You are on page 1of 189

MODELING ISOTHERMAL AND NON-

ISOTHERMAL FLOWS IN POROUS MEDIA

by

Ehsan Mohseni Languri

A Dissertation Submitted in

Partial Fulfillment of the

Requirements for the Degree of

Doctor of Philosophy

in Engineering

at

The University of Wisconsin-Milwaukee

May 2011
UMI Number: 3462824

All rights reserved

INFORMATION TO ALL USERS


The quality of this reproduction is dependent upon the quality of the copy submitted.

In the unlikely event that the author did not send a complete manuscript
and there are missing pages, these will be noted. Also, if material had to be removed,
a note will indicate the deletion.

UMT
Dissertation Publishing

UMI 3462824
Copyright 2011 by ProQuest LLC.
All rights reserved. This edition of the work is protected against
unauthorized copying under Title 17, United States Code.

ProQuest LLC
789 East Eisenhower Parkway
P.O. Box 1346
Ann Arbor, Ml 48106-1346
MODELING ISOTHERMAL AND NON-

ISOTHERMAL FLOWS IN POROUS MEDIA

by

Ehsan Mohseni Languri

A Dissertation Submitted in

Partial Fulfillment of the

Requirements for the Degree of

Doctor of Philosophy

in Engineering

at

The University of Wisconsin-Milwaukee

Graduate School Approval Date

ii
ABSTRACT

MODELING ISOTHERMAL AND NON-

ISOTHERMAL FLOWS IN POROUS MEDIA


By
Ehsan Mohseni Languri
The University of Wisconsin-Milwaukee, 2011
Under the Supervision of Professor Krishna M Pillai

A complete understanding of the physics of flow and heat transfer

phenomena in porous media is vital for accurate simulation of flow processes in

industrial applications. In one such application pertaining to liquid composite

molding (LCM) for manufacturing polymer composites, the fiber preforms used in

LCM as reinforcements are limited not only to the single-scale porous media in

the form of random fiber-mats, but also include dual-scale porous media in the

form of woven or stitched fiber-mats. The conventional flow physics is not able to

model the resin filling process in LCM involving the dual-scale porous media. In

this study, the flow in dual-scale porous media is studied in order to predict the

permeability of these fiber mats. The effect of aspect ratio of the fiber preform on

the accuracy and flow during permeability estimation in single- and dual-scale

porous media is analyzed experimentally and numerically.

iii
Flow of liquid in a free channel bounded on one side by porous medium is

studied next, and two well-known boundary conditions of stress continuity and

stress jump at the interface of the two regions are evaluated numerically. A point-

wise solution for Stokes flow through periodic and non-periodic porous media

(made of cylindrical particles) adjacent to the free channel is presented using the

finite element based CFD software COMSOL. The efficacy of the two interfacial

conditions is evaluated after volume averaging the point-wise velocity using a

long averaging volume, also called the representative elementary volume or REV,

and then comparing such a volume-averaged velocity profile with the available

analytical solution. The investigation is carried out for five different porosities at

three different Reynolds numbers to cover a wide range of applications. The

presence of randomly-placed cylinders during the creation of non-periodic porous

media damps out spatial fluctuations in the averaged velocity observed in periodic

porous media. The analytical solutions obtained after applying the stress-

continuity and stress-jump boundary conditions are found to work well at low

porosities, which is in contradiction with the results achieved earlier by other

researchers.

The traditional approach of using averaged equations in the regions of sharp

gradients in porous media to describe flow and transport is theoretically untenable

and perhaps inaccurate. A novel ensemble averaging method is being proposed to

test the accuracy of the volume averaged or smoothed description of flows in

iv
porous media in the regions of sharp gradients. In the new method, the flow in a

certain arrangement of particles (called a realization) is averaged using a small

unit cell, much smaller than the REV. Then such an averaged flow variable is

further averaged over a whole gamut of randomly-generated particle realizations.

This method of averaging will allow one to simulate the ensemble averaged flow

in the sharp-gradient regions of porous-media flows and thus compare the result

with the flow predicted by the traditional volume-averaged set of flow equations.

First the accuracy of the ensemble averaging method was tested by comparing the

permeability of an artificially generated porous medium obtained by the proposed

method against the permeability predicted by some established theoretical models

of permeability. The proposed method was found to be quite accurate. Later the

ensemble average method was applied to the open-channel porous-medium

interface region characterized by a sharp gradient in the flow velocities. It was

discovered that the volume averaged description of such flows, characterized by

the use of the Brinkman equation along with the stress-continuity and stress-jump

conditions, is quite accurate for a range of Reynolds numbers.

The non-isothermal transport during flow in porous media is examined next.

The main focus in this area of research is the thermal dispersion term found in the

heat transfer equation for single- and dual-scale porous media. Most of the

previous efforts on modeling the heat transfer phenomena in porous media were

devoted to isotropic porous media. However, for the anisotropic porous media

v
widely in many industrial applications (such as the woven or stitched fiber mats

used in LCM), not much research on the dispersion tensor is available. A new

combined experimental/numerical approach to estimating the dispersion tensor is

introduced and applied for both isotropic (single-scale) and anisotropic (dual-

scale) porous media. The equivalence between the heat and mass transfer is

exploited and a 1 -D flow experimental setup is employed to study the spreading of

a dye. Later the mathematical model for such a spreading of concentration

(equivalent to the temperature) around a point input in a constant velocity field is

solved using the finite element based software COMSOL. Thus obtained

numerical spreading pattern is fitted onto the experimentally observed one using

the dispersion matrix (tensor) as a fitting parameter. A few cases of single- and

dual-scale porous media are studied and the dispersion tensors are reported for

each individual case.

A chapter-wise summation of this Ph.D. thesis proceeds as follows. The

first chapter is allocated to study the effect of aspect ratio on the permeability

measurement for both single- and dual-scale porous media. In the same chapter,

flow-front of the test fluid during mold-filling in a 1-D mold for both random and

unidirectional fibrous porous media was studied experimentally and numerically.

In the next chapter, an evaluation of the boundary conditions employed to describe

the free-channel porous-medium interface is proposed after using a long REV

close to the interface to average the numerically simulated point-wise flow inside

both periodic as well as non-periodic porous media. Later, a new ensemble

vi
averaging method is introduced which can be used to study the efficacy of the

volume averaged equations in the regions of sharp gradients. In the last chapter,

the non-isothermal flow of liquid through the porous media studied with emphasis

on estimating the thermal dispersion term.

VH/u
Major Professor Date

vn
TABLE OF CONTENTS
CHAPTER 1. INTRODUCTION

1.1. Literature Review on Unsaturated Flow through Dual-scale Porous 5

Media

1.2. The Scope of Research 8

References 11

Chapter 2. EFFECT OF ASPECT RATIO ON MEASURED PERMEABILITY

AND FLOW-FRONT PROGRESS

2.1 Permeability Measurements 16

2.1.1 1-D Flow Experiment 17

2.1.2 Radial Flow Experiment 18

2.2 Effect of Preform Aspect Ratio on the Permeability Measured through 19

the 1-D Flow Experiment

2.2.1 Theory 21

2.2.2 Experiment 25

2.2.3 Results and Discussions 26

2.3 Effect of Preform Aspect Ratio on the Unidirectional Flow Front 31

during Mold-filling in the LCM

2.3.1 Results and Discussion 33

2.4 Conclusion 37

vni
References 38

CHAPTER 3 Evaluation of Boundary Conditions at the Clear-Fluid Porous-Medium

Interface

3.1 Introduction 43

3.2 Design of Numerical Experiments 47

3.2.1 Numerical Solution 49

3.3 Open Channel Flow Model for the Theoretical Solution 50

3.4 Theoretical Solution of the Porous Medium Flow 51

3.5 Stress-Continuity Interfacial Condition 51

3.6 Stress-Jump Interfacial Condition 54

3.7 Results and Discussions 56

3.8 Summary and Conclusion 70

References 72

Chapter 4. PREDICTING THE AVERAGE FLOW VARIABLES IN POROUS

MEDIA USING ENSEMABLE AVERAGING METHOD

4.1 Introduction 76

4.2 A Conceptual Description of the Ensemble Averaging Method 81

4.3 Validation 88

4.3.1 Permeability Estimation in a Porous Medium 88

4.3.1.1 Permeability Measurement Using Ensemble Averaging 91

Method

4.3.1.2 Results and Validation 93

4.3.1.3 New Permeability Models 97

ix
4.3.2 Clear-Fluid and Porous Media Interface Problem 99

4.3.2.1 Results and Validation 103

4.4 Conclusion

References 108

CHAPTER 5 NON-ISOTHERMAL MODELING OF POROUS MEDIA FLOW

5.1 Introduction 114

5.2 Proposed Work 116

5.3 Governing Equations of Non-Isothermal Flow in Porous Media 117

5.3.1 Macroscopic Level 117

5.3.2 Microscopic Level 121

5.4 Dispersion Tensor 121

5.4.1 Dispersion in Isotropic Porous Media 124

5.4.2 Dispersion in Anisotropic Porous Media 128

5.5 A New Combined Experimental/Numerical Methodology to Assess 133

Dispersion Tensor

5.5.1 Theory 132

5.5.2 Experimental Measurements and Devices 134

5.5.3 Numerical Modeling 139

5.5.4 Combined Experimental/Numerical Methodology 139

5.6 Results and Discussions 141

5.7 Summary and Conclusion 149

References 151
CHAPTER6 SUMMARY, CONTRIBUTION, AND FUTURE WORK

6.1 Summary 156

6.2 Contributions 158

6.3 Future Work 159

APPENDIX 161

CV 163

XI
LIST OF FIGURES

Number Description Page

1.1 Representative elementary volume (REV) 2

1.2 Some types of woven mats used in LCM 6

1.3 Schematic view of flow of fluid through dual-scale porous medium 7

2.1 Schematic view of 1 -D flow experiment 18

2.2 Schematic view of radial flow experiment 19

2.3 Schematic description of the three 1-D flow experiments 23

2.4 Radial flow experiment at flow-rate of 0.00256 ml/s 26

2.5 The pressure history for three different orientations for the aspect 27

ratio 1

2.6 Permeability measurement error, K, - K 2 , versus the aspect ratio 28

2.7 ^ , .,. K.-K, , . 29


Permeability measurement error, — , versus the aspect ratio

2.8 Setup for the 1-D flow experiment 32

2.9 Unsaturated flow front at different mat orientations for aspect ratio 33

of 4

2.10 Flow front of the filling process in the dual-scale porous media for 35

(a): Numerical modeling at different time steps (second) (b):

Experimental results

2.11 Figure 2.11 Unsaturated flow-front at four different aspect ratios 36

xu
of 1, 2, 3 and 4, from left to right

2.12 Pressure history profile for both cases of single scale and dual 36

scale porous media

3.1 Modeling creeping (Stokes) flow through the hybrid (open-channel 48

porous-medium) region: (A) A schematic of the velocity profile through

the channel and the porous medium, (B) The geometry of the hybrid

region with a periodic porous medium; (C) The geometry of the hybrid

region with a non-periodic, random porous medium. The last two figures

show typical REVs (Representative Elementary Volumes) employed for

averaging flow velocities inside the periodic and irregular porous media.

3.2 Effect of increase in Reynolds number on the streamline pattern 58

observed adjacent to the open-channel porous-medium interface for

the periodic porous medium. The porosity for the porous medium is

e = 0.6.

3.3 Streamline pattern at the interface of clear-channel and porous wall 59

for non-periodic porous media at f = o.5 a n d Re = l x i o 3 .

3.4 Plots of normalized average-velocity inside the periodic porous 60

medium at three different Re numbers for five different porosities: a)

£ = 0.5,b) s = 0.6,c) s = 0.7, d) s = 0.8 ande) a = 0.9.

3.5 An example of a weakly-random arrangement of cylinders employed 62

for the creation of an irregular porous medium adjacent to the

xiii
channel. The porosity is 0.7 (i.e., s = 0.7) for the porous medium.

3.6 Schematic view of six different REVs used for the REV-size 63

independence study (not in the actual dimensions).

3.7 Results of the REV size-independency analysis for the irregular 64

porous medium at £=0.5 for the six REVs listed in Table 3.3.

3.8 Results of the mesh independence study. (Details of the three grids 65

or meshes are given in Table 3.4.)

3.9 Repeatability of the solution by comparing three different random 66

realizations' results

3.10. Comparison of average velocity using FEM for both regular and 67

random cases with the Brinkman models at Re= 1 x 10"4 for

A)e = 0.5 ; B) e = 0.6 , Q e = 0.7 , D) e = 0.8 a n d E) * = 0-9

3.11 Actual velocity profile versus dimensionless channel height for 8=0.6 68

and Re=lx 10-4.

3.12 Velocity contours for the non-periodic porous media at E = 0.5 and 69

Re = lxio-3: A) velocity contour and B) volume-avearged velocity

3.13 Effect of variation in porosity of the irregular porous medium on the 70

distribution of the volume averaged velocity near the open-channel

porous-medium interface.

xiv
4.1 Variation of averaged quantities with respect to the REV size 77

4.2 Dimensionless vertical velocity and temperature profiles versus 79

similarity variable for natural convection on a vertical heated surface

filled with porous media (Cheng and Minkowycz, 1977).

4.3 Liquid composite molding process with hot walls of mold on sides 80

and the cold resin.

4.4 Anticipated variation in (¥{) through an increase in; a) Volume, b) 82

Number of realizations

4.5 Figure 4.5 Schematic view of random locations of cylinders inside 84

cells in different realizations

4.6 Fixed number of particles totally inside a cell, case 1 86

4.7 A typical distribution of a fixed number of particles inside a cell 87

for case 2.

4.8 (a): The geometry of the problem, (b): Details of a sample 91

realization with square unit cells each including a solid cylinder.

4.9 (a) Two different realizations randomly selected (b) Square 95

arrangement of periodic fiber mat array.

4.10 Comparison of the obtained permeability values from ensemble 95

average method with six other results.

4.11 Dependency of the permeability value on the number of realization 99

used for estimating the ensemble average.

xv
4.12 Creeping flow through the hybrid clear-fluid-porous medium 100

interface: (a) schematic view and (b) Ideal geometrical model

4.13 Velocity distribution in the clear-fluid porous medium interface; 104

( a ) Re = 0.01 a nd (b) Re = 0.001

4.14 Figure 4.13. a) Dependency of the Ensemble average method on 106

number of realizations; b) Part A; c) Part B.

5.1 Schematic view of the initial impregnation of a dual-scale porous- 118

medium by a liquid [19]

5.2 Mixing as a result of obstruction 123

5.3 Dispersion of a point injection displaced a distance L 125

5.4 The 1-D flow experiment setup: A) open mold cavity and B) the 135

assembled mold

5.5 The flow injection and flow-meter setup 136

5.6 Measuring devices for A) Density, B) Viscosity, and C) Porosity 137

5.7 Effect of dispersion in the form of an elliptical dye expansion in a 138

single-scale porous medium

5.8 Schematic view of the geometry, the boundary and initial 139

conditions

5.9 Effect of pure molecular diffusion of thermal dispersion by time 141

5.10 Random fiber-glass mat as a representative of single-scale porous 142

media

xvi
5.11 Matching the experimental and numerical results for a single-scale 143

porous medium (random fiber mat) at Q=2ml/s at three different

time frames of A) 7 s, B) 33 s and C) 51 s

5.12 The longitudinal and transverse increases of the initial circular 145

dye-patch with time

5.13 Stitched fiber-glass mat as a representative of dual-scale porous 145

media.

5.14 Comparison of the experimental and numerical modeling for an 146

anisotropic but periodic medium at Q = 2 ml/s : A) experimental

results at t = 0 s, B) numerical results at t = 0 s, C) experimental

results at t = 35 s, D) numerical results at t = 35 s, E) experimental

results at t = 71 s and F) numerical results at t = 71 s

5.15 Unidirectional fiber-glass as a representative of dual-scale porous 148

media

5.16 Unidirectional fiber-glass in A) initial unsaturated flow and B) 148

saturated flow

5.17 Experimental observation of flow of tracer through a dual-scale 149

porous medium made of the unidirectional fiber mats

Al Schematic view of locating the solid cylinder at the center of the 161

unit cell in order to generate the periodic porous media

A2 Schematic view of location of the solid cylinder at horizontal and 162

xvii
vertical distances from the center of unit cell in order to generate

the non-periodic porous media

A3 No overlapping for the irregular porous media 162

xviii
LIST OF TABLES
2.1. Details of experiments 25

2.2. Coefficients of two equations fitted to Kx - K2 error data. 30

23 K —K 31
Coefficients of two equations fitted to— error data.
K
\

3.1. 50
Details of volume averaging moving step for each porosity.
3' 2 ' Values of effective viscosity JU', coefficient /?, and slip coefficient a at 57
different porosities
3.3. 63
Details of the six REVs used in the REV-size independence study.

3.4. Details of the elements and nodes of three different meshes employed to 66
conduct the mesh independence study
4.1. Five different permeability models 94

4.2. 96
Details of permeability values at different porosities for each model.

4.3. Error's details for each model's results based on the numerical method's 97
results.

xix
NOMENCLUTURE

Roman Letters

A Area [m2]

b A mapping parameter

c Specific heat coefficient [J/kg.K]

D Diffusivity [m2/s] and Diameter [m]

f A position vector function

h Channel height [m]

i An integer number

K Permeability [m2]

k Dispersion tensor [m2/s]

L Length of channel [m]

m,n Parameters

N An integer number

q" Heat conduction [w/m2]

S A sink term

T Temperature [K]

t Time [s]

V Volume [m3]

Pe Pecklet number

Re Reynolds number

Greek Letters

XX
a Thermal diffusivity [m2/s]

P A coefficient in Brinkman equation

5 Thermal conductivity [W/mK]

5c Boundary layer thickness [m]

e Porosity

I An axis

P Density [kg/m3]

H Viscosity [pa.s]

9 Porosity

V A flow variable

a A function

Subscripts

ave Average

eff Effective

f Fluid and fiber

g Gap

P Pressure

R Reaction

real Realization

s Solid

t tow

Superscripts

clear Clear flow

xxi
porous Porous-media flow

Other Symbols

<> Volume-averaged quantity

< >f Pore-averaged quantity

xxii
ACKNOWLEDGMENTS

I would like to express my deepest gratitude to my advisor, Professor Krishna

Pillai, for his continuous guidance, patience, support, care. Professor Pillai provided an

excellent atmosphere for research and teaching during these years. I am proud of having

such an exceptional supervisor and teacher in my Ph.D. study. His high standards on

research and teaching and his commitment towards students made it a privilege to work

with him.

I sincerely appreciate my dissertation committee members: Professor Michael

Lovell, Professor Sam Helwany, Professor Ilya Avdeev, and Professor Arash Mafi for

their support, comments and insights. I am also grateful for their time and efforts in

evaluating this research work. I would like to thank Professor Tien-Chien Jen, Professor

John Riesel, Mr. Dan Beller, for their support, help and motivation during these years.

I am indebted to my B.Sc. and M.Sc. supervisors and teachers, Professor Hessam

Taherian, Professor Davood Domairry Ganji, Professor Mofid Gorji Band-Pay, Professor

Kamel Hooman and Professor Halimi who cheered me up and supported me during last

ten years of study.

I would like to thank my friends, Dr. Reza Masoodi, Rahi Abouk, Nima Jalali,

Sahar Hosseini, Sepideh Maleki, Reza Jahani, Dr. Hua Tan, Saman Beyhaghi, Andrew

Vechart, Mike Verhagen, Grover Bennett, Russell Moore, Ehsan Ghotbi and Dr.

Mohammad Habibi who were always willing to help and give their best suggestions and

motivation.

xxiii
I would also like to thank my parents, Ali Mohseni Languri and Mahdokht Tirgar,

my sisters, Elham, Elahe and Azadeh, and in-law family. They were always supporting

me and encouraging me with their best wishes and support.

Above all, I would like to thank my wife, Maedeh Sedaghati. She was always there

cheering me up and stood by me through the good times and bad. This success was not

possible without her support and strength during last two years of my study.

xxiv
1

Chapter 1

INTRODUCTION

Composite materials have been widely used in industries such as

automotive, shipbuilding and civil due to their light weight, high strength, high

corrosion resistance and flexibility in design. Such composite materials are made

of polymer resins as matrix and glass fibers or carbon fibers as reinforcements for

the plastics. There is a variety of manufacturing processes for making polymer

composites, but liquid composite molding (LCM) is one of the most widely used.

The LCM process includes several technologies, such as Resin Transfer Molding

(RTM), Vacuum-Assisted Resin Transfer Molding (VARTM), and Seemann

Composites Resin Infusion Molding Process (SCRIMP). The following major

steps are common for all LCM processes: first the reinforcing fibers are shaped

similarly to the final shape, which is called preform, placed in a closed mold [1].

Then matrix materials, in the form of liquid polymer, infiltrate the preform either

by injection or by vacuum. The matrix material later undergoes a solidification

process wherein the thermoset type of resins cure with the help of an increased

temperature. The final parts are removed when the solidification process is

complete.

There is a great need for numerical modeling of these processes for

industrial applications. The preforms are assumed to be a porous medium in this


2

case. A volume averaging needed in order to average flow variables like

temperature over a representative volume at the pore level. The averaging volume

should be sufficiently large so that it can capture any fluctuation in that region.

This volume is called the representative elementary volume (REV), Fig. 1.1. The

REV must be far smaller than the size of the entire domain; otherwise the resulting

average through the whole domain cannot represent what happens at any

individual points in the domain. On the other hand, the size of the REV must be

larger than the size of single pore that each REV includes a sufficient number of

pores to have a meaningful averaging process [2].

Figure 1.1 Representative elementary volume (REV)

By a further assumption of fully saturated flow behind the flow front of the resin,

one can model the mold filling stage of LCM using the Darcy's law as
3

<jf>=--.V<pf>f (1.1)
M

where <vf > is the volume-averaged velocity of resin in the preform, < pf >f is

the pore-averaged pressure of resin, K is the permeability tensor for fiber preform

and |i is the resin velocity. In the case of isotropic porous media, the permeability

tensor become a scalar and the resin can be assumed incompressible. Therefore,

continuity equation can be written as

V.<vf>=0 (1.2)

By substituting Eqn. (1.1) into Eqn. (1.2), one can obtain an elliptic partial

differential equation (the Laplace equation) that has only pressure as the unknown

variable.

V.(K t .V<p f > f ) = 0 (1.3)

Tucker [3] used the volume averaging method to derive the energy equation for

the single- scale porous media, as

{erPrCpf +E s p,C p ,J^(T) + pfCp,f(vf).V{T} = V.(Klh,f +Klh,s).V(T) + £fPr H R (f £ ) f (1.4)

+ Uf(vr).K;'.(vf)

where ef and es are the volume fraction of the fluid and solid, pf and pi are

densities of the fluid and solid, CpJ and Cpsare specific heats of the fluid and

solid, T is temperature for both the fluid and solid phases obtained through the

local thermal equilibrium assumption, HRis the heat of reaction per unit mass of

the fluid, and fc is the reaction rate for curing of the resin. KthJ and Kths, the
4

effective thermal conductivities of the fluid and solid phases, are expressed as

Kth,f =k f e f § + ^ j n f s b d A - ^ ^ JvfbdV (1.5)


V V
f A„ f V,

K ^ = k , e . 5 + ^ ^ Jnf.bdA (1.6)

where kf and ki are thermal conductivities of fluid and solid phases, Vf and Vs

are the volume of fluid and solid phases in REV, and A^is the interfacial area

between fluid and solid phases within the REV. The vector function bthat

transforms the gradient of the average temperature into the local variations of the

deviation is given by

f f =b.V(T) (1.7)

The next step to modeling the whole plastics processing process is to set the

balance equation for the chemical reaction of the resin which may also called by

curing. The balance equation for chemical reaction, as derived by Tucker [3], is

ef | ( c f > f +(vf).V(cf)f =V.{Df.V(cf)f}+ef(fc)f (L8)

where cf is the dimensionless extent of chemical reaction and D f , the effective

diffusivity tensor as following

Df =D f £ f 8 + ^ J n h f d A - ^ - JvffdV (1.9)
V V
f Afs f Vf

where / is a vector function of position within the fluid phase that transforms the

average concentration gradients into the point-wise concentration deviation


5

c f =f.V(c f ) f (1.10)

It should be noted that partially saturated flow front may occur due to the

geometrical characteristics of some fiber mats. In such situation, the numerical

modeling for single-scale porous media is not applicable and the importance of

numerical modeling for dual-scale porous media is often felt.

