlet concentration less than 1 is observed. Thus, for slug sizes grater
or equal to 2, a complete saturation of the rock is guaranteed. Figure 4.8 (b) shows the impact
of slug size on adsorption. For values of slug size above 1, saturation is observed. The use of
slug sizes higher than those needed to reach the adsorption saturation is strongly
recommended in order to obtain reliable results and to avoid non-saturation being confused
with porosity effects.

1.0 0.5
SS-0.5
0.8 SS-1 0.4
Adsorption (mg/g)

SS-1.5
0.6 0.3
SS-2
C/C0

SS-2.5 0.2
0.4 SS-3
0.2 0.1

0.0 0.0

0 1 2 3 4 5 6 0 1 2 3
PV Slug size (PV)

(a) (b)
Figure 4.8. Model prediction of (a) outlet concentration using different slug sizes (SS: Slug
Size or injection size) and (b) adsorption obtained using different slug sizes.

94
Non-Local Equilibrium

Figure 4.9 shows surfactant concentration as a function of pore volume for varying Nst. Figure
4.9 (a) shows results for Nst 1,4 = Nst 2,5 while Figure 4.9 (b) keeps Nst 1,4 = 1 and varying Nst
2,5. The values of Stanton number can be found in the legends of Figure 4.9 following the Nst
tag. Figure 4.9 (a) shows that, as both Stanton numbers increase, the response approaches
that of instantaneous equilibrium. In Figure 4.9 (b) it is observed that changing Nst 2,5 by a
factor of 100, has little impact on the outlet concentration. Hence, Nst 1,4 seems to dominate
in reaching the equilibrium state. This may be because the mass transfer in the macro porosity
section seems to regulate the overall adsorption due to it being the section in direct contact
with the fluid phase, causing a bottleneck when its Stanton number (Nst14) is lower than that
of the meso porosity section (Nst25). Stanton numbers are principally controlled by the
velocity (v1) and the specific area (av) of the solid surface.

1.0 1.0
NE1-T NE2-T
0.8 NE1-S-Nst0.1 0.8 NE2-S-Nst0.1
NE1-S-Nst0.5 NE2-S-Nst10
0.6 0.6
NE1-S-Nst1 Instant equilibrium
C/C0

C/C0

0.4 NE1-S-Nst10 0.4


Instant equilibrium
0.2 0.2

0.0 0.0

0 1 2 3 4 5 6 0 1 2 3 4 5
PV PV

(a) (b)
Figure 4.9. Model prediction of effluent concentration for tracer (NE1-T and NE2-T) and
surfactant (NE2-S and NE2-S) as a function of pore volume (PV). (a) Using the same value for
Nst14 and Nst25 in non-equilibrium and (b) using a constant value of Nst14 = 1 while varying
Nst25 in non-equilibrium. In both legends, the value for the Stanton number that varies
appears after the Nst tag.

Porosity

The effect of microporosity on the model outlet concentration prediction to a rectangular


pulse injection is shown in Figure 4.10, for tracer and surfactant. The figure depicts four
results identified in its legend as follows. For tracer (T) and for surfactant (S), the response
obtained from a medium that is composed of only macro (M) and meso (m) porosity, is
compared to that resulting from considering macro (M), meso (m) and micro (μ) porosity.

95
The flowing fraction (f1) in all cases presented in Figure 4.10 remains the same (i.e., same
macro porosity), and when the case of microporosity being present is considered (P-M,m,µ),
both tracer and surfactant penetrate to a lesser degree. Additionally, the output signal takes
more time to reach zero concentration. This may be caused by the mechanism of
adsorption/desorption in the meso-micro porosity interface, where only diffusion occurs, so
that the surfactant penetrates slowly into the micro porosity and leaves it at the same rate,
taking more time to completely leave the system. When only macro and meso porosity were
considered (P-M,m), the concentration response had a lower maximum than when micro
porosity is included in the model. This may be caused by the diminishing meso porosity
fraction (from 50% to 25%) when including micro porosity. As result, the ratio of macro to
meso porosity fraction doubles and the flowing phase enters and leaves the system faster,
causing a higher maximum concentration.

