You are on page 1of 250

IMPROVING PERFORMANCE AND DRAINAGE OF COALESCING FILTERS

A Dissertation

Presented to

The Graduate Faculty of The University of Akron

In Partial Fulfillment

of the Requirements for the Degree

Doctor of Philosophy

Shagufta Usman Patel

August, 2010
IMPROVING PERFORMANCE AND DRAINAGE OF COALESCING FILTERS

Shagufta Usman Patel

Dissertation

Approved: Accepted:

_________________________ _________________________
Advisor Department Chair
Dr. George G. Chase Dr. Lu-Kwang Ju

_________________________ _________________________
Committee Member Dean of the College
Dr. Lingyun Liu Dr. George K. Haritos

_________________________ _________________________
Committee Member Dean of the Graduate School
Dr. H. Michael Cheung Dr. George R. Newkome

_________________________ _________________________
Committee Member Date
Dr. Kevin L. Kreider

_________________________
Committee Member
Dr. Subramaniya I. Hariharan 

ii
ABSTRACT

Pure air or gas is very critical to many industrial applications. Gas streams contain

impurities in the form of solid and liquid aerosols of micron and submicron sizes. It is

very important to remove these aerosols for improving the economy and reliability of

industrial processes and equipment as well as protecting our health and environment.

Among different filters being used, coalescing filters are effectively used to remove

liquid aerosols from gas streams and hence has numerous industrial applications. The

performance of the filter affects the economy of the process.

A coalescing filter captures the oil droplets and the captured liquid typically

drains from the filter by action of gravity. The saturation or hold-up of liquid in the filter

constricts the gas flow, increases pressure drop, and increases the operating costs of the

filter. The filter loaded with liquid droplet indicates limited filter life and needs to be

replaced which increases the cost of process. Filter performance and filter life can be

improved if the liquid saturation is reduced without reduction in capture efficiency. In

this research work filter media are modified with drainage structures to reduce saturation

and to reduce drag resistance. The media are tested in horizontal and vertical orientations

to determine whether their orientation with gravity influences the performance. The

experimental results show that with no drainage channels the media oriented with flow

vertically downward operates the best whereas with drainage channels the horizontally

iii
oriented media had the best performance. The results also show filter geometries

developed by embedding low surface energy woven drainage channels at 45 degree

downward angles have the overall best performance. Adding nanofibers to improve the

performance of the filter media is an effective way of improving filter media’s capture

efficiency with moderate increase in pressure drop. Nanofiber augmented filter media

indicated higher pressure drop but when the drainage channels are incorporated in the

filter, the increase in the pressure drop is significantly low. The pressure drop increase

will be significant for a nanofiber augmented filter media without the drainage channels.

Filter geometries developed by incorporating drainage channels at downward angles

indicate significantly low pressure drop, the drainage channels create a path of lower

resistance and hence maximum flow goes through the drainage channel improving the

filter performance and drainage.

This research work allows to develop cost effective filter geometries which will

significantly increase filter performance as compared to the glass fiber filter media which

are commonly used in the industry.

iii
DEDICATION

I would like to dedicate this work to my parents Mr. Usman C. Patel and Mrs. Badrunisa

U. Patel.

iv
ACKNOWLEDGEMENTS

I would like to sincerely acknowledge and thank the following people who helped

me achieve my goal. I would like to express my sincere gratitude towards Professor

George Chase for extending support and encouragement throughout the course of this

degree. He has been a great mentor and guide throughout the course of my research. I

would like to thank my committed members; Dr. Lingyun Liu, Dr. H. Michael Cheung,

Dr. Kevin L. Kreider and Dr. Subramaniya I. Hariharan for their valuable advice,

assistance, and encouragement. I will like to thank Mr. Frank Pelc for helping me in

building my experimental assembly and helping me cutting my filter samples. Mr. Pelc’s

help and suggestions in cutting the filter media are greatly appreciated. I would like to

thank all the multiphase group members for their help. I would like to acknowledge

Coalescence Filtration Nanomaterials Consortium for their financial support and

encouragement.

I would like to thank my parents and my brother Mr. Sarfaraz U. Patel for their

constant support, unwavering belief in my abilities, and encouragement. Special thanks to

my friend Mr. Prashant S. Kulkarni who was with me since the beginning of my journey

and encouraged me at every stage. He was always there when I needed him.

v
TABLE OF CONTENTS

Page

LIST OF TABLES ........................................................................................................... xiii

LIST OF FIGURES ......................................................................................................... xiv

CHAPTER

I. INTRODUCTION ...................................................................................................1

1.1. Theory of filtration .................................................................................1

1.2. Types of filtration ...................................................................................2

1.3. Coalescing filters ....................................................................................2

1.4. Hypothesis ...............................................................................................3

1.4.1. Flow orientation .........................................................................3

1.4.2. Drainage channel incorporated filter media ...............................5

1.4.3. Composite filter media ...............................................................7

1.5. Research Objectives ...............................................................................8

1.6. Benefits of current work .........................................................................9

1.7. Dissertation outline ...............................................................................10

II. LITERATURE REVIEW ON COALESCENCE FILTRATION .........................12

vi
2.1. Aerosol filtration ..................................................................................12

2.2. Sources of liquid aerosol ......................................................................13

2.1.1. Mechanical atomization ...........................................................13

2.1.2. Evaporation–condensation .......................................................13

2.1.3. Entrainment by gas flow in liquid–gas contactors ...................14

2.1.4. Crankcase ventilation (CCV) ...................................................14

2.3. Coalescing filters ...................................................................................15

2.4. Capture mechanisms..............................................................................17

2.5. Coalescence filtration and drainage channel design..............................20

2.6. Fiber orientation and filter performance ...............................................22

2.7. Saturation of filter media .......................................................................23

2.8. Nanofibers in coalesce filtration............................................................25

2.9. Modeling of fibrous filters ....................................................................26

III. DEVELOPING AND CHARACTERIZING FILTER MEDIA ............................27

3.1. Introduction ..........................................................................................27

3.2. Steps involved in filter media preparation ...........................................27

3.2.1. Glass fiber media ....................................................................27

3.2.2. Nanofiber augmented micro glass fiber filter media ..............31

3.2.3. Composite filter media ............................................................32

3.3. Vacuum molding set-up .......................................................................33

vii
3.4. Filter media characterization technique ...............................................35

3.4.1. Porosity measurement .............................................................36

3.4.2. Permeability measurement ......................................................38

3.4.3. Hardness or strength measurement .........................................41

IV. DEVELOPING FILTER GEOMETRIES .............................................................43

4.1. Introduction ..........................................................................................43

4.2. Filter media cutter saw .........................................................................43

4.3. Drainage channels ................................................................................46

4.4. Filter geometries ....................................................................................48

4.4.1. Horizontal orientation .............................................................48

4.4.2. Vertical orientation .................................................................50

4.5. Composite filter media ..........................................................................51

V. COALESCENCE FILTRATION SETUP .............................................................52

5.1. Experimental set-up and procedure for coalescence tests ....................52

5.1.1. Prefiltration ..............................................................................52

5.1.2. Aerosol generation ..................................................................55

5.1.3. Filter holder .............................................................................56

5.1.3.1. Horizontal flow orientation filter holder ...............56

5.1.3.2. Vertical flow orientation filter media ...................57

5.1.4. Measurement equipments ........................................................59

viii
VI. EXPERIMENTAL RESULTS OF GLASS FIBER FILTER GEOMETRIES......62

6.1. Polypropylene woven drainage channels .............................................62

6.1.1. Capture efficiency ...................................................................64

6.1.2. Pressure drop ...........................................................................66

6.1.3. Quality factor ..........................................................................68

6.1.4. Saturation ................................................................................70

6.1.4.1. Saturation profile of no-drainage filter media .......74

6.1.5. Performance comparison of filter geometries developed


with woven drainage channel ................................................77

6.1.6. Factors affecting coalescence filtration ...................................79

6.1.7. Effect of number of drainage channels in inclined angle filter


media and upward inclination .................................................81

6.1.8. Effect of pore size and filament thickness of drainage

channel ....................................................................................87

6.1.9. Filter bed length ......................................................................90

6.2. Nonwoven drainage channels ...............................................................93

6.2.1. Capture efficiency ...................................................................94

6.2.2. Pressure drop ...........................................................................95

6.2.3. Quality factor ..........................................................................97

6.2.4. Saturation ................................................................................99

6.2.5. Effect of basis weight of drainage channel ...........................102

ix
6.2.6. Performance comparison of filter geometries developed
with woven and nonwoven drainage channels ......................106

6.3. Effect of surface wettabilty of drainage channel ................................108

6.3.1. Capture efficiency ................................................................111

6.3.2. Pressure drop .........................................................................114

6.3.3. Quality factor ........................................................................115

6.3.4. Saturation ..............................................................................117

6.3.5. Performance comparison of filter geometries ........................118

6.4. Composite filter media .......................................................................121

VII. NANOFIBER AUGMENTED FILTER GEOMETRIES ...................................125

7.1. Introduction .......................................................................................125

7.2. Electrospinning ...................................................................................126

7.2.1. Electrospinning setup .............................................................127

7.2.2. NanospiderTM .........................................................................129

7.2.2.1. Principle of NanospiderTM technology.................129

7.3. Production of nylon nanofibers ..........................................................132

7.4. Experimental results of nanofiber augmented filter media modified


with woven drainage channels of varying surface wettability ...........134

7.4.1. Capture efficiency .................................................................134

7.4.2. Pressure drop .........................................................................137

7.4.3. Quality factor ........................................................................139

x
7.4.4. Saturation ..............................................................................141

7.4.5. Performance comparison of nanofiber augmented filter


geometries .............................................................................143

VIII. MODELLING OF DRAINAGE CHANNEL INCORPORATED FIBROUS

FILTER MEDIA ..................................................................................................145

8.1. Multiphase theory in porous media .....................................................146

8.2. Volume average theory .....................................................................146

8.3. Conservation equations ......................................................................148

8.3.1 Mass and momentum balaces ...............................................148

8.3.2 Mass and momentum jump balances .....................................150

8.4 Assumptions .......................................................................................151

8.5 Permeability of media and Darcy’s law .............................................153

8.6 Simplified conservation equations .....................................................158

8.6.1 Mass and momentum balances ..............................................158

8.6.2 Mass and momentum jump balances .....................................159

8.7 Boundary conditions............................................................................160

8.8 Grid generation ...................................................................................160

8.8.1 Goals of grid generation.........................................................161

8.8.2 Grid mapping .........................................................................162

8.9 Algorithm ............................................................................................166

xi
8.10 Results .................................................................................................168

8.10.1 Optimizing the grid size .........................................................168

8.10.2 Velocity profile of isotropic filter media without drainage

channel ...................................................................................170

8.10.3 Velocity profile of anisotropic drainage channel


incorporated filter media .......................................................172

8.10.4 Parametric study.....................................................................174

8.10.4.1 Varying the angle of drainage channel ................174

8.10.4.2 Changing the thickness of drainage channel ........175

8.10.4.3 Exchanging porosity and permeability of


media and drainage channel ................................176

IX. CONCLUSION ....................................................................................................181

X. FUTURE WORK ................................................................................................188

BIBILOGRAPHY ............................................................................................................190

APPENDICES

APPENDIX A. EXPERIMENTAL RESULTS FOR DRAINAGE


CHANNEL INCORPORATED FILTER MEDIA MADE BY USING CARBOSET 560
BINDER ..........................................................................................................................201
 

APPENDIX B. FORTRAN CODE FOR DRAINAGE CHANNEL


INCORPORATED FILTER MEDIA .............................................................................203
 

APPENDIX C. NOMENCLATURE ...............................................................................225 

xii
LIST OF TABLES

Table Page

3.1 Filter media performance and amount of binder and starch .....................................29

3.2 Composite filter media design ..................................................................................32

6.1 Factors affecting coalescence filtration ....................................................................80

6.2 Pore sizes of polypropylene spunbond fabric .........................................................103

7.1 Electrospinning process parameters for producing nylon-6 nanofibers ................... 132

8.1 Volumetric flow rates of filter geometry with drainage channels at different
angles ................................................................................................................................ 175