1.1 Literature Review on Unsaturated Flow through Dual-scale

Porous Media

Some recent studies [4-13] have shown that resin flow through certain types

of porous material cannot be modeled using the conventional flow physics. The

micro-structure of fiber mats, such as woven or stitched fabrics, is the main reason

for such a discrepancy between the experiments and numerical predictions of mold

filling (Fig. 1.2). This is a fact that the fiber tows made of thousands of individual

fiber filaments either woven or stitched. The gap within the intra-tow region is of

the order of micrometers. The distance between the fiber bundles, called inter-tow

region, is of the order of millimeters. The difference in the orders of magnitude of

pore length-scales in a medium is assumed the main reason for partial saturation of

such porous media. These types of porous media are classified under the category

of dual-scale porous media.


6

Biaxial Iteaue Triaxial Weaue Knit Nultiaxial Multilayer


Harp Knit

Figure 1 2 Some types of woven mats used in LCM

In dual-scale porous media, the injected resin passes through the inter-tow

channels without impregnating due to the high flow-resistance inside tows. Hence,

such flow is much quicker than the flow impregnating the tows. This delay in tow

impregnation leads to partially saturated region in the flow-front, Fig. 1.3. This

dual-scale nature of a preform has certain influences on the mold filling process

such as void formation [4-6, 9 and 14-17], creation of visible partially-saturated

region [4, 7-11, 13], and dropping of inlet-pressure history under constant

injection rate [7-13, 18].

The flow model in dual-scale porous media can be divided into me two

main categories based on their length scale. One is the tow-impregnation model

and the second is the global impregnation model. The former focuses on the void

formation mechanism due to its micro-macro flow interaction, while the latter

studies the flow characteristics during mold filling. In the tow-impregnation

models, a single computational domain studies the flows in the inter-tow and intra-

tow, separately.
Fully saturated area Partially saturated region Unsaturated area

Macro-flow front

OnO QOOOo^
O O O O O O r, O o °
O O O Q O O O eft O
Single fiber
O n oO 0 ° o ° 0 ° n O
o°o°oO o°o°o
u
oo o 0 o°o o o 0
Figure 1.3 Schematic view of flow of fluid through dual-scale porous medium

The computational domain distinguishes the intra-tow and inter-tow regions

through the mesoscopic fabric architecture with certain number of fiber tows and

their adjacent ieter-tow regions [6-14]. On the other hand, the global impregnation

models do not need detailed local (the tow scale level) flow information- a lumped

quantity is used to study the mass exchange between the inter-tow and intra-tow

regions. The delay in impregnation within the tows at the micro-level acts as sinks

of liquid in the macroscopic flow field. The modified continuity equation at


8

macroscopic level was proposed by including a non-zero sink term [4, 8, 11, 15,

19 -27]

V.v = -S (1.11)

where S is the sink term, equal to the volumetric rate of resin absorption by

tows per unit volume. One can combine the above equation with Darcy's law to

complete the modeling of dual-scale porous flow using the global impregnation

model.

1.2 The Scope of Research

The primary goal of this research is to advance the understanding the

transport phenomena in isothermal and non-isothermal fluid flow through the

porous medium.

In chapter 2, the focus is on the important term of permeability in dual-scale

porous media. The conventional methods for measuring the permeability have

been introduced. The effects of some parameters such as aspect ratio of the

preform in both steady and unsteady states have been studied experimentally.

In chapter 3, a comprehensive study is conducted to model flow near the

interface of a free channel and a porous-medium channel. A review of different

boundary conditions available in the literature has been done which mainly

focuses on regular arrangements for solid phases in the porous medium. In this

study, flows in both the regular and random arrangements were studied

numerically. Both cases were compared to see the deviation from the flow in an
9

ideal regular arrangement.

Chapter 4 is devoted to the development of a new method of ensemble

averaging to predict the average flow parameters in the porous-medium problem.

We propose a new method to estimate the flow parameters like velocity,

permeability, etc. while avoiding the conventional representative elementary

volume (REV) approach. In the REV method, the size of the REV has always been

problematic since it is difficult to quantify the size of this volume, especially when

the problem involves a sharp gradient of a flow variable. Using the proposed

ensemble averaging method, one can confidently deals with porous-media

problems without worrying about the size of the REV, especially in complex

geometries. Two problems have been solved using the ensemble averaging method

and the results are compared to the available results of the REV method to show

the simplicity and the accuracy of the method.

Chapter 5 focuses on the thermal transport during fluid flow in porous

media where the governing equation for non-isothermal flow is considered. One of

the main issues with LCM in the non-isothermal flow problems of dual-scale

porous media found is prediction of the dispersion tensor. Most of the previous

efforts in this area have been devoted to random fiber mats (a single-scale porous

medium). In many industrial applications, however, the dual-scale porous media in

the form of woven and stitched fiber mats are used. A new technique is proposed

in this dissertation to predict the dispersion tensor in the thermal governing

equation in dual-scale porous media. This method is a combination of a finite-


10

element based solution and an experimental method to find the dispersion tensor.

Knowing the dispersion tensor in dual-scale porous media, one can numerically

model the heat transfer aspects of flow through porous media in the composite

industry.

Finally, in chapter 6, the summary is presented.

$
11

References

1. Rudd, CD., long, A.C., Kendall, K.N. and Mangin C.G.E., Liquid

molding technologies, Woodhead Publishing Ltd, 1997.

2. Parnas, R.S., A.J. Salem, T.A.K. Sadiq, et al., Interaction between

Micro- and Macro-Scopic Flow in RTM Preforms. Composite

Structures, 1994. 27(1-2): p. 93-107.

3. Tucker III, C.L. and Dessenberger, R.B., Governing equations for

flow and heat transfer in stationary fiber bed, Flow and rheology in

polymer composites manufacturing, Elsevier, Amsterdam, pp. 257-

323, 1994.

4. Parnas, R.S. and F.R. Phelan, Jr., Effect of Heterogeneous Porous

Media on Mold Filling in Resin Transfer Molding. SAMPE Quarterly,

1991. 22(2): p. 53-60.

5. Sadiq, T.A.K., S.G. Advani, and R.S. Parnas, Experimental

Investigation of Transverse Flow through Aligned Cylinders.

International Journal of Multiphase Flow, 1995. 21(5): p. 755-774.

6. Binetruy, C , B. Hilaire, and J. Pabiot, Interactions between Flows

Occurring inside and Outside Fabric Tows During RTM. Composites

Science and Technology, 1997. 57(5): p. 587-596.


12

7. De Parseval, Y., K.M. Pillai, and S.G. Advani, Simple Model for the

Variation of Permeability Due to Partial Saturation in Dual Scale

Porous Media. Transport in Porous Media, 1997. 27(3): p. 243-264.

8. Slade, J., K.M. Pillai, and S.G. Advani, Investigation of Unsaturated

Flow in Woven, Braided and Stitched Fiber Mats During Mold-Filling

in Resin Transfer Molding. Polymer Composites, 2001. 22(4): p. 491-

505.

9. Dungan, F.D. and A.M. Sastry, Saturated and Unsaturated Polymer

Flows: Microphenomena and Modeling. Journal of Composite

Materials, 2002. 36(13): p. 1581-1603.

lO.Breard, J., Y. Henzel, F. Trochu, et al., Analysis of Dynamic Flows

through Porous Media. Part I: Comparison between Saturated and

Unsaturated Flows in Fibrous Reinforcements. Polymer Composites,

2003. 24(3): p. 391-408.

ll.Babu, B.Z. and K.M. Pillai, Experimental Investigation of the Effect

of Fiber-Mat Architecture on the Unsaturated Flow in Liquid

Composite Molding. Journal of Composite Materials, 2004. 38(1): p.

57-79.

12.Kuentzer, N., P. Simacek, S.G. Advani, et al., Permeability

Characterization of Dual Scale Fibrous Porous Media. Composites


13

Part A: Applied Science and Manufacturing, 2006. 37(11): p. 2057-

2068.

13.Tan, H., T. Roy, and K.M. Pillai, Variations in Unsaturated Flow with

Flow Direction in Resin Transfer Molding: An Experimental

Investigation. Composites Part A: Applied Science and

Manufacturing, 2007. 38(8): p. 1872-1892.

14.Binetruy, C , B. Hilaire, and J. Pabiot, Tow Impregnation Model and

Void Formation Mechanisms During RTM. Journal of Composite

Materials, 1998. 32(3): p. 223-245.

15.Chan, A.W. and R.J. Morgan, Tow Impregnation During Resin

Transfer Molding of Bi-Directional Nonwoven Fabrics. Polymer

Composites, 1993. 14(4): p. 335- 340.

16.Chang, C.-Y. and L.-W. Hourng, Numerical Simulation for the

Transverse Impregnation in Resin Transfer Molding. Journal of

Reinforced Plastics and Composites, 1998. 17(2): p. 165-182.

17.Lekakou, C. and M.G. Bader, Mathematical Modelling of Macro- and

Micro- Infiltration in Resin Transfer Moulding (RTM). Composites -

Part A: Applied Science and Manufacturing, 1998. 29(1-2): p. 29-37.

18.Pillai, K.M. and S.G. Advani, A Model for Unsaturated Flow in

Woven Fiber Preforms During Mold Filling in Resin Transfer


14

Molding. Journal of Composite Materials, 1998. 32(19): p. 1753-

1783.

19.Phelan, F.R., Jr. Modeling of Microscale Flow in Fibrous Porous

Media. 1991. Detroit, MI, USA: Publ by Springer-Verlag New York

Inc., New York, NY, USA.

20.Pillai, K.M., Governing Equations for Unsaturated Flow through

Woven Fiber Mats. Part 1. Isothermal Flows. Composites - Part A:

Applied Science and Manufacturing, 2002. 33(7): p. 1007-1019.

21.Pillai, K.M. and M.S. Munagavalasa, Governing Equations for

Unsaturated Flow through Woven Fiber Mats. Part 2. Non-Isothermal

Reactive Flows. Composites Part A: Applied Science and

Manufacturing, 2004. 35(4): p. 403-415.

22.Wang, Y. and S.M. Grove, Modelling Microscopic Flow in Woven

Fabric Reinforcements and Its Application in Dual-Scale Resin

Infusion Modeling. Composites Part A: Applied Science and

Manufacturing, 2008. 39(5): p. 843-855.

23.Wang, Y., M. Moatamedi, and S.M. Grove, Continuum Dual-Scale

Modeling of Liquid Composite Molding Processes. Journal of

Reinforced Plastics and Composites, 2008: p. 1469-1484.


15

24.Jadhav, R.S. and K.M. Pillai, Numerical Study of Heat Transfer

During Unsaturated Flow in Dual-Scale Porous Media. Numerical

Heat Transfer; Part A: Applications, 2003. 43(4): p. 385-407.

25.Pillai, K.M. and R.S. Jadhav, A Numerical Study of Nonisothermal

Reactive Flow in a Dual-Scale Porous Medium under Partial

Saturation. Numerical Heat Transfer; Part A: Applications, 2005.

47(2): p. 109-136.

26.Williams, J.G., C.E.M. Morris, and B.C. Ennis, Liquid Flow through

Aligned Fiber Beds. Polymer Engineering & Science, 1974. 14(6): p.

413-419.

27.Gutowski, T.G., Z. Cai, S. Bauer, et al., Consolidation Experiments

for Laminate Composites. Journal of Composite Materials, 1987.

21(7): p. 650-669.
16

Chapter 2

EFFECT OF ASPECT RATIO ON MEASURED

PERMEABILITY AND FLOW-FRONT

PROGRESS

2.1 Permeability Measurements

Permeability is described as the ease with which a fluid flows in a porous

medium-material where a larger permeability indicates the fluid flows more easily.

Measuring the permeability of a porous medium is performed by using the Darcy's

law, Eqn. (1.1), along with flow experiments. Two famous methods for measuring

the permeability of a preform are the 1-D flow experiment and the radial flow

experiment [1-28]. Both of the mentioned methods can be carried out under

either steady state or transient conditions. The saturated permeability is the

measured permeability where the preform is fully wetted by the flowing fluid

under steady-state conditions. The unsaturated permeability is the permeability

measured using the transient mold-filling flow which is marked by partial

saturation of the fiber mat near the liquid-front. Some researchers have shown that

unlike saturated permeability, the unsaturated permeability is not only a function

of a fabric's architecture, porosity and diameter, but also a function of the fluid

properties as well as the flow-rate [6,10-12,15,19-21,23,25,28].


17

2.1.11-D Flow experiment

The 1-D flow experiment is conducted in a rectangular mold with its length

much larger than the width. The sample preform is placed inside the mold and the

test fluid is injected either at constant flow-rate or under constant injection-

pressure along the mold length through the preform [1-5, 7-12, 17-25, 27]. In

order to obtain the in-plane permeability tensor of an anisotropic porous medium,

one needs to conduct the 1-D flow experiments three times with different

orientations with respect to flow direction [9,10,19-21,4,25,27]. A schematic view

of 1-D flow experiment setup is shown in the Fig. 2.1. The unidirectional flow is

created within the preform inside the mold cavity when the test liquid is injected

from one end. The mold consists of two aluminum plates that sandwiches the

optical grade- acrylic sheet between them. The 1" thick acrylic sheet has very high

flexural-stiffness to minimize deflections under the proposed high-pressure test

conditions. The mold cavity is a 3/8" deep and 7" wide pocket which is sealed

with two o-rings along the cavity edge where the acrylic plate rests. There are

thirty five evenly spaced bolts which clamp the mold's plates to ensure the mold's

sealing and avoid plate deflection due to high pressure.


18

computer
8
^ ^ f *Woodata stepper-motor

\ Gear ounu
cant rallied bjf
stepper motor

piston
cyl-lndei:

Figure 2.1 Schematic view of 1-D flow experiment

2.1.2 Radial Flow Experiment

Radial flow experiment is mainly injection of a test liquid from an inlet port

centered in the mold where it creates a radial flow in the mold cavity [6, 13-28].

For obtaining the permeability tensor using this method, one single experiment is

enough and involves observing the shape of the radial flow-front during the fluid

impregnation [13-28]. A schematic view of the radial flow experimental setup is

shown in Fig. 2.2. The mold is composed of three parts. The top and bottom parts

are 53cmx53cmx2.54cm-transparent polycarbonate plates. The center part is a

0.9525 cm thick steel spacer. The two thick transparent plates sandwich the steel

spacer forming a cavity for packing the fiber mat. Sixteen equally spaced nuts and
19

bolts along the edges ensure enough compression of the fiber mat. Four pressure

transducers placed on the top surface of the setup read pressure at four different

locations.

Figure 2 2 Schematic view of radial flow experiment

2.2 Effect of Preform Aspect Ratio on the Permeability Measured

through the 1-D Flow Experiment

Numerous studies have reported the measured permeability values for

various types of fiber mats. This physical quantity has been measured by several

measurement techniques, such as 1-D flow and radial flow methods [1-9]. The

microstructure of most structural fiber mat materials consists of a network of fiber

bundles called tows. Arrangements of these tows range in configuration from a

simple, unidirectional alignment to complex weaves and braids. Similarly, the

steady-state (saturated) flow method and the transient (unsaturated) flow method
20

have been used as two separate ways of estimating the permeability [1-9].

Although measurement of permeability of different kinds of fiber mats is

crucial for an accurate mold-filling simulation for an LCM process, very little

effort has been made to investigate the accuracy of permeability measurement

using the 1-D flow method. Lundsrtom et al. [8] used the multi-cavity parallel

flow technique to measure the in-plane permeability of anisotropic fiber mats.

They reported that two ratios can influence the accuracy of 1-D flow basic

permeability measurements method. One of these ratios is the ratio of preform's

length L to its width W (henceforth to be called the aspect ratio LIW). The other

is the ratio between two in-plane principal permeabilities, K{ and K2 (henceforth

to be called the anisotropy ratio K{ /K2). These two ratios influence the accuracy

of the theoretical solution on which evaluation of the permeability through the 1-D

flow experiment is based. Lundstrom et al. [8] ran numerical simulations to

estimate the error in the effective permeability estimation through this

unidirectional approximation. Their results showed that the accuracy of the

permeability measurement techniques depends on the aspect ratio, the anisotropy

ratio, and the angle of the principal permeability direction with respect to the flow

direction. Tan and Pillai [29] ran numerical mold-filling simulations in a

rectangular 1-D flow mold for a given permeability tensor under constant

injection-rate and uniform inlet-pressure conditions to investigate how the inlet

pressure history and the flow-front patterns are influenced by the aspect ratio, the

anisotropy ratio, and the principal permeability angle. Their results showed that
21

these three factors have strong influence on the inlet pressure history as well as the

final steady-state pressure achieved. They also found that the simulated flow-

fronts differ significantly (were observed to be tilted) from the straight flow-fronts

predicted by the 1-D flow model.

In this reported work, experimental mold-filling investigations were carried

out in a rectangular 1-D flow mold under constant injection rate condition to study

the effect of the aspect ratio, ratio of the preform's length to its width, on the

accuracy of permeability measured through the 1-D flow experiment.

2.2.1 Theory

For an anisotropic fiber mat, the permeability is a tensor rather than a scalar

quantity. However, during the 1-D flow in an anisotropic fiber-mat, the Darcy's

law is given as

<v>=-^V<R>f (2.1)
H

where < v > is the volume-averaged flow velocity, < Pf >f is the pore-averaged

liquid pressure, and Keffis the effective permeability along the flow direction in the

1-D mold. It is essential to know the relation between the effective permeability

and the principal values of the permeability tensor to determine all the components

of the permeability tensor.

Darcy's law given in Eqn. (1.1) can be expanded in a matrix form as


22

a< P f >f
l<v>>( Ku K12 d\{
(2.2)
[< v 2 >J v'
JV
v' 3<p f >'
_ 21 ^22 _
3x2

where < v, > and < v2 > are the volume-averaged velocity components along the

flow (x,) and the transverse (x2) directions, respectively. It is assumed that the

unidirectional flow observed during the 1-D flow experiment implies that the

velocity along the x2 direction is uniformly zero in the flow domain, i.e.,

< v 2 >= 0 (2.3)

By substituting < v2 > from Eqn. (2.3) into Eqn. (2.2) and simplifying the

equation, the following relation between the effective permeability and

components of the permeability tensor can be established.

K"2
K — ir' 12 (2.4)
22

If the principal axis of the permeability tensor is orientated at an angle 6 to the

flow direction, all components of the permeability tensor, in terms of the two

principal permeability components and the orientations, are as follows:

K' =K,cos 2 e+K,sin 2 9

K' = K , sin 2 6+K,cos 2 9 (2.5)

K
i2 = K 2i = ( K 2 - K , ) sinG cosG

Here K, and K2 are the two in-plane principal permeability components and 6 is
23

the angle of orientation of K, with respect to the flow direction.

Finally, upon substituting Eqn. (2.5) in Eqn. (2.4), the effective permeability K

can be expressed as a function of K,, K2 and d as

l)2
Kerf =K,.cos : l + ^ t a n 2 9 - - ' (2.6)
K, K.
+1
K, tan2 9

Hence, three different mat-orientation experiments need to be done to determine

the principal permeabilities, K^andK.,, and the orientation0, Fig. 2.3.

Ill, Y

I,X

Figure 2.3 Schematic description of the three 1-D flow experiments.

As shown in the Fig. 2.3, these three directions are defined as I, II and III for the

angle of#, = 0, du = 0 + 45° and 6m =0 + 90° respectively. Based on Eqn. (2.5),

the effective permeability along these three directions are


24

KXK2
K, (2.7)
KxSin 6,+K2Cos2dl
2

K
KXK2
A (2.8)
// Kx Sin 6u + K2Cos2du
2

KXK2
V (2.9)
Kx Sin 0,„ + K2Cos2em
2

By defining

A=K,+Kl" (2.10)
2

D=K'~K'" (2.11)
2
Using above relations, one can transform Eqns. (2.7)-(2.9) to

Kx =K,——cos20 (2.12)
1
' A-D

K2=Km^-^cos20 (2.13)
2
'" A + D

^=-tan-'{A2~D2--} (2.14)
2 K„D D

Four different aspect ratios were investigated experimentally to establish the

dependence of the accuracy of the 1-D flow based permeability measurement

technique on the aspect ratio. Because the higher aspect ratios should follow the 1-

D assumption < v2 >= 0 more closely than lower aspect ratios, results from the

higher aspect ratio experiments should be more accurate. To support this idea, a

radial flow experiment with the same bi-axial fiber mat was performed. The
25

completely circular flow front occurring in that experiment showed that AT, and

K2 should be the same. Using experimental data in Eqn. (2.11), one obtains three

equations with three unknowns, Ki,, K22 and 0, that are essential for constructing

the full permeability tensor.

2.2.2 Experiment

Details of the preform used in this investigation are presented in Table 2.1.

Bi-axial fiber mat was selected for the preform in the experiments. The test liquid,

motor oil SAE 10W-40, wetted the preform at a constant flow rate. The mold was

so designed that it can be used for different aspect ratios. Details for each aspect

ratio are presented in the Table 2.1. Nine layers of reinforcement were placed in

the cavity to give the desired value of fiber volume fraction. Three preforms

oriented in three different directions were prepared for each aspect ratio.

Table 2.1 Details of experiments

Aspect ratio Dimensions Layers Number of experiments

1.0 5"x5" 9 3

2.0 10"x5" 9 3

3.0 15 x5 9 3

4.0 20"x5" 9 3

A pressure transducer was coupled to a digital voltmeter for recording the pressure
26

history. In order to ensure the repeatability, we often conducted the experiments

multiple times to obtain more than one measured values. The results show a good

repeatability in experiments.

2.2.3 Results and Discussions

Recalling the assumption used in the 1-D flow experiment, < v2 >=0, it is

clear that the longer aspect ratios should lead to more accurate results in

permeability measurements. The radial flow experiment was done using a bi-axial

fiber mat. Completely circular flow-front of the liquid penetrating the preform

shows that the principal permeability components should be the same (i.e. Kx = K2

). The flow front for the radial flow experiment is shown in the Fig. 2.4.

Circularity of the flow front is demonstrated by the fact that the three pressure

transducers placed symmetrically and at equal distance from the central injection

hole.

Figure 2.4 Radial flow experiment atflow-rateof 0.00256 ml/s.


27

Running three experiments along three different orientations of fiber-mat,

one can estimate the principal components of the permeability tensor. Fig. 2.5

shows the pressure plot as a function of time for an aspect ratio of 1 for the 3

different directions. Similar results were observed for the other aspect ratios with a

slight difference that the pressure histories of 0 degree and 90 degree orientations

became more similar as the aspect ratio increased.

70-
Aspect Ratio: 1

0 Degree — 45 Degree - - - 90 Degree


60

50 -

«T
^ 40 _ _ —-
/ S —
3
(A
2 30
a.

20
§ J*

10

if
0 121
Time (Sec)

Figure 2.5 The pressure history for three different orientations for the aspect ratio 1.

Figs. 2.6 and 2.7 show the decrease in the permeability measurement error

with an increase in the aspect ratio. Two different error criteria are used here to

focus more on the effect of the longer aspect ratios decreasing the error

incorporated with the permeability measurement using 1-D flow experiment. In


28

Fig. 2.6, the difference between the two principal directions of the permeability is

used as a criterion for accuracy. According to the results from the radial flow

experiment, the two principal directions of the permeability should be the same,

but different 1-D flow experiments show that the difference between these two

values decreases with increasing aspect ratio. The difference between Kl and K2

is 1.17 e -9 for an aspect ratio of 1. Fig. 2.6 shows that difference decreases to a

low of 4.26 e-10 for an aspect ratio of 4 which indicates that the accuracy

increases with aspect ratios.