1.0
P-M,m-S
0.8 P-M,m,-S
P-M,m-T
0.6 P-M,m,-T
C/C0

0.4

0.2

0.0

0 1 2 3 4 5 6 7 8 9 10 11 12 13 14 15
PV

Figure 4.10. Model prediction of outlet surfactant and tracer concentrations as a function of
pore volume for different porous network scenarios. Continuous lines represent the surfactant
(S) concentration and the dashed lines represents tracer (T) concentration. The porous
network considered is described in the legend as: Macro (M), meso (m) and micro (μ)
porosity.

4.2.2. Adsorption in heterogeneous porous media 2-D


The results for axisymmetric simulations are shown in the next section. The effect
permeability contrast in a porous media and dispersion/diffusion effect are analyzed. A
heterogeneous permeability as well as a variation in the dispersion/diffusion may affect the
flow patterns of the surfactant, having a significant impact in the adsorption process.

96
Finally, the effect of having a surfactant blend where each component interacts with the rock
independently is analyzed. When more than one surfactant is injected, they can interact with
each other, affecting the outlet concentration lecture.
Permeability contrast in heterogeneous porous media
1.0
PC-Low
PC-Middle
PC-High
0.8

0.6
C/C0

0.4

0.2

0.0
0 1 2 3 4 5 6 7 8
PV

Figure 4.11. Shape of the concentration signal at the effluent of the porous media.

Figure 4.11 shows the behavior of the surfactant in the outlet when different permeability
contrast were used. For the case of high permeability the surfactant presented segregation,
causing two peaks, both with concentration ≈ 0.4. For the middle permeability contrast case,
a different behavior is shown, the surfactant presents just a peak and shows a slow decrement
compared with the first case, the maximum concentration of this case is ≈ 0.38. In the third
case, a low permeability was used, and as we can see in figure 4, the behavior of this case is
almost similar to the case of middle permeability.

In Figure 4.12 a time lapse of the low permeability contrast case is shown, here we can see
how through dimensionless time (Figure 4.12 a to f) the flow path follows the zones with the
higher permeability first, passing through all the zones with lower permeability last, what is
expected behavior on this kind of media.

97
a) b) c)

d) e) f)

Figure 4.12. Time lapse for a 2-D simulation of a heterogeneous porous media. Maximum
normalized concentration (1) is represented with red color, while zero concentration is in
blue. Figure a) is for PV=0.5, b) for PV=1.5, c) for PV=2.5, d) for PV=4, e) for PV=5 and f) for
PV=6

Figure 4.13 a) shows the magnitude of velocity field and equipotential lines, Figure 4.13 (b)
indicates the velocity field direction, both profiles are from the low permeability case. In
Figure 4.13 a) we can see how the velocity field is higher in the regions of higher permeability
and lower in the regions of lower permeability, hence, we can see how the velocity field is
related to the permeability of the zones.

(a) (b)

Figure 4.13. Model showing velocity field during the simulation with lower permeability, Red
zone is the higher pressure, and blue the lowest pressure. a) shows the velocity magnitude and
direction while b) indicates just the velocity direction.

98
Dispersion/ Diffusion effect in a heterogeneous porous media

1.0
DDE-Low
DDE-Middle
DDE-High
0.8

0.6
C/C0

0.4

0.2

0.0
0 1 2 3 4 5 6 7 8
PV

Figure 4.14. Output signal for heterogeneous porous media under the assumption of different
levels of dispersion/diffusion.

Figure 4.14 shows the behavior of the surfactant in the outlet when different values of
dispersion/diffusion were used. For the case of high dispersion the surfactant presented a
typical behavior, which is, an increment and a decrement of the concentration creating a peak
with a maximum value of ≈ 0.6. For the middle dispersion case, a different behavior is shown,
the outlet concentration of the surfactant presents a plateau on its maximum value, instead of
a peak as in first case, and a lower maximum concentration of ≈ 0.4. In the case with low
dispersion the surfactant at the outlet presents two concentration peaks with a maximum
value similar to case of middle dispersion.

99
Effect of having a surfactant blend that each component interacts with the rock
independently.

1.0
SBI-T
SBI-S1
0.8 SBI-S2

0.6
C/C0

0.4

0.2

0.0
0 1 2 3 4 5 6 7 8
PV

Figure 4.15. Response of the concentration downstream for tracer, and two different
surfactants. The two surfactants present no interaction in this simulation.