8.2 Volumetric flow rates of filter geometry with drainage channels of varying
thickness ........................................................................................................................... 176

A.1 No-drainage channel media in horizontal orientation .............................................201

A.2 Inlet-drainage channel media in horizontal orientation ..........................................201

A.3 Outlet-channel drainage media in horizontal orientation........................................202

A.4 No-drainage channel media in vertical orientation ..................................................... 202

xiii
LIST OF FIGURES

Figure Page

1.1. Coalescing filter ......................................................................................................2

1.2. Horizontal flow orientation .....................................................................................4

1.3. Vertical flow orientation ......................................................................................... 4

1.4. Drainage channel incorporated fibrous filter media in horizontal flow

orientation ...............................................................................................................5

1.5. Drainage channel incorporated fibrous filter media in vertical flow orientation ....7

2.1 Single fiber capture mechanisms (a) Direct interception, (b) Inertial impaction,
(c) Brownian diffusion (d) Gravitational capture .................................................18

3.1 Mixing tank assembly to make slurry of fibers ....................................................31

3.2 Schematic of vacuum molding setup .................................................................... 33

3.3 Lab scale setup of vacuum molding process .........................................................35

3.4 Lab scale pycnometer assembly............................................................................ 37

3.5 Porosities of glass fiber filter media ..................................................................... 37

3.6 Frazier® Differential Pressure Air Permeability Measuring Instrument ............... 40

xiv
3.7 Components of Frazier® Differential Pressure Air Permeability Measuring
Instrument (a) Sample holder (b) Manometer to read pressure drop (c) Orifice (d)
Pump and pump control (e) Different orifice opening...........................................40
 

3.8 Air permeability of glass fiber filter media........................................................... 41

3.9 Lab scale Durometer ............................................................................................. 42

4.1 Filter media band saw and cutter assembly (a) band saw (b) filter holder ........... 44

4.2 Filter holder to cut filter sample by using the cutter saw...................................... 45

4.3 Filter media sample (a) Thick filter media (b) Filter media cut by using filter

media cutter saw ...................................................................................................45

4.4 Woven drainage channels (a) Polypropylene circular shaped woven drainage
channel (b) Microscopic image of woven polypropylene drainage channel ........ 46
 

4.5 Nonwoven drainage channels (a) Polypropylene circular shaped nonwoven


drainage channel (b) Microscopic image of nonwoven polypropylene drainage
channel .................................................................................................................. 46
 

4.6 (a) Nylon nanofibers electrospun on Teflon® woven mesh of 500 µm pore
opening (b) Microscopic image of Nylon nanofibers on Teflon® woven mesh ...48
 

4.7 Filter geometries in horizontal orientation (a) No-drainage (b) Inlet-drainage (c)
Middle-drainage (d) Outlet-drainage (e) Both-End-Drainage (f) Inclined-angle-
drainage .................................................................................................................49
 

4.8 Filter geometries in vertical orientation (a) No-drainage (b) Inlet-drainage (c)
Middle-drainage (d) Outlet-drainage (e) Both-End-Drainage (f) Inclined-angle-
drainage .................................................................................................................50
 

4.9 Composite filter media .......................................................................................... 51

xv
5.1 Schematic of the coalescing filter testing set up ................................................... 53

5.2 Lab scale coalescence filtration setup ................................................................... 54

5.3 (a) Surge tank (b) Prefilters................................................................................... 54

5.4 (a) Laskin nozzle (b) Mixing chamber.................................................................. 55

5.5 Heater .................................................................................................................... 56

5.6 Horizontal flow orientation filter holder (a) Filter holder parts (b) Filter holder
assembly ................................................................................................................ 57
 

5.7 Filter holder assembly in vertical orientation (a) Filter holder assembly (b) Filter
holder fixed to the filter holder assembly (c) Filter holder and the wire mesh .....58
 

5.8 (a) Pressure gauge (b) Photometer ........................................................................ 60

5.9 (a) Rotameter (b) Downstream HEPA filter ......................................................... 60

5.10 SMPS and CPC ..................................................................................................... 61

6.1 Upstream and downstream droplet size distribution of no-drainage channel filter
media ..................................................................................................................... 63

6.2 Pressure and downstream concentration profile of no-drainage channel filter


media ..................................................................................................................... 64

6.3 Capture efficiency of filter geometries with polypropylene woven drainage


channels................................................................................................................. 65

6.4 Pressure drop of filter geometries with polypropylene woven drainage


channels................................................................................................................. 67

xvi
6.5 Quality factor of filter geometries with polypropylene woven drainage
channels................................................................................................................. 69

6.6 Saturation of filter geometries with polypropylene woven drainage channels ..... 71

6.7 Tested filter media (a) no-drainage channel filter media in horizontal orientation
(b) Both-end-drainage filter media in horizontal orientation. The green colored
oil is collected near the bottom edge of the filter medium ......................................  72

6.8 Quality factor versus average saturation of all filter geometries with
polypropylene woven drainage channels .............................................................. 74

6.9 Measurement of saturation profile of no-drainage filter media


in horizontal orientation ........................................................................................ 75

6.10 Saturation profile of no-drainage channel filter media in horizontal orientation . 76

6.11 Relative quality factor of no-drainage filter media in horizontal and vertical
orientation ............................................................................................................. 77

6.12 Relative Quality Factor of filter geometry with polypropylene woven drainage
channels in horizontal and vertical orientation ..................................................... 78

6.13 Effect of number of channels on capture efficiency of 450 downward inclination


filter geometry in horizontal orientation ............................................................... 81

6.14 Effect of number of Channels on pressure drop of 450 downward inclination filter
geometry in horizontal orientation ........................................................................ 82

6.15 Effect of number of channels on quality factor of 450 downward inclination filter
geometry in horizontal orientation ........................................................................ 82

xvii
6.16 Effect of number of channels on saturation of 450 downward inclination filter
geometry in horizontal orientation ........................................................................ 83

6.17 Filter geometry developed with three equally spaced polypropylene woven
drainage channel at 450 upward inclination .......................................................... 84

6.18 Capture efficiency of three drainage channels at 45 degree upward inclination and
experimental control ............................................................................................. 85

6.19 Pressure drop of three drainage channels at 45 degree upward inclination and
experimental control ............................................................................................. 85

6.20 Quality factor of three drainage channels at 45 degree upward inclination and
experimental control ............................................................................................. 86

6.21 Saturation of three drainage channels at 45 degree upward inclination and


experimental control ............................................................................................. 86

6.22 Capture efficiency of filter media equipped with woven drainage channel of
different pore openings ......................................................................................... 88

6.23 Pressure drop of filter media equipped with woven drainage channel of different
pore openings ........................................................................................................ 88

6.24 Quality factor of filter media equipped with woven drainage channel of different
pore openings ........................................................................................................ 89

6.25 Saturation of filter media equipped with woven drainage channel of different pore
openings ................................................................................................................ 89

6.26 Flow pattern in the filter geometry with three equally spaced drainage channels at
450 downward inclination ..................................................................................... 90

xviii
6.27 Capture efficiency of filter geometries with reduced filter bed length ................. 91

6.28 Pressure drop of filter geometries with reduced filter bed length ......................... 91

6.29 Quality factor of filter geometries with reduced filter bed length ........................ 92

6.30 Saturation of filter geometries with reduced filter bed length .............................. 92

6.31 Capture efficiency of filter geometries with polypropylene nonwoven drainage


channels ..........................................................................................................................  94

6.32 Pressure drop of filter geometries with polypropylene nonwoven drainage


channels................................................................................................................. 96

6.33 Quality factor of filter geometries with polypropylene nonwoven drainage


channels................................................................................................................. 98

6.34 Saturation of filter geometries with polypropylene nonwoven drainage


channels................................................................................................................. 99

6.35 Quality factor versus average saturation of all filter geometries with
polypropylene nonwoven drainage channels ...................................................... 102

6.36 Capture efficiency of filter geometries for varying basis weight of polypropylene
spunbond fabric ................................................................................................... 103

6.37 Pressure drop of filter geometries for varying basis weight of polypropylene
spunbond fabric ................................................................................................... 104

6.38 Quality factor of filter geometries for varying basis weight of polypropylene
spunbond fabric ................................................................................................... 104

xix
6.39 Saturation of filter geometries for varying basis weight of polypropylene
spunbond fabric ................................................................................................. 105

6.40 Relative Quality Factor of filter geometry with polypropylene nonwoven drainage
channels in both horizontal and vertical orientation ........................................... 107

6.41 RQF of woven and nonwoven drainage channel incorporated filter geometries in
horizontal and vertical orientation ...................................................................... 108

6.42 Sullube® 32 oil contact angle on plane surfaces (a) Nylon, (b) Polypropylene and
(c) Teflon®  .................................................................................................................... 110

6.43 Sullube 32® oil contact angle on woven surfaces (a) Polypropylene and
(b) Teflon®  .................................................................................................................... 111

6.44 Capture efficiency of filter geometries developed with varying surface wettability
of drainage channels ........................................................................................... 113

6.45 Pressure drop of filter geometries developed with varying surface wettability of
drainage channels ................................................................................................ 114

6.46 Quality factor of filter geometries developed with varying surface wettability of
drainage channels ................................................................................................ 116

6.47 Saturation of filter geometries developed with varying surface wettability of


drainage channels ................................................................................................ 117

6.48 RQF of filter geometries developed with woven and nonwoven drainage channels
of varying surface wettability ............................................................................. 119

6.49 Height between drainage channels for different angles of the inclined angle

xx
experiments ..................................................................................................................  120

6.50 Capture efficiency of composite filter media...................................................... 121

6.51 Pressure drop of composite filter media ............................................................. 122

6.52 Quality factor of composite filter media ............................................................. 122

6.53 Saturation of composite filter media ................................................................... 123

7.1. Schematic of single jet electrospinning .............................................................. 128

7.2. Lab scale set up of single jet electrospinning ..................................................... 128

7.3. Lab scale NanospiderTM module ......................................................................... 131

7.4. Wire electrode and solution bath ........................................................................ 131

7.5. SEM images of nylon nanofibers (a) 6 wt% (b) 8 wt% (c) 12 wt% (d) 20 wt% 133

7.6. SEM image of nanofiber augmented glass fiber filter media ............................. 134

7.7. Capture efficiency of nanofiber augmented glass fiber filter media incorporated
with nylon and polypropylene woven drainage channels ................................... 135

7.8. Pressure drop of nanofiber augmented glass fiber filter media incorporated with

nylon and polypropylene woven drainage channels ........................................... 137

7.9. Quality factor of nanofiber augmented glass fiber filter media incorporated with
nylon and polypropylene woven drainage channels ........................................... 140

7.10. Saturation of nanofiber augmented glass fiber filter media incorporated with
nylon and polypropylene woven drainage channels ........................................... 141

7.11. Quality factor versus average saturation of all nanofiber augmented filter
geometries with nylon and polypropylene woven drainage channels ............... 143

xxi
7.12. Relative Quality Factor of nanofiber augmented filter geometry with nylon and
polypropylene woven drainage channels ............................................................ 144

8.1 Interface of media and drainage channel ............................................................ 151

8.2 Anisotropic permeability .................................................................................... 154

8.3 Principle and system axes for drainage channel incorporated filter media ........ 156

8.4 Physical and logical space .................................................................................. 161

8.5 Physical and logical space conversion ................................................................ 163

8.6 Nine point-solver ................................................................................................ 164

8.7 Physical and logical space of drainage channel incorporated filter media and
equations used to calculate the pressure profile in the logical space .................. 166

8.8 Algorithm ............................................................................................................ 167

8.9 Error analysis at different grid sizes ................................................................... 169

8.10 (a) Velocity magnitude of isotroipc media without drainage channel


(b) velocity direction of isotroipc media without drainage channel .................. 171