1.4E-09

1.1737E-09
1.2E-09

•J.02567E-09
1E-09 -

8E-10
CM

6E-10
^V 5.04731 E-10

4E-10 -
4.26465E-10

2E-10

1 2 3
Aspect Ratio

Figure 2.6 Permeability measurement error, K, - K 2 , versus the aspect ratio.

Fig. 2.7 discusses another accuracy criterion for the permeability


29

K —K
measurement using 1-D flow experiment. The parameter —! indicating the %

error is selected as another criterion. The plot shows that this error value falls

gradually as the aspect ratio increases. The same conclusion as before can be

obtained from this figure: higher aspect ratios lead to more accurate permeability

measurements in the 1-D flow apparatus.

Ky —K2
Figure 2.7 Permeability measurement error, , versus the aspect ratio.

Some error predictions have been done for other aspect ratios in 1-D flow

experiments. Using the data from the specified aspect ratios used in these

experiments and fitting two different curves, one linear and one quadric, to these

data, one can estimate the permeability measurement error for other aspect ratios.
30

For measuring the error, it is needed to replace the aspect ratio number by X, and Y

would then be the predicted error of permeability measurements for that aspect

ratio. The correlation coefficient, r, is also presented for each curve fitting to

ensure the accuracy of the prediction equations. Using the general linear equation

Y - aX + b and general quadric equation Y = a + bX + cX2 and fitting these two

equations to the A", - K2 error data, results are obtained as shown in Table 2.2.

Table 2.2 Coefficients of two equations fitted to A", - K2 error data.

Y = aX + b Y = a + bX+cX2

a 1.473 e-9 a 1.561 e -9

b -2.763 e -10 b -3.635 e -10

— — c 1.744 e-11

r 0.958 r 0.959

Fitting the same equations to the —l- error data obtained through the

experiments, the following results in Table 2.3 can be used for the error prediction

of other aspect ratios.


31

Table 2.3 Coefficients of two equations fitted to error data.

Y = aX +b Y = a + bX +cX2

a 1.485 e-9 a 2.660 e -9

b -0.366 e -10 b -1.541 e -10

— — c 0.235 e -11

r 0.855 r 0.986

2.3 Effect of Preform Aspect Ratio on the Unidirectional Flow-Front

during Mold-Filling in the LCM

The main apparent characteristics of unsaturated flow through dual-scale

porous media are:

(1) A drooping inlet-pressure history is observed for woven or stitched fiber mats

while the pressure history is linear for single-scale porous media.

(2) Unlike the full-saturation observed behind the resin front for single-scale

porous media, a partially saturated region behind the resin front is observed for

dual-scale porous media.

The presence of unsaturated region in dual-scale porous media is due to the

dual-length scale pore space created in porous media by fiber tows. Pores present

between tows in such fiber mats are of a higher size compared to the microscopic

pores present inside fiber tows. Such a dual-scale cavity causes the invading resin
32

to preferentially go around the fiber tows and fill up the large pores between tows

before penetrating the fiber tows. Such delayed impregnation of tow acts like a

sink of liquid in the macroscopic flow field. A sink model based on adding a

nonzero sink function to the equation of continuity is proposed to describe such an

unsaturated flow in dual-scale porous media [1-2, 7]:

V.v=-S (2.15)

Here S , the sink term, is equal to the volumetric rate of liquid absorption.

The presence of the sink term in the continuity equation successfully explains the

drop in the inlet-pressure history seen during the unsaturated flows in dual-scale

porous media.

Fig. 2.8 shows the setup used for the 1-D flow experiments with fiber

performs of different aspect ratios. Experiments are done for four different aspect

ratios with the same flow rate and same number of layers.
33

2.3.1 Results and Discussion

Unsaturated flow front of the resin in dual-scale porous media is presented

here for the aspect ratio of 4 at the three different orientations of 0, 45 and 90

degrees (Fig. 2.9). The slides were taken at the intervals of 20 seconds.

(A) 0 Degree

(B) 45 Degree

(C) 90 Degree
Figure 2.9 Unsaturated flow-front at different mat orientations for aspect ratio of 4.

Fig. 2.9 shows that the unsaturated flow-front for the orientation of the 45
34

degree of mat is straighter than the two other orientations of 0 and 90 degrees. The

two other mat orientation lead to more curved unsaturated flow-fronts at the fluid

flow in a main principal direction of the fiber mat.

Fig. 2.10 shows that the flow-front of the filling process in RTM vary

significantly for the single-scale and dual-scale porous media. The numerical

simulation is done through the PORE-FLOW® code owned by the Laboratory for

Flow and Transport Studies in Porous Media at University of Wisconsin-

Milwaukee. This is a finite element/control volume code to simulate resin flow in

single-scale and dual-scale porous media. This 2D/3D code can accept structured

and unstructured mesh. The flow can be governed by either Darcy's law or

Brinkman equation in porous media and Stokes/Navier-Stokes equations in fluid

flow problems. The simulation of the flow-front for the single-scale porous media

shows that the flow-front is straight and tilted while the flow-front for the bia-axial

fiber-mat reveal the curvature characteristic of the dual-scale porous media. It can

be easily concluded that one cannot easily use the single-scale analysis for the

dual-scale case.
flow front
141 00
131.64
122.29
112.93

'imii-ill'III
103.57
1 94.21

• ! J.
11. 11! 84.86

111 75.50
66.14
56.79
47.43
38.07
28 71
19 36
10.00

Figure 2.10 Flow-front of the filling process in the dual-scale porous media for (a): Numerical
modeling at different time steps (second) (b): Experimental results

the Fig. 2.1:


36

Figure 2.11 Unsaturated flow-front at four different aspect ratios of 1, 2, 3 and 4, from left to right.

Fig. 2.12 shows the differences in the pressure histories for the single-scale

and dual-scale porous media. It can be observed that the pressure profile is curved

for the bia-axial mat while it is linear for single-scale porous media. This plot

shows a large difference in the pressure histories of the single- and dual-scale fiber

mats.

- Experimental result of dual scale porous media

-Numerical result of single scale porous media

0 5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 85 90 95 100 105 110 115 120 125 130 135 140


Time (sec)

Figure 2.12 Pressure history profile for both cases of single scale and dual scale porous media
37

2.4 Conclusion

Permeability measurement technique used in 1-D flow experiments has been

investigated experimentally for four different aspect ratios both in steady-state and

transient conditions. Two different error criteria have been defined to gauge the

accuracy of permeability measurement, and linear and quadric equations have

been fitted to those data to predict the error in permeability measurement due to

perform aspect ratio using a biaxial fiber mat in 1-D flow experiments. The

quadric equation fitted to the error data has a better correlation coefficient than the

linear equation. Results, using both criteria, indicate that permeability

measurement accuracy for biaxial fiber mat in ID flow experiments increases with

increasing aspect ratio. Curvature pressure history profile and curvature flow-front

show the huge difference between the single-scale and the dual-scale porous

media characteristics. Experimental results reveal that using higher aspect ratios of

the preform leads to a straighter flow-front. Numerical simulation for single-scale

porous media can not be used to predict inlet-pressure histories for dual-scale fiber

mats.
38

References

1. Gauvin, R., M. Chibani, and P. Lafontaine, Modeling of Pressure

Distribution in Resin Transfer Molding. Journal of Reinforced Plastics

and Composites, 1987. 6(4): p. 367-377.

2. Trevino, L., K. Rupel, W.B. Young, et al., Analysis of Resin

Injection-Molding in Molds with Preplaced Fiber Mats. Part 1:

Permeability and Compressibility Measurements. Polymer

Composites, 1991. 12(1): p. 20-29.

3. Gebart, B.R., Permeability of Unidirectional Reinforcements for

RTM. Journal of Composite Materials, 1992. 26(8): p. 1100-1133.

4. Young, W.B. and S.F. Wu, Permeability Measurement of

Bidirectional Woven Glass Fibers. Journal of Reinforced Plastics and

Composites, 1995. 14(10): p. 1108-1120.

5. Ferland, P., D. Guittard, and F. Trochu, Concurrent Methods for

Permeability Measurement in Resin Transfer Molding. Polymer

Composites, 1996. 17(1): p. 149-158.

6. Hammond, V.H. and A.C. Loos, The Effects of Fluid Type and

Viscosity on the Steady-State and Advancing Front Permeability

Behavior of Textile Preforms. Journal of Reinforced Plastics and

Composites, 1997. 16(1): p. 50-72.


39

7. Gebart, B.R. and P. Lidstrom, Measurement of in-Plane Permeability

of Anisotropic Fiber Reinforcements. Polymer Composites, 1996.

17(1): p. 43-51.

8. Lundstrom, T.S., B.R. Gebart, and E. Sandlund, In-Plane Permeability

Measurements on Fiber Reinforcements by the Multi-Cavity Parallel

Flow Technique. Polymer Composites, 1999. 20(1): p. 146-154.

9. Ma, Y.L. and R. Shishoo, Permeability Characterization of Different

Architectural Fabrics. Journal of Composite Materials, 1999. 33(8): p.

729-750.

lO.Shojaei, A., F. Trochu, S.R. Ghaffarian, et al., An Experimental Study

of Saturated and Unsaturated Permeabilities in Resin Transfer

Molding Based on Unidirectional Flow Measurements. Journal of

Reinforced Plastics and Composites, 2004. 23(14): p. 1515-1536.

11.Adams, K.L., B. Miller, and L. Rebenfeld, Forced in-Plane Flow of an

Epoxy Resin in Fibrous Networks. Polymer Engineering and Science,

1986. 26(20): p. 1434-1441.

12.Chan, A.W. and S.T. Hwang, Anisotropic in-Plane Permeability of

Fabric Media. Polymer Engineering & Science, 1991. 31(16): p.

1233-1239.
40

13.Chan, A.W., D.E. Larive, and R.J. Morgan, Anisotropic Permeability

of Fiber Preforms: Constant Flow Rate Measurement. Journal of

Composite Materials, 1993. 27(10): p. 996-1008.

14.Hammami, A., F. Trochu, R. Gauvin, et al., Directional Permeability

Measurement of Deformed Reinforcement. Journal of Reinforced

Plastics and Composites, 1996. 15(6): p. 552-562.

15.Parnas, R.S. and A.J. Salem, A Comparison of the Unidirectional and

Radial Inplane Flow of Fluids through Woven Composite

Reinforcements. Polymer Composites, 1993. 14(5): p. 383-394.

16.Parnas, R.S., A.J. Salem, T.A.K. Sadiq, et al., The Interaction between

Microscopic and Macroscopic Flow in RTM Preforms. Composite

Structures, 1994. 27(1-2): p. 93-107.

17.Parnas, R.S., J.G. Howard, T.L. Luce, et al., Permeability

Characterization. Part 1: A Proposed Standard Reference Fabric for

Permeability. Polymer Composites, 1995. 16(6): p. 429-445.

18.Ahn, S.H., W.I. Lee, and G.S. Springer, Measurement of the 3-

Dimensional Permeability of Fiber Preforms Using Embedded Fiber

Optic Sensors. Journal of Composite Materials, 1995. 29(6): p. 714-

733.
41

19.Wang, T.J., C.H. Wu, and L.J. Lee, In-Plane Permeability

Measurement and Analysis in Liquid Composite Molding. Polymer

Composites, 1994. 15(4): p. 278- 288.

20.Gauvin, R., F. Trochu, Y. Lemenn, et al., Permeability Measurement

and Flow Simulation through Fiber Reinforcement. Polymer

Composites, 1996. 17(1): p. 34-42.

21.Lekakou, C , M.A.K. Johari, D. Norman, et al., Measurement

Techniques and Effects on in-Plane Permeability of Woven Cloths in

Resin Transfer Moulding. Composites Part a-Applied Science and

Manufacturing, 1996. 27(5): p. 401-408.

22.Lai, Y.H., B. Khomami, and J.L. Kardos, Accurate Permeability

Characterization of Preforms Used in Polymer Matrix Composite

Fabrication Processes. Polymer Composites, 1997. 18(3): p. 368-377.

23.Parnas, R.S., K.M. Flynn, and M.E. DalFavero, A Permeability

Database for Composites Manufacturing. Polymer Composites, 1997.

18(5): p. 623-633. 28.

24.Han, K.K., C.W. Lee, and B.P. Rice, Measurements of the

Permeability of Fiber Preforms and Applications. Composites Science

and Technology, 2000. 60(12- 13): p. 2435-2441.


42

25.Lundstrom, T.S., R. Stenberg, R. Bergstrom, et al., In-Plane

Permeability Measurements: A Nordic Round-Robin Study.

Composites - Part A: Applied Science and Manufacturing, 2000.

31(1): p. 29-43.

26.Kim, S.K. and I.M. Daniel, Observation of Permeability Dependence

on Flow Rate and Implications for Liquid Composite Molding.

Journal of Composite Materials, 2007. 41(7): p. 837-849.

27.Parnas R.S. and Salem A.J. (1993), A comparison of the

unidirectional and radial flow of fluids through the woven composite

reinforcements, Polymer composites, 14: 383-397.

28.Han K.K., Lee C.W. and Rice B.P. (2000), Measurements of the

permeability of fiber preforms and application, Composites science

and technology, 60 (12): 2435-2441.

29.Tan H. and Pillai K.M., (2007), Effects of fiber-mat anisotropy on 1-D

mold filling in LCM: A numerical investigation, Polymer composites,

DOI 10.1002/pc.20463.
43

Chapter 3

Evaluation of Boundary Conditions at the Clear-

Fluid Porous-Medium Interface

3.1 Introduction

Transport phenomena in a fluid channel bound by a porous medium layer

are of great importance in a wide range of industrial, environmental and medical

applications. Among the several simulation applications, one can mention

modeling of liquid composite molding (a composites manufacturing process), fuel-

cell transport, filtration of contaminants, drying processes, ground water pollution,

and overland-flow infiltration into soil during rainfalls, as well as applications in

the medical field such as simulation of flows in biological tissues, drugs and

nutrients transport in the body, etc. Transfer of shear stress from the open channel

into the interstitial fluid domains of the porous medium leads to the formation of a

boundary layer inside the porous medium at the open-channel porous-medium

interface. Though Beavers and Joseph [1] had demonstrated that mass flow-rate

through open channels bound by porous walls is higher than for channels bound

by solid walls, thereby inferring the existence of such a boundary layer indirectly,

it was only recently that Tachie et al. [2] showed conclusively through
44

experiments that such a boundary layer region does exist.

In order to solve this problem numerically, one need to establish appropriate

governing equations in the clear-fluid and porous-medium domains. The Navier-

Stokes equations are used to model the flow through the free channel, while the

Brinkman-Forchheimer equation (a.k.a. the Darcy-Brinkman-Forchheimer

equation) is used to model the fluid flow through the porous wall [3-7]. The latter

models the volume-averaged flow through a porous medium close to its clear fluid

boundary. For the low particle-based Reynolds-number flows (i.e. Re < 1), one

can neglect inertial effects in the Brinkman-Forchheimer equation and reduce it to

the well-know Brinkman equation [1]. Additional simplifications can be

introduced by neglecting the second-order viscous term in the Brinkman equation

for the flow of fluid far away from the interface; then the Brinkman equation

reduces to the well-known Darcy's law for modeling flows in ordinary porous

media.

The boundary conditions at the clear-fluid porous-medium interface have

been investigated by several researchers in the past [8-15]. In 1967, Beavers and

Joseph [1] introduced the slip-velocity boundary condition at this interface, which

has later been used by other researchers [8-11] to solve such problems

numerically. Ochoa-Tapia and Whitaker [12-14] introduced a different interfacial

boundary condition which assumed a jump in stress at the interface. They used the

rigorous volume averaging method to show the inconsistency in the stress


45

estimation at the clear-fluid/porous-medium interface. Valdes-Parada et al. [15]

developed the work done by Ochoa-Tapia and Whitaker further to provide an

expression of the stress jump boundary condition at the free channel/porous wall

interface that does not involve adjustable coefficients. They proposed an

approximate methodology to estimate the mixed stress tensor of the stress jump

boundary condition. In another work, they [16] used the same boundary condition

to model the mass transfer at the clear-flow/porous-medium interface. Recently,

Chandesris and Jamet [17,18] developed the stress jump boundary condition using

the matched asymptotic expansion method. They used a two-step up-scaling

approach for three different interface descriptions to identify the characteristic

length scales associated to each level of interface. Later in 2009, the same authors

[19] used the identical method to recover the empirical slip coefficient of Beavers

and Joseph [1] through the multi-scale approach. Duman and Shavit [20] used the

apparent interface concept to develop the stress jump boundary condition at the

clear-fluid/porous media interface. The stress jump boundary condition introduced

by Ochoa-Tapia and Whitaker [12] is very sensitive to the value of the parameter

P and exact location of the interface. They used an alternative method to avoid

the necessity of finding the exact location of the interface; their method needs only

one measurable parameter, the maximum velocity or the flow-rate. Tan et al. [21]

used boundary element method (BEM) to compare the three different interfacial

boundary conditions: Beavers-Joseph's slip-velocity condition, Brinkman's stress-

continuity condition and Whitaker's stress jump condition. They found that the
46

stress-jump condition gives more accurate results for high porosities; however, all

the three conditions are inaccurate at lower values of porosities. Later, Tan and

Pillai [22] used the Galerkin finite element based simulation to model flow at the

clear channel/porous media channel interface using both slip flow and stress jump

boundary conditions. They used a second order adjustable tensor to implement the

stress jump condition for three dimensional flows.

Some of the above listed papers (Larson and Higdon [10,11], Tan et al. [21])

as well as the papers listed later on (Sahraoui and Kaviany [23]) pertain to the

use of flow simulations in a virtual porous medium to either test the efficacy of

various boundary conditions or to estimate relevant parameters for them. Most of

these efforts have been done on periodic models of porous media that assume an

ideal, square-lattice arrangement of cylinders within a rectangular domain. On the

other hand, a non-periodic porous medium with a random placement of cylinders

is closer to the real porous media found in nature. To the best of our knowledge,

no ejfort has been made till date to investigate the boundary conditions adjacent

to a free channel flow on non-periodic or irregular porous-media. Our focus in

this chapter is to explore the efficacy of different boundary conditions at the clear-

fluid/porous-medium interface in a non-periodic porous medium. Our hope is that

such an approach to test the accuracy of the various boundary conditions is more

realistic and should yield results that are observed in nature as well. For the

purposes of comparison, we will also present the simulation results involving the
47

periodic porous media.

3.2 Design of Numerical Experiments

In this chapter, we will restrict our analysis to the creeping or Stokes flow

since the majority of flows involving porous media fall in that regime. Such a flow

will be studied adjacent to a porous medium in a domain to be called a 'hybrid'

domain (Fig. 3.1 A). In order to model the hybrid region including both periodic

and non-periodic porous regions, two different rectangular geometries have been

created. The porous medium is made of a square array of cylinders (representing

the solid phase) for the periodic case (Fig. 3.1 B), and is made of a random

arrangement of cylinders for the non-periodic case (Fig. 3.1 C). In both cases,

cylinders of radii a=10 |^m are used and the depth of the clear fluid region is h=20

um. The porosities of these geometries can be changed from 0.5 to 0.9 by varying

the inter-cylinder spacing, e.

In order to create a periodic porous medium with a specific porosity,

cylindrical-solid phase particles are located at fixed horizontal and vertical

distances from each other in the rectangular geometry.

On the other hand, to create a random distribution of cylinders in the non-

periodic porous medium at a specific porosity, a FORTRAN code is developed to

randomly place the solid-phase cylinders inside the rectangular geometry; the code

estimates the horizontal and vertical distances between cylinders using a random

number generator after ensuring no overlap of cylinders (see appendix for details).
48

ys///y/y/ss//!&///y//y^^ Y//y////A'Awy//y/AV^^^^
channel

(A)

wall

000gQOOj000O0QQ0Oi30O0QQ0O0O00QQ
Inlet ooooo^oi^oaeoo^oTDiDcToooo^Trcraooeo" outlet
oooooooooooooooooooooooooooooo w
oooooooooooooooooooooooooooooo
oooooooooooooooooooooooooooooo
oooooooooooooooooooooooooooooo
wall

(B)

wall

ooooo0o0o0o0o009Pn^P°oo'ooOoooo outlet
Inlet
Hb o o o b o o o o o
OOO O O O Q O O o o o o w
^ _ o o o o o p o OOOOoo
OOOQOO Q O Q O OO o o p o o o o Q O O Q O O OOO
wall

(C)

Figure 3.1 Modeling creeping (Stokes) flow through the hybrid (open-channel porous-medium) region: (A)
A schematic of the velocity profile through the channel and the porous medium, (B) The geometry of the
hybrid region with a periodic porous medium; (C) The geometry of the hybrid region with a non-periodic,
random porous medium. The last two figures show typical REVs (Representative Elementary Volumes)
employed for averaging flow velocities inside the periodic and irregular porous media.

The no-slip boundary condition is applied at the top and bottom walls ( y =
49

h, y = -W) along with the cylinder surfaces. A pressure difference Ap is imposed

on the two vertical boundaries of the inlet and outlet. The resultant pressure-driven

flow experiences the pressure gradient of — = 1— along the flow direction. For

Ax m

the sake of consistency and simplicity, velocity and channel's height are

normalized by the interface velocity and cylinder radius, respectively.

3.2.1 Numerical Solution

The Stokes equation governs the flow of fluid through the clear channel as well as

through the interstitial spaces between cylinders inside the porous-medium. These

governing equations are solved through the finite element method (FEM) based

software, COMSOL, which employs the Galerkin based weak formulation. The

generated numerical result furnishes the point-wise velocity distribution within the

pore space of our computational domain.

Since one has to deal with averaged quantities within a porous medium, the

volume averaging method is applied to average the point-wise velocity distribution

within the porous region. A rectangular representative elementary volume (REV)

with a fixed size is used to average the flow velocities in the domain (see Fig. 3.1).

The averaged velocity is assigned to the centroid of the REV. The size of the REV

is kept fixed while moving it along the y direction. The steps of moving the REV

along the y direction depends on the porosity of the porous medium. As mentioned

earlier, the radius of cylinders remain constant for all porosities while the vertical

and horizontal distances between cylinders change by changing the porosities.


50

Setting each moving step along the y direction equal to a quarter of the vertical

distance between cylinders for the case of periodic media, one can simply

calculate the moving steps for individual porosity, Table 3.1.

Table 3.1 Details of step size for moving REVs corresponding to each porosity.
Porosity (e) 0.5 0.6 0.7 0.8 0.9
Moving step (J) 6.3x10"" 7.0xl0" 6 8.1xl0" 6 9.9X10"6 1.4xl0" 5

Following describes unit cells and moving steps, the imaginary interface

line between the clear channel and porous region is defined at the fifth unit cell

from the bottom of geometry for both periodic and non-periodic cases. In the

periodic case, there is no overlapping of solid cylinders with the imaginary line is

allowed. On the other hand, there is a chance for overlapping of solid-phase and

interface-line for the non-periodic case due to the randomly placed cylinders.

3.3 Open Channel Flow Model for the Theoretical Solution

The domain of the flow problem can be divided into two sections: the clear-

fluid region on top and the homogeneous porous-medium region at the bottom.

The continuity and the Stokes equations governing the steady laminar creeping

flow of a Newtonian fluid through the empty channel are

^ =0 (3.1)
dx

(^2 32 \
a u o u =0 (3.2)
dx
51

where u is velocity, p is pressure and ju is viscosity. The No-slip boundary

condition is used on the top impermeable boundary of the channel. The interface

boundary condition of the clear-fluid region with the porous media region is

discussed below.