Figure 4.15 shows the outlet behavior of two surfactants and a tracer, all with continuous
injection, different adsorptions were used for each one of the surfactants. The tracer reaches
the outlet first, just as expected, because it presents no adsorption, so it reach the outlet in PV
= 1, and saturation in PV = 4. The surfactant 1 has a lower adsorption than surfactant 2, so it
is adsorbed in less proportion, hence it goes out first, at PV = 1.9 and reach saturation at PV
= 5. For surfactant 2 a higher adsorption was used, so it is adsorbed in a higher proportion,
retarding the outlet flow to PV = 2.6 and reaching saturation at PV ≈ 6.

100
1.0

SBI-T
Surfactants blend
0.8

0.6
C/C0

0.4

0.2

0.0
0 1 2 3 4 5 6 7 8
PV

Figure 4.16. Tracer and surfactant blend concentration normalized.

Figure 4.16 shows the outlet behavior of tracer and the surfactants blend from Figure 4.15
normalized, all with continuous injection. The tracer behavior is the same as in Figure 4.15,
but the surfactant blend behavior indicates that if heterogeneities are present in the
permeability, the output signal of surfactant blend concentration may present several maxima
and minima slopes.

4.2.3. Axisymmetric 2-D simulations in porous media.


Heterogeneities effects

1.0

AC1-T
AC1-S
0.8

0.6
C/C0

0.4

0.2

0.0
0 1 2 3 4 5 6
PV

Figure 4.17. Surfactant and tracer output normalized concentration for axisymmetric
simulation (Case 1).

101
In Figure 4.17 the results for an axisymmetric simulation is shown, there we can observe the
surfactant and tracer concentrations in the outlet. The tracer goes out first, at PV=1 as
expected, reaching a maximum concentration of 0.9 and a zero concentration approximately
at PV = 5. From the surfactant is possible to observe that it presents adsorption, reaching the
outlet at PV=2.5, and a near zero concentration at PV = 5.

1.0

AC2-T
AC2-S
0.8

0.6
C/C0

0.4

0.2

0.0
0 1 2 3 4 5 6
PV

Figure 4.18. Surfactant and tracer signal for axisymmetric simulations with double peak
(Case 2).

The simulation results for case 2 are shown in Figure 4.18 and indicates that if heterogeneities
are present in the permeability, the output concentration may present a maximums and a
minimums, as in the 2-D Simulations.

102
a) d)

b) e)

c) f)

Figure 4.19. Time lapse for an axisymmetric simulation of a heterogeneous porous media
using a rectangular impulse of surfactant in the injection (Case 1). Maximum normalized
concentration (1) is represented with red color, while zero concentration is in blue. Figure a)
is for PV=0.25, b) for PV=0.5, c) for PV=1.2, d) for PV=2, e) for PV=2.5 and f) for PV=3.

Figure 4.19 shows a time lapse of the tracer of case 1 flowing through the porous media. In
Figure 4.19 (a) we can observe how the surfactant enters in the media. From Figure 4.19 (b)
to (c) the tracer flows to the outlet, is possible to see how flows go first through the regions
of higher permeability, and least in the zones of lower permeability. In Figure 4.19 (d) to (f)
the tracer reach the outlet and begins to leave the media, leaving first the zones of higher
permeability and last the zones of lower permeability.

103
4.3. Results of axisymmetric simulations for
displacements in capillary conducts
4.3.1. Experimental results and comparison
The comparison of the experimental results and the models used for the case of the blue oil,
is shown in Figure 4.20.

50

40

30
h (mm)

20 Fluent results
One-Dimensinal model
Experimental results
10

0
0 200 400 600 800 1000 1200 1400
t (s)

Figure 4.20. Experimental results for blue oil for displacement in capillary conducts and the
results obtained using Fluent model and an one-dimensional model.

In Figure 4.20 we can observe the difference between the one dimensional model and Fluent
model. At first, both results are similar to the experimental results, but after 600 s, the one
dimensional model tends to separate from the experimental results, while the fluent model
stays closer to those results. The IFT was obtained varying it in the one-dimensional model
until its results coincided with the experimental results (in the most part), resulting an IFT of
5.5e-3 N/m. This IFT was used to obtain the Fluent results. The correlation coefficients are
shown in Table 4.6. Is possible to observe that Fluent results have a more similar to the
experimental results than the one-dimensional model.