8.11 (a) Magnitude and (b) direction of velocity profile in drainage channel
incorporated anisotropic filter media with an angle of 45 deg. The drainage
channels are marked inside of the rectangular boxes. ...........................................  173

8.12 Close-up view of Figure 8.10. Velocity magnitude of drainage channel near the
drainage channel and filter media interface ........................................................ 174

8.13 Velocity magnitude near the drainage channels when the drainage channel when
the drainage channel porosity and permeability is less than the filter media ..... 177

xxii
8.14 Velocity direction near the drainage channels when the drainage channel porosity
and permeability is less than the filter media...................................................... 177

8.15 Velocity (a) magnitude and (b) directions in drainage channel and media when
porosity and permeability of filter media is higher than drainage channels ....... 178

8.16 Drop motion through drainage channel and media ............................................. 179

xxiii
CHAPTER I

INTRODUCTION

1.1 Theory of filtration

Filtration is a mechanical process in which a dispersed phase is separated from a

continuous phase by allowing it to flow through a porous material. The dispersed phase

can be solid or liquid and continuous phase can either be a liquid or a gas. The filtration

process can be used to remove unwanted impurities as well as to recover expensive

dispersed phase from the fluid. When the continuous phase flows through the filter

medium, the dispersed phase fluid or solid is captured either on the surface when the

dispersed phase particle size is larger than the pores or on the pores inside the filter

medium when the particle size is smaller than the pores. The filter medium is the core

part of any filtration process. It is defined as a permeable medium which allows the

continuous phase to pass through and captures the dispersed phase i.e. the liquid or solid

impurities. The flow through the filter medium is due to the pressure difference between

the inlet and outlet surfaces. The efficiency of the filtration process depends mainly on

the effectiveness of the filter medium in removing the impurities. The selection of filter

medium is an important aspect of filtration.


 
1.2 Types of filtration

Filtration can be classified as surface and depth filtration. When the dispersed

phase particle size is bigger than the pore size of the filter medium, most of the filtration

occurs at the inlet surface of the medium. No particle can pass through the medium. In

depth filtration, the dispersed phase particle size is often less than the pore size of the

filter media [1, 2]. The particles penetrate into the medium and get captured due to the

mechanisms like interception, impaction, diffusion, gravitational deposition etc.

Depending on the application, filtration is divided into solid-gas, liquid-gas, solid-liquid,

and liquid-liquid filtration.

1.3 Coalescing filters

Gas flow in
Gas flow out

Gravity

Drainage

Figure 1.1 Coalescing filter.


 
Coalescence is the agglomeration and growth of small liquid droplets to form

bigger drops. The coalescing filtration is a depth filtration process in which liquid is

removed from a gas stream or another liquid stream. When a liquid is removed from a

gas stream, it is called air or gas filtration. When a liquid is removed from another liquid

stream, it is called liquid-liquid filtration. In coalescing filtration, the dispersed liquid

droplets enter with the continuous phase, strike the fibers due to various capture

mechanisms, coalesce with other drops, grow big and eventually drain out of the filter

medium as shown in Figure 1.1. The dominating mechanisms of this filtration process are

direct interception, inertial impaction, diffusion deposition, gravity settling and

electrostatic capture.

1.4 Hypothesis

1.4.1 Flow orientation

In the coalescing filter, when the droplets are captured on the fibers, coalesce and

move from fiber to fiber, they experience gravitational force. Hence the droplets moving

from fiber to fiber in horizontal flow orientation do not follow the straight line path

(Figure 1.2). When the flow is in direction of gravity, i.e. in the vertical flow orientation,

the droplets follow straight line path (Figure 1.3). Available experimental or theoretical

information is unable to tell which of the flow orientation will perform better. When the

flow is in the direction of gravity, the gravitational force can help to drain the big

coalesced drops faster and can enhance the performance of the filter in terms of quality

factor and saturation.


 
Gas flow in
Gas flow out

Gravity

Drainage

Figure 1.2 Horizontal flow orientation.

Gas flow

Gas flow out


Gravity Drainage

Figure 1.3 Vertical flow orientation


 
1.4.2 Drainage channel incorporated filter media

As the liquid droplets are captured in the void space of the filter, the filter media

loads up with the oil. Hence the saturation of the filter increases and the filter porosity

decreases which constricts the air flow resulting in pressure drop increase. Increase in

pressure drop results in decrease in the quality factor. When the filter loads with the oil it

often needs to be replaced which makes the filtration process expensive.

Gas flow in
Gas flow out

Drainage channel

Gravity

Drainage

Figure 1.4 Drainage channel incorporated fibrous filter media in horizontal flow

orientation.


 
Drainage channels can be incorporated in the filter media to enhance drainage and

reduce saturation of the filter media. The drainage channels are open and porous

structures of varying surface energy. The drainage channel provides a path of low

resistance hence the oil droplets can drain very easily from the filter media. Hence the

filter media will have higher drainage and lower saturation. This filter media can remain

porous for longer time and can capture more droplets. Hence the filter can remain

functional for longer time and filter life will be enhanced. The schematic of drainage

channel incorporated filter media is horizontal flow orientation is shown in Figure 1.4.

The drainage channel incorporated filter media in vertical orientation shows

improved performance because of the available gravitational force will help in draining

the bigger coalesced drops faster from the filter media and usually has a reduced

saturation. Hence the filter will remain porous for a longer time and will have a higher

quality factor and longer filter life.


 
Gas flow in

Drainage channel

Gas flow out

Gravity

Drainage

Figure 1.5 Drainage channel incorporated fibrous filter media in vertical flow orientation.

1.4.3 Composite filter media

The composite filter media is a filter media design which has bigger pores along

the filter bed length. As the droplets move along the filter bed length they coalesce to

form bigger drops. If a filter medium does not have big pores then these big droplets can

split due to the air drag and secondary aerosol generation can take place. A filter media

can be developed by using fiber diameters from smaller to bigger fiber sizes along the

filter bed length. This filter media design will be capable of capturing a variety of droplet

sizes. Hence it is expected to have a lower pressure and higher drainage rate and it can

help in reducing the secondary aerosol generation.


 
1.5 Research objectives

Following are the primary objectives of the research work.

a. Study the effect of gravitational force to improve the quality factor and drainage

of coalescing filters.

b. Develop different filter geometries by incorporating woven and nonwoven

drainage channels in the coalescing filters for air liquid coalescence process. The

filter media used to develop drainage channel incorporated filter geometries will

have similar properties.

c. Evaluate the performance of the drainage channel incorporated filter media

experimentally.

d. Study the parameters affecting on the performance of the drainage channel

incorporated filter media and improve the filter media design.

e. Develop nanofiber augmented filter media and study the performance of

nanofiber augmented filter geometries. The nanofiber augmented filter media will

have similar properties.

f. Develop a composite filter media design of varying fiber diameters and evaluate

its performance experimentally.

g. Develop a mathematical model of the drainage channel incorporated filter

geometry to predict pressure and velocity profile with no oil drops present at

steady state.


 
1.6 Benefits of current work

Coalescing filters are widely used in industry to remove liquid droplets from air

stream. The coalescing filters are fibrous and hence they have internal pores of different

sizes. When theses filters capture the liquid droplets these pores are filled with the liquid.

Hence the filter media saturates with the liquid droplets and possess higher pressure drop.

The filter having higher pressure drop requires more energy for the air to flow through

the media making the process energy expensive. Once the filter loads up with the liquid

droplet it has to be replaced. The time period for which the filter is functional is known as

filter life. Hence the filter life is dependent upon pressure drop and saturation.

Current research work utilizes the available gravitational force to improve the

performance and drainage of filter media. Effect of gravitational force on the

performance of filter media is studied by changing the air flow direction. The

gravitational force will help removing the bigger coalesced drop faster from the filter

media. Hence the filter media will have low saturation and low pressure drop. Filter can

remain porous for longer time and hence can have higher capture efficiency as well as

longer filter life.

Current research work is also associated with developing a filter media design

which will effectively drain the saturated liquid from the filter media. These filter designs

are developed by incorporating woven and nonwoven drainage channels. These drainage

channels are porous open structures and possess lower surface energy as compared to the

glass fiber filter media. Hence these drainage channels will improve the liquid drainage

from the filter media and the filter media will have low saturation and hence low pressure


 
drop with a high capture efficiency. This will help in keeping the filtration process energy

efficient. The filter geometry having lower saturation and lower pressure drop will

remain porous for longer time. Hence the filter can capture more droplets and remain

functional for longer time and will have longer filter life.

1.7 Dissertation outline

a. Chapter I gives an introduction to coalescence filtration theory, hypothesis and

objectives of this work, and its benefits.

b. Chapter II reviews the background information on aerosol filtration, prior

literature on coalescing filter media with drainage channels and industrial

applications of coalescence filtration.

c. Chapter III explains fabrication and characterization of fibrous filter media.

d. Chapter IV explains the woven and nonwoven drainage channels and procedure

of developing different filter geometries.

e. Chapter V describes the experimental setup of coalescence filtration and

procedure to test the filter geometries.

f. Chapter VI presents the experimental results and performance comparison of

woven and nonwoven drainage channel incorporated filter geometries. It also

explains the effect of different parameters on the performance of the filter

geometries.

g. Chapter VII describes the process of making nanofibers, experimental results

and performance comparison of nanofiber augmented woven and nonwoven

drainage channel incorporated filter geometries.

10 
 
h. Chapter VIII explains the modeling approach and modeling results of drainage

channel incorporated filter media with no oil drops at steady state.

i. Chapter IX presents the conclusion of this dissertation work based on the

experimental and modeling results.

j. Chapter X presents the future direction of experimental and modeling work of

filter geometries.

11 
 
CHAPTER II

LITERATURE REVIEW ON COALESCENCE FILTRATION

This chapter reviews the aerosol filtration, sources of liquid aerosol generation,

prior work on coalescing filters and drainage channel incorporated coalescing filter, and

the industrial applications of these types of filter media.

2.1 Aerosol filtration

An aerosol is a suspension of solid particles or liquid droplets in air. Air is the

bulk transportation medium for transmission of particulate contaminants. The control

over airborne solid and liquid contaminants, hazardous biological agents, allergens and

pollutants is a key issue in many industries like the metal-cutting, automobile,

semiconductor, food, pharmaceuticals and biotechnology processes as these businesses

require centralized air conditioning in production environment, clean gases and

effluent/waste treatment. The particle size of particulate matter is determined by the

process that generates the particles. The processes that generate liquid aerosols are

described below.

12 
 
2.2. Sources of liquid aerosol

Liquid aerosols mainly originate from following different processes.

2.2.1 Mechanical atomization

Liquids in contact with high rotation rate equipment acquire mechanic energy

high enough to be sheared into small droplets [1] and are differentiated from droplets

generated by impaction and centrifugal forces [2]. Metal working fluids are used to

lubricate the components of metal cutting processes [3, 4]. Droplets from these lubricants

disperse into the air from the metal cutting operations. The oil droplets are formed during

shearing action between moving surfaces. The average droplet diameter of metal working

fluid is about 5-8 μm [5, 6].

Oil is also used to lubricate and remove excess heat generated in compressed air

systems. During the compression process, tiny droplets of oil are introduced in the air

stream. Even with no oil lubrication, oil droplets are introduced in the air stream as the

atmospheric air contains 20-30 ppm of hydrocarbon aerosols [7]. Apart from these

applications, removing liquid aerosols from a gas stream is important in other industries

like, petroleum refineries, natural gas industries, and the nuclear power industries.

2.2.2 Evaporation–condensation

At sufficiently high temperature conditions, a liquid can evaporate and

spontaneously re-condense around liquid nuclei in locations of slightly lower temperature

[8]. The heat is generated during high temperature metal working processes which

evaporates the oil. The oil vapors can condense on dust particles to form small oil

droplets [9]. For instance, combustion particles are usually in the 10-50 nm size range,

but when they combine with other particles like the oil, water or solvent mist they

13 
 
agglomerate to form larger particulates. The agglomerate particles may be broken down

into smaller particles and released into air. It is difficult to break down such particles

smaller than 0.5 µm [10].