3.4 Theoretical Solution of the Porous Medium Flow

In order to solve the interface problem, the Stokes and Brinkman equations

are used to model the steady, one-dimensional creeping flow in the channel and

porous medium, respectively. The interface between the free channel and porous

medium is located at y = 0. The width of the channel is taken to be h. By

considering both of the well-known boundary conditions of stress continuity and

stress jump at the interface, we have the following classical results.

3.5 Stress-Continuity Interfacial Condition

The Brinkman equation can be used to model the flow of fluid through the

porous media when the pore-scale Reynolds number is less than 1 [24,25]. The

superiority of the Brinkman equation over the classical Darcy law is due to its

second-order viscous term which can model the non-linear velocity profile,

especially close to boundaries with gradients [22]. The volume-averaged form of

the continuity and Brinkman equation are described, respectively, as


52

3 <p> f Id2 <u >~\ [i


•+ H ,2 •£- < u >= 0 (3.4)
3x 3y
K

where < u > is the volume-averaged velocity, <p>* is the pore-averaged pressure,

while P and ^ are the effective viscosity and permeability, respectively. The

permeability K can be either a scalar for isotropic porous medium or a tensor for

anisotropic porous medium [22]. For our case, the permeability corresponds to the

flow perpendicular to cylinders. Though the dependency of P on the porous

medium characteristic and fluid viscosity has been studied by some researchers

[23,26-27], n' = n is assumed for the sake of simplicity.

In this case, the continuity of velocity boundary condition mandates that

uclear\i=<u>porous\i (3.5)

where superscripts clear and porous indicate the regions of clear-fluid and

porous medium, respectively. The normal and shear stress continuity at the

interface is also conducted—the former results in the equality of pressure,

while the latter results in the equality of shear stress that is established in the

present case through the relation

j clear •. , _ _ porous
du . 1a <u>
~T '«=7—1— '« (3-6)
dy e dy

By considering the stress continuity interfacial condition, the velocity and


53

stress are assumed to be continuous at the interface of the clear-fluid and the

porous-medium domains. The following velocity profiles in the clear fluid and

porous medium regions are obtained for the case of the pressure-driven creeping

flow [28]:

2 \
K_dp
u{y)-- 1+ M y 2j^4=+y for 0<y<h (3.7)
M JK 2ju dx K\MJK K

(u)(y) = {u)i + ({u)D -(u)i j 1 - „ p ^ | Z - ^ L for -oo< v < 0 (3.8)

where the interface velocity, <«>,, is defined as

K d(p)f
2 i^-^+hl
2ju dx M 4K K (3.9)
<«>.=- H' h
1+
M 4K

Defining the boundary-layer thickness Sc as the width of the region inside the

porous medium near the interface where (u) > 1.0l(u) , one may obtain the boundary

layer thickness through the relation («)| _ =1.01(K)D [23] as

50 * l - 2
K (3.10)
H' h
1+
M JK

According to this relation, one can estimate the boundary-layer thickness to be on


54

the order of K05 (which in turn, is on the order of diameter of cylinders used in the

present study, and hence quite small).

There is no boundary layer inside the porous medium modeled with Darcy's law,

while a boundary-layer regime exists inside the porous medium modeled with

Brinkman equation due to the nonlinear term in the Brinkman governing equation.

Thus the two approaches can be correlated.

3.6 Stress-Jump Interfacial Condition

To model flow inside porous medium near the clear-fluid porous-medium

interface more accurately, Ochoa-Tapia and Whitaker [12] used the volume

averaging method to propose the continuity equation and a more accurate form of

the Brinkman governing equation as

d<u>
=0 (3.11)
dx

3<p> f \l(d2<u>
^<u>=0 (3.12)
dx e1 3y K

where e is the porosity of porous medium. The new equation, to be called

the modified Brinkman equation in the present chapter, is similar to the original

Brinkman equation given in Eqn. (3.4), except for the fact that it dispenses with

the parameter effective viscosity -" which is often used as a fitting parameter in

boundary flows. A comparison of the two equation reveals that ^' = it .


e
The idea that the stress may not be continuous at the interface as the solid
55

may be bearing part of the shear stress transmitted from the channel fluid is an old

one. Ochoa-Tapia and Whitaker used the rigorous volume averaging method to

derive a mathematical relation to describe the stress jump at the clear-fluid porous-

medium interface. For the one-dimensional flow parallel to the interface, the

stress-jump interfacial condition simplifies to

1 d(u) du
e dy ydT "dy ?=o +
=£<->**> (3.13)

where /? is a dimensionless coefficient whose value is on the order of one.

The second boundary condition is the continuity of velocity at the interface,

Eqn. (3.5). Using the stress-jump boundary condition given by Eqn. (3.13) as well

as the continuity of velocity condition, solutions to the Stokes and modified

Brinkman equations, Eqns. (3.2) and (3.12), can be worked out to be

K_dp_ for 0<y<h (3.14)


u{y) = (u) 1 +
^
-p 2ju dx •Je JK K

("}O0 = <"), +(<«>„ -("),) l-exp for - oo < y < o (3.15)

where the interface velocity, (u) , is given as

K d(Py ( 2 h . h2^
2jU dx 4E4K + -K (3.16)
w =•
h
re -p
I+

Once again, using the boundary layer definition, (M)| = l.oi(w) , the
> ' Iy=—§ \ ' D
56

thickness of boundary-layer in the porous region can be found to be

U2 2
h | 2hj3 .
50
<?.= *ln ^KK /KJK / (3.17)

One should note that the above solutions involve volume averaged quantities and,

therefore are applicable to the porous medium only.

It is easy to obtain the solution of the modified Brinkman Equation with the

stress-jump boundary condition by assuming /?=0 (stress continuity) and

substituting e with,«///' in Eqns. (3.14)-(3.17).

3.7 Results and Discussions

In order to compare the theoretically predicted flow variables using various

boundary conditions with the FEM simulated ('experimental') velocity, the

effective viscosity fi' and the adjustable coefficient (3 must be known. The

parameter fx' and /? are computed by adjusting ju' or /? to match the total flow-rate in

the channel calculated from Eqns. (3.7)-(3.8) and Eqns. (3.14), (3.15),

respectively, with that of the FEM results is similar to the experimental procedure

employed by Beavers and Joseph to calculate their slip coefficient [1]. The
5
Beavers-Joseph slip coefficient gets the form a = (uV/O , when the flow-rate

predicted by the Darcy law with the slip-velocity boundary condition at the

interface is equated to the flow-rate predicted by the Brinkman equation in

conjunction with the stress-continuity condition. Values of numerically estimated


57

p' I p., P and a are listed in Table 3.2 based on our calculations. The numerical

results for the periodic case are identical to the values estimated by Hua et al. [21].

Table 3.2 Values of effective viscosity^', coefficient/?, and slip coefficient a at different
porosities
e = 0.5 e = 0.6 e = 0.7 e = 0.8 £ = 0.9

IX'Ip 1.2 1.7 3.2 6.3 14.4

p 0.3 -0.1 -1.0 -1.8 -2.2

a 1.1 1.3 1.8 2.5 3.8

In order to understand the micro-scale flow patterns inside the porous

region near the open-channel porous-medium interface, streamlines for the

pressure-driven channel flow over the periodic porous medium are plotted in the

Fig. 3.2 at four different Reynolds numbers for the porosity £ - 0.6.

Re = 1x10^ Re = lxl0" 5
58

Re slxKT 4 Re = lxlO -3
Figure 3.2 Effect of increase in Reynolds number on the streamline pattern observed adjacent to the open-
channel porous-medium interface for the periodic porous medium. The porosity for the porous medium is
£ = 0.6.

The Reynolds number is calculated based on the appropriate characteristic

length inside the porous region. The cylinder diameter, 2a, is considered as the

characteristic length and the characteristic velocity, the pore-averaged velocity

inside the porous region, is calculated from the Darcy law by employing the

pressure gradient between the two ends of the channel. Density of p = 900kg/m3

and viscosity of // = 220xi0"3/5a.s are assumed to be the fluid properties (close to

the motor oil properties) for entire problem.

The streamlines above the imaginary interface (i.e., the tangent to the upper

layer of cylinders) shows a slight perturbation in the free channel flow due to the

presence of cylinders below. The streamline pattern below the interface is marked

by the presence of identical recirculation patterns between cylinders in the first

row, which is caused by the transmittal of momentum through shear stress from

the open channel to the porous medium. The recirculation pattern remains constant

despite a four orders of magnitude increase in the flow velocity. However,


59

streamlines at a higher Reynolds numbers, such as Re = lxlO~3, are more dense

compared to the lower Reynolds-number plots. It is clear that the pattern of

streamlines do not change with the Reynolds numbers in the Stokes flow regime.

Streamline pattern for a typical irregular configuration is plotted in Fig.

2-45 25 2,55 26 205 27 2,5 2a 295 2 <J 2 95 3 i05 J! 315 32 325

Figure 3 3 Streamline pattern at the interface of clear-channel and porous wall for non-periodic porous
media at e = 0 5 add Re - 1 x 10 ^

One should note that due to the irregularity, no trend is expected. Fig. 3.3

shows that at the interface, the resistance for the flow is higher in the porous

region and hence, more streamline exist in the clear-channel compared to the

porous region.

In Fig. 3.4, the volume averaged velocity inside the periodic porous

medium at the interface is shown for five different porosities at three different

Reynolds numbers (At any porosity, there are three different average velocity
60

profiles corresponding to three different Reynolds numbers.)

0.1

-0.1

-0.3 £=0.6
e = 0.5
-0.5 ©0
-*>-Re=1E-5 -0,7
_—--*£-^ • - v
©0 — Re=1E-5
-*~Re=1E-4
«-Re=1E-4 • •- Ra=1E-3
-0.9
•- Re=1E-3

4.E-03 6.E-03 4.E-03 6.E-03


u/u,'Interlace

(a) (b)

-- — - " 1
0.1 -- - - - -

<f
-0.1

e = 0.7 £ = 0.8
oo -0.3
o o
<x'° oo —»-Re=1E-5
-0.5

•0.7
c S)°'
o o —-Re=1E-5

:
^-Re=1E-4 -»-Re=1E-4

3
• • - Re=1E-3 • • • Re=1E-3
-0.9

-1 1
•05 5.E-03 1.E-02 2.E-02 2.E-02 3.E-02 1.E-05 l.E-02 2.E-02 3.E-02 4 E-02 5.E-02
U/Uintertaco

(c) (d)

0.1

-0.1

-03
£ = 0.9
O o
-0.5

o o
-0.7 -»-Re=1E-5
—>-R8=1E-4
-0.9 • • - Re=1E-3

-1.1
1.E-05 2.E-02 3.E-02 5.E-02 6.E-02 8.E-02
U/UintBr)Hce

(e)
Figure 3.4 Plots of normalized average-velocity inside the periodic porous medium at three different Re
numbers for five different porosities: a) £ = 0.5 , b) £ = 0.6,c) £ = 0.7,d) £ = 0.8 and e)
£ = 0.9.

The x-direction velocity is averaged using the REV shown in Fig. 3.1 B,
61

while the average velocity is normalized with the help of the interface velocity in

Fig. 3.4. We observe that the velocity varies in a sinusoidal manner—this happens

because the thin REV shown in Fig. 3.1 B alternates between row of cylinders and

empty space during its translation along the y axis. (The former leads to a dip in

the velocity while the latter is responsible for the peak.) One can study the effect

of an increase in the Reynolds number on velocity profile in the porous medium:

as expected, the average velocity increases with an increase in the Reynolds

number. However, one can note that the normalized velocity at higher porosity

(such as e = 0.9) is not increased by the increase in Reynolds number. The reason is

that the Darcy velocity is almost similar to the interface velocity at high porosity

values and hence, increase in Reynolds number does not lead to an increase in the

normalized velocity (see Fig. 3.4).

The geometry of an irregular arrangement of the cylinders to model the non-

periodic porous media at porosity 0.7 is presented in Fig. 3.5. The volume-

averaged velocity, obtained using the REV shown in Fig. 3.1C on this weakly-

disordered porous medium, is expected to have smaller fluctuations compared to

the periodic porous medium. This is because arbitrary variations in local velocity

fields caused due to randomness in cylinder locations is expected to damp out the

possible fluctuations in the averaged flow caused due to the presence of 'channels'

between cylinder rows in a periodic porous medium. The horizontal rows of

narrow cells, which constitute REVs corresponding to various y locations within


62

the porous region (Fig.3.5), were employed to volume average the flow velocity

within the porous medium.

&Mdii *MlM •it&i m± >4

• •
• • • • • • • • • • • • • • • • • • • • <

Figure 3.5 An example of a weakly-random arrangement of cylinders employed for the creation of an
irregular porous medium adjacent to the channel. The porosity is 0.7 (i.e., £ = 0.7) for the porous
medium.

The REV independency check was conducted on the irregular case to

ensure that the obtained results are independent of the REV size. This is because a

too-small REV causes fluctuations in the averaged quantity as it moves about

within the porous medium; hence the REV should be sufficiently big so that no

such spurious fluctuations are witnessed within a homogeneous porous medium

[29]. (It is obvious that there is no need tc conduct such an analysis for the

periodic porous media due to its periodic boundary conditions for Stokes flow.)

Using this analysis for the non-periodic porous medium, we plan to optimize the

REV size such that it is rendered suitable for averaging.

Fig. 3.6 shows the sample REV sizes used in this REV-size independence

test. Details of the six REVs used for the purpose of such a test are presented in

Table 3.3.
63

REV 1
,d 3^^tS5CK3Bw«jl^i^Ji
^ - Q O O O O'p Q oQ
REV 2

OPOOOP eff£o REV 3

000o0000obo "oonno^/^e,-—^^
•©OQ w —^~^-~
rtOO OOOQ REV 4

Do PopoOo 0 , 9 0 0c ^ 0 "dPOOQOo OOP OOP REV 5

OOOpOQ Q O Q O o o o t o - g g i & o o - o Q O O O O Q O O P

Figure 3.6 Schematic view of six different REVs used for the REV-size independence study (not in the
actual dimensions).

The largest REV called REV 1 contains 34 cells (Note that the cell size

varies by change in porosity, due to change in the distance between solid

cylinders.) The smallest REV is called REV 6 and has 4 cells.

Table 3.3. Details of the six REVs used in the REV-size independence study.

REV Number of cells Area ( m 2 )

REV1 34 5.36e-9

REV 2 28 4.41e-9

REV 3 22 3.47e-9

REV 4 16 2.52e-9

REVS 10 1.58e-9

REV 6 4 6.30e-10

Fig. 3.7 shows the effect of change in REV size on variations in the normalized

volume-averaged x-direction velocity. One can see that REV 1 and REV 2 seem to

achieve an optimum size for the REV as they minimize fluctuation while at the

same time attain convergence in the velocity plot (especially away from the

interface). The REV 6 is the smallest and perhaps the worst REV for averaging
64

because the velocity distribution diverges significantly from the other, larger-REV

solutions. Because of these results, all the upscaling done in the present chapter is

done using REV 2.

Y/a -3

O.E+00 5.E-02 1.E-01 2.E-01 2.E-01 3.E-01 3.E-01


U/Ujnterface

Figure 3.7 Results of the REV size-independence analysis for the irregular porous medium at £ = 0.5 for

the six REVs listed in Table 3.3.

Due to a large velocity gradient at the interface, a flat line can be observed

near the clear-channel porous-medium interface. This reconfirms the fact that

velocities in the free channel and porous region have several orders-of-magnitude

difference and cannot be modeled through the same governing equation.

Mesh independence analysis was also conducted to ensure that our

numerical results are independent of the number of elements and nodes employed
65

in the finite element mesh. The effects of three different meshes, the details of

which are listed in the Table 3.4, were analyzed while estimating the volume

averaged velocity for our geometry. The normalized velocity distribution along the

y direction has been plotted for the three meshes in Fig. 3.8. Results show a

0.000E+00 -r

-2.000E-05

-«-Grid1

— a - Grid 3
-4.000E-05
- ©- Grid 2
Y/a

-6.000E-05

-8.000E-05

-1.000E-04

O.E+00 5.E-12 1.E-11 2.E-11 2.E-11 3.E-11 3.E-11


U/Uj n t e r f a c e
Figure 3.8. Results of the mesh independence study. (Details of the three grids or meshes are
given in Table 3.4.)

negligible dependence on mesh size and one can use the smaller mesh to minimize

computational time while maintaining sufficient accuracy. Repeatability of this

method in the non-periodic porous media has been analyzed by comparing the

results of three different realizations (Fig. 3.9). The pore-averaged velocity is

plotted along the channel including the interface for three randomly generated

porous geometries.
66

Table 3.4 Details of the elements and nodes of three different meshes employed to conduct the
mesh independence study.

Grid Number of elements Number of nodes


1 54177 29779
2 82313 44301
3 3467328 1755930

The pore-scale results show a similar trend for velocity profile for all three

realizations. This comparison indicates that long enough REV in an irregular

arrangement of solid phase for all different realizations damp out local fluctuations

and upscale to yield similar trends.

. —^» — — 1 ~ _ -
E
a.
•«©- Realization 1
Q.
O - • • Realization 2
- A - Realization 3
c
n

0.0E+00 4.0E-11 8.0E-11 1.2E-10 1.6E-10 2.0E-10


Pore-Averaged Velocity (m/s)

Figure 3.9 Repeatability of the solution by comparing three different random realizations' results

Effect of porosity variation on the normalized velocity distribution along the

y direction is compared with the two analytical solutions in Fig. 3.10. All the

computations were done for the non-periodic case at Re = lxl0" 4 . Through Fig.

3.10, one can observe that the higher porosity cases have a higher normalized
67

-2
£=0.5 Re=lxl0 4
-4 • e = 0.6 Re = lxlfJ- 4
—e— Brinkman Eq Stress Continuity Condition
Y/a -6 Bnnkman Eq. Stress Continuity Condition
'v —a— Brinkman Eq Stress Jump Condition -8 Bnnkman Eq Stress Jump Condition
- ©» Regular Arrangement
-10 « Regular Arrangement
~~°- Irregular Arrangement
-12
*- - Irregular Arrangement
0 000 0.001 0.002 0.003 0.004 0.005 0.006 0.007 0.006 0.009 (
-14
U/Uintortace
0.00 0- <u>,/<u>„ 0.02

B
2
0
-2
e = 0.7 Re = lxl0-<
-4
£ = 0.8 Re = 1x10-"
Y/a
-6
Bnnkman Eq. Stress Continuity Conqit Y/a Bnnkman Eq Stress Continuity Cond tion
-8
Brinkman Eq Stress Jump Condition Qr--a * Bnnkman Eq Stress Jump Condition
Regular Arrangement 10
's-A
-12 «~*y-s Regular Arrangement
Irregular Arrangement 1
ijl-o
-14 — *- — Irregular Arrangement
O.E+00 2.E-03 4.E-03 6.E-03 8.E-03 l.E-(
0.00 0.01 nm oiw nnd not 0.06 0.07 0.08
<U> f /<U>,„ t e r ( a e e <u>,/<u> ,„,.„,„

£= 09
R e = l « 10 •*

Y/a
—e—Brinkman Eq Stress Continuity Condition
—i>- Brinkman Eq Stress Jump Condition
-•**• Regular Arrangement
- o~ Irregular Arrangement

0.05 010 0.15 0.20 0.25 0 30 0.35 0.40 0.45 0.50

<U >, /< U> MM,,

Figure 3.10 Comparison of average velocity using FEM for both regular and random cases with the
Brinkman models at Re= lxlO"4 for A) £ = 0.5 ; B) £ = 0.6 Q £ = 0.7 D) £ = 0.8 and E) £ = 0.9

velocity as well as a smoother transition to interface velocity. As expected, there


68

are smaller fluctuations in velocity for the irregular arrangement of cylinders

compared to the regular one. It is clear that the irregular arrangements of cylinders

helps to damp out the fluctuations in the averaged velocity due to the 'blocking' of

any clear flow path between rows of cylinders.

Similar to Fig. 3.10 which shows normalized velocity versus normalized

height, Fig. 3.11 depicts the actual velocity profile on the logarithmic scale versus

normalized channel height. In this plot, one can see the actual velocity at the

interface of clear channel and porous medium rather than the normalized velocity.

-2

E=0.6
-4 Re=l<104

Y/a

-8 —B—Brinkman Eq. Stress Continuity Condition

—a •Brinknian Eq. Stress Jump Condition


-10
- e -regular modified
-12 - o -Irregular Arrangement

-14
1. E-12 l.E-10 l.E-08 l.E-06 l.E-04 l.E-02 1.E+00

Pore-averaged velocity (m/s)

Fig. 3.11 Actual velocity profile versus dimensionless channel height for £=0.6 and R e = l x l 0 " .

The comparison agrees with the fact that the results of weakly disordered

cylinders are in a better agreement with the Brinkman's model in conjunction with

the stress-jump condition. The main reason for the better, less 'oscillatory'
69

prediction of the irregular arrangement, compared to the regular arrangement, is

the damping out of velocity fluctuations in long REVs due to blocking of gaps

between cylinder rows.

Fig. 3.12 depicts the velocity contour and volume-averaged velocity profile

for the whole geometry for the non-periodic porous media at £ = 0.5 and

Re = lxlCT3. The position of the imaginary interface between clear region and

porous wall is clear in both volume-averaged and contours plots.

65
*I***W
45
'""••i*
*s
,gf 1?
:r* 00
"
v*" ^
O.E+00 5.E-07 l.E-06

A B

Figure 3.12 Velocity contours for the non-periodic porous media at e — 0.5 and Re = lxlO"3: A) velocity
contour and B) volume-avearged velocity

Effect of variation in porosity on volume averaged velocity near the

interface for irregular porous media is plotted in Fig. 3.13.


70

/
Re = lxl0" 4

^ e = 0.5
^ e = 0.6
• * - e = 0.7
-"- e = 0.8
-*>- e = 0.9
1 1

0.04 0.06 0.08 0.10


"'^interface

Figure 3.13 Effect of variation in porosity of the irregular porous medium on the distribution of the volume
averaged velocity near the open-channel porous-medium interface.

As expected, the higher porosity leads to a higher normalized velocity inside the

porous region.

3.8 Summary and Conclusion

Two important interfacial boundary conditions at the interface of free

channel-porous medium are evaluated using a numerical simulation of point-wise

flow based on the finite element method. Two types of porous media modeled by

regular and irregular cylinder arrangement are assumed for the lower porous

channel. While developing the analytical solution, the Stokes and Brinkman

equations are employed for modeling flow through the free channel and the porous
71

medium, respectively. The analytical solutions of the Brinkman equation with the

stress-jump and stress-continuity boundary conditions are presented in details and

compared with the results obtained in this chapter for periodic and non-periodic

cases. Mesh independence and REV independence checks were conducted to

ensure the accuracy of the results. The numerical results of the periodic porous

media at different porosities show that the Darcy velocity gets closer to the

interface velocity for higher-porosity porous media compared to the lower-

porosity one. The numerical results of velocity for non-periodic porous media

were found to be more damped (closer to the Brinkman equation analytical

solutions) compared to the sinusoid velocity profile observed for the periodic

porous media. Increase in porosity of the non-periodic porous media yields better

results in terms of predicting the interface velocity as compared to periodic porous

media.
72

References

[1] Beavers GS, Joseph DD. Boundary condition at a naturally permeable wall. J

Fluid Mech 1967;30:197-207.

[2] Tachie MF, James DF, Currie IG. Velocity measurements of a shear flow

penetrating a porous medium. J Fluid Mech 2003;493:319-43.

[3] Joseph DD, Nield DA, Papanicolaou G. Nonlinear equation governing flow in

a saturated porous medium. Water Res Res 1982;18:1049-52.

[4] Hsu CT, Cheng P. Thermal dispersion in a porous medium. Intl J Heat Mass

Transfer 1990;33:1587-97.

[5] Kladias N, Prasad V. Experimental verification of Darcy-Brinkman-

Forchheimer flow model for natural convection in porous media. J

Thermophysl991;5:560-76.