104
Table 4.6. Correlation coefficients and IFT for the models used in the blue oil experiments

Model R2 IFT (N/m)


Fluent 0.9345 5.5e-3
One-Dimensional 0.9467 5.5e-3

The IFT was also obtained from the drop formed in the outlet of the capillary conduct, and
this results was compared with the IFT obtained with the one-dimensional model. The drop
analyzed is shown in Figure 4.21.

Figure 4.21. Experimental drop analyzed for the blue oil case.

The IFT obtained through Barry et al code was 4.7e-3 N/m. The comparison of the IFT
obtained through both methods is shown in Table 4.7.

Table 4.7. Comparison of IFT’s obtained through two different methods.

Method IFT (N/m) % Difference


One-dimensional model 5.5e-3
17.02%
Barry et al. code 4.7e-3

The comparison of the experimental results and the models used for the case of the red oil,
is shown in Figure 4.22.

105
50

40

30
h (mm)

20 Fluent results
One-Dimensinal model
Experimental results
10

0
0 20 40 60 80 100 120 140
t (s)

Figure 4.22. Experimental results for red oil for displacement in capillary conducts and the
results obtained using Fluent model and an one-dimensional model.

In Figure 4.20 we can observe the difference between the one dimensional model and Fluent
model. At first, both results are similar to the experimental results, but after 60 s, the one
dimensional model tends to separate from the experimental results, while the fluent model
stays closer to those results. The IFT was obtained in the same way as in the past case,
resulting an IFT of 4e-3 N/m. This IFT was used to obtain the Fluent results. The correlation
coefficients are shown in Table 4.8. Is possible to observe that Fluent results have a more
similar to the experimental results than the one-dimensional model.

Table 4.8. Correlation coefficients and IFT for the models used in the blue oil experiments

Model R2 IFT (N/m)


Fluent 0.9337 4e-3
One-Dimensional 0.9463 4e-3

As in the past case, IFT was also obtained from the drop formed in the outlet of the capillary
conduct, and this results was compared with the IFT obtained with the one-dimensional
model. The drop analyzed is shown in Figure 4.23.

106
Figure 4.23. Experimental drop analyzed for the red oil case.

The IFT obtained through Barry et al code was 4.5e-3 N/m. The comparison of the IFT
obtained through both methods is shown in Table 4.9.

Table 4.9. Comparison of IFT’s obtained through two different methods.

Method IFT (N/m) % Difference


One-dimensional model 4e-3
11.11%
Barry et al. code 4.5e-3

4.3.2. Conduct radius effect


As mentioned before, different conduct radius were used to address its effects in the capillary
conduct emptying. Radius of 0.1, 0.2 and 0.3 mm were used with a constant interfacial
tension of 5e-5 N/m. In the next paragraphs the results are shown.

a) b) c)

Figure 4.24. Drop shape for varying conduct radius. a) drop shape for r=0.1 mm, b) r= 0.2
mm and c) r=0.3mm

107
Figure 4.24 shows the drop shape for different conduct radius, using the same IFT (5e-5 N/m).
When a smaller radius is used (Figure 4.24 a), the drop formed tends to be smaller. When an
intermediate radius is used (Figure 4.24 b), an elongation of the drop can be observed, as
well as an increment in the drop size and a longer drop neck. For a bigger radius (Figure 4.24
c), a bigger and more elongated drop tends to form, including a longer drop neck. This is an
expected behavior for a bigger conduct radius when IFT keeps constant.

50

40

30
h (mm)

20
r = 0.1 mm
r = 0.2 mm
r = 0.3 mm
10

0
0 50 100 150 200 250 300 350 400
t (s)

Figure 4.25. Emptying speed for varying conduct radius

In Figure 4.25 is possible to observe the oil remaining in the conduct vs time. The conduct
has a total height of 50 mm. The emptying speed increases dramatically when the radius
changes from 0.3 mm to 0.2 mm. However, when radius changes from 0.2 mm to 0.1 mm,
there is a lesser increment in the emptying speed of the conduct. Also, is observable that the
emptying is faster at the beginning and slower as more oil leaves the conduct. This may be
due to the decreasing of the buoyancy forces while the adhesion forces remain the same. The
same applies for the increment in the emptying speed. As a wider radius is used, there are
bigger buoyancy forces compared with the adhesion forces.