2.2.3 Entrainment by gas flow in liquid–gas contactors

Flowing gases in spray towers, cooling towers, and plate columns, can entrain

droplets from the liquid phase [6]. Compressed air is important in many industrial

applications. Some of the main applications include pneumatic conveying, spray paint

equipment, breathing air, laboratory air use, gas separator systems, aeration in

pharmaceutical and chemical processes, air bearings for mechanical power transmission

etc. The typical water drops are 0.05-10 µm and oil droplets are 0.1-10 µm in diameter

[11]. The liquid droplets are carried with the air in the above mentioned applications and

are harmful for the workers. The impurities in the air affect the downstream processes

and process equipments as well. Filtration is required to improve the overall quality of

the processes and the reliability of the process equipments.

2.2.4 Crankcase ventilation (CCV)

Liquid aerosols are also generated in internal combustion engines and are released

into the environment through crank case ventilation. The crankcase encloses many of the

moving parts of the engine that are lubricated by oil circulated from the oil pan. The

impurities are introduced in the gas in two ways. First, oil droplets are formed due to the

shearing action of moving parts in the crankcase. Second, the gases inside the combustion

chamber leak out of the piston and enter crankcase. The droplet size distribution is 0.1-10

µm and soot particles are 0.3-0.5 µm [12]. Regulations require removal of these

impurities to achieve overall emission standards for the engines. The other applications of

14 
 
filtration in automotive industry are tank venting, engine intake air, engine exhaust gas,

cabin air, coolant system, gearbox oil and engine oil [13].

The typical aerosol droplets of interest have sizes ranging from about 0.01 µm to

50 µm. Aerosols generated during the above mentioned processes represent a significant

waste stream as well as a health hazard to humans. The metal working fluids affect the

health of the people when they are exposed for long duration of time. Medical evidence

has linked worker exposure to cutting fluid mist with respiratory ailments and several

types of cancers [4]. Respiratory illnesses associated with inhalation of above mentioned

aerosols include respiratory irritation, bronchitis, occupational asthma, and loss of lung

function. Several epidemiological studies have also shown statistically significant

increases in cancer of the esophagus, stomach, pancreas, larynx, colon, and rectum due to

prolonged exposure to cutting fluid mists [14]. To reduce worker exposure common mist

control strategies include enclosing the machine tool, using air filters and mist collectors,

and adding antimisting agents to the cutting fluid.

2.3 Coalescing filters

Coalescing filters are used to remove small liquid droplets from air streams. They

have numerous industrial applications e.g. dehumidification, cabin air filtration, gas

compressors, removal of liquid aerosols from metal cutting, CCV and agricultural

processes [4, 10, 11, 14]. Coalescence filtration is a process in which droplets carried by

a flowing gas are captured by the fibers of a filter medium. Ensuing droplets carried by

the gas collide and coalesce with the prior drops. Coalescence occurs when two or more

liquid droplets come into contact with each other with sufficient energy to overcome

surface tension for the drops to merge [15]. The coalesced drops grow in size in the filter

15 
 
medium until drag of the moving gas or gravity force are strong enough to cause the

enlarged drops to drain out of the filter [16]. The critical size of the drop after which it

detaches from the fiber depends on the local velocity, fiber size, interfacial tension and

contact angle. It has also been observed that if the contact angle of the drop on the fiber is

small the drop does not break away from the fiber but is conducted along the fiber

direction of fluid stream [17-21].

Increasing the gas-liquid separation efficiency of a filter is usually accompanied

by an increase in pressure drop across the filter [22]. The increase in the pressure drop

can be divided into three stages. When a critical oil mass is trapped in the filter, the

pressure drop rises quickly and then stabilizes and a steady state characterized by an

equilibrium between the drainage and liquid flow entering in the filter is established.

Finally, the pressure drop becomes constant and a steady state is reached. From this

moment on, the oil mass stored in the filter is constant and the rate of drainage exiting the

filter balances the liquid droplet capture rate [23]. The separation efficiency depends

upon the drop size, liquid viscosity, gas velocity, pressure, temperature of the gas,

structure of the filter medium (fiber diameter, fiber orientations and packing density),

surface properties of the fibers, binder content, and filter thickness. The surface tension of

the liquid has a crucial influence on the formation of the droplets, although it is difficult

to separate the influences of the surface tension and the viscosity of the liquid because

they vary proportionately with temperature [24]. Purchas [25] provides a comprehensive

list of filtration-specific properties, machine-oriented properties, and application-oriented

properties that are important in the selection of filter media.

16 
 
The filter performance is characterized by the combined performance of pressure

drop and separation efficiency. The performance measure has several equivalent

definitions such as filtration index [26], figure of merit [27, 28], or quality factor [29, 30].

Liquid gas coalescers are very effective in removing liquid droplets down to 0.1

μm with higher gas velocity than other methods and much lower pressure drop with

effective drainage mechanism. The liquid droplets of an aerosol are caught on the fibers,

coalesce with each other and drain out of the filter media. A variety of fibers are used for

various applications. High efficiency particulate air, or HEPA, filter is a type of high-

efficiency air filter made up of randomly arranged glass fibers. According to the standard

adopted by most industries, it removes at least 99.97% of airborne particles of 0.3 µm in

diameter. The fibrous filters capture particles by following capture mechanisms as

described below.

2.4 Capture mechanisms

As the aerosol passes through the filter media the particles are captured in the

depth of the fiber media. There are several mechanisms that contribute to the capture of

particles [29]. Submicron size droplets and particles are captured by four dominant

capture mechanisms as shown in Fig. 2.1. These mechanisms are called the single fiber

capture mechanisms. The capture on a single fiber is considered to be dependent upon the

local flow conditions around the fiber and in when the filter properties are uniform the

capture on a single fiber is a representation of the entire filter medium. The main

assumption of the single fiber concept is that all fibers are identical [24]. The single fiber

mechanism has been widely accepted by the filtration industry as it leads to comparable

results with experiments. All of these mechanisms occur simultaneously at any point of

17 
 
time. One or more of these mechanisms dominate depending on the droplet size of the

liquid.

Figure 2.1 Single fiber capture mechanisms (a) Inertial impaction, (b) Direct interception,

(c) Diffusion, (d) Gravitational deposition.

a. Inertial impaction: When a dense or large particle is transported in the gas

stream, the particle’s inertia may be large enough that the particle’s trajectory

will deviate from the air streamline as the air bends around a fiber. If the

particle’s path carries the particle within one particle radius to the fiber

surface, the particle will collide with the fiber and become captured. Aerosols

usually larger than 1 µm follow this mechanism of particle capture.

18 
 
b. Direct interception: Particles in the range of 0.3-1 µm in diameter usually

follow the air streamline; they will intercept a fiber if the distance of stream

line approaches to less than the particle’s radius to the fiber surface.

c. Diffusion: Particles smaller than 0.3 µm have very little mass. As a result they

travel in random motion superimposed upon the stream line flow due to

collisions with air molecules, similar to molecular diffusion. The random

motion increases the likelihood that they will come in contact with the fibers

and will be captured.

d. Gravitational deposition: When the velocities within the filter medium are

very low, the aerosol droplets of 20-50 µm in diameter may fall out of the air

stream before reaching the media. If they do reach the media they continue to

fall as they pass through the media. This vertical motion increases the

likelihood that they will bump into the fiber and be captured.

While the capture mechanisms described here are most effective on particles on a

certain size range, an increase or decrease in the air velocity can cause the capture

efficiency to vary [29]. The comprehensive view of fibrous filtration and its theory,

efficiency of fibrous filters, comparison of fibrous filters with other filtration, loading

characteristics of solid particles, and forces controlling the filtration process is given by

Brown [29] and Davies [30,31]. Stenhouse and Trotter [32] studied the loading of

submicron solid particles on fibrous filters.

The fiber sizes, fiber orientation, surface energy of the fibers and the filter,

surface tension of the liquid droplets, aerosol flow rate are key parameters which affects

the performance of the filter media. Liquid saturation and pressure drop of the filter

19 
 
media are important parameters which need to be measured and controlled for the filter

media to indicate higher filter life. Filter life and the filter media performance can be

improved by lowering the saturation and pressure drop with high capture efficiency. One

of the easy ways of improving filter life is to incorporate drainage channels in the

coalescing filters.

2.5 Coalescence filtration and drainage channel design

Incorporating drainage channels in the coalescing filters have been reported in the

literature. Miller et al. [25] designed a composite filter media made up of coalescing and

drainage layers at the inlet and outlet surfaces of the filter. The coalescing layer was

made up of micro glass fibers of 0.5 µm to 10 µm diameters and the drainage layer was

made of coarser fibers than the coalescing layer. Both of these layers were positioned in

face to face contact with each other with the coalescing layer facing the flow. The

coalescing layer and drain layer were separated by an open metal mesh having 70 % open

area. The composite medium had capture efficiency greater than 99% for capture of 0.01

µm to 10 µm size oil droplets with an inlet concentration of 400 milligrams per cubic

meter. The drainage layer received the oil from the coalescing layer and provided a path

for oil to flow by gravity from the filter. The drainage layers were either an open plastic

foam of polyester, polyethylene or polyurethane with 80 pores per inch or a felt of fibers

with diameters in the range of 10-20 µm loosely bonded together with an oleophobic

binder. The coalescing layer was bonded with an oleophilic binder. Experiments were

conducted with or without the drainage layers impregnated with a low surface energy

fluorocarbon material. The fluorocarbon impregnated drainage layers had better

20 
 
performance in terms of reduced oil carry over and in removing the oil from the filter

media by providing a flow path for oil to flow by gravity from the filter.

Cylindrical pleated filter media made up of three layers of fibers having different

fiber diameters and surface energies are discussed in references [33, 34]. In their designs

the upstream and downstream layers had larger pore openings as compared to the

intermediate layer made up of fibers having diameters from 0.1 µm to 20 µm. The filter

media were made of polyolefins or fluorochemical materials to modify the surface

properties. The fluorochemical treated and untreated media were tested for separating

mineral oil droplets from an air stream and the pressure drop of the filter media was

reported. Their experimental results indicated that the untreated filter structure had a

much higher pressure drop as compared to the treated filter structure. The filters without

surface treatment had high oil concentrations in the downstream air which at some point

was above the accurate measurement range of the sampling equipment while the surface

treated media had much lower oil carry over [34].

Hunter [35, 36] developed a filter media design with a coalescing layer of

borosilicate glass microfibers and a drainage layer made up of open-celled foam plastic

material. The open-celled foam plastic material had a lower surface energy as compared

to the glass microfibers. He observed that drainage channel incorporated filter media

have better drainage and a lower oil wetted area which lowered the possibility of

reentrainment of oil drops into the gas stream.

Spencer [37] developed a tubular filter medium with a coalescing layer and a low

surface energy drainage layer located downstream of the coalescing layer. The coalescing

layer was made up of 0.5-10 µm size microglassfibers while the drainage layer was made

21 
 
of polyester spunbond 40-60 micron diameter fibers. The polyester fibers have a lower

surface energy than the glass fibers. The coalescing layer used an acrylic binder while the

drainage layer used 15 % bicomponent polyethylene terephthalate fibers as binder which

had a lower melting point than the polyester fibers. The drainage layer was a 5 mm sleeve

made from 200 g/m2 spunbond fabric which surround the coalescing layer. The

spunbond fibers, being larger than the glass fibers, had larger pore openings that reduced

capillary action to retain liquid in the medium.

Waltl [38] modified the filter design by incorporating inner and outer drainage

layers. The inner drainage layer was made of two layers of spunbond polyester while

outer drainage layer was made of polyester bi-component nonwoven fleece material. The

authors detected the oil carry over by using Gelman filter paper and observed that the low

surface energy drainage channel equipped filter media has significant reduced oil carry

over and pressure drop.