[6] Vafai K, Kim SJ. Fluid mechanics of the interface region between a porous

medium and a fluid layer-an exact solution. Intl J Heat Fluid Flow 1990; 11:254-6.

[7] Nield DA. The limitations of the Brinkman-Forchheimer equation in modeling

flow in a saturated porous medium and at an interface. Intl J Heat Fluid Flow

1991;12:269-72.

[8] Saffman PG. On the boundary condition at the surface of a porous medium.

Stud Appl Maths 1971;50:93-101.

[9] Taylor GI. A model for the boundary condition of a porous material: part 1. J

Fluid Mech 1971;49:319-26.

[10] Larson RE, Higdon JJL. Microscopic flow near the surface of two-
73

dimensional porous media, part 1: axial flow. J Fluid Mech 1986;166:449-72.

[11] Larson RE, Higdon JJL. Microscopic flow near the surface of two-

dimensional porous media, part 2: transverse flow. J Fluid Mech 1987; 178:119-

36.

[12] Ochoa-Tapia J A, Whitaker S. Momentum transfer at the boundary between a

porous medium and a homogeneous fluid I: theoretical development. Int J Heat

Mass Transfer 1995;38:2635^16.

[13] Ochoa-Tapia J A, Whitaker S. Momentum transfer at the boundary between a

porous medium and a homogeneous fluid II: comparison with experiment. Int J

Heat Mass Transfer 1995;38:2647-55.

[14] Ochoa-Tapia J A, Whitaker S. Momentum jump condition at the boundary

between a porous medium and a homogeneous fluid: inertial effects. J Porous

Media 1998;1:201-17.

[15] Valdes-Parada FJ, Goyeau B, Ochoa-Tapia JA. Jump momentum boundary

condition at a fluid-porous dividing surface: derivation of the closure problem.

Chem Eng Sci 2007;62(15):4025-39.

[16] Valdes-Parada FJ, Goyeau B, Ochoa-Tapia JA. Diffusive mass transfer

between a microporous medium and an homogeneous fluid: jump boundary

conditions. Chem Eng Sci 2006,61(5): 1692-704.

[17] Chandesris M, Jamet D. Boundary conditions at a planar fluid-porous

interface for a poiseuille flow. Int J Heat Mass Transfer 2006;49(13-14):2137-50.

[18] Chandesris M, Jamet D. Boundary conditions at a fluid-porous interface: an a


74

priori estimation of the stress jump coefficients. Int J Heat Mass Transfer

2007;50(17-18):3422-36.

[19] M. Chandesris and D. Jamet, Jump Conditions and Surface-Excess Quantities

at a fluid /Porous Interface: A Multi-scale Approach, Transport in Porous Media,

Volume 78, Number 3, 419-438.

[20] Tomer Duman and Uri Shavit, An Apparent Interface Location as a Tool to

Solve the Porous Interface Flow Problem, Transport in Porous Media, Volume 78,

Number 3, 509-524.

[21] Tan H, Chen X, Pillai KM, Papathanasiou TD. Boundary conditions at the

interface between the clear-fluid and porous medium domains'. In: Ninth

international conference on flow processes in composite materials, Montreal;

2008.

[22] Hua Tan, Krishna M. Pillai, Finite element implementation of stress-jump and

stress-continuity conditions at porous-medium, clear-fluid interface, Computers &

Fluids 38 (2009) 1118-1131.

[23] Sahraoui M, Kaviany M. Slip and no-slip velocity boundary conditions at

interface of porous, plain media. Int J Heat Mass Transfer 1995;35:927^-3.

[24] Reddy JN, Gartling DK. The finite element method in heat transfer and fluid

dynamics. New York: CRC; 2000.

[25] Donea J, Huerta A. Finite element methods for flow problems. England: John

Wiley & Sons Ltd.; 2003.

[26] Martys N, Bentz DP, Garboczi EJ. Computer simulation study of the effective
75

viscosity in Brinkman's equation. Phys Fluids 1994;6:1434-9.

[27] Kaviany M. Principles of heat transfer in porous media. New York: Springer;

1991.

[28] Neale G, Nader W. Practical significance of Brinkman's extension of Darcy's

law: coupled parallel flows within a channel and a bounding porous medium. Can

J Chem Engng 1974;52:475-8.

[29] Bear J., Dynamics of fluid in porous media, Dover Publications, 1988.
76

Chapter 4

PREDICTING THE AVERAGE FLOW

VARIABLES IN POROUS MEDIA USING

ENSEMBLE AVERAGING METHOD

4.1 Introduction

Most of the current studies in the area of porous media are based on using a

suitable representative elementary volume, REV, which is a domain used for

averaging the variables [1]. The problem associated with using REV is that the

size of the REV is needed to be defined for each type of problem. For example,

using a smaller REV for those parts of the problem with larger gradient of changes

while bigger REV for some sections of the computational domain with no sudden

changes. There are no definite standards and specified criteria for the selecting

size of the REV: converged variables tend to oscillate if very small sample is used

while the oscillations begin to dampen out with an increase in the sample size.

Eventually if the sample size is large enough, we begin to get a consistent reading,

and such an REV is called a suitable REV (Fig. 4.1).


77

Domain of
microscopic effects
Inhomogeneous
Averaged
quantity Homogeneous

P,y.(vf),e
REV's volume

Figure 4.1 Variation of averaged quantities with respect to the REV size

Recently, some work has been done for estimating the optimum size of REV

[1-3]. Wang et al. [1] used a new methodology based on estimating the 3-D

hydraulic conductivity tensor for a fractured rock mass to determine the REV size.

They studied the directional hydraulic conductivity of different cubic blocks and

calibrated the hydraulic parameters to quantify the REV size [1]. Nordahl and

Ringrose [2] proposed a criterion based on a normalized standard deviation, to

determine the REV. They showed that the size of the REV depends on both the

property measured and the correlation lengths of the lithological elements. Based

on this, they identified three flow upscaling regimes that each requires a different

method for upscaling. A new insight in to the determination of REV for

permeability estimation at the lithofacies scale and its relation to sedimentological

parameters is given [4].

The REV must be much smaller than the size of the entire domain;

otherwise the resulting averaged description through the whole domain cannot
78

represent what happens at individual points in the side domain. On the other hand,

the size of the REV must be sufficiently larger than the size of a single pore such

that each REV includes a sufficient number of pores to generate a meaningful

average [4]. By assuming ^ t o be flow variable in a computational domain, the

main assumption in using the REV technique is that the values of *P are not

supposed to change significantly in that REV while in porous media problems,


cm m
specially in industrial porous media of smaller dimensions (i.e. from to ),

strong gradients in ¥ renders the use of REVs problematic.

One of the practical examples of such a problem is the boundary layer

formation (Fig. 4.2, [5]) during the convective heat transfer in porous medium

adjacent to a heated vertical flat plate. A thin thermal boundary layer is formed

when the Raleigh number, Ra, takes large values. The governing equations for 2-D

boundary layer [5] can be expressed as

3<«> 8 < v> (4.1)


dx dy

K d<P (4.2)
<u >= '> Pg/3(<T>-T„)
M dx

d<P'> (4.3)
= 0,
dy

3<T> d<T> d<T> d2<T> (AA\


a +<u> +<v> = am ;— VT-^V
6t dx dy dy

where the subscript °° denotes the reference value at a large distance from the

heated boundary and P denotes the difference between the actual static pressure

and the local hydrostatic pressure.


79

Porouss Medium | T = T,
w

T=T

Figure 4.2 Dimensionless vertical velocity, (a), and temperature profiles plotted against the similarity
variable r|, (b), for natural convection adjacent to a vertical heated surface placed in porous media (Cheng
and Minkowycz, 1977).

The Oberbeck-Boussinesq approximation is assumed in the governing

equations and the Darcy law is valid for this problem. One can see that a large

gradient in the velocity and temperature is present in a very thin layer in the

porous media region adjacent to the heated wall. The question arises as to whether

it is possible to study the velocity and temperature variations in that large-gradient

thin-layer region through the volume averaged equations, Eqns. (4.1) to (4.4). Is

there any other way to avoid the assumption of constant property of a fluid in the

REV while averaging? (Note that the fluid properties are assumed constant within

an REV during averaging, while it is clear that a large temperature gradient in the

boundary layer will induce a sharp gradient in the viscosity as well.


80

Another example of weakness of using REV approach in industrial

problems with larger gradients is modeling the flow of resin in dual-scale porous

media found in the liquid composite molding process, LCM, where the resin flow

happens in thin molds (typically few millimeters) with fiber bundles of the size of

about a millimeter [6-8]. The main question arising in this problem is that are the

average temperature equations valid in such thin geometries where the temperature

values change significantly across the thickness of such a thin mold [6-8]. (Note

that often a temperature difference on the order of 100 °C exist between mold

walls and the incoming resin flowing between the walls.

'cold ^

•hot

Figure 4.3 Liquid composite molding process with hot walls of mold on sides and the cold resin.

In this chapter, a new method to be called the ensemble averaging method

will be introduced for testing the accuracy of averaged equations in the region of

sharp gradients. The main limitation of volume averaging method happens in

regions with large gradients in flow variables such as velocity, temperature, etc.

The situation gets worse when a property depends on another property such as the

fluid viscosity being a function of temperature, i.e., // = ju(T). The overall aim of

this work is to use weakly disordered porous media to test the accuracy of the

averaged equations.
81

In the current chapter, we seek to first use the ensemble averaging method

for the estimation of permeability of our porous medium at different porosities.

Later we use the method to obtain the average-velocity profile at the porous-media

clear-channel flow interface and compare it with the results of previous models

available for such flows. Using this technique, one can easily find if the governing

equations obtained for much larger system with clearly defined REVs are

applicable to regions of porous media with sharp gradients.

4.2 A Conceptual Description of the Ensemble Averaging Method

Using the concept of ensemble averaging available in statistical mechanics

[9], the mean of a quantity that is a function of a porous system's micro-state is

deduced from the ensemble of possible states based on the random distributions on

its solid and liquid phases. A random medium is characterized by an ensemble of

disordered states. Therefore, the mathematical expression of variables varies from

realization to realization. However, the mean of all realizations given for a

physical quantity is independent of the ensemble chosen. In general, the ensemble

average value of some physical quantity such as x at a state / can be defined by

Eqn. (4.5) [9] as

(*(?,)) = i«n—y> (,, a,) (4-5)

A realization in our case is one arrangement of randomly placed cylindrical

fibers to recreate porous medium. Each ensemble average of a flow variable is


82

obtained by volume averaging the flow variables over several realizations. The

phase average of a flow variable ¥ can be described as:

(4.6)
W^J*'^

where V and Vf are the total volume of REV and the pore volume within the

REV, respectively. Subscript/refers to the fluid phase present inside pores. It can

be noted that as a small averaging volume moves around in an REV, the particles

'move' about in a random manner, creating a new realization in the averaging

volume. Thus splitting an REV into N small averaging volumes is perhaps

equivalent to creating N random realizations and then averaging the flow

variables in a small averaging volume within each realization.

*/> *,)

(a) (b)

Figure 4.4 Anticipated variation in (*Pf \ through an increase in a) Volume of REV and b) Number of

realizations

One can rewrite Eqn. (4.6) as


83

v(v,)=\vfdv (4-7)
v
,

The following analysis can be done for two general cases. In case one, each cell

has a specific number of particles or cylinders all of which remain inside the cell.

In case two, assuming a fixed number of particles in each cell, at the most a half of

each cylindrical particle can be located outside the cell borders. By assuming N

cells inside the REV, Eqn. (4.7) can be rewritten as


V (*¥,) = \mfdV+ ft,dV+...+ l^fdV (4.8)
V V V
A f2 f»

such that the REV volume can be split into N cell volumes as

V=V,+V2+... + VN (4.9)

Similarly, the total pore volume within the REV can be split into pore volumes

within the cell as

V
f=VA+Vf>+- + V
fN (4-10)

By expressing Eqn. (4.8) as

v(^ / ) = ^ jyfdv+^- [vfdv+...+^- fydv (4-H)


V V V
\ V, l V,2 N V,N

and then dividing both sides of Eqn. (4.11) by the total volume, we get

\ fl v vlv{ ' v vlvifj ' v vNV{ !

Assuming all cells volumes used for averaging to be equal, one can write
84

V =v = =v =v (4.13)
vl — v2 — ... — v N — v c
Using Eqn. (4.13) in Eqn. (4.12), we get

(4.14)
<*/>-£ J«F,dv+i- ft,dv+...++- ftfdv

Knowing that

(4.15)
V N

and on defining

(4.16)
hlvlv
one can rewrite the Eqn. (4.8) as

(4.17)

Now, instead of assuming each cell to be a part of the REV, we assume each cell

corresponds to the cell used in a realization for averaging, Fig. 4.5.

• • • •

• • •""
" • •
• •
• # / A,
,A '•'" •

' *'l' -'•.
A Realizations
• •
§u~ 'A

A"
r ' > "i •
• •
• %"' Realization 2 |


• •
• •
Realization 1

Figure 4.5 Schematic view of random locations of cylinders inside cells in different realizations
85

In other words,

(4.18)

such that

haf) = Cell phase average for ith realization = — f^rfV (4.19)


c
v,,

Fig. 4.4 shows that as V increases, the OPJ is expected to converge to a fixed

value. Similarly, same pattern is likely to happen for (*¥f) when N , the number

of realizations, increases.

We will conduct a similar analysis here to relate the intrinsic phase average

(uff) of a flow quantity \\r to its ensemble average of the cell intrinsic averages

lwf\ / . By introducing the intrinsic phase average of x¥{ as

W-jrJ*/^'/ V,
(4.20)

one can rewrite the above equation as

(4.21)

which can be further changed to

Vfh¥jY = frfdV + jy¥fdV+...+ \VfdV (4.22)

/ vf v. , vf , vf ,
(4.23)
^yfhaf)f =-*- J*, dv+-^ Jyf d v+...+-^ j^ f dv
V v v
f> v,, f2 v [2 fN vfN

Now, we can compare two possible cases that can happen in a unit cell.
86

Case 1:

If a cell has fixed number of cylinders and all cylinders remain inside the cell

completely (Fig 4.6), then it is easy to show that

Figure 4.6 Fixed number of particles totally inside a cell, case 1

V =Vf =... = Vf =Vf (4.24)

and hence

Vf=VA+Vf2+- + V
f„=N-Vfc (4.25)

On using Eqns. (4.24) and (4.25), one can show that

(4.26)
Vf vf Vf V
f N

On further using Eqn. (4.26) in Eqn. (4.23), it is easy to see that

frfdv+±jyfdV+^+J- frfdV
W-£ r
/ . Vf, f V
>H /» v„ (4.27)

where

v/ 1 (4.28)

If we define
87

<+«.,+•-«,„] (4.29)
\ // N

The final form of Eqn. (4.27) reduces to

(N , , ^
*,)'=- (4.30)
N

where

*y.)f - Cell instrinsic phase-average for ith realization= J_ |\p jy (4.31)


/. K,

From Eqn. (4.29), it is clear that an increase in the number of realizations will lead

to the convergence of ( ^ ) to a fixed value.

Case 2:

We now assume that a cell has a fixed number of particles, such that up to a half

of any particle can go outside the cell boundary during the generation of a

realization of our porous medium (Fig. 4.7).

Figure 4.7 A typical distribution of a fixed number of particles inside a cell for Case 2.
88

In such situation, Vf,Vf^,...,Vf may or may not be equal to each other. If the

cells are created by translation of a cell within a REV, then

v
/=^/,+^+- + ^ (4.32)

n • >u * V, «Vf (i.e., Vc <<: v


), we get
f f
On assuming that -
V,
-L~. 1 (4.33)
Vf ~ N

Using Eqn. (4.33) with Eqn. (4.23), it is easy to obtain

[xV,dV + — \*¥,dV +...+— hf.dV (4.34)


*, > / - L —
"/, v h v v
f„ v, v
h h n

after imposing an approximation that

V
/ i ~Vf2 a
-VjN (4.35)

N has to be large enough to obtain the following approximate relation

(4.36)

Unlike the exact relation, Eqn. (4.30), for Case 1, the intrinsic phase average is

only approximately equal to the averaging of the cell intrinsic-phase averages.

This approximation is likely to improve with a increase in the several theoretical

models based on idealized geometries are available in the literature.

4.3 Validation

4.3.1 Permeability Estimation in a Porous Medium


89

The constant of proportionally between the flow-rate and the pressure

gradient in the Darcy law in a porous medium is related to permeability.

Permeability measurement is one of the main concerns in the study of porous

media. Permeability is typically a function of the geometry of the porous medium.

In 1927, Kozeny [10] developed the first model for permeability based on

an idealized porous-medium geometry with tortuous capillaries. Then Carman [11]

used the theory of specific surface area of the medium particles to modify

Kozeny's model. The resulting expression became the famous Kozeny-Carman

equation [12] which expresses permeability as function of porosity and

particle/fiber diameter (More information on some older permeability models can

be obtained from [4, 12]). In recent years, numerous research efforts have been

devoted to the development of a permeability model. Brushke and Advani [13]

proposed a relation for permeability for the case of fluid flow across a periodic

array of cylinders in 1993. Papathanasiou [14] did a computational investigation

on the effect of macro and micro scale porosities on the effective transverse

permeability of square arrays of permeable fiber tows. He described the

permeability as a power law function of the effective intra-tow porosity. In 2006,

Chen and Papathanasiou [15] developed a model for the axial permeability of

fibrous media by considering the underlying microstructure and its variability

numerically. They used unit cells consisting of 600 randomly placed fiber cross-

sections for modeling Stokes flow in a unidirectional fiber array. They considered
90

a large REV to represent a large spectrum of disordered fiber distributions. Monte-

Carlo (MC) procedure, similar to the method used here, was used for generating

an ensemble of non-overlapping disks [16] and Stokes flow through a large

number of unidirectional, disordered fiber arrays simulated.

One of the applications of the proposed ensemble averaging method as

applied to the study of flow and transport in porous media is to estimate

permeability. A simple 1-D flow through a porous medium is being modeled and

solved using this new technique (Fig. 4.8a). The boundary conditions for the

shown rectangular geometry are uniform velocity at the inlet, zero gage-pressure

at the outlet, and symmetry conditions on sides. For modeling this problem via the

ensemble averaging technique, the computational domain is divided into square

unit cells, in which each unit cell contains a solid cylinder inside it (Fig. 4.8b).

Using the ensemble averaging technique, one should solve for different

realizations such that each realization should have a random location for the

cylinder inside any square unit cell. The computational domain is then solved for

velocity and pressure values in each unit cell via the Stokes equation using the

finite element method based commercial software COMSOL [17].

The permeability value for each realization will be estimated through the

Darcy Law, Eqn. (4.40), as applied to the macro-flow geometry shown in Fig.

4.8a. The average of different realization is called the ensemble average of


91

permeability. Applying this technique to problems of flow in thin porous media is

recommended.

Symmetry boundary condition Outlet pressure


Uniform inlet condition
velocity

(a)

0.09"
0.08
0.07
D o0ooo o
O Oy • 0 o
o o o
0.06
0.05

oo o o o o
Y X
0.04 O
0.03
0.02 O o o o o oo
0.01
0
o o oo o o c
X
-0.01 .
0.02 0.04 0.06 0.08 0.1 0.12

(b)
Figure 4.8 (a): The geometry of the problem, (b): Details of a sample realization with square unit cells each
including a solid cylinder.

On applying this method, we anticipate that one can easily converge to

correct results without any concern about the REV size vis-a-vis the domain.

4.3.1.1 Permeability Measurement Using Ensemble Averaging

Method
92

We recall the definitions of the phase average y¥f) and the intrinsic phase

average ( ^ ) for any flow property ^ as

(«P)=i[Wv (4.37)

<*,)'-£#* (4-38)

By assuming an isotropic porous medium in this 1-D flow case, the continuity

equation and the Darcy law can be written, respectively, as

V.(uf) = 0 (4.39)

(S ) = _K V ( P f ) f (4.40)
n

On substituting Eqn. (4.40) into Eqn. (4.39), one can easily show from the Laplace

equation for average pressure and imposed pressure boundary conditions that the

pressure gradient should be linear with respect to x and can be shown to be

* w _<',>;-w: (4.41)
dx

where (Pf) and (Pf) are the intrinsic phase average pressure from two different

unit cells at the same horizontal level at the inlet and outlet, respectively, and L is
93

the horizontal distance of these two unit cells. By substituting Eqn. (4.41) in Eqn.

(4.40), one can easily find the relation for permeability through the Darcy law as

(4 42)
K=-A-r'(«r)i -
fc>;-<Pr>:

u.) is the mean value of the x-component velocity of the phase averaged fluid
' I ave

u
velocity in the porous medium, i.e., /„ \ =\ f/i + \ u '/z where (Uf) and /„ \ are the

averaged velocities in control volumes (cells) close to the inlet and outlet,

respectively. The same governing equation and boundary conditions apply to all

different realizations, Fig. 4.8a. The ensemble average value for permeability can

be obtained using different realizations as

1 ^
TJZ<«
N W
r
f
>
ave.i
Kave = M L < r M ' = =f (4.43)
f <P
N .X
i=l
<Pf> 2,i-X
i=l
f>^

where Krealn represents the value of permeability obtained from the n*

realization.

4.3.1.2 Results and Validation

For the sake of validation, the ensemble average result is compared with

several models for permeability including the Brushke-Advani [18], Gebart [19],

Davies [20], Kozeny Karman [10], Rumpf and Gupte [20] (see Table 4.1).
94

The finite element method is also used to calculate the permeability for the

reference case of square arrangement of cylinders using the commercial software

ofCOMSOL4.1.

Table 4.1 Five different permeability models


Model Formula Comment

•, 4
2 2
l2=-(l-£)
K 1(1-/ ) tan-'cVjJ) e n
Bruscke-Advani [18]
r2"3 V (J/
VT^ 7 2+1)
r is the fiber radius

^min = 1 " ^ for square


K. 16 ll-£mn %
Gebart[19]
arrangement

Davies [20] K—
64(1 - £)^{l + 56(1 -ef]

M ,5 = - for sphere
d
3
C £
Kozeny Carman [10] K—
M2(\-£)2
Ms=— for cylinder
d

Dp2 is surface average


p
Rumpf and Gupte[20] K—
5.6
sphere diameter

The square arrangement involves placing circular cross-sections of the fibers

in a square array in the rectangular region of Fig. 4.9b.


95

(a) (b)
Figure 4.9 (a) Two different realizations randomly selected, (b) Square arrangement of periodic fiber mat
array.

Figure 4.10 shows the results of different permeability estimation methods for the

square arrangement of fibers in porous medium. The validation is done for the

range of porosities from 0.4 to 0.9 with 10 realizations. Note that the permeability

is rendered dimensionless using the diameter of cylinders.

•Ensemble Average Method


0.500 Square Array
Gebart
— - - Kozeny-Karman
— - Davies
0.375 — - Rumpf-Gupte
—x— Brushke-Advani
K/D2
0.250

0.125

0.3 0.4 0.5 0.6 0.7 0.8 0.9

Figure 4.10 Comparison of the obtained dimensionless permeability values from ensemble average method
with six other results.
96

There is a perfect agreement between our results and other results for porosities

from 0.4 to 0.6 and then, the ensemble results getting closer to the square-array

numerical results for porosities from 0.6 till 0.8. The ensemble average

permeability for the porosity of 0.9 is almost in the middle of the other methods'

predictions.

In Table 4.2, details of the permeability values for different porosities

through different methods are listed. One can easily notice that the ensemble

results are very close to the square array results for all porosities.

Table 4.2 Details of permeability values at different porosities for each model.