108
4.3.3. Interfacial tension effects
Different interfacial tensions were used to address its effects in the capillary conduct
emptying. Interfacial tensions of 5e-5, 5e-4 and 5e-3 were used with a constant conduct radius
of 0.1 mm. In the next paragraphs the results are shown.

a) b) c*)
Figure 4.26. Drop shape for varying interfacial tension. a) IFT = 5e-5 N/m, b) IFT=5e-4 N/m
and c*) IFT =5e-3 N/m. For the c*) case, injection was used to form the drop, because the oil
could not go out by itself due to the high IFT.

Figure 4.24 shows the drop shape for different IFTs, using the same conduct radius (0.1 mm).
When a low IFT is used (Figure 4.24 a), a smaller and longer drop is formed. When an
intermediate IFT is used (Figure 4.24 b) a larger drop with a narrow neck is formed.
Moreover, when a larger IFT is used (Figure 4.24 c), a significantly larger and rounder drop
is formed. There are no drop neck elongations when using a high IFT as in the case of radius
variation, but a significant size increment of the drops.

109
50

40

30
h (mm)

20
 = 5e-5 N/m
 = 5e-4 N/m
 = 5e-3 N/m
10

0
0 50 100 150 200 250 300 350 400
t (s)

Figure 4.27. Emptying speed for varying IFT

In Figure 4.27 the oil height in the capillary conduct for different times and using different
interfacial tensions is shown. For higher IFT, the oil takes longer time to leave the conduct,
while, when a lower IFT is used, the oil tends to leave faster. Also, there is not a total
emptying for an IFT of 5e-4 and for an IFT 5e-3 there is no displacement at all. Moreover, the
emptying time increases greatly for an IFT higher than 5e-4. This may be caused by the
adhesion forces, because a higher IFT increases those forces, making it difficult for the oil to
leave the conduct.

4.3.4. Model comparison


For the model comparison the emptying curves were used. First a comparison of the curves
for a variable radius are compared. This comparison is shown in Figure 4.28.

110
50

40

30
Fluent r = 0.1 mm
h (mm)

Fluent r = 0.2 mm
Fluent r = 0.3 mm
20 EES r = 0.1 mm
EES r = 0.2 mm
EES r = 0.3 mm

10

0
0 50 100 150 200 250 300 350 400
t (s)

Figure 4.28. Emptying speed models comparison for varying conduct radius

Figure 4.28 shows the oil height in the capillary conduct for different times and using
different conduct radius for fluent and Jose Lopez model results. Is notorious that fluent
results differs from those obtained using Jose Lopez model. Fluent tends to predict shorter
emptying times. Jose Luis model predicts an emptying time with a large increment in the
end. This difference may be caused by the simplifications of the one dimensional model. In
the axisymmetric simulations, wall adhesion causes that some oil remains trapped inside the
conduct, while when considering a one dimensional model, this residuary oil is neglected.
But, in general terms, both models predict the same behavior.

50

40

30
Fluent  = 5e-5 N/m
h (mm)

Fluent  = 5e-4 N/m


Fluent  = 5e-3 N/m
20 EES  = 5e-5 N/m
EES  = 5e-4 N/m
EES  = 5e-3 N/m
10

0
0 50 100 150 200 250 300 350 400
t (s)

Figure 4.29. Emptying speed models comparison for varying IFT

111
Figure 4.29 shows the oil height in the capillary conduct for different times and using
different IFTs for fluent and Jose Lopez model results. As in the varied radius case, Fluent
tends to predict shorter emptying times. Jose Luis model predicts an emptying time with a
large increment in the end. Just as in the past analysis, this difference may be caused by the
simplifications of the one dimensional model. It’s very notorious that IFT has a main role in
the emptying of capillary conducts but to observe its effects. But conducts radius has a higher
impact, because just a little variation in the radius can modify dramatically the emptying
time, while IFT needs to vary in a magnitude order to have important effects.

112
4.4. References
[1] J. L. Lopez-Salinas, “Transport of Components and Phases in a Surfactant/Foam,”
Rice University, 2013.