2.6 Fiber orientation and filter performance

The fibers in nonwoven filter medium tend to lie with their axes approximately

perpendicular to gravity but randomly oriented in the horizontal plane when made by

using the vacuum molding process which is also known as a wet-laid process in the

industry. Because of this, filter media cut at different angles from a large vacuum formed

filter cake have different orientations (angles) relative to the inlet surface. Filter

performance relative to the fiber orientation have been reported [39-41].

Chokdeepanich [39] experimented with media cut from a large filter cake of glass

micro fibers to construct media of the same fibers and same porosity but with different

fiber orientations. He tested filter samples made of two sandwiched layers of equal

22 
 
thickness in liquid aerosol filtration. The layers were combinations of capture and

drainage layers. The capture layers were constructed of glass micro fibers with the fibers

predominately perpendicular to the direction of gas flow. Drainage layers were

constructed of the same glass microfibers with the same packing fraction (porosity) but

with the fibers oriented with their axes at angles more closely aligned with the direction

of flow. The fiber orientation in the drainage layer had less flow resistance for the

captured liquid drops and less flow resistance to the gas flow. Filters were assembled as

combinations of capture and drainage layers or two layers of the same type. The layers

were oriented normal to the flow, with the flow passing through each layer in succession.

The filter samples with a capture layer at the inlet surface followed by a drainage layer

gave better quality factor performance compared to the other combinations.

2.7 Saturation of the filter media

Liquid saturation of the filter medium is the volume fraction of the pores filled by

the liquid phase. It is dependent upon many of the above parameters and it directly

affects the pressure drop and local gas velocity within the medium. The mass of a liquid

captured by a filter medium is typically measured gravimetrically by subtracting the mass

of a filter medium before an experiment from the mass of the medium at the end of an

experiment. The liquid mass is converted to volume using the liquid density and the

average saturation is determined as the ratio of liquid volume to void volume. By

constructing a filter medium as a stack of thin layers, Andan [42, 43] experimentally

demonstrated that local saturation in a filter varies with position. A continuum model

shows that the average saturation and the variation in local saturation affect the pressure

drop [44]. In general, the greater the saturation the higher is the pressure drop. However,

23 
 
it is plausible that a locally high saturation can cause a high pressure drop in a filter

medium that has a low average saturation.

Local saturation has been investigated by models and experiments [42-44]. In

their continuum model Andan et al. [44] show that the local saturation profile affects the

pressure drop across the filter medium by restricting the flow and causing an increased

local gas velocity. Experimental measurements show the saturation in a medium of

uniform properties is higher near the inlet and outlet surfaces and lower in the interior of

the medium [42-46]. They propose this is due to the small size of the drops carried by

the gas at the inlet and the smaller drops tend to move slower than larger drops on fibers

as those found in the interior [43] The increase in saturation at the exit is not well

understood but it is suspected that it is due to the capillary forces holding the liquid to the

fibers and the resistance to flow of the liquid to drain vertically down the outside of a

horizontal flow oriented filter, causing the saturation to be higher at the exit surface. The

saturation was observed to decrease as the flow velocity increased [46-49] as expected

because of higher drag forces that cause the drops to move faster and the mass continuity

balance requires a decrease in the saturation at higher drop flow rates. In the interior of

the filter, larger drops will flow faster than the drops at the inlet which due to the mass

balance should similarly cause a decrease in the saturation in the interior as indicated

above [43, 44]. Contal et al [21] measured the saturation with respect to time during

various stages of liquid loading. Authors [48-50] describe the saturation and drainage

processes in terms of the Washburn equation. They consider that there is more drainage

when the saturation is high. At the exit boundary, saturation builds up due to a barrier

effect. The liquid drops have high resistance at the exit boundary due to surface tension.

24 
 
The resistance acts as a barrier for the liquid flow and increases the saturation at the

boundary.

2.8 Nanofibers in coalescence filtration

Nanofibers are small sized solid fibers with diameters less than 1 μm. The

nonwoven industry generally considers nanofibers as having a diameter of less than one

micron, although the National Science Foundation (NSF) defines nanofibers as having at

least one dimension of 100 nanometer (nm) or less [51]. Nanofibers possess large surface

area per unit mass and form mats of small pore size. Nanofiber augmented filter media

showed lower energy consumption and easier maintenance, higher filter life, increased

contaminate holding capacity and enhance filtration efficiency [16, 52-59]. The

nanofiber based filtering media, made up of fibers of diameter ranging from 100 to 1000

nm, can be conveniently produced by electrospinning technique [60] and the fiber

diameter depends upon solution properties and concentration, hydrostatic pressure in

capillary tube, electric potential at the capillary tip, the tip-to-collector distance [57-60].

Zhang et al. [61] studied effect of nylon nanofiber diameter fabricated using the

electrospinning process on the performance of the filter media. Kim et al. [62] showed

the possible application of electrospinning in producing nylon nanofiber filtering media

specifically designed for capturing particles smaller than 50 nm. Fine-tuning of fiber size

distribution by controlling the spinning process is thought to facilitate the application in

this area [52, 58]. The pressure drop of a Nylon 6 nanofiber filter linearly increases with

the increasing face velocity [62]. An electrospun Nylon 6 nanofiber filter (mean fiber

diameter: 100 nm) shows a much lower pressure drop performance relative to the

commercial HEPA filter media when the filtration efficiency of the Nylon 6 nanofiber

25 
 
filter and the HEPA filter are over 99.98% with test particles of 0.02-1.0 m in diameter.

The particle capture efficiency decreases with the increasing fiber diameter.

2.9 Modeling of fibrous filters

The pressure drop caused by fibrous filters has been studied for many years and

numerous analytical, numerical and empirical correlations are available for such media.

In almost all of these models, a filter is assumed to be made up of fibers of one size [64-

71]. Authors [71, 72] studied the effect of fiber diameter and developed a correlation for

pressure drop of filter media with fiber diameter which is a function of fiber diameter

ratio and their number fractions. The capture mechanisms for solid particles and liquid

droplets are quite similar, but the efficiency and pressure drop profiles are different for

both processes [21]. Brown [73] developed a model to predict pressure drop when the

particle are captured by inertial impaction by assuming all fibers are symmetric and

arranged parallel to each other. Authors [73-77] divided coalescence filtration in to four

stages according to the changes in pressure drop. The pressure drop increases during the

first three stages and during the final stage, equilibrium is reached between liquid

loading, drainage and re-entrainment and the pressure drop does not change thereafter.

Also, he made a comparison between the pressure drop profiles of filters with solid and

liquid aerosols. They also found that the fiber size, filtration velocity and surface tension

effect the clogging of fibrous filter.

26 
 
CHAPTER III

DEVELOPING AND CHARACTERIZING FILTER MEDIA

3.1 Introduction

Developing and characterizing filter media for developing different filter

geometries is objective of this research work. This chapter focuses on filter media

preparation recipe, technique and characterization techniques.

3.3 Steps involved in filter media preparation

All the filter samples are prepared by vacuum molding process following the below

mentioned steps.

3.2.1 Glass fiber filter media

Glass fiber filter media are widely used in industries for various coalescing

filtration applications. Different grades of glass fibers are available for a variety of

applications. Silica is the base ingredient for most forms of glass fibers. Higher flow rates

and effective filtration of micron/submicron droplets can be achieved through glass fiber

media. B-Glass media has been widely used for coalescing type filtration applications

and the same media has been used for this experimental work. Glass fibers of average

fiber diameter 3 µm are used to prepare the nonwoven glass fiber filter media by using

27 
 
the vacuum molding process. The various steps involved in making the glass fiber filter

media are as follows.

1. Selecting a binder and starch: Carboset® 560 (The Lubrizol Coorporation) was

previously used as a test binder. Carboset® 560 is an acrylic polymer dispersed in

water which has milky white appearance [78]. It has a density of 1.0-1.2 kg/m3 [78].

The disadvantage of working with Carboset® 560 is that it requires mixing for 24

hours to interact with the glass fibers and to avoid clogging the metal screen during

the vacuum molding operation. The filter media made using Carboset® 560 softens

after about 2 hours of exposure to the Sullube 32® liquid aerosol in the coalescence

filtration experiment [74]. Filter media showed significant decrease in the structural

strength after coalescence filtration testing which indicated that the filter media is not

inert to the aerosol. On an industrial scale, the filter media must perform for many

hours after attaining steady state. Hence Megasol® S50 (Wesbond Corporation) was

selected as the binder along with the Westar+ starch (Wesbond Corporation).

Westar+ is a cationic flake corn starch which is specifically developed for bonding

silica based fibers to form vacuum formed composites and shapes [79]. This starch is

pre-cooked, disperses evenly and dissolves readily in cold water; hence it can create

more uniform structures [79]. Westar+ requires 30% less colloidal silica for optimum

flocking hence less binder is required, which reduces the cost of making filter

samples [79]. Megasol® S50 is a colloidal silica binder with the colloidal particles of

size 70 nm [80]. Megasol® S50’s unique physical properties result in higher packing

densities and stronger bonds [80]. Westar+ and Megasol® S50 both can react well

with the glass fibers at pH 7. Hence there is no need to maintain a specific pH which

28 
 
saves addition of acid or base. Hence filtered tap water was used to make the filter

sample which further reduces the cost of filter media making. A numbers of trial

experiments are run to formulate the ideal binder content. Table 1 indicates the

amount of binder and starch content and the media performance.

Table 3.1 Filter media performance and amount of binder and starch

Amount Amount Amount Capture Pressure Quality Comments

of glass of starch of Binder Efficiency drop factor

fibers Westar+ Megasol® (%) (kPa) (1/kPa)

(gm) (gm) S50 (ml)

5 0.15 1 70.24* 7.58* 0.13* Filters did

5 0.15 1.5 72.34* 8.16* 0.13* not reach

5 0.15 2.0 75.67* 8.28* 0.13* steady

5 0.15 2.5 76.89* 8.31* 0.13* state

5 0.15 3.0 77.16* 8.58* 0.13*

5 0.15 3.5 78.12* 8.81* 0.14*

5 0.15 4.0 82.23* 9.13* 0.16*

5 0.15 4.5 85.12* 9.28* 0.16*

5 0.15 5.0 87.87 9.64 0.18 Filters

5 0.15 5.5 90.12 12.78 0.14 reached

5 0.15 6.0 92.34 14.45 0.12 steady

state

* indicate the final capture efficiency, pressure drop and quality factor values since

the filters did not reach steady state.

29 
 
Every data point in table 3.1 is an average of three filter samples made with the

respective amount of starch and binder and tested for three hours in the lab scale

coalescence filtration setup. Filters labeled with an asterisk did not reach steady state.

This possible reason can be insufficient amount of binder. The fibers were not bonded

well together and were not stationary. They might have relative motion. These filters

had less structural stability and had fiber shedding problem as compared to the filters

which reached steady state. The amount of starch content was kept constant based on

the fiber to starch ratio of 1: 0.03 by weight as suggested by the company.

2. To make the slurry of fibers, 5 grams of B-glass fibers are soaked in 3 liters of

water for about 30 minutes. Uniform diluted slurry of fibers is made by using the

mechanical blender and adding 6 liters of water. 0.15 grams of Westar+ starch

was added and well dispersed by using the mechanical stirrer for 5 minutes.

Surface charge, good dispersion, and water solubility of Westar+ significantly

reduces the stirring time. 5 milliliter of Megasol® S50 was added and the slurry

was agitated for another 10 minutes. A mixing tank with mechanical stirrer to

make the slurry of fibers is shown in Figure 3.1. From Fig. 3.1, it can be seen that

the mixing tank is completely covered to prevent external particle contamination

of the slurry. This slurry is poured into the vacuum molding tank of the vacuum

molding process for making disc shaped filter media.

3. To form the filter media discs, a vacuum molding process is used to form the disc

shaped filters of 6 cm in diameter and 1.4 cm thick. The wet medium is dried in

an oven at 1500 C for 3 hours to remove the moisture and bind the media. The lab

scale vacuum molding assembly is described in section 3.3.