Porosity

Permeability 0.4 0.5 0.6 0.7 0.8 0.9

Ensemble Average 3.37E-07 1.02E-06 2.12E-06 3.73E-06 6.65E-06 1.31E-05

Square array 3.63E-07 9.58E-07 1.96E-06 3.44E-06 6.21E-06 1.15E-05

Gebart [19] 1.54E-07 5.27E-07 1.33E-06 9.97E-07 3.17E-06 1.45E-05

Kozeny Karman [10] 3.62E-07 8.49E-07 8.15E-07 1.32E-06 1.97E-06 1.12E-05

Brushki- Advani [18] 1.60E-07 5.63E-07 1.48E-06 1.17E-06 4.03E-06 1.77E-05

Davies [20] 5.02E-07 9.00E-07 1.76E-06 1.26E-06 4.00E-06 1.55E-05

Rumpf and Gupte [20] 2.26E-07 6.43E-07 1.40E-06 8.33E-07

The error of each method's result based on deviation from the squared array

result is tabulated in Table 4.3. The last column represents the average error of
97

each result through the all porosities. It shows that the ensemble averaging method

has the average error of 8.6% which is quite good in comparison to the other

methods.

Of the permeability models considered, the Davies and the Kozeny Karman

methods give the most accurate results with respect to the squared array result

with the average error of 31.5% and 33.7% respectively.

Table 4.3 Error's for each permeability model with respect to the square array results.
Porosity

Error with respect to

square array 0.4 0.5 0.6 0.7 0.8 0.9 Error Average

Ensemble Average 7.3% 6.5% 8.1% 8.5% 7.1% 14.4% 8.6%

Gebart 57.6% 45.0% 32.0% 71.0% 49.0% 26.4% 46.8%

Kozeny Karman 0.4% 11.4% 58.4% 61.7% 68.3% 2.2% 33.7%

Brushki- Advani 55.9% 41.2% 24.5% 66.0% 35.1% 54.7% 46.2%

Davies 38.1% 6.0% 10.3% 63.5% 35.5% 35.6% 31.5%

Rumpf and Gupte 37.8% 32.8% 28.4% 75.8% 43.7%

4.3.1.3 New Permeability Models

Two new models of permeability can be proposed through the curve fitting of

results obtained from our proposed method in this section. One can estimate the

ensemble-average permeability value using

^V=7.3xl0-'f5740 (4.44)
D)
98

where permeability, K, is a power function of fiber's diameter, Df, and porosity,s.

This formula has a very low standard error (0.0151502) and very good correlation

coefficient (0.9961584) which clearly shows that the curve fitting is very accurate.

Another useful form of the permeability relation is an exponential function of the

form

— = 5.4X1(TV340£ (4.45)
D
)

which has a very low standard error value of 0.0199310 and an accurate

correlation coefficient value of 0.9933420. These two fitted relations were

introduced in this chapter for the first time and one can use these relations for

permeability estimation from the porosity and fiber diameter data.

Figure 4.11 shows the dependence of the ensemble averaging method on the

number of realizations to obtain accurate results. It is obvious that increasing the

number of realization leads to a more accurate result where the ensemble results

are compared with the semi-empirical Kozeny-Carman [10] method.


99

3.66E-07

3.64E-07

<^W>s-^<>>i^>i<>*-<<S*a^-^<^^
3.62E-07

R 3.60E-07
£ • One Realization
a 3.58E-07 • Two Realizations
A Three Realizations
3.56E-07
x Four Realizations
OH

x Five Realizations
3.54E-07
QlSquare Array

3.52E-07
0 1 2 3 4 5 6
Number of realizations

Figure 4.11 Dependence of the ensemble-average permeability value on number of realizations used for
estimating the ensemble average.

4.3.2 Clear-Fluid and Porous Media Interface Problem

Problems of fluid flow at the boundary of the porous medium and clear

channel flow are of the great importance in industrial applications [21]. One can

use the Stokes equation, Eqn. (4.46), and the Darcy's law, Eqn. (4.47) to model

the clear flow region and the flow in the isotropic porous medium, respectively:

-Vp + nV2u = 0 (4.46)

(u f )=-fv( Pf ) f (4.47)

where u and (ij represent the point-wise and volume-averaged velocities


100

while p and (p)f represent the point-wise and pore-averaged pressure,

respectively. K is the isotropic permeability of the porous medium and ju is the

fluid viscosity. Figure 4.12 shows the geometry of the porous media-clear fluid

boundary problem. There is the no-slip boundary condition at the y = h , on the

wall, and the constant pressure in the inlet (x - 0) and zero pressure at the outlet (

x = L). We have assumed both boundary conditions of stress continuity and stress

jump for the interface area.

H
X Imaginary Interface

(a) (b)
Figure 4.12 Creeping flow through the hybrid clear-fluid-porous medium interface: (a) schematic view and
(b) Ideal geometrical model

The slip velocity boundary condition at the interface problem modeled by

Beavers and Joseph [22] as:

±U-£e<»-M.> (4-48)

Here u is the x-direction tangential channel velocity beyond the interface for

y>0, M(0) is the tangential channel velocity at the interface y=0, (u)D is the
101

tangential volume-averaged velocity inside the porous media calculated by the

Darcy law, K is the permeability and a is the slip coefficient. Then, Saffman [23]

estimated the slip coefficient mathematically and showed that the ex depends on

the location of the interfacial boundary. Sahraoui and Kaviany [24] showed the

dependency of the slip coefficient on more factors such as porosity, Reynolds

number, channel size and interface location.

The Brinkman equation, Eq. (4.49), is a partial differential equation with a

second-order viscous term acts as a momentum equation for the flow through the

porous media. Using the Brinkman equation, Eqn. (4.49), for conserving the

momentum governing fluid flow in the porous medium, both the velocity and the

shear stress are often assumed to be continuous at the interface.

-v(py+Mv^-^=o (4.49)

Here ju and ju' represent the fluid viscosity and the effective viscosity,

respectively. By using the Brinkman equation in the hybrid (porous medium- clear

channel) region, it is assumed that stress is continuous at the interface [25-31]. The

solution in terms of velocity distributions in clear channel and porous medium

channel can be described as

K_dp
«(>')={"), i + . l * - - y 2l^^U-y fOr 0 < y < h (4.50)
M JK, 2H dx fi 4K Kj

(14)(y) = ( M ) i + ( ( M ) f l - ( „ > j l - e x ^ - ^ for - oo < y < o


(4.51)

where the interface velocity is defined as


102

K d(p)f
(4.52)
H 2(i dx VM- V K + K 1/ 1 + -\(^ V K

In 1995, Ochoa-Tapia and Whitaker [32] proposed the following equation for

transport of flow through the porous media close to the free channel-porous

medium interface:

-V<„)'+|v»-|(u) = 0 (4.53)

They have found that there is continuity of velocity at the clear-flow porous media

interface while there may not be such continuity for the stress. They used the

rigorous volume average method, to derive an expression for stress jump at this

interface. Using the condition of (u)\ =1.0l(w) , the boundary layer thickness

can be given by

S.=.
£ ta V ^ (4.54)
£ . . h ( I
l+
7K{Te-ft
where /? is the stress continuity coefficient (i.e, for /? = 0, stress is continuous at

the interface). Ochoa-Tapia and Whitaker introduced the stress-jump boundary

condition at y = 0 for the 1-D parallel flow case as

ld(u) du P
€ dy j=0"
"dy
y=0*
(4.55)

Knowing the ft to be of the order of one and using the Eqn. (4.55), the velocity

profiles for the Stokes and modified Brinkman governing equations can be
103

derived:

K^dp
u{y) = (u) 1 +
4s )4K 2/n dx K4e JK K (4.56)

{u){y) = {u)i + (( M ) D -( M )^i-ex^-^]


(4.57)

The interface velocity is defined as

K d{p)' [ 2 h h2
2// dx \-[e 4K K (4.58)
( « > , = •

Many researchers used the stress-jump boundary condition which gives good

results [33-39]. Tan and Pillai [40] recently implemented a finite element solution

for modeling flows near the porous medium-clear fluid interface.

4.3.2.1 Results and Validation

Results of the Ensemble averaging method is compared with the solution of

the Brinkman equation with both the stress-continuity interfacial condition [41]

and the stress jump interfacial condition [32, 42-43]. The averaging of the flow

variables (i.e., velocity) were done through rows of unit cells in a same height. The

ensemble average of several realizations gives the ensemble averaging result for

geometry studied. These comparisons have been done for two different Reynolds

numbers of Re =0.01 and Re =0.001 , Fig. 4.13.


104

Y[m]

0.12 0.14

-*- Ensemble Average

-•- Brinkman_Slress Continuity

-*- Brinkman_Stress Jump

Velocity [m/s]

(a)

Re = 0.001

-*- Ensemble Average

-*- Brinkman_Stress Continuity, Re=0.001

-*- Brinkman_Stress Jump, Re=0.001

Velocity [m/s]

(b)

Figure 4.13 Velocity distribution in the clear-fluid porous medium interface: (a) Re = 0.01 and (b)
Re = 0.001

The results of the velocity profile along the thickness of the hybrid channels

for the cases of Re = 0.001 and Re = 0.01 are in a good agreement with the solution

of the Brinkman equations [32, 41-42]. Fig. 4.14 shows the dependence of the
105

ensemble averaging method on the number of realizations to obtain reasonable,

converged results.

velocity [m/t]

(a)

velocity [m/s]

(b)
106

•0 05
0 Oi *00

\)l

III
I II
—•— Ensemble Average
4*<H»
i ii --B--1 Realization
ii
- -A- • 2 Realizations
-•K--3 Realizations
T T
I I - - * - - 4 Realizations
I I
I I
I I - -»•- • 5 Realizations
I I
I III —&— 6 Realizations
I
I
- -o- • 7 Realizations
I II
i n - — • 8 Realizations
i i
i i hi
i «i

velocity [m/s]

(C)

Figure 4.14 a) Dependency of the Ensemble average method on number of realizations; b) Part A;
c)PartB.

It can be observed that having only one or two realizations can cause

unrealistic fluctuations inside the porous medium. As the number of realizations

increases, the Ensemble average results become more accurate. It can be seen that

the solution for this problem was obtained accurately with maximum 8 realizations

after attaining a suitably converged solution.

4.4 Conclusion

In this study, a new averaging technique through the concept of ensemble

averaging is introduced and applied to study of two problems of flow in porous

media. The important benefit of this method is that the weakness of REV's size

dependence in the volume averaging method is omitted and it yields an approach


107

to evaluate the effectiveness of governing equations established using the volume

averaging method in regions with sharp gradients of a parameter. In the first

application, ten realizations were studied and averaged to estimate the

permeability of a rectangular domain filled with porous medium for different

porosities. The ensemble average results validated by comparison with five

different models for permeability, show good agreement. In the second

application, the ensemble averaging technique is used to model the boundary layer

of porous medium-clear fluid interface with two different boundary conditions as

an example of a problem with a large gradient of variables in a small domain. The

results are compared with the available results in the literature and show a

reasonable agreement. Such reasonable agreements in both applications show the

strength of the proposed ensemble average method as a powerful mathematical

tool to evaluate the effectiveness of the conventional volume-averaging based

governing equations in porous media with sharp gradients.


108

References

1. Wang M., Kulatilake P.S.W., UM J., Narvaz J., Estimation of REV size and

three-dimensional hydraulic conductivity tensor for a fractured rock mass

through a single well packer test and discrete fracture fluid flow modeling,

International journal of rock mechanics and mining sciences,

2002, vol. 39, no7, pp. 887-904.

2. Kjetil Nordahl and Philip S. Ringrose, Identifying the Representative

Elementary Volume for Permeability in Heterolithic Deposits Using

Numerical Rock Models, Mathematical Geosciences, Volume 40, Number

7/October, 2008.

3. Torquato S., Random heterogeneous materials. New York: Springer; 2001.

4. Jacob Bear, Dynamics of fluids in porous media, Dover publications, INC.,

1972.

5. Donald A. Nield, Adrian Bejan, Convection in porous media, Springer; 3rd

edition.

6. Ratle F., Achim V., Trochu F., Evolutionary operators for optimal gate

location in liquid composite moulding, Applied soft computing, Vol.9,

Issue 2, pp:817-823.

7. Luthy T., Ermanni P., Flow monitoring in liquid composite molding in

linear direct current sensing technique, Polymer composites, Vol.24, Issue

2, pp:249-262.
109

8. Umer R., Bickerton S., Fernyhough A., Modelling liquid composite

moulding preocesses employing wood fiber mat reinforcements, Key

engineering materials, Vol.334-335, pp:113-l 16.

9. Norman Davidson, Statistical mechanics, McGraw-Hill book company,

INC. 1962.

10. Kozeny, J., Uber kapillar leitung das wassers im bpden, Sitzungsberichte

wiener akademie der wissenschaft, Abt Ha 136, 271-276 (1927).

11. Carman, P. C , Fluid flow through granular beds, Trans Inst. Chem. Eng.

15,150-166(1937).

12. Scheidegger, A. E., Physics of flow through porous media,3rd ed.,

University of Toronto press, Toronto, 1974.

13.M.V. Brushke and S. G. Advani., Flow of generalized Newtonian fluids

across a periodic array of cylinders, J. Rheo., 37 (3) 1993.

14. T.D. Papathanasiou, On the effective permeability of square arrays of

permeable fiber tows, Int. J. Multiphase Flow, Vol 23, No. 1, pp.81-92,

1997.

15.Xiaoming Chen and T.D. Papathanasiou, Micro-scale modeling of axial

flow through unidirectional disordered fiber arrays, Composites science and

technology, 67 pp. 1286-1293 (2007).

16.Torquato S., Random heterogeneous materials. New York: Springer; 2001.

17. Www.comsol.com
110

18.M. V. Brushki and S.G. Advani, Flow of generalized Newtonian fluids

across a periodic array of cylinders, J. Rheol. 37 (3), 479-498 (1993).

19. B. R. Gebart, Permeability of unidirectional reinforcements for RTM,

Journal of composite material, Vol. 26, No. 8 (1992) 1100-1133.

20. Porous Media Fluid transport and pore structure, F. A. 1. Dullien, Academic

press, INC., 1991.

21.D.D. Joseph and L. N. Tao, Lubrication of a porous bearing-Stokes'

solution, J. Appl., 753-760 (1966).

22.G.S. Beavers and D.D. Joseph, Boundary conditions at naturally permeable

wall, J. Fluid Mech., 30, 197 (1967).

23.P.G. Saffman, on the boundary condition on the surface of a porous

medium, Stud. Appl. Maths, 50, 93 (1971).

24. M. Sahraoui and M. Kaviany, Slip and No-Slip Velocity Boundary

Conditions at Interface of Porous, Plain Media, Int. J. Heat Mass Transfer.

35, 927 (1992).

25. Salinger AG, Aris R, Derby JJ. Finite element formulation for large-scale

coupled flows in adjacent porous and open fluid domains. Int J Numer

Meth Fluids 1994; 18: 1185-209

26. Gartling DK, Hilckox CE, Givler RC. Simulation of coupled viscous and

porous flow problems. Comput Fluid Dyn 1996: 7:23-48.


Ill

27.Neale G, Nader W. Practical significance of Brinkma's extension of

Darcy's law: Coupled parallel flows within a channel and bounding porous

medium. Can J Chem Engng 1974;52:475-8.

28.Martys N, Bentz DP, Garboczi EJ. Computer simulation study of the

effective viscosity in Brinkman's equation. Phy Fluids 1994;6:1434-9.

29. Sahraoui M, Kaviany M. Slip and no-slip velocity boundary conditions at

interface of porous, plane media. Int J Heat Mass Transfer 1995;35:927-43.

30. James DF, Davis AM. Flow at the interface of a model fibrous porous

medium. J Fluids Mech 2001;426:47-72

31. Costa VAF, Oliveira LA, Baliga BR, Sousa ACM. Simulation of coupled

flows in adjacent porous and open domain using a control-volume finite -

element method. Numer Heat Transfer, Part A 2004;45:675-97.

32. Ochoa-Tapia JA, Whitaker S, Momentum transfer at a boundary between a

porous media and homogenous fluid I: theoretical development. Int J Heat

Mass Transfer 1995;38:2635-46

33. Valdes-Parada FJ, Goyeau B, Ochoa-Tapia JA. Jump momentum boundary

condition at a fluid-porous dividing surface derivation of the closure

problem. Chem Eng Sci 2007; 62(15):4025-39.

34. Valdes-Parada FJ, Goyeau B, Ochoa-Tapia JA. Diffusive mass transfer

between a microporous medium and an homogenous fluid: jamp boundary

conditions. Chem Eng Sci 2006; 61 (5): 1692:704.


112

35.Kuznetsov AV. Influence of the stress jump condition at the porous-

medium/clear fluid interface on a flow at a porous wall. Int. Commun.

Heat Mass Transfer 1997;24:401-410.

36. Kuznetsov AV. Analytical investigation of Coette flow in a composite

channel partially filled whith a porous medium and partially filled with a

clear fluid. Int. J Heat Mass Transfer 1998;41:2556-60

37. Chandesris M, Jamet D. Boundary conditions at a planar fluid-porous

interface for a poiseuille flow. Int. J Heat Mass Transfer 2006;49(13-

14):2137-50.

38. Chandesris M, Jamet D. Boundary conditions at a fluid-porous interface: an

a prior estimation of the stress jump coefficients. Int. J Heat Mass Transfer

2007;50(17-18):3422-36.

39. Tan H, Chen X, Pillai KM, Papathanasiou TD. Boundary conditions at the

interface between the clear-fluid and porous medium domains. In: Ninth

international conference on flow processes in composite materials,

Montreal; 2008.

40. Hua Tan, Krishna M Pillai, Finite element implementation of stress-jump

and stress-continuity conditions at porous medium, clear-fluid interface,

Computers and Fluids, Volume 38, Issue 6, June 2009, Pages 1118-1131.

41. G. Neale and W. Nader, Practical significance of Brinkman's extension of

Darcy's law: coupled parallel flows within a channel and a bounding

porous medium, Can J. Chem. Eng., 52, 475-478 (1974).


113

42. Ochoa-Tapia J A, Whitaker S, Momentum transfer at a boundary between a

porous media and homogenous fluid II: Comparison with experiment. Int J

Heat Mass Transfer 1995;38:2647-55.

43. Ochoa-Tapia J A, Whitaker S, Momentum jump condition at the boundary a

porous media and a homogenous fluid: inertial effects. J Porous Media

1998; 1:201-17.
114

Chapter 5

NON-ISOTHERMAL MODELING OF POROUS

MEDIA FLOW

5.1 Introduction

One of the most important manufacturing processes of polymer composites

is the liquid composite molding (LCM), which includes several other technologies

such as: resin transfer molding (RTM), vacuum assisted resin transfer molding

(VARTM) and Seeman composite resin infusion molding (SCRIM) [1]. In such

technologies, the composites are created by impregnation of a preform with resin

injected into the mold's inlet. Some thermoset resins may undergo the cross-

linking polymerization, called curing reaction, during and after the mold-filling

stage. Thus, the heat transfer effect due to exothermal polymerization reaction of

resin may not be neglected in the mold-filling modeling of LCM. One should keep

in mind that the viscosity of the resin depends on the temperature as well as the

degree of cure. Therefore, the physics of the flow is coupled with the curing of

resin. This shows the importance of heat transfer and curing governing equations

in the non-isothermal flow in both single-scale and dual-scale porous media.


115

Generally, the energy balance equations can be derived using two different

approaches: (1) two-phase or thermal non-equilibrium model [2-5] and (2) local

thermal equilibrium model [6-18]. There are two different energy balance

equations for resin and fiber separately in the two-phase model, and the heat

transfer between these two equations occur via the heat transfer coefficient. In the

thermal equilibrium model, we assume that the resin and fiber reach local

thermodynamic equilibrium once the fiber is impregnated. Therefore, only one

energy equation is needed as the thermal governing equation [3,5]. Firstly, we

consider the heat transfer governing equation for the simple situation of isotropic

porous media. Assume that radioactive effects, viscous dissipation, and the work

done by pressure are negligible. We do further simplification by assuming the

thermal local equilibrium that Ts =Tf =T where Tsand rfare the solid and

fluid phase temperature, respectively. A further assumption is that there is a

parallel conduction heat transfer taking place in solid and fluid phases.

Taking the average over a representative elementary volume (REV) of the porous

medium, we have the following governing equation for the solid and fluid phases

(l-(p)(pc)sd<ls> =(\-<p)V.(ksV<Ts >°) + {\-<p)q: (5.1)


at

< f>
(P{pcP)f d( +(pcP)fv.V<Tf >f = (pV.(kfV<Tf >f) + (pqmf (5.2)

where c is the specific heat of the solid and cpis the specific heat at constant

pressure of the fluid, kis the thermal conductivity coefficient and q"is the heat
116

production per unit volume. Volume averaging of any arbitrary function OF) is

shown by <*F> operator and it can be taken either over fluid phase or solid phase

using <*F>f and <¥> s notations, respectively. By assuming the thermal local

equilibrium, setting Ti =Tf = T, one can add Eqns. (5.1) and (5.2) to obtain a single

equation for the volume averaged temperature as

(pc)m ^ p + (pc)f v.V <T >= V.(kmV <T» + q"m (5.3)

where (pc)m, km and q"m are the overall heat capacity per unit volume, the overall

thermal conductivity, and the overall heat conduction per unit volume of the

porous medium, respectively. They are defined as follows:

{pc)m = (\-<p){pc)s+(p{pcp)f (5.4)

km=<\-<p)k,+<pkf (5.5)

q><\-<P)q:+9qmf (5-6)

5.2 Proposed Work

In this chapter, our aim is to focus on the thermal dispersion term arising in

the governing equation for non-isothermal flow through porous media. The

thermal dispersion term exists as a result of both the micro-structure of the porous

material as well as the heat convection effect. A combined experimental/numerical

approach is introduced for the estimation of the dispersion tensor in the heat-

transfer governing equation of non-isothermal flow through porous media. The

experimental setup and the finite element method (FEM) procedure will be
117

defined in the following sections. Using the experiment data in numerical

simulation leads to an accurate estimation of dispersion tensor in the single-scale

and dual-scale porous media. This is first such attempt to explore thermal

dispersion term through a combined experimental/numerical method, especially in

dual-scale porous media characterized by a complex micro-structure.

5.3 Governing Equations of Non-Isothermal Flow in Porous Media

5.3.1 Macroscopic Level

Pillai and Munagavalasa [19] used a volume averaging method with the

local thermal equilibrium assumption to derive a set of energy and species

equations for dual-scale porous medium. The schematic view of the initial

impregnation of such a medium by a liquid is presented in the Fig. 5.1. Unlike the

single scale porous media, there is an unsaturated region behind the moving flow-

front in dual-scale porous media. The reason for such partial saturation can be

mentioned as the flow resistance difference between the gap and the tows where

the flow goes faster in the gaps rather than the wicking into the tows. Pillai and

Munagavalasa [19] applied the volume averaging method to the dual-scale porous

media. Using woven fiber mat in the LCM, they considered the fiber tows and

surrounding gaps as the two different phases.


118

Resin front

/
Sink effect ^ Gas bubble

Figure 5.1 Schematic view of the initial impregnation of a dual-scale porous-medium by


a liquid [19]

The point-wise microscopic energy balance and species equations for resin

inside the gaps is proposed at first, and then the volume average of these equation

is taken. This way, they developed the macroscopic energy and species balance

equations.

The macroscopic energy balance equation in dual-scale porous medium is given

by

£
P.C
g^p.i *Tdtt(T*Y+ (v*>-v(r*Y = V K *- v ( r *>' + e.p.H*f<+ Q«»-Q-* (5 7)
-

where the pg and Cp are the resin density and specific heat respectively. Tg is

the temperature of resin in the gap region, e g is gap fraction, HR is the heat

reaction and fc is the reaction rate. The egpgHRfc term represents the heat source

term caused due to the exothermic curing reaction of a thermosetting resin. The

term Kth is the thermal conductivity tensor for dual-scale porous medium defined
119

as

K =k e 5+
* g s v In^dA-£sv&I^SdV (5 8)
-
A gl g

where k g , 8 and vg are the thermal conductivity of the resin, a unit tensor, and

the fluctuations in the gap velocity with respect to the gap averaged velocity,

respectively. The vector b relates temperature deviations in the gap region to the

gradient of gap-averaged temperature in a closure. Considering the temperature

closure formulation as fg = b.V < Tg >s , the local temperature deviation is related

to the gradient of the gap-averaged temperature through the vector b [19].