[2] K. Mannhardt and J. J. Novosad, “Chromatographic effects in flow of a surfactant


mixture in a porous sandstone,” J. Pet. Sci. Eng., vol. 5, no. 2, pp. 89–103, Feb. 1991.

[3] K. Mannhardt, L. L. Schramm, and J. J. Novosad, “Adsorption of anionic and


amphoteric foam-forming surfactants on different rock types,” Colloids and Surfaces,
vol. 68, no. 1–2, pp. 37–53, Nov. 1992.

113
Chapter V: Conclusions and future work
5.1. Conclusions
There are several ways to obtain approximations of the behavior of a two-phase system in a
porous medium because there are a large number of equations that help predict different
properties, such as permeability, capillary pressure, and capillary number, among others.
Being some equations more useful than others, depending mainly on what fundamentals and
conditions these equations are developed and the type of porous medium in which the
simulation of the phenomenon is being carried out. In general, it could be said that it is
possible to obtain the necessary properties from capillary pressure, which is one of the most
important variables in these systems since the flow and displacement of the phases occur
mainly due to this. However, there is little information available about equations of state that
allow us to relate these properties. The equations with which the properties are predicted in
Kulkarni's work usually work only for the system in which they were developed, generally
governed by the type of rock used. When trying to use these same equations in another
system, the equations lose precision. This is mainly because the equations used use
parameters adjusted to the experimental data of that system, under which, they can give quite
satisfactory results. However, they will only work efficiently when working in the same
system. However, the ideal case would be to use models that predict the properties under any
conditions, or at least, under a wider range of conditions.

Regarding the Kulkarni article analyzed in this essay, it could be said that quite acceptable
results were obtained. Although this was mainly because the prediction of the properties was
carried out directly from experimental data of the analyzed system, which ensures good
results, but not a good model. The coefficients and equations obtained in this article would
only allow good predictions when the experimental data of capillary pressure and relative
permeabilities of the phases of the system are known. This is a common problem in the
analysis of this type of systems, so several scientists have been given the task of making
equations to obtain satisfactory predictions of capillary pressure in various types of rock
through the Leverett equation. The Central focus has been capillary pressure since there is a
relationship between permeabilities and this variable.

114
Regarding the model for transport of solid. It is possible to obtain a good fit between
experimental and calculated results using the model suggested in this work. This model
addresses the impact of key effects on the adsorption phenomena from the theoretical point
of view.

Using an isotherm with a maximum before the plateau region (which can be achieved with
the Redlich-Peterson isotherm but not with the Langmuir isotherm) contributes to the final
adsorption results. In general, for adsorption analysis, maximum adsorption is considered to
have been reached in the plateau, leaving aside any analysis of isotherm peak. The correct
selection of the appropriate isotherm type is determinant in obtaining accurate agreement
with experimental data.

For systems with large Peclet number, the outlet boundary condition appears to be
unimportant, while for small Peclet number systems, the outlet BC has a great impact on the
downstream response. Therefore, if an analytical solution is used to fit the experimental
parameters for the tracer, the appropriate boundary condition should be used.

It was observed that the characteristics of the input signal (slug size, rectangular pulse)
proved to be of greater importance for smaller signals than that which is required to meet the
adsorption saturation. It is highly recommended, therefore, that greater signals be used than
those needed to reach saturation.

A local equilibrium assumption is only valid where the flowing conditions result in a Stanton
number higher than 10. In order for the instant local equilibrium assumption to be applied
correctly, high velocity flows (v1 for macro porosity) must be considered, making sure that
the specific area (av) of the solid surface is higher than the specific area of the connecting
necks of the macro and meso porous regions.

The proposed model agrees closely with the experimental results when macro and meso
porosity are considered. It is strongly recommended that at least two types of porosity be
used for modeling surfactant transport through porous media. When micro porosity was
included, a behavior that resembles irreversible adsorption appeared. This may help
differentiate irreversible adsorption from reversible adsorption present in micro porosity. If

115
micro porosity is overlooked, the model results may be misinterpreted as adsorption in the
porous media.

When a high permeability contrast is present in the system, segregation of the surfactant may
occur, causing two concentration peaks in the outlet. This segregation effect may not be
confused with the dispersion/diffusion effect. When a low permeability contrast is present,
no significant impact may be observed in the surfactant outlet concentration. To avoid data
misinterpretation, systems with low permeability contrasts should be used (or homogeneous
systems).