30 
 
Mechanical Stirrer

Mixing tank 

Figure 3.1 Mixing tank assembly to make slurry of fibers.

3.2.2 Nanofiber augmented micro glass fiber filter media

A vacuum molding process is used to make a nylon nanofiber micro glass fiber

filter media. Nylon nanofibers are made by dissolving nylon-6 (Sigma-Aldrich) in

formic acid (Sigma-Aldrich) and the polymer solution is electrospun by using

NanospiderTM and single jet electrospinning as described in Chapter VII. Electrospun

nylon nanofibers are measured and roughly chopped by scissor and then are finely

chopped and mixed with water using a mechanical blender. This nanofiber slurry is

mixed with the thick slurry of micro glass fibers soaked in 3 liters of water. Nylon

nanofiber micro glass fiber filter media are made by following the above mentioned

procedure. Srinivasan [81] showed that the optimum filter performance occurs with a

nanofiber to microfiber area ratio of about one. The area ratio is the ratio of the external

surface area of nanofibers to surface area of glass fibers estimated from the diameters of

31 
 
the fibers, fiber densities, and mass of fiber used to construct the media. Hence nylon

nanofibers of average fiber diameter 100 nm, 300 nm and 600 nm were added in

proportionate amounts to obtain the area ratio of one.

3.2.3 Composite filter media

Composite filter media is made up of glass fibers of different fiber diameters. The

different fiber sizes used are given in Table 3.2. Slurries of fibers along with starch (1:

0.03 wt %) and binder (1:1 wt %) were made following the above mentioned procedure.

Slurries of different fiber sizes were prepared separately and then poured into the vacuum

molding tank one after another starting with the biggest fiber size. Largest to smallest

fiber size order was selected to make a filter cake with minimal resistance for flow

through the metal screen.

Table 3.2 Composite filter media design

Fiber diameter (µm) Thickness of filter

3 6 10 39 media (inch)

Amount of 0.5 0.5 0.5 0.5 0.25

fibers (gm) 1 1 1 1 0.50

1.5 1.5 1.5 1.5 0.75

32 
 
3.3 Vacuum molding setup

R
AI

A
IR
R AI
AI R

Figure 3.2 Schematic of vacuum molding setup.

Schematic and lab scale set up of vacuum molding process is shown in Figure 3.2

and 3.3 respectively. The vacuum molding tank, mold to make 6 cm diameter disc filters

and a vacuum pump are three major parts of the vacuum molding setup. The slurry is first

33 
 
poured in to the vacuum molding tank of the assembly. Air is passed though the slurry

through various ports available in the vacuum molding tank to agitate the slurry and to

prevent the fibers from settling down. The mold to make 6 cm disc filter is fixed in at the

bottom of the vacuum molding tank along with a metal screen to prevent fibers from

passing through. The mold is then connected to the vacuum pump through a vacuum tank

which serves as a temporary storage of drain liquid. The vacuum pump can generate

around 100 to 500 mm of mercury vacuum pressure. The vacuum is applied through the

vacuum tank to the slurry tank. The liquid is sucked through the metal screen and stored

in the vacuum tank. The metal screen is covered with the filter paper on which the disc

shaped filter cakes are build up. The vacuum is controlled to maintain uniform structure

of filter cake. Higher vacuum will create a non-uniform as well as less porous filter cake

while lower vacuum will not generate a proper driving force for liquid to pass through the

buildup cake of fibers. The mold is removed from the slurry tank and the filter medium is

removed from it. The liquid stored in the vacuum tank is drained out to get it ready for

the next run. The filter media made using the process is a filter disc 6 cm in diameter and

1.4 cm in thickness. The nonwoven micro B-glass fiber filter media were made using the

vacuum molding process described above had irregular shaped pores on the order of 10 to

20 microns in diameter.

34 
 
Vacuum molding Vacuum pipe
tank  attached to vacuum  

Tubes for creating air bubbles

Vacuum tank

Filter mold parts


Vacuum pipe
attached to Mixing
tank
Metal screen

Figure 3.3 Lab scale setup of vacuum molding process.

3.4 Filter media characterization techniques

There are various techniques to characterize filter media. Filter media

characterization are important to understand the properties of media. The filter media are

characterized to obtain porosity, air permeability and strength. The experimental

procedures to measure porosity by using pycnometer, air permeability using The

Frazier® Differential Pressure Air Permeability Measuring Instrument and strength or

hardness by using the Durometer are described below.

35 
 
3.4.1 Porosity measurement

Porosity of the filter, ε , is the ratio of the volume occupied by the pores to the

total volume of the filter medium. The porosity varies from zero to one. A good filter

medium has a higher porosity that results in a lower pressure drop. The bulk porosity of

all the media are measured by using a special made pycnometer. Fig. 3.4 shows the

pycnometer and its components. There are two chambers in the pycnometer operated by

using the ball valve which is in between the two chambers. Initially the ball valve is

closed. The filter sample is loaded in the top chamber and bottom chamber is

pressurized to 15 psig. The top chamber is designed to fit thick filter samples and a

number of spacers having different thicknesses can be used to fill the remaining gap.

The valve between the two chambers is then opened to equalize the pressure. Boyle’s

law states that pressure is inversely proportional to volume. Void volume is determined

by the difference in pressure in the sample chamber before and after opening the valve.

Three different sets of pressure readings are taken for each sample. In first set of

readings, three readings are taken with just the spacer in the top chamber. In second set

of readings, three readings are recorded with a calibration cube with known dimensions

in the top chamber and in third set of readings are recorded with filter sample placed in

the top chamber. The readings in triplicate are averaged. The average values are used to

determine the porosity using Boyle’s law. The porosities of the glass fiber filter media

are similar. The typical porosities of the glass fiber filter and composite filter media are

0.96 +/- 0.01 and nanofiber augmented glass fiber filter media are 0.94 +/- 0.01 where

+/- error is the standard deviation of three or more measured samples. Porosities of the

glass fiber filter media are plotted against filter geometry as shown in Figure 3.5. All the

36 
 
filter media indicate similar porosities. It is desirable to study the performance of filter

geometries and effect of gravitational force on the performance of the filter media

solely. The porosity of drainage channels is obtained by using the density bottle. The

porosity of the drainage channel is 0.986.

Pressure gauge

Sample chamber
Ball valve

Vent

Air inlet

Figure 3.4 Lab scale pycnometer assembly.


0.96

0.96

0.96

0.96

0.96

0.96

0.96

0.96

0.96

1.00
0.95

0.95

0.95

0.95

0.95

0.95

0.95

0.95

0.95

0.95

0.95

0.95

0.95

0.90

0.80

0.70

0.60
Porosity

0.50

0.40

0.30

0.20

0.10

0.00
H-N H-I H-M H-O H-B . H-0 H-15 H-30 H-40 H-45 H-50 H-60 V-N V-I V-M V-O V-B . V-0 V-15 V-30 V-45 V-60 .

Filter geometry

H-HORIZONTAL ORIENTATION
H-HORIZONTAL ORIENTATION H-0 : Drainage at 00 (three cut) V-VERTICAL ORIENTATION
H-N : No drainage channel V-VERTICAL ORIENTATION
H-15 : Drainage at 150 (three cut) V-0 : Drainage at 00 (three cut)
H-I : Inlet drainage channel V-N : No drainage channel
H-30 : Drainage at 300 (three cut) V-15 : Drainage at 150 (three cut)
H-M : Middle drainage channel V-I : Inlet drainage channel
H-40 : Drainage at 400 (three cut) V-30 : Drainage at 300 (three cut)
H-O : Outlet drainage channel V-M : Middle drainage channel
H-45 : Drainage at 450 (three cut) V-45 : Drainage at 450 (three cut)
H-B : Drainage channel at Both ends H-50 : Drainage at 500 (three cut) V-O : Outlet drainage channel
V-B : Drainage channel at Both ends V-60 : Drainage at 600 (three cut)
H-60 : Drainage at 600 (three cut)

Figure 3.5 Porosities of glass fiber filter media.

37 
 
3.4.2. Permeability measurement

Permeability is a measure of the ease of a fluid that passes through the filter

medium. A number of techniques can be used to measure the permeability. Frazier®

Differential Pressure Air Permeability Measuring Instrument (Frazier Precision

Instrument Company, Inc.) is used for permeability measurement. A filter medium should

be highly permeable to the fluid flow. If the filter medium has low fluid permeability,

then filter media will indicate high pressure drop. Generally, permeability is measured by

passing a fluid through the medium at a known pressure drop and observing the flow rate.

For fibrous filters, the fluid permeability is given by Darcy’s Law [3, 81].

Q Gk ΔP
= (3.1)
A μ L

where Q = volumetric flow rate

A = cross section area of the exit surface

G = shape factor

k = permeability

ΔP = pressure drop

L = media thickness

In this work, the Frazier® Differential Pressure Air Permeability Measuring

Instrument, shown in Figures 3.6 and 3.7, is used to measure air permeability of all filter

samples. Air permeability of the filter media has a unit of m2. The filter medium is

placed in the sample holder. The filter medium must fit snuggly in the sample holder to
38 
 
avoid any air flow along the sides of the sample. Air flow along the sides of the medium

will produce erroneous results because the air is not flowing through the medium itself.

Different orifice openings are tested and specific orifice opening is selected to obtain the

pressure drop at 0.5 psig as the test is designed to produce accurate result at 0.5 psig

pressure drop. Three sets of readings are taken for each sample. The Frazier® Differential

Pressure Air Permeability Measuring Instrument is designed to measure air permeability

of thin samples where the shape factor is 1. For other samples, the shape factor is read

from a correlation and is incorporated in the permeability calculation. The air

permeability (measured with flow perpendicular to the mesh) of Nylon, Polypropylene

and Teflon® woven mesh and B-glass fiber filter media are 3.39x10-9 m2 +/- 0.04x10-9

m2, 5.67x10-9 m2 +/- 0.02x10-9 m2, 8.67x10-9 m2 +/- 0.01x10-9 m2 and 2.24x10-10 m2

+/- 0.03x10-10 m2, respectively where the +/- error is the standard deviation of three or

more measured samples. The air permeability of polymeric woven meshes were ten times

higher than the air permeability of the glass microfiber filter medium. The air

permeability of Nylon and Polypropylene spunbond fabric was 4.54x10-9 m2

+/- 0.01x10-9 m2 and 6.54x10-9 m2 +/- 0.02x10-9 m2 while the Nylon electropsun

nanofiber on Teflon® mesh drainage channel had an air permeability of 4.81x10-10 m2

+/- 0.03x10-10 m2.

39 
 
Manometer

Different orifice

Pump control

Pump

Figure 3.6 Frazier® Differential Pressure Air Permeability Measuring Instrument.

a b c

d e

Figure 3.7 Components of Frazier® Differential Pressure Air Permeability Measuring

Instrument (a) Sample holder (b) Manometer to read pressure drop (c) Orifice (d) Pump

and pump control (e) Different orifice opening.

40 
 
2.50

2.26

2.26

2.26
2.25

2.25

2.25

2.25

2.25
2.24

2.24
2.24

2.24
2.24
2.24

2.24

2.24
2.24

2.24

2.24
2.24

2.24
2.24
2.00

1.50
Air Permeability * E-10 (m^2)

1.00

0.50

0.00
H-N H-I H-M H-O H-B . H-0 H-15 H-30 H-40 H-45 H-50 H-60 V-N V-I V-M V-O V-B . V-0 V-15 V-30 V-45 V-60 .

Filter geometry
H-HORIZONTAL ORIENTATION
H-HORIZONTAL ORIENTATION H-0 : Drainage at 00 (three cut) V-VERTICAL ORIENTATION V-VERTICAL ORIENTATION
H-N : No drainage channel H-15 : Drainage at 150 (three cut) V-N : No drainage channel V-0 : Drainage at 00 (three cut)
H-I : Inlet drainage channel H-30 : Drainage at 300 (three cut) V-I : Inlet drainage channel V-15 : Drainage at 150 (three cut)
H-M : Middle drainage channel H-40 : Drainage at 400 (three cut) V-M : Middle drainage channel V-30 : Drainage at 300 (three cut)
H-O : Outlet drainage channel H-45 : Drainage at 450 (three cut) V-O : Outlet drainage channel V-45 : Drainage at 450 (three cut)
H-B : Drainage channel at Both ends H-50 : Drainage at 500 (three cut) V-B : Drainage channel at Both ends V-60 : Drainage at 600 (three cut)
H-60 : Drainage at 600 (three cut)

Figure 3.8 Air permeability of glass fiber filter media.