Qconvin the Eqn. (5.8) is the heat source term due to release of resin heat prior to

the absorption of surrounding tows given by

Qconv=PgCP,gSg[(Tg)8-(Tg)8t] (5.9)

where Sg , the sink term and areal average of temperature on the tow-gap

interface, <Tg>gt, are expressed as

(5.10)

and

W-7-JT.dA
g
' (5.11)

Qcond is the heat sink term caused by conductive heat loss to the tows and is given

by
120

Qoo„c=^jk g (-VT 8 ).n gt dA (5.12)

Using the analogy between heat and mass transfer to derive the gap-averaged cure

governing equation following the Tucker and Dessenberger [6] approach, one can

derive the following equation for the pore-averaged resin cure:

e
* f (c«}' +
( v
g )- V
( c
8 ) S
= V
- D
- V
( 0 E + e
s f c +Mco„v - M d l f f (5.13)

8
Here is the degree of cure in a resin for which the value of 0 and 1 correspond

to the uncured and fully cured resin respectively, D is the diffusivity tensor for the

gap flows and is given by

D = D ieg 5 + ^ | n g t b d A - ^ - J v g b d V (5.14)

where D,is the molecular diffusivity of resin. In Eqn. (5.13), Mconv the convective

source created due to the release of resin cure when absorbing into tows as a

results of the sink effect, is given by

Mconv=Sg[(cg)8-(cg)8t] (5.15)

where /c g ) 8 , the areal average of temperature on the tow-gap interface, is

expressed as

gt
c-r=-J-Jc,dA
*/ ~ T ~ J Vg "~ (5 16)
-

Mdiff, the cure sink term created as a result of the diffusion of cured resin into
121

tows, is given by

M diff =^|D 1 (-Vc g ).n gt dA (5.17)

It should be noted that the only way to compute theQconv,Qcond, Mconvand Mdiffis

by solving for flow and transport inside the tows.

5.3.2 Microscopic Level

Phelan et. al. [20] showed that the conventional volume averaging method can be

directly used to derive the transport equation for thermo-chemical phenomena

inside the tows for single-scale porous media. The final derivation for microscopic

energy equation is

[e.fccj, +(l-e t XpC P ) f ]-^- + (pCP)Ivt.VTt = V.KtM.VTt +etp,HRfc (5.18)

where the subscript t refer to tows. The microscopic species equation is given by

e t - ^ - + vt.Vct =V.e,Dt.Vc, +etfc (5.19)


dt

The complete set of microscopic and macroscopic energy and species equations as

well as the flow equation should be solved to model the unsaturated flow in a

dual-scale porous medium.


5.4 Dispersion Tensor

In some cases, a further complication arises in the thermal governing

equation due to the presence of thermal dispersion [21] which happens due to the

spreading of heat at the pore scale. Such spreading is mainly due to the molecular
122

diffusion of heat as well as due to the hydrodynamic mixing caused by the random

motion of fluid in a porous medium. Greenkorn [22] mentioned the following nine

mechanisms for most of the thermal spreading at pore scale:

1. Molecular diffusion: important in the case of sufficiently long time scales.

2. Mixing due to obstructions: The flow channels in porous medium are

tortuous means that fluid elements starting a given distance from each other

and proceeding at the same velocity will not remain the same distance apart

after a certain time, Fig. 5.2.

3. Existence of autocorrelation in flow paths: Knowing all pores in the porous

medium are not accessible to the fluid after it has entered a particular fluid

path.

4. Recirculation due to local regions of reduced pressure: The conversion of

pressure energy into kinetic energy gives a local region of low pressure.

5. Macroscopic or megascopic dispersion: Due to non-idealities which change

gross streamlines.

6. Hydrodynamic dispersion: Macroscopic dispersion is produced in capillary

even in the absence of molecular diffusion because of the velocity profile

produced by the adhering of the fluid wall.

7. Eddies: Turbulent flow in the individual flow channels cause the mixing as

a result of eddy migration.

8. Dead-end pores: Dean-end pore volumes cause mixing in unsteady flow.

The main reason is that as the solute rich front passes the pore, diffusion
123

into the pore occurs due to molecular diffusion. After the front passes, the

solute will diffuse back out and thus, disperse.

9. Adsorption: It is an unsteady-state phenomenon where a concentration front

will deposit or remove material and therefore tends to flatten concentration

profiles.

\ —
I
)
II

I
< 1

J
Figure 5.2 Mixing as a result of obstruction

Generally, the first seven mechanisms are included in the dispersion coefficient.

Rubin [23] generalized the thermal governing equation as

(pel ~ + (pc)f v.VT = V.(k,VT) + C (5.20)

where k m is a second-order tensor called the dispersion tensor.

Two dispersion phenomena that have been extensively studied in as part of the

transport phenomena in porous media, are the mass and thermal dispersions. The

former involves the mass of a solute transported in a porous medium, while the

latter involves the thermal energy transported in the porous medium. Due to the
124

similarity of mass and thermal dispersions, they can be described using common

dimensionless transport equation as

a
< ^ •/+„ \ a < " > _ 1 3 f D, 3(ft)
vW
(5.21)
30 " 3X,. Pe3X, ° », ,
where (£2) is either the volume averaged concentration for mass dispersion or the

volume averaged dimensionless temperature for thermal dispersion, 0 is

dimensionless time, (£/,•) is averaged velocity vector, Pe is Peclet number, Dy is

dispersion tensor of 2n order. It should be noted that Pe = — for mass dispersion

andPe-— for thermal dispersion where u and L are characteristic velocity and
a

length, respectively. D and a are molecular mass and thermal diffusivities,

respectively.

5.4.1 Dispersion in Isotropic Porous Media

Most studies on the dispersion tensor so far have been focused on the isotropic

porous media. Nikolaveskii [24] obtained the form of dispersion tensor for

isotropic porous media by analogy to the statistical theory of turbulence. Bear [25]

obtained a similar result for the form of the dispersion tensor on the basis of

geometrical arguments about the motion of marked particles through a porous

medium. Bear studied the relationship between the dispersive property of the

porous media as defined by a constant of dispersion, the displacement due to a

uniform field of flow, and the resulting distribution. He used a point injection
125

subjected to a sequence of movements. The volume averaged concentration of the

injected tracer, C0, around a point which is displaced a distance L = utin the

direction of the uniform, isotropic, two dimensional field of flow from its original

position is considered in his research.

C(x,y;x 0 ,y 0 ) = m n (5.22)
27i7tx oy | 2CTV 2a,

where L is the distance of mean displacement, u is the uniform velocity of flow, t

is the time of flow, ax and cy are standard deviations of the distribution in the x

and y directions, respectively and, finally m and n are the coordinates of the point

(x,y) in the coordinate system centered at (£,r|) given by m = x - ( x 0 + L ) and

n = y - y 0 , Fig. 5.3. This figure shows a point injection as a result of subsequence

movement where initially circle tracer gets an elliptic shape at L = ut.

m H
(><.y)
fi— y
1—i—Ts— * ' ' 7~ ', 7 " "/

Figure 5.3 Dispersion of a point injection displaced a distance L

Standard deviations are defined by ax =(2DIL)05 and a =(2DnL)05 where D, and


126

D„ are the longitudinal and transverse constants of dispersion in porous media,

respectively. One should note that the D! and D„ used in the Bear work depend

only upon properties of the porous medium such as porosity, grain size,

uniformity, and shape of grains1. From Eqn. (5.22), it follows that, after a uniform

flow period, lines of the similar concentration resulting from the circular point

injection of the tracer take the ellipse shape centered at the displaces mean point

and oriented with their major axes in the direction of the flow:

" 4 + ^ =1 (5.23)

Bear conjectured that the property which is defined by the constant of dispersion,

DiJkl, depends only upon the characteristics of porous medium and the geometry of

its pore-channel system. In a general case, this is a fourth rank tensor which

contains 81 components. These characteristics are expressed by the longitudinal

and lateral constants of dispersion of the porous media. Scheidegger [26] used the

dispersion tensor Dy in the following form

D^a^Yr (5-24)
M
where v is the average velocity vector, vk is the kth component of velocity vector,

aijhn is a fourth rank tensor called geometrical dispersivity tensor of the porous

medium. Bear demonstrated how the dispersion tensor relates to the two constants

He also considered a case when at the end of period t , , the velocity, remaining horizontal, is suddenly
changed from u , to u r
127

for an isotropic medium: «,, = longitudinal dispersion , and a±= transversal

dispersion2. Scheidegger [26] has shown that there are two symmetry properties

for dispersivity tensor

a a a n d a
Vbn= j*m ,ihn=av* (5'25)

Therefore, only 36 of 81 components of fourth rank tensor al]kl are independent.

For an isotropic porous medium, the dispersivity tensor must be isotropic. An

isotropic fourth rank tensor can be expressed as

««*» = *3,<L + P5Jim + Y5J3k (5.26)

where a, (3, y are constants and 51} is Kronecker symbol. Because of symmetry

properties expressed by Eqn.(5.23), we get

P= Y (5.27)

So the dispersivity tensor can be written as

at =aSS, +B(5,S +5 8A (5.28)


v f
ijkm tj km r* \ tk jm tm jk J

On substituting Eqn. (5.28) into Eqn. (5.24), we can obtain the dispersion tensor as

If we define a,_=alvl, all-a1=2yfflvl and n, =v,/lvl (n, is the mean flow direction), then

dispersion tensor Dl} can be written as

D
,j=aAj+{an-ai)n,nj ( 5 - 3 °)

2
The longitudinal direction is along the mean flow velocity in porous media, whereas the transverse
direction is perpendicular to the mean flow velocity.
128

From Eqn. (5.28), it is quite clear that the three principle directions of dispersion

tensor D are orthogonal to each other (due to the symmetry of Di}), and one

principle direction is along the mean flow direction (n) and the other two are

perpendicular to the mean flow direction. Therefore, for isotropic medium, the

dispersion tensor can be expressed by longitudinal and transverse dispersion

coefficients. If we consider the mean flow is along x-axis, Dtj can be written as

<h 0 0
D = 0 a± 0 (5.31)
0 0 a,

Therefore, transport equation can be written as

3(a) . ,3(n)_ I [ a2(Q) d2(a) 1 d2(a.)


flu — L y L + ax ——r - + ax (5.32)
de ^ xl ax, pe oXl oX2 oX 3 j

5.4.2 Dispersion in Anisotropic Porous Media

It has been shown that one of the principle axes of the dispersion tensor in

isotropic porous medium is along the mean flow direction. Unlike the isotropic

media, there are nine independent components in the dispersion tensor for the case

of anisotropic porous media. Bear [25] noted that the dispersion problem in a non-

isotropic material still remains unsolved. He suggested to distinguishing between

various kinds of anisotropics and doing statistical analysis with different frequency

functions for the spatial distribution of channels in each case. Unless some specific

types of porous media, such as the axisymmetric or transversely isotropic porous

media are considered, it is not possible to simplify the form of dispersion tensor.
129

In 1965, Poreh [27] used the theory of invariants to develop a dispersion tensor for

the axisymmetric porous media. The average properties of an axisymmetric porous

medium which affect the macroscopic dispersion pattern are invariants to rotation

about given line. He establish the general form of D^ with two arbitrary vectors R

and S as

DIJRlSJ = BlSIJRlSJ + B2vIvJRlSJ+B3AiAJRlSJ + B4vlRlAJSJ + B5ZlR,vJSJ (5.33)

where X. is the axis of symmetry, Bu B 2 , B3, B 4 , and B 5 are arbitrary functions of v2

and vklk. For arbitrary R and S and symmetric DtJ, one can have

DtJ = V „ + B2vlVj + B^k, + B4 (v,^ + Xy,) (5.34)

The dimensionless form of Eqn. (5.32) is obtained as

f?\ ' I A
7 = GA+G2 v,vy+G3^+G4 (VAJ+^VJ) (5.35)
D, \D0J \P*J

where G\, G2, G3 and G4 are dimensionless functions of (vl/D0) , (vl/v) . D0 is the

molecular diffusivity coefficient, / is the length characterizing the size of pores,

and v is the kinematic viscosity. The Final form of the dispersion tensor for

axisymmetric porous media can be given as

(*\ ^
Dn A+A2 D 2
,
*„+A V,Vj + 44+A ^ M,+4v,) (5.36)
v°oy V u
o J vA,v

where p, and p4 are dimensionless numbers, p 2 , P3 and P5 are even functions of

cos 0), and P6 is an odd function of cos GO

By assuming no motion within an axisymmetric medium, D^ is simplify to


130

^ =A3,+M^ (5-37)

Eqn.(5.37) indicating that one of the principal axes of D1} is, in this case, co-

directional with X. He followed these arguments for sufficiently large Reynolds

number, the dispersivity tensor for axisymmetric porous medium can be expressed

as

D e,v v e. (v X + v X)
-f- = ex8t) + ^ + eMj + ' J ' ' (5-38)
IV V V

where s,, E 2 , £ 3 , and £4 are parameters determined by the dimensionless

geometry of the medium and depends slightly on the Reynolds number and value

of cosft).On should note that the above derivation, primarily based on symmetry

considerations, can not reveal the scalar nature of general dispersivity tensor. Bear

[25] also noted that his analysis was based on an unproved assumption that D

may be expanded in a power series.

In 1967, Whitaker [28] applied pointwise volume-averaging method for transport

equation in anisotropic porous media and obtained the dispersion tensor DtJ as

D„ = D0 (Su + RB,J) + Clkjvk + Elkmivkvm (5.39)

where the second order tensor Bl} is a function of tortuosity vector

T = JQrijds (5.40)
s

The third-order tensor Clkj is given by


131

Cikj= VTTTVT (5-41)


5(v,)a
v ^ j

And the fourth-order tensor Eikmj is introduced as

a 3 (nv
v
;/
^ = ;T. , _ ^ (5-42)
a(a)
3<vt>a(v->a 3x
V , 7

where Q. is deviation of concentration or temperature from the average and v; is

the velocity deviation:

Q. = fi -(Q) f and v, = v,. - (v,.}f (5.43)

One should keep in mind that from Eqns. (5.41) and (5.42), we know that Cikj and

Ejhnj are completely symmetrical.

On comparing Eqn. (5.38) with Eqn. (5.24), one can note that there are both

third- and fourth-order symmetric tensors associated with velocity in the

Whitaker's derivation, while Nikolaveskii [24], Bear [25], and Scheidegger's [26]

derivations only contain fourth-order symmetric tensor. Patel and Greenkorn [29]

suggested that Whitaker's expression for dispersion tensor is the correct one for

anisotropic media. It is shown in section 5.4.1. that there are two distinct

components of dispersion tensor for isotropic medium while the Whitaker's

expression for dispersion tensor resulted in only one component for isotropic

media.

The diffusion term becomes less important at higher velocities and as a result,
132

Eqn. (5.39) reduces to

D ~ C i v, + E .v, v (5.44)
ij ^ikjk ikmj k m

For isotropic media, the tensors Clkj and Elkm] must be isotropic. Hence, Clk] =0, and

Elkmj is a linear combination of the Kronecker deltas as expressed in Eqn. (5.26).

Since Elkm] is completely symmetric, Eqn. (5.42), the tensor Elkm] can be shown to

be

E. =a(S.S +S S.+SS. ) (5.45)


ikmj \ ik mj im kj ij km }

Therefore Eqn. (5.39) reduces to

D y =a(2v,v,+3,|v| 2 ) (5.46)

Assuming 1-D flow in Cartesian coordinate frame where

v{ = u v2 = v3 = 0 (5.47)

then, eqn. (5.46) can be written as

"3 0 0"
D = au 0 1 0 (5.48)
0 0 1

Equation 5.48 shows that the longitudinal coefficient of dispersion tensor in this

case, is three times the transverse coefficient. Greenkorn [29] showed

experimentally that the ratio J\ I Dx varies approximately from lower value of 3 to

the higher value of 60. He showed that this ration is a function of the flow

velocity.

This ratio of longitudinal to transverse dispersion coefficients is shown to be a


133

function of the velocity of flow. Although the Whitaker's method is at variance

with Greenkorn experimental results, it still does give the correct lower limit result

for a homogeneous, uniform, isotropic medium.

5.5 A New Combined Experimental/Numerical Methodology to Assess


Dispersion Tensor

5.5.1 Theory

A new experimental-numerical method for predicting the dispersion tensor is

introduced and applied in this study. Using this method, one can estimate the

thermal dispersion tensor for different types of fiber mats characterized as single-

scale and dual-scale porous media. In this method, a numerical modeling comes

along with an experimental setup where the combination of both results leads to

prediction of the dispersion tensor. Since mass and thermal dispersions are very

similar in the transport mechanism in porous media, they can be described using

the dimensionless transport equations, Eqn. (5.21). Due to the similarity of the

transport equations for temperature and concentration, we are employing the

analogy between the concentration governing equation (in our experimental setup)

and the temperature governing equation (in our numerical modeling) to identifying

the dispersion tensor. In this prediction, the growth of the initially circular tracer

(dye) with time is studied. The results of the dye expansion at different time

frames are compared with results from our numerical model. Once the numerical

model's result is adjusted the same as the one from experiments at a specific time,
134

the dispersion tensor is obtained from the corresponding tensor components used

in the numerical simulation. Therefore, the dispersion tensor in the numerical

modeling is the filling parameter used in order to match the numerical results with

the results obtained from the experiments.

5.5.2 Experimental Measurements and Devices

The experimental section of this combined experimental/numerical method is done

using the 1-D flow experiment setup. In order to run the setup, a stack of fiber-

glass mat (called preform) is placed inside the mold cavity (Fig. 5.4 A). Two sides

of the preform in the horizontal plane are sealed using the removable glue in order

to avoid the leaking and race tracking along the sides during the experiment. The

mold is then closed and sealed tightly. The top layer of the mold is made out of a

Polycarbonate material which is clear and allows observation of the flow behavior

in the porous media. The test fluid is injected from one end of the mold and an

outlet is designed at another end of the mold. A pressure transducer is connected at

the top of the mold to record the inlet pressure history (see Fig. 5.4 B).
Figure 5.4 The 1-D flow experiment setup: A) open mold cavity and B) the assembled mold

A video camera is located on the top of the mold with the clear top mold to record

the flow process during the experiments. The camera equipped with a digital timer
136

and a measuring bar located along the mold length and facing to the camera are

employed to estimate the velocity of tracer liquid at different times. A syringe

filled with the dye (Methylene Blue) is placed at the flow inlet to inject the colored

dye into the steady flow of liquid through the fiber mats.

The test fluid injection system is designed to inject the liquid at a constant volume

flow-rate that can be monitored through a computer. The flow injection system

consists of a steel cylinder floating in a larger polycarbonate cylinder filled with

the test liquid. The floating cylinder is connected to a level gage for monitoring

the level of test liquid inside the large cylinder. A feedback system using the level

gage is employed on this flow-meter setup in order to keep the injection at a

steady level for a constant flow-rate experiment (see Fig. 5.5).

Figure 5 5 The flow injection andflow-metersetup

The test liquid is composed of 50% corn syrup and 50% water. The corn syrup is

white and hence, the blue dye injected at the inlet of the mold has a good color
137

contrast with the test liquid. The density of the corn syrup and water mixture is

measured by weighing a specific volume of liquid using a weighing balance, Fig.

5.6 A. The dynamic viscosity of the test liquid is obtained by using a Brookfield

viscometer, Fig. 5.6 B. The porosity of the porous material inside the mold cavity

is estimated by submerging the porous material inside a graduated cylinder filled

with water, Fig. 5.6 C (the increment of the water level caused by inserting of the

porous material into the water gives us the solid volume of the porous material. On

the other hand, by knowing the volume of the mold cavity, one should be able to

calculate the porosity of the porous medium inside the mold cavity using the

r , Mold volume-Solid volume . _ , . . . . ^ .


formula: = ). These physical measurements are conducted
Mold volume
several times and averaged in order to have a reasonable property value.

(A) (B) (C)


Figure 5.6 Measuring devices for A) Density, B) Viscosity, and C) Porosity.
Once the whole 1-D flow experiment setup is ready, it can be run at different flow
138

rates with different porosities as well as different porous media micro-structures

corresponding to the single and dual-scale porous media. The blue dye is injected

when the test liquid is passing through the porous material under steady-state

condition. The dispersion of injected dye is visible due to the contrast of its color

with the test liquid's color. According to the literature, the initial circular spot of

injected dye expands to an elliptical shape with time for the homogeneous single-

scale porous medium. Fig. 5.7 shows the ellipse generated as a result of dispersion

in a single-scale porous medium (this picture shows details of the dye injection

into a fluid moving through the random fiber-glass mats).

0JOHNSOIVJ LEVEL f i T O O L . Ts.


I «• C-* *.*. r * ? H l ta ft to RT

Figure 5.7 Effect of dispersion in the form of an elliptical dye expansion in a single-scale porous medium.

In order to complete this combined experimental and numerical method, the

numerical results should accompany the experimental data so that the dispersion

terms can be estimated.


139

5.5.3 Numerical Modeling

Our numerical results are needed to match with the experimental data in

order to estimate the dispersion tensor associated with the flow through the fibrous

porous media. The transient energy equation is solved by means of the finite

element method using COMSOL multiphysics software. Fluid properties as well

as the porous-material porosity are used as parameters to solve the transient energy

equation. Boundary conditions and initial conditions are assumed for the geometry

created (see Fig. 5.8).

The initial temperature (or initial concentration) distribution in the form of a

circle is assumed inside the geometry, close to the inlet. The test fluid (corn syrup-

water mixture) is injected from left inlet boundary which flows out from the right

outlet boundary.

Corrective heat flux


Convective heat flux
\imei; (Outlet)

k Initial temperature (initial


concentration)

Figure 5.8 Schematic view of the geometry, the boundary and initial conditions.

5.5.4 Combined Experimental/Numerical Methodology

The dye can be injected either under the steady-state condition when the
140

porous medium is fully wet and the test fluid is passing at a constant flow-rate or

under transient conditions when the porous medium is partially wet and the

injected dye is carried forward by the liquid front. The injected dye initially in the

form of a circle is converted downstream along with the injected liquid.

Depending on the micro-structure of the fiber mat, the flow rate, the porosity, etc.,

the expansion pattern of the initial circle of dye varies widely from one type of

fiber mat to another. For a single-scale porous medium (random fiber mat), the

initially circle of dye will change to an elliptical shape when the dye moves along

the mold direction. As long as the major axis of the elliptical tracer is along the

flow direction and the minor axis of the ellipse is perpendicular to it, one should

be able to model this dispersion as a tensor.

A finite element method (FEM) is used to model the problem. A

rectangular geometry (Fig. 5.8) is taken with one end as the inlet and the other end

as the outlet. The initial circular tracer at a location close to inlet is defined in the

FEM by defining some elements initially with a known temperature (or

concentration). The temperature governing equation is used in the FEM and the

thermal conductivity value, K , in this governing equation is the dispersion tensor.

One can simply change the values for the tensor K in the form, of 2x2 matrix in

order to fit the experimental data by comparing the dye expansion with time in

both the experimental and numerical results.