When the dispersion/diffusion effects are negligible, the heterogeneities in the porous media
may produce a segregation of the surfactant in the macroporous region of each section. To
avoid this effects of segregation, using a high dispersion and/or a constant injection of the
surfactant is recommended.

Is recommended to be carefully when realizing the analysis of the surfactant outlet


concentration, when more than one surfactant is used, to avoid misinterpretation of the data.
A surfactant blend may be misinterpreted as a surfactant with several maxima and minima
slopes in outlet concentration. So, if more than one surfactant is present in the system, a
suitable analysis method should be used.

Axisymmetric simulations and 2-D simulations show similar results as in both cases
heterogeneous media can produce segregation of the surfactant because of the different flow
velocities caused by the heterogeneous permeability. These simulations may be helpful
understanding the adsorption mechanisms, and proving that heterogeneities are the cause of
segregation in the surfactant but not in the tracer. Thus, the simulation may help to design
better experiments related to adsorption phenomena.

The proposed model is a tool for guiding the design of dynamic adsorption experiments.
Also, this model will help understanding how rock heterogeneities influence interpretation
of experimental results.

Concerning the axisymmetric capillary conducts simulations is possible to infer that


axisymmetric simulations may reproduce the capillary displacement phenomena accurately
but at a high computational cost. But there are simplified one-dimensional models for this

116
phenomenon which are significantly less time consuming and let us obtain a good prediction
of the capillary displacement. However, if a deep analysis of the capillary displacement
phenomena is necessary, axisymmetric simulations are strongly suggested, because more
analyzable data can be obtained, regarding the effects of wall adhesion and drop formation,
both of which can be of vital importance in some cases.

Regarding the computational analysis of the software used. COMSOL is quite inefficient for
fluid flow analysis taking up to 50 times more time to solve the same systems as ANSYS
Fluent. However, COMSOL has an important advantage, in this software is possible to easily
specify any function to be resolved, in other words is a more “user friendly software”, while
doing this in ANSYS Fluent is a lot more complex. So, in conclusion, COMSOL is an optimal
tool for simple systems (1-D, 2-D or axisymmetric) where we need to specify our equations
or to combine our equations with embedded COMSOL equations. Moreover, ANSYS Fluent
is a perfect tool to solve complex systems related to fluid mechanics, although, if we want to
introduce our own equations, further coding is needed.

5.2. Future work


There is still a lot of work regarding this models and simulations. More experimental tests
should be done to obtain a better validation of the results under different circumstances. For
to obtain a simple and accurate model is the ultimate goal for model development. Better
predictions of these phenomena would allow us to minimize the costs of many processes. It
would be ideal to develop a software specialized in the solution of any fluid mechanics
problem using finite volume method without the limitations of ANSYS Fluent.

In conclusion, the next future work is necessary:

 More experimental tests should be done to obtain a better validation of the results
obtained.
 To obtain a more simple and accurate model for the prediction of transport of
phases and species in porous media.
 Develop a software specialized in the solution of any fluid mechanics problem
using finite volume method without the user friendly constrains of ANSYS
Fluent.

117
 Complete the modeling of different transport phenomena in porous media.
 Expand the cases studied in this work to complex capillary network.
 Study the effect of non-Newtonian fluids in porous media.
 Study flow in fractured porous media.
 Study triple permeability and triple porosity models.
 Study of flow in porous media of complex fluids (Foams, emulsions, colloidal
dispersions, etc.)

118
ANNEXES
The COMSOL module for differential equations resolution allow us to manipulate the next
differential equation through the use of its coefficients.

𝜕2 𝑢 𝜕𝑢
𝑒𝑎 𝜕𝑡 2 + 𝑑𝑎 𝜕𝑡 +𝛻∙ (−𝑐𝛻𝑢 − 𝛼 𝑢 + 𝛾) + 𝛽 ∙ 𝛻𝑢 + 𝑎𝑢 = 𝑓 (1a)

For the flow in saturated porous media case, the next procedure was used.

With a mathematical comparison of COMSOL differential equation and our model, it’s
possible to define the coefficient values in the next form.