3.4.3 Hardness or strength measurement

Strength or hardness of the media indicates its structural strength. The filter media

need to have structural integrity to minimize the fiber shedding as well as to withstand at

high flow rates. Strength is indirect measure of amount of binder in the medium. A filter

medium without a binder is too soft and structurally weak. Hardness of the sample is

measured using Durometer (Rex Gauge, Type 0 and Model 1600) as shown in Figure 3.9.

Durometer has a scale of 1 to 100 with 100 being the hardest.

41 
 
Durometer

Filter media

Figure 3.9 Lab scale Durometer.

The filter medium is placed on a hard surface. The pointer at the end of the

Durometer is pressed against the surface of the filter medium and the measurement is

taken directly from the scale. Strength measurement can also be done after the

coalescence filtration test as an indication of whether the binder interacts and softens in

the presence of the aerosol liquid. If hardness measurement of the filter before and after

coalescence filtration test changes significantly, it indicates that the filter media is not

inert to the aerosol.

Glass fiber filter media made by using Megasol® S50 binder have a hardness of

around 20 while the Nylon nanofiber micro glassfiber filter media have a hardness of 30.

Filter media of relatively similar porosity, air permeability and strength are used to

develop different filter geometries by incorporating drainage channels.

42 
 
CHAPTER IV

DEVELOPING FILTER GEOMETRIES

4.1 Introduction

Developing different filter geometries to improve the performance and drainage

of coalescing filter is objective of this research work. Different filter geometries are

developed by incorporating woven and nonwoven drainage channels of varying surface

energies in the glass fiber filter media. The effect of pore opening as well as filament

thickness of woven drainage channel is studied on the performance of filter media. The

effect of different basis weight of nonwoven drainage channel on the performance of the

filter media is also studied. The effect of wettability or surface energy of the woven and

nonwoven drainage channel on the performance of the filter geometries is also studied.

The effect of gravitational force on the performance of the filter media is studied by

changing the flow orientation.

4.2 Filter media cutter saw

Nonwoven disc shaped glass fiber filter media are made by using the vacuum

molding process as described in Chapter 3. These filter samples are cut by using a

dedicated band saw at various locations along the filter bed length and drainage channels

are incorporated to develop different filter geometries. The filter media cutter saw is

shown in Figure 4.1. A hard back carbon steel blade of 1/8 inch thickness and 32 teeth

43 
 
per inch is used to cut the filter sample. This is the thinnest and finest blade available in

the market which cuts the filter samples without ripping off fibers from the media. The

filter media is spongy in nature and squeezes easily hence it is important to use a fine

blade to cut the filter media without losing its structural stability.

Band saw
b

Blade

Filter holder

Figure 4.1 Filter media band saw and cutter assembly (a) band saw (b) filter holder.

A plexiglass filter holder is made to hold the filter samples of different

thicknesses. The filter holder is shown in Figure 4.2. Filter holder is equipped with a

screw which allows adjustment to change the thickness of the sample holder and hence

enables a filter medium to be cut at the desired thickness. The filter holder is also

equipped with a compass and a guiding ruler which helps in keeping the holder parallel

as well as at different angles respective to the cutter saw blade. Filter media are cut by

44 
 
using the cuttter saw wheen the filterr holder is parallel
p to thhe cutting bllade as show
wn in

Figure 4.3. Figure


F 4.3 inndicates thatt the filter samples
s cann be cut of desired
d thickkness

w
without losing
g the structuural strength of the filter medium.

S
Screw

Coompass
G
Guiding rulerr

S
Sample cham
mber

Figure 4.2 Fillter holder too cut filter saample by usiing the cutteer saw.

a b

Figure 4.3 Fiilter media sample


s (a) Thick
T filter media
m (b) Fillter media cut by using filter

m
media cutter saw.
s

45 
 
4.3 Drainage channels

Drainage channels are woven and nonwoven polymeric materials. Drainage

channels are open porous structures and the materials have low surface energy compared

to the glass fiber filter media. Polypropylene woven mesh, (Spectrum Laboratories) is

used as a drainage channel shown in Figures 4.4 (a) and 4.4 (b). All of the filter media

samples are cylindrical in shape; hence many drainage channels are circular in nature as

shown in Figure 4.4 (a).

a b

Figure 4.4 Woven drainage channels (a) Polypropylene circular shaped woven drainage

channel (b) Microscopic image of woven polypropylene drainage channel.

Figure 4.4 (b) is the microscopic image of the polypropylene woven drainage

channel indicating uniform pore opening of about 500 µm and filament thickness of 600

µm. Polypropylene spunbond (Spunfab Ltd.) used as drainage channels is shown in

46 
 
Figures 4.5 (a) and (b). Because the filter media samples are disc shaped, the nonwoven

drainage channels are circular as shown in Figure 4.5 (a). Figure 4.5 (b) is the

microscopic image of the polypropylene nonwoven fabric commonly known as spunbond

which shows non-uniform pores with the filament thickness of the 50 µm. Along with

polypropylene, woven meshes of nylon (Dexmet Corporation) and Teflon® (Dexmet

Corporation) as well as nonwoven nylon spunbond (Spunfab Ltd.) were used as drainage

channels to develop filter geometries to study the effect of surface properties and pore

opening of drainage channels on the performance of filter media.

a b

Figure 4.5 Nonwoven drainage channels (a) Polypropylene circular shaped nonwoven

drainage channel (b) Microscopic image of nonwoven polypropylene drainage channel.

Figure 4.6 (a) indicates circular shaped nylon (NF) +Teflon® drainage channel

which is made up by electrospinning nylon-6 (Sigma-Aldrich) nanofibers on the surface

of the Teflon® woven mesh of 500 µm pore opening and 600 µm filament thickness.

47 
 
Figure 4.6 (b) shows the non uniform pore opening of the (NF) + Teflon® drainage

channel. The pores of the nylon nanofiber mat are significantly smaller as compared to

the pores of polypropylene woven and nonwoven drainage channels.

aa b

Figure 4.6 (a) Nylon nanofibers electrospun on Teflon® woven mesh of 500 µm pore

opening (b) Microscopic image of Nylon nanofibers on Teflon® woven mesh.

4.4 Filter geometries

4.4.1 Horizontal orientation

Filter media are cut with the help of the filter media cutter saw and woven as well

as nonwoven drainage channels are incorporated in the filter media and different filter

geometries are developed. These filter geometries are made up of a nonwoven coalescing

layer of glass fibers having smaller pores and woven as well as nonwoven drainage

channels having big pores. The glass fiber coalescing media has a higher surface energy

than the materials forming the drainage channels. Glass fiber filter media is an oleophilic

48 
 
and hydrophilic, hence it aids in coalescence. Smaller drops coalescence to form big

drops in the glass fiber media and these big drops drain very effectively from the filter

media through the low surface energy porous drainage channels. Different filter

geometries developed and tested in horizontal orientation are shown in Figure 4.7.

Drainage channels

G G G G G G f
θ
L L L L L L
Gas flow θ
A A A A A A

S S S S S S θ
Gravity
S S S S S S

a b c d e f

Figure 4.7 Filter geometries in horizontal orientation (a) No-drainage (b) Inlet-drainage

(c) Middle-drainage (d) Outlet-drainage (e) Both-End-Drainage (f) Inclined-angle-

drainage.

“No-drainage” channel filter media are glass fiber filter media without any

drainage channels. Filter geometries labeled as “Inlet,” Outlet” and “Middle” drainage

each have one circular drainage channel at the inlet, outlet and in the middle of the filter

media. A filter with drainage channel at both the inlet and outlet surfaces is referred to as

“Both-End-Drainage”. The last design has three equally spaced rectangular drainage

channels inserted at inclined angles within the media. All the drainage channels are

49 
 
physically placed touching face to face at the inlet, outlet, middle or both the ends of the

filter. For the inclined angle experiments in horizontal orientation, the filter media are cut

at angles 00, 150, 300, 400, 450, 500, and 600 and rectangular shaped drainage channels

were inserted physically touching face to face with the glass fiber sections of the filter

media. All the filter geometries were held together by a compressive force when

assembled into the filter holder. When the filter geometries tested with the air flow

direction perpendicular to gravitational force, it is termed as “horizontal orientation”.

4.42. Vertical flow orientation

When the air flow is in the direction of gravity it is termed as “vertical

orientation”.

Gas flow Gas flow


Gravity

G L A S S
G L A S S

a
d

G L A S S G L A S S
b
e
G L A S S
θ θ θ

G L A S S

c Drainage channel
f

Figure 4.8 Filter geometries in vertical orientation (a) No-drainage (b) Inlet-drainage (c)

Middle-drainage (d) Outlet-drainage (e) Both-End-Drainage (f) Inclined-angle-drainage.

50 
 
Inlet, outlet, middle, both-end and inclined angle experiments were performed in

vertical orientation to study the effect of gravitational force on the performance of the

filter media. For vertical flow orientation experiments the different angles studied were

00, 150, 300, 450, and 600. Filter geometries in vertical orientation are shown in Figure 4.8.

4.5 Composite filter media

Composite filter media is a filter media design which is made up of different fiber

sizes of glass fibers. The different fiber sizes used are 3 µm, 6 µm, 10 µm, and 39 µm.

The filter media is developed to design a filter which will have bigger pores along the

filter bed length. The composite filter media is developed by layering the fibers in the

order of 3 µm, 6 µm, 10 µm, and 39 µm with 3 µm fibers facing the flow. As the droplets

moves along the filter bed length they coalesce to form bigger drops. This filter design

will have smaller to bigger pores along the filter bed length. Hence it will be capable of

capturing variety of droplet sizes and will have lower pressure and higher drainage and it

can help in reducing the secondary aerosol generation. The composite filter media design

is shown in Figure 4.9. Figure 4.9 indicate a filter media design made up of glass fibers of

fiber sizes 3 µm, 6 µm, 10 µm, 39 µm in section 1, 2, 3, and 4 respectively.

1 2 3 4

Figure 4.9 Composite filter media.

51 
 
CHAPTER V

COALESCENCE FILTRATION SETUP

Construction of filter media, their characterization, and developing different filter

geometries by incorporating drainage channels are discussed in previous chapters. This

chapter discusses experimental procedure and setup to test the various filter geometries.

All of the filter geometries as well as the “No drainage” channel filter media are tested in

the lab scale coalescence filtration setup. The coalescence filtration set up is equipped to

obtain the inlet and outlet droplet size distribution, in-situ upstream and downstream

concentration monitoring using a photometer, pressure drop across the filter media, and

flow rates.

5.1 Experimental setup and procedure for coalescing tests

The schematic of the coalescing filter testing set up is shown in Figure 5.1. A

photo of the lab scale setup of coalescence filtration and its major components are shown

in Figures 5.2 and 5.3 respectively. The process is described in four sections in this

chapter.

5.1.1 Prefiltration

Atmospheric air contains various impurities and water vapor. Those impurities are

removed before challenging the air to the coalescing filter medium. A surge tank is used

to dampen effects of fluctuations in house air pressure on the experiment. The

52 
 
compressed air initially passes through a drier to remove particulates moisture, then

through HN2L-10CD coarse filter, HN2L-6CD fine filter; DN4L-SG4 silicon gel

desiccator and HN2L-AUD filter to remove micron and submicron size dust and dirt

particles. All these filters are Parker Hannifin Corporation filters. The surge tank and

prefiltration system are shown in Figure 5.3.