141

5.6 Results and Discussions

Study of Thermal Dispersion in Single-Scale Porous Media

The dispersion in porous media occurs due to the combined effects of

molecular diffusion and convection of flow in pore spaces. In order to have a

better understanding of the effect of convection and molecular diffusion, effect of

each term is studied individually. The Fig. 5.9 shows the effect of molecular

diffusion on thermal dispersion in porous media where there is no convection

during heat transfer. As it is expected, the initial temperature (or concentration)

disperses in its initial location as there is no convection. A 2x2 unit matrix is

assumed as the dispersion coefficient for this case which leads to a full circular

expansion due to similar diffusion in both horizontal and vertical directions.

o © o Q

t=0s t=6s t=13 s t=40s


Figure 5.9 Effect of pure molecular diffusion of thermal dispersion by time

To gain physical insights of the aforementioned method, the dispersion results of

the experimental modeling and finite element method are compared. Constant

flow-rate of Q=2 ml/s through the random fiber mat, a single-scale porous medium

(Fig. 5.10) with porosity of s=0.8 at three different time frames are considered

here.
=6 "*•
5e

Figure 5 10 Random fiber-glass mat as a representative of single-scale porous media

In order to estimate the dispersion tensor in such a single-scale porous medium,

002 0 0 02 0 04 0O6 0 08 01 012 014 016 019 02

- ^ f

0 02 0 002 0 0 1 0 0 6 0.08 01 012 014 0 16 0 18 02


143

E
002 0 002 004 006 0 08 01 012

F
01-1 016 018 02 I
Figure 5.11 Matching the experimental and numerical results for a single-scale porous medium (random
fiber mat) at Q=2ml/s at three different time frames of A) 7 s, B) 33 s and C) 51 s.

the experimental observation has to be matched with the numerical simulation

results. Three different time frames are assumed for this analysis.

The mentioned combined experimental/numerical method tries to estimate

the dispersion tensor by matching the experimental data for longitudinal and

transverse terms of dispersion by those in the numerical simulation. Results of the

proposed experimental/numerical approach give the dispersion tensor of D=

for the random porous media as a representative of a


(J o.looc — oJ

single-scale porous medium, Fig. 5.11. For this single-scale porous medium, the

flow direction along the mold horizontal direction and the corresponding vertical

direction are the longitudinal and transverse directions, respectively for the

dispersion tensor. Our results for three different flow-rates (Q=lml/s, Q=2 ml/s

and Q=3 ml/s) show that the dispersion tensor remains the same for this type of

porous media.
144

For the sake of validation, a comparison has been done between our results and

the results from an available model [32]as

u A t 2
n = Lim
D„ j - '

2x
*— (5.49)

Eqn. (5.49) models the longitudinal dispersion coefficient of the injected tracer

particles at the inlet of the porous medium sample. In this equation, U is the mean

tracer velocity which covers distance x in At seconds. In order to validate our

results with this model, two different time frames were selected and the velocity,

At, and the distanced traveled are substituted in Eqn. (5.49).

The porous medium is assumed to be Isotropic for this section. The tracer is

moving with a speed u=0.0002m/s at the initial location of x=0.070 m at t= 36 s to

the next location of x=0.090 at t=40 s. Following Eqn. (5.49), the longitudinal

dispersion is D = 3.20 xlO "7 while the longitudinal dispersion predicted by our

combined experimental/numerical method is D=3.19 xlO"7. One can see that there

is a good match between our results and those obtained from Eqn. (5.49) for the

longitudinal dispersion in the random fiber mat.

A study on the changing dimensions of the expanding tracer-generated ellipse in

the considered single-scale porous medium is presented in Fig. 5.12. This picture

shows the increase in the longitudinal and transverse dimensions of the initial

circular dye-batch with time. One can notice a rapid expansion along the

longitudinal direction compared to the transverse direction due to the present

convection effects along the flow direction.


145

_ — — -s
0 Lon
' Longitudinal Dimension (L)
— B - Trar
Transverse Dimension (D)

20 40 60 80 100

Time (sec)

Figure 5.12 The longitudinal and transverse increases of the initial circular dye-patch with time.

Study of Thermal Dispersion in Dual-Scale Porous Media

Similar study was carried out on a dual-scale porous medium for the purpose of

estimation of the dispersion tensor.

Figure 5.13 Stitched fiber-glass mat as a representative of dual-scale porous media.

In this study, a procedure similar to the one applied for the single-scale case, is

applied to a cavity filled with dual-scale porous media, Fig. 5.13. Experimental
146

observations along with the numerical simulation results are matched in order to

estimate the dispersion tensor.

i^:
1
<Hm 6 » bCt ti UM W»
U
B

i
I
E F
Figure 5.14 Comparison of the experimental and numerical modeling for an anisotropic but periodic
medium at Q = 2 ml/s : A) experimental results at t = 0 s, B) numerical results at t = 0 s, C) expenmental
results at t = 35 s, D) numerical results at t = 35 s, E) experimental results at t = 71 s and F) numerical
results at t = 71 s.
147

An important characteristic of the stitched fiber-glass (as a type of dual-scale

porous media used in the experiments) is its symmetry along the tows and the

corresponding vertical direction. This anisotropic porous medium is assumed as a

periodic porous medium in the two directions. Analysis on this type of porous

media through the experiments and computer simulations are presented in Fig.

5.14.

Based on the symmetry present in this anisotropic periodic porous medium,

the dispersion of heat (or concentration) has to be symmetric with respect to the

direction of the flow. The prediction for this case gives

D _ r3.186£ - 7 0 1
\- 0 1.062E-8J'

Another anisotropic porous medium studied in this chapter is the unidirectional

glass-fiber mat shown in Fig. 5.15. The dual-porosity effect in this unidirectional

porous medium causes an unsaturated flow-front to appear during the mold-filling

process. The micro-structure of the unidirectional mat is such that the flow

initially moves between tow bundles, Fig. 5.16 A. Once the channel between the

tow bundles is filled with the test fluid, it begins to wet the tow bundles from sides

which slows the liquid front. However, one tows are saturated, the injected liquid

flows in the gap region only (Fig. 5.16 B).


148

Figure 5.15 Unidirectional fiber-glass as a representative of dual-scale porous media.

Based on aforementioned reasons, the actual flow-rate in channels under steady-

state conditions is much more than the Darcy velocity.

channel
-Q/ > > U
^ A ~~ Darcv
cs, channel cs,cavity (5.50)
•> • >

I
• >

•*

• »
Wet *• •*
\ [ '

X •A:"•A.A
Wet

Dry

A B
Figure 5.16 Unidirectional fiber-glass in A) initial unsaturated flow and B) saturated flow.

Results obtained from experiments confirm this theoretical prediction. The

velocity of the dye inside the channels between the tow bundle (ucAa„„ep0.068 m/s)

is an order of magnitude larger than theoretical Darcy velocity (uDarcy=0.0050


149

m/s). Fig. 5.17 depicts the tracer movements through the unidirectional mat

channels with a much higher velocity compared to the Darcy velocity.

t=0 s t=2 s t=6 s


Figure 5.17 Experimental observation of flow of tracer through a dual-scale porous medium
made of the unidirectional fiber mats

According to above-mentioned reasons, the dispersion tensor for this type of

fiber glass is not estimated here. The Darcy's law is not obeyed in the entire

domain. This type of problem should be handled separately by assuming two

flows (i.e., gap and tow region flows) with different velocities.

5.7 Summary and Conclusion

In this chapter, the non-isothermal flow through porous media is studied

experimentally and numerically. An effort was made to quantify the thermal

dispersion effect which enhances heat transfer in porous-media flows. A combined

experimental/numerical approach to estimate the thermal dispersion during the

non-isothermal flow in porous media is described. Both single-scale and dual-

scale porous media were considered. Studies were done over three different mass

flow-rates. The estimated longitudinal thermal dispersion value in single-scale

porous media is compared with available results in the literature. The validation
150

study showed a very good match, and hence, motivated us to explore this method

for the dual-scale porous media. Preferential flow through channels frustrated our

efforts to apply the proposed experimental/numerical approach to model thermal

dispersion in the unidirectional stitched mat.


151

References

1. Rudd CD, Long AC, Kendall KN, Mangin CGE. Liquid molding

technologies. Woodhead Publishing Ltd; 1997.

2. Chan, A.W. and S.-T. Hwang, Modeling Nonisothermal Impregnation of

Fibrous Media with Reactive Polymer Resin, Polymer Engineering &

Science, 1992. 32(5):p. 310-318.

3. Chiu, H.-T., B. Yu, S.C. Chen, et al., Heat Transfer During Flow and Resin

Reaction through Fiber Reinforcement. Chemical Engineering Science,

2000. 55(17): p. 3365-3376.

4. Lee, L.J., W.B. Young, and R.J. Lin, Mold Filling and Cure Modeling of

RTM and Srim Processes. Composite Structures, 1994. 27(1-2): p. 109-120.

5. Lin, R.J., L.J. Lee, and M.J. Liou, Mold Filling and Curing Analysis Liquid

Composite Molding. Polymer Composites, 1993. 14(1): p. 71-81.

6. Tucker, C.L. and R.B. Dessenberger, Governing Equations for Flow

through Stationary Fiber Beds, in Flow and Rheology in Polymer

Composites Manufacturing, S.G. Advani, Editor. 1994, Elsevier.

7. Lam, Y.C., S.C. Joshi, and X.L. Liu, Numerical Simulation of the Mould-

Filling Process in Resin-Transfer Moulding. Composites Science and

Technology, 2000. 60(6): p. 845-855.


152

8. Wu, C.H., H.-T. Chiu, L.J. Lee, et al., Simulation of Reactive Liquid

Composite Molding Using an Eulerian-Lagrangian Approach. International

Polymer Processing, 1998(4): p. 398-397.

9. Bruschke, M.V. and S.G. Advani, Numerical Approach to Model Non-

Isothermal Viscous Flow through Fibrous Media with Free Surfaces.

International Journal for Numerical Methods in Fluids, 1994. 19(7): p. 575-

603.

10. Dessenberger, R.B. and C.L. Tucker, Thermal Dispersion in Resin Transfer

Molding. Polymer Composites, 1995. 16(6): p. 495-506.

ll.Kang, M.K., W. Lee, II, J.Y. Yoo, et al, Simulation of Mold Filling

Process During Resin Transfer Molding. Journal of Materials Processing

and Manufacturing Science, 1995. 3(3): p. 297-313.

12.Liu, B. and S.G. Advani, Operator Splitting Scheme for 3-D Temperature

Solution Based on 2-D Flow Approximation. Computational Mechanics,

1995. 16(2): p. 74-82.

13. Mai, O., A. Couniot, and F. Dupret, Non-Isothermal Simulation of the

Resin Transfer Moulding Process. Composites - Part A: Applied Science

and Manufacturing, 1998. 29(1-2): p. 189-198.

14.Ngo, N.D. and K.K. Tamma, Non-Isothermal '2-D Flow/3-D Thermal'

Developments Encompassing Process Modeling of Composites:

Flow/Thermal/Cure Formulations and Validations. American Society of


153

Mechanical Engineers, Applied Mechanics Division, AMD, 1999. 233: p.

83-102

15. Shojaei, A., S.R. Ghaffarian, and S.M.H. Karimian, Simulation of the

Three- Dimensional Non-Isothermal Mold Filling Process in Resin Transfer

Molding. Composites Science and Technology, 2003. 63(13): p. 1931-

1948.

16. Shojaei, A., S.R. Ghaffarian, and S.M.H. Karimian, Three-Dimensional

Process Cycle Simulation of Composite Parts Manufactured by Resin

Transfer Molding. Composite Structures, 2004. 65(3-4): p. 381-390.

17. Young, W.-B., Three-Dimensional Nonisothermal Mold Filling

Simulations in Resin Transfer Molding. Polymer Composites, 1994. 15(2):

p. 118-127.

18. Young, W.-B., Thermal Behaviors of the Resin and Mold in the Process of

Resin Transfer Molding. Journal of Reinforced Plastics and Composites,

1995. 14(4): p. 310.

19.Pillai, K.M. and M.S. Munagavalasa, Governing Equations for Unsaturated

Flow through Woven Fiber Mats. Part 2. Non-Isothermal Reactive Flows.

Composites Part A: Applied Science and Manufacturing, 2004. 35(4): p.

403-415.

20.Phelan, F.R., Jr. Modeling of Microscale Flow in Fibrous Porous Media.

1991. Detroit, MI, USA: Publ by Springer-Verlag New York Inc., New

York, NY, USA.


154

21. Donald A. Nield and Adrinan Bejan, Convection in Porous Media,

Springer, 3rd edition.

22. Robert A. Greenkorn, Flow phenomena in porous media, Marcel Dekker,

INC., New York and Besel, 1983.

23. Rubin, H., Heat dispersion effect on thermal convection in a porous

medium layer. J. Hydrol. 21, 1074,P: 173-184.

24. Nikolaevskii, V.N., Convective diffusion in porous media, Journal of

applied mathematics and mechanics, 23, 6, 1492-1503, 1959.

25. Bear, J., On the Tensor Form of Dispersion in Porous Media. Journal of

Geophysical Research, 1961. 66: p. 1185-1197.

26. Scheidegger, A.E., General Theory of Dispersion in Porous Media. Journal

of Geophysical Research, 1961. 66(10): p. 3273-3278.

27.Poreh, M., The Dispersivity Tensor in Isotropic and Axisymmetric

Mediums. Journal of Geophysical Research, 1965. 70: p. 3909-3913.

28.Whitaker, S., Diffusion and Dispersion in Porous Media. AIChE Journal,

1967. 13(3): p. 420-427.

29.Patel, R.D. and R.A. Greenkorn, On Dispersion in Laminar Flow through

Porous Media. AIChE Journal, 1970. 16(2): p. 332-334.

30. Beavers, G.S. and Joseph D.D., Boundary condition at a naturally

permeable wall, J. Fluid Mech,. Vol. 30, pp. 197-207, 1967.


155

31.Sahraoui, M. and Kaviany, M., Slip and no-slip boundary condition at

interface of porous plain media, Int. J. Heat Mass Transfer, Vol. 35, pp.

927-943, 1992.

32.Bouchaus J.P., GorgeJ.P. Bouchaud, A. Georges, Anomalous diffusion in

disordered media: statistical mechanisms, models and physical applications,

Phys. Rep. 195 (1990) 127-293.


156

Chapter 6

SUMMARY, CONTRIBUTION, AND FUTURE

WORK

6.1 Summary

In this Ph.D. thesis, transport and flow phenomena in porous media were

studied from several different points of view. We have first studied the effect of

preform aspect ratio on the permeability measurement in both single- and dual-

scale porous media under steady and transient conditions. Both 1-D flow and

radial flow experiments were conducted. Four different aspect ratios were

considered for both the experimental study as well as the numerical modeling. It is

concluded the longer aspect ratio leads to more accurate results which agrees with

the previous studies. Two different criteria were introduced in this work to capture

error in permeability measurements and two models for these errors were obtained

by curve fitting linear and quadratic equations to experimental data. Transient

mold-filling model showed huge differences in the pressure profile and flow-front

shape between the single- and dual-scale porous media. The mold filling

simulation with PORE-FLOW® and experimental data emphasized the fact that

flow in dual-scale porous media can not be predicted by the simulation based on
157

the single-scale flow physics.

The interface boundary condition problem between the clear-flow channel

and porous wall studied in a separate chapter. The main focus of this study is the

boundary condition estimation at the interface for both periodic and non-periodic

porous media. The results have been carried out with different REVs to ensure the

accuracy in the results. Our finite element method's results are then compared

with available results in the literature.

One of the chapters of this proposal is dedicated to a new method proposed

in this work on predicting different flow variables in porous media independent of

the REV size. The ensemble averaging method successfully proposed and applied

to some problems in porous-media flow. Using this method, one can solve

problems of flow in porous media without any concern about the size of the REV

especially with complex geometry or in a problem with sharp gradients in some

parameters. Two different practical applications of such method are presented and

solved. The comparison results showed a reasonable agreement between the

ensemble averaging method and the other available results.

Heat transfer aspect of a flow through porous media studied with particular

stress in thermal dispersion term. The governing equations are derived and

presented for both micro and macro scales. Most of the work done in the

dispersion so far is dedicated to the single-scale porous media. In this work, a


158

combined experimental/numerical approach is proposed to predict the dispersion

tensor in both single-scale and dual-scale porous media. Using this method, one

can predict the dispersion tensor correctly in different types of porous media and

use this term in non-isothermal flow through porous media such as the thermal

modeling of different composite plastics processing.

6.2 Contributions

The selected scientific contributions resulting from this Ph.D. thesis, which

have either already been published in conference proceeding or to be submitted for

publication in technical journal papers, are as follows:

• Study of the aspect ratio variations on the accuracy of permeability

estimation under transient and steady-state conditions of flow through

porous media.

• Proposing two new models based on our experimental data for the

estimation of error in permeability estimation through 1-D flow

experiments.

• Emphasizing the fundamental difference between the flow modeling of

single- and dual-scale porous media through the experimental and

numerical comparisons of the flow front evaluation.

• Evaluating the accuracy of various boundary conditions at the clear

channel-porous medium interface.


159

• First numerical study on the efficacy of boundary conditions at the interface

of a non-periodic porous medium and clear channel by employing a weakly

disordered porous medium created by random distribution of cylinders.

• A new method based on ensemble averaging for flow variable estimation in

porous media flows with sharp gradients.

• Successfully applying the proposed ensemble averaging method in two

industrial applications for validation purposes.

• A new combined experimental/numerical approach for estimating the

thermal dispersion term during non-isothermal flows in porous media.

• Successfully employing the new proposed technique for the estimation of

the thermal dispersion tensor for flows in dual-scale porous media.

6.3 Future Work

Isothermal and non-isothermal problem of flow through porous media

have many industrial and medical applications, each requiring a special

treatments. In this Ph.D. thesis, appropriates theories were developed for the

study of flow through porous media in some of these applications. The

following are the suggestions for improvement in the current work for future

studies:

• One can develop a better numerical modeling of the role of aspect ratio

on the measurement of permeability of porous media in order to have a


160

better comparison under both the steady-state and transient flow

conditions.

• Experimental modeling of the clear-channel porous-media interface

problem through several velocimetry techniques in order to have an

exact experimental results for the velocity profile at the interface (One

may use a simple color- tracer technique for the observation of flow of a

liquid at the interface of the clear channel and porous medium).

• Applying ensemble averaging method to more applicable, large-flow-

gradient problems (such as study of natural convection in porous media

adjacent to a hot flat plate, porous media flows with temperature

dependent viscosity, forced convective heat transfer during the flow of

resin through fiber-glass mats in resin transfer molding process, etc.)


161

Appendix

The irregular porous medium is made out of fixed number of cylinders is

treated as the solid phases of porous medium in the lower channel. The cylinders

are located randomly. A program written in FORTRAN randomly generates the

horizontal and vertical locations of the cylinder centers for each unit cell inside the

porous region. The simple case is that all cylinders located at the centers of the

unit cells, such that a periodic porous media is generated (see Fig.Al).

Figure Al Schematic view of locating solid cylinders at the center of the unit cells in order to generate a
periodic porous media.

On the other hand, in order to locate the cylinders randomly, Ax and Ay

defined as the horizontal and vertical distances of a cylinder's center from the

unit-cell center, respectively, are generated randomly (see Fig. A2).

An important constraint in this process is that no overlapping of cylinders is

allowed. In this case, each cylinder is allowed to go beyond the unit cell at

maximum half of the cylinder ( Axmax = Cylinder radius or Aymax=Cylinder


radius).

Figure A2 Schematic view of location of the solid cylinder at horizontal and vertical distances from the
center of unit cell in order to generate the non-periodic porous media.

Further, if location of the randomly generated cylinders overlapped, another

random location for the overlapped cylinder is generated so that the "no

overlapping" rule is implemented (see Fig. A3).

Figure A3 No overlapping rule applied to two cylinders generated in wo adjoining unit cells which
generating the irregular porous medium.
163

EHSAN MOHSENILANGURI
Address: 4042 North Wilson Drive, Phone: (414)364 5561
Unit 4 Email: Ehsan@uwm.edu,
Milwaukee, WI 53211 mohseni esn@vahoo.com

Education

2007- Ph.D. University of Wisconsin-Milwaukee (UWM), USA, Mechanical


2011 Engineering

Dissertation Title: Isothermal and Non-Isothermal Rows in
Porous Media
• Advisor: Associate Professor Krishna M. Pillai
2005- M.Sc. University of Mazandaran (UMI), Iran, Mechanical Engineering
2007 • Thesis Title: Study of solar airs collector in Building Energy
Analysis
• Advisor: Associate Professor Hessam Taherian
2001- B.Sc. University of Mazandaran (UMI), Iran, Department of Mechanical
2005 Engineering
• Senior Project Title: Experimental Study on Electro-Magnetic
Row-meter
• Advisor: Associate Professor Hessam Taherian

Publications
• Mohseni Languri E., Moore D. R., Masoodi, R., Pillai, K., Sabo, R., An Approach to
Model Resin Flow Through Swelling Porous Medium using Natural Fibers, 10th
International Conference on Flow Processing in Composite Materials (FPCM 10), July
2010, Ascona, Switzerland.
• Mohseni Languri E., Bennett III G. L., Masoodi R., Pillai K. M., A Reference Porous
Medium Made by Rapid Prototyping as a Calibration Tool for Permeability, 10th
International Conference on Flow Processing in Composite Materials (FPCM 10), July
2010, Ascona, Switzerland.
• Mohseni Languri E., Tan H. and Pillai K. M., Effect of Preform Aspect Ratio and
Anisotropy on the Transient 1-D Mold Filling in LCM, 9th International Conference
on Flow Processes in Composite Materials (FPCM 9), 2008, Montreal, Canada.
• Tan H., Mohseni Languri E. and Pillai K. M., Permeability prediction of dual-scale
fibrous structure using the unit cell, 9th International Conference on Row Processes in
Composite Materials (FPCM 9), 2008, Montreal, Canada.
• Mohseni Languri E., Vechart A., Tan H. and Pillai K. M., "Effect of Preform Aspect
Ratio on Permeability measured through ID Flow Experiments, 9th International
164

Conference on Flow Processes in Composite Materials (FPCM 9), 2008, Montreal,


Canada.
• Mohseni Languri E. and Pillai K.M., Estimating the Average Flow Variables in
Porous Media using Ensemble Averaging Method, Journal of Porous Media, 2010.
• Mohseni Languri E. and Pillai M. K., Numerical Study of Stress-Jump and Stress-
Continuity Conditions at Random Porous-Medium-Clear Fluid Interface Problems,
Physics of Fluid, 2010.
• Mohseni Languri E. and Pillai M. K., Study of Thermal Dispersion Term in Dual-
Scale Porous Media, International Journal of Heat and Mass Transfer, 2010.
• Ganji D.D., Jannatabadi M., Mohseni Languri E., Application of He's Variational
Iteration Method to Nonlinear Jaulent-Miodek Equations and Comparing it with
ADM, Journal of Computational and Applied Mathematics, Vol. 207 , Issue 1,
Pages 35-45, 2007.
• Ganji D.D., Nourollahi M., Mohseni Languri E., Application of He's Methods to
Nonlinear Chemistry Problems, Computers & Mathematics with applications,
Vol. 54, Issues 7-8, Pages 1122-1132, 2007.
• Ganji D.D., Jannatabadi M., Mohseni Languri E., The Comparison of Homotopy
Perturbation Method with Adomian's Decomposition Method in Coupled Jaulent-
Miodek Nonlinear Equations, Far East Journal of Applied Mathematics, Vol. 29,
Issue 1, Page 19-30, 2007.

Teaching Experience

2010 Fall Introduction to Fluid Mechanics (Adjunct Instructor) UWM


2009 Summer Introduction to Fluid Mechanics (Adjunct Instructor) UWM
Spring Engineering Fundamentals II (Teaching Assistant) UWM
2008 Fall Fundamental of Physics (Teaching Assistant) UWM
Fall Engineering Fundamentals I (Teaching Assistant) UWM
2007 Fall Introduction to Fluid Mechanics (Grader) UWM
Spring Thermodynamic Laboratory (Lab Assistant) UMI
2006 Fall Fluid Mechanics Laboratory (Lab Assistant) UMI
Spring Fluid Mechanics Laboratory (Lab Assistant) UMI
2005 Fall Fluid Mechanics Laboratory (Lab Assistant) UMI
Fall Thermodynamic Laboratory (Lab Assistant) UMI

Major Professor Date


W/AOI

You might also like