𝜕𝑆𝑤 𝜕 𝑑𝑗(𝑆𝑤 ) 𝜕𝑆𝑤 𝜕𝐹 𝜕𝑆𝑤


𝜙 + 𝛼 𝜕𝑧 (𝐹𝑤 𝑘𝑟𝑜 ) + 𝜕𝑆𝑤 =0 (2a)
𝜕𝜏 𝑑𝑆𝑤 𝜕𝑧 𝑤 𝜕𝑧

𝜕𝑆𝑤
𝑑𝑎 + 𝛻 ∙ (−𝑐𝛻𝑆𝑤 ) + 𝛽 ∙ 𝛻𝑆𝑤 = 0 (3a)
𝜕𝑡

Where

𝑑𝑎 = 𝜙 (5a)

𝑑𝑗(𝑆𝑤 )
𝑐 = −𝛼 𝐹𝑤 𝑘𝑟𝑜 (6a)
𝑑𝑆𝑤

𝜕𝐹
𝛽 = 𝜕𝑆𝑤 (7a)
𝑤

This allow us to introduce our equation in COMSOL. For the inlet boundary condition, the
Flux/Source condition is applied. This boundary conditions has the next form.

−(−𝑐𝛻𝑢 − 𝛼 𝑢 + 𝛾) = 𝑔 − 𝑞𝑆𝑤 (8a)

Where we need to specify g and q. Just as in the analysis of eq (2a), we obtain the value of
the coefficients. Simplifying in terms of the boundary condition we obtain.

𝜕𝑆𝑤
−(−𝑐𝛻𝑢 − 𝛼 𝑢 + 𝛾) = −(−𝑐𝛻𝑆𝑤 ) = (𝑐 ) = 𝑔 − 𝑞𝑆𝑤 (9a)
𝜕𝑧

𝜕𝑆𝑤
∴ 𝑔 = (𝑐 ) + 𝑞𝑆𝑤 (10a)
𝜕𝑧

119
Knowing that our inlet boundary condition in the system is

𝜕𝑆𝑤 1 𝜕 −1
= −𝑘 [ 𝑗(𝑆𝑤 ) ] (11a)
𝜕𝑧 𝑟𝑜 𝛼 𝜕𝑆 𝑤

Then we have

1 𝜕 −1
𝑔 = (𝑐 (− 𝑘 [ 𝑗(𝑆𝑤 ) ] )) (12a)
𝑟𝑜 𝛼 𝜕𝑆 𝑤

𝑞=0 (13a)

For the outlet boundary condition we use a Dirichlet boundary condition, which has the next
form

𝑆𝑤 = 𝑟

Here, we use r = 1.

For the case of the transport of surfactant through porous media a similar procedure was used

Via comparison of the mathematical model and the COMSOL equations, the parameters for
the simulator take the form:

𝜕𝐶1 (1−𝜑) 𝑓 𝜕2 𝐶1 𝜕𝐶1


−𝑓1 (1 + 𝐹 ′1 )+𝑁1 − 𝑏1 = 𝑁𝑆𝑡12 (𝐶1 − 𝐶2 ) (14a)
𝜕𝜏 𝜑 𝑃𝑒1 𝜕𝜒2 𝜕𝜒

𝜕𝐶1
𝑑𝑎 + 𝛻 ∙ (−𝑐𝛻𝐶1 ) + 𝛽 ∙ 𝛻𝐶1 = 𝑓 (15a)
𝜕𝑡

Where

(1−𝜑)
𝑑𝑎 = −𝑓1 (1 + 𝐹 ′1 ) (16a)
𝜑

𝑓
𝑐=𝑁1 (17a)
𝑃𝑒1

β = −𝑏1 (18a)

𝑓 = 𝑁𝑆𝑡12 (𝐶1 − 𝐶2 ) (19a)

120
In this way we introduce our equations to COMSOL, the same approach applies for the
equations describing the transport to the meso and microporous network.

Nomenclature:

a = Adsorption coefficient

c = Diffusion/dispersion coefficient

𝑑𝑎 = Accumulation coefficient

𝑒a = Mass transfer coefficient

f = Source coefficient

g = Boundary flux

q = Boundary absorption

r = Dirichlet coefficient

u = System variable

α = Convective flux coefficient

β = Convection coefficients vector

γ = Generation flux

121

You might also like