Figure 5.1 Schematic of the coalescing filter testing set up.

53 
 
Filter holder: Horizontal orientation Flowmeter

HEPA filter
Mixing chamber

Photometer

Laskin nozzle

Surge tank
Pressure transducer

Figure 5.2 Lab scale coalescence filtration setup.

a b Prefilters

Surge tank

Dryer

Figure 5.3 (a) Surge tank (b) Prefilters.

54 
 
5.1.2 Aerosol generation

After removal of impurities and moisture the purified airline is then divided in to

two streams. The main stream goes directly to the filter holder through a heating

assembly and the side stream goes through a Laskin nozzle to generate liquid droplets in

to the air stream is shown in Figure 5.4 (a). A Pressure drop of 25 psig needs to be

maintained across the Laskin nozzle to generate droplets of average diameter of 300 nm.

Aerosol is generated by dispersing fine droplets of compressor oil Sullube® 32 which is

manufactured by Dow Chemical Company and contains 60-70 % of polypropylene glycol

with a density 982.7 kg/m3 [83]. The oil droplets generated in the Laskin nozzle are

mixed with the main air stream upstream of the coalescing filter assembly in the mixing

chamber and the aerosol is challenged to the filter.

a Laskin nozzle b
Main line

Mixing chamber
Laskin nozzle line

Figure 5.4 (a) Laskin nozzle (b) Mixing chamber.

55 
 
Heater

Figure 5.5 Heater.

The heating assembly (Refer Figure 5.5) can be used to run the experiments at

elevated temperature. In current research work all the experiments are carried out at room

temperature.

5.1.3 Filter holder

The effect of gravitational force on the performance of the filter media was

studied by changing the flow orientation. The filter geometries were tested in horizontal

as well as vertical flow orientation.

5.1.3.1 Horizontal flow orientation filter holder

The filter holder assembly in horizontal orientation and its parts are shown in

Figure 5.6. The horizontal orientation filter holder is made up of stainless steel so that the

experiment can be carried out at elevated temperature. The filter holder is equipped with

a spacer; hence the filter samples of different thicknesses can be tested. A stainless steel

wire mesh is placed behind the filter medium to protect the medium from deforming due
56 
 
to air flow as well as to maintain parallel air flow geometry throughout the medium. A

threaded aluminum plug is fixed at the end to keep the filter medium in its position. The

liquid collected in the filter holder is drained using drainage ports at various locations in

the filter holder. The drainage is collected through these ports at regular intervals of time

during the experiment and all the oil drained is measured at the end of the experiment.

The filter holder assembly without a filter was tested to understand particle loss

and change in pressure drop. The filter holder without a filter indicated atmospheric

pressure drop and similar upstream and downstream droplet size distribution. The loss in

the droplets was not detectable.

a b
Spacer Filter holder

Aluminum plug

Drain ports

Drain ports

Figure 5.6 Horizontal flow orientation filter holder (a) Filter holder parts (b) Filter holder

assembly.

5.1.3.2 Vertical flow orientation filter holder

The vertical flow orientation filter holder assembly is shown in Figure 5.7. The

filter holder is developed to orient the flow in the direction of gravity hence the flow is
57 
 
diverted twice in the filter holder assembly. This filter holder assembly is made up of

plexiglass so that the drainage and carrying over of big droplets in downstream can be

visualized. The cylindrical shaped filters are fixed in the filter holder and the holder is

attached to the filter holder assembly. The drained liquid is collected in oil collecting dish

and the downstream is vented. The Filter holder assembly in vertical flow orientation

without a filter was tested to detect changes in pressure drop and droplet concentration.

The filter holder indicated less than 1% loss in droplet size distribution which is

negligibly small and negligible pressure drop for the experimental flow rates.

a b
Upstream Downstream

Filter holder
c

Filter holder
Filter holder Wire mesh
Oil collecting dish

Figure 5.7 Filter holder assembly in vertical orientation (a) Filter holder assembly (b)

Filter holder fixed to the filter holder assembly (c) Filter holder and the wire mesh.

58 
 
5.1.4 Measurement equipments

An electronic pressure gauge (Figure 5.8 (a)) is used to measure the pressure drop

across the filter media. Pressure drop is continuously monitored and recorded after every

10 minutes during the experiment. The photometer (Air Techniques International Model

No TDA-2G) is used for in-situ monitoring of the concentration of droplets in the air

stream is shown in Figure 5.8 (b). It can measure the upstream or downstream aerosol

concentration when connected to the appropriate sampling ports. During the experiments

the photometer readings are recorded after every 10 minutes and once the photometer

readings are stable for about one hour it is considered as the filter has reached steady

state. During the experiment normally the pressure drop reaches to steady state in about

one hour from the beginning of the test while the photometer indicates steady

downstream droplet concentration in about two hours from the beginning of the test.

Hence all the tests are run for at least three hours to obtain steady state performance of

the filter media.

The air flow is measured by using a rotameter. The downstream air passes

through the HEPA filter and then is vented. The air flow rate is maintained 5.66 m3/hr at

standard conditions (200 SCFH) at 0oC and 1 atm pressure to obtain the highest face

velocity of 0.6 m/s.

59 
 
a b
Photometer

Pressure transducer

Figure 5.8 (a) Pressure gauge (b) Photometer.

a b

Rotameter
HEPA filter

Figure 5.9 (a) Rotameter (b) Downstream HEPA filter.

A Scanning Mobility Particle SizerTM (SMPS) (TSI Inc. Model No. 3080) is used

to monitor the upstream and downstream droplet size distributions. The SMPS is setup to

complete one droplet size distribution measurement in 135 seconds and it can count up to

1 million droplets at the same time. To use the equipment below its threshold limit the

60 
 
sample stream is diluted with air at a known dilution ratio using rotameters. The TSI

3080 has three major components known as Electrostatic Classifier, Differential Mobility

Analyzer (DMA) and Condensation Particle Counter (CPC). The intake flow rate of

Electrostatic Classifier is termed as sheath flow rate while the sample air flow rate is

termed as the aerosol flow rate. The sheath flow rate is always 1/10 times the aerosol

flow rate. The electrostatic Classifier normalizes the charge on the droplets, if any, by

using a krypton-85 bipolar charger. The charged droplets then enter the DMA. The

droplets are segregated according to their electrical mobility. The DMA passes one size

of droplets at one time to the CPC depending on their electrical mobility. The CPC

counts the number of droplets of a particular size that are sent from the DMA. The CPC

uses N- Butanol to condense on the droplets to increase their size to a size that can be

detected by a laser. The air stream from CPC is further vented.

Flowmeter

Electrostatic Classifier

Differential Mobility analyzer


Condensation Particle Counter

Pump

Figure 5.10 SMPS and CPC.

61 
 
CHAPTER VI

EXPERIMENTAL RESULTS OF GLASS FIBER FILTER GEOMETRIES

6.1 Polypropylene woven drainage channels

The coalescence filtration set up is described in Chapter V. All of the filter media

are tested in the coalescence filtration setup. The experimental results for the filter

geometries developed by inserting polypropylene woven mesh of 500 µm pore opening

and 600 µm filament thickness as well as “No drainage” filter media in both horizontal

and vertical orientation are expressed in terms of capture efficiency, pressure drop,

quality factor and saturation. The filter geometry is plotted against capture efficiency,

pressure drop, quality factor and saturation with no-drainage filter media being the

experimental control. The filter performance is characterized by the combined

performance of pressure drop and separation efficiency. The filter performance is

characterized by a figure of merit (quality factor) which accounts for capture efficiency

and pressure drop [29]. A higher quality factor indicates better filter performance. The

quality factor (QF) is defined by

C
- ln Cout
in
QF= (6.1)
∆P

where Cout and Cin are the outlet and inlet droplet concentrations respectively and ∆ is

the pressure drop across the filter. The capture efficiency is given by

C C
η (6.2)
C

62 
 
where the inlet and outlet (upstream and downstream) concentrations are calculated from

the particle size distribution

∑     (6.3)

where C is the total mass concentration, Ni is the number of droplets of diameter di, and

ρOil is the liquid mass density. The number of droplets in the inlet and outlet is obtained

by using the SMPS. The upstream and downstream droplet distribution for the no-

drainage channel filter media is shown in Figure 6.1.

160000

140000

120000
No. of Droplets (#/cm^3)

100000

80000

60000

40000

20000

0
0 100 200 300 400 500 600 700 800 900 1000
-20000
Droplet Diameter (nanometer)

Cout Cin

Figure 6.1 Upstream and downstream droplet size distribution of no-drainage channel

filter media.

63 
 
The pressure drop and upstream and downstream droplet size distribution are

monitored and recorded regularly during the experiment. The pressure drop and

downstream concentration profile of no-drainage channel filter media are shown in

Figure 6.2. The pressure drop and concentration both increases exponentially and then

becomes steady indicating the filter has reached steady state. The upstream and

downstream oil concentration is calculated by using equation 6.3.

12 3.5E-08

10 3E-08

2.5E-08
Pressure Drop (kPa)

Cout (gm/cm^3)
2E-08
6
1.5E-08
4
1E-08

2 5E-09

0 0
0 20 40 60 80 100 120 140 160 180 200
Time (minutes)

Pressure drop Cout

Figure 6.2 Pressure and downstream concentration profile of no-drainage channel filter

media.

6.1.1 Capture efficiency

Filter geometries are plotted against capture efficiency as shown in Figure 6.3.

All filter media were constructed in triplicate and the experimental results show the

64 
 
steady state values averaged over the three samples. The error bars in the plots indicate

one standard deviation of the three averaged points.

98.31 98.62 98.72 99.48


100.00
93.28 93.45
87.87 88.92
90.00

80.00

70.00
63.90
61.23
59.01 60.52 57.32
60.00 56.29
Efficiency (%)

53.26
50.00 45.81
38.57 37.59
40.00 36.20 35.06
29.00 27.45
30.00

20.00

10.00

0.00
H-N H-I H-M H-O H-B . H-0 H-15 H-30 H-40 H-45 H-50 H-60 V-N V-I V-M V-O V-B . V-0 V-15 V-30 V-45 V-60 .
Filter geometry

H-HORIZONTAL ORIENTATION H-HORIZONTAL ORIENTATION V-VERTICAL ORIENTATION V-VERTICAL ORIENTATION


H-N : No drainage channel H-0 : Drainage at 00 (three cut) V-N : No drainage channel V-0 : Drainage at 00 (three cut)
H-I : Inlet drainage channel H-15 : Drainage at 150 (three cut) V-I : Inlet drainage channel V-15 : Drainage at 150 (three cut)
H-M : Middle drainage channel H-30 : Drainage at 300 (three cut) V-M : Middle drainage channel V-30 : Drainage at 300 (three cut)
H-O : Outlet drainage channel H-40 : Drainage at 400 (three cut) V-O : Outlet drainage channel V-45 : Drainage at 450 (three cut)
H-B : Drainage channel at Both ends H-45 : Drainage at 450 (three cut) V-B : Drainage channel at Both ends V-60 : Drainage at 600 (three cut)
H-50 : Drainage at 500 (three cut)
H-60 : Drainage at 600 (three cut)

Figure 6.3 Capture efficiency of filter geometries with polypropylene woven drainage

channels.

The plots are divided in two sections. The first section is horizontal flow

orientation and the second section is vertical flow orientation. Each of these are further

divided into two subsections. Horizontal orientation data with and without drainage

channels in the order of no-drainage, inlet-drainage, middle-drainage, outlet-drainage and

both-end-drainage are plotted in first subsection. Inclined-angle-drainage channel

experiments in horizontal orientation are plotted in second subsection in the order of 00,

150, 300, 400, 450, 500, and 600. The 00 horizontal orientation case study filter media is

equipped with three equally spaced drainage channels in the direction of flow. The next

subsection shows vertical orientation data with and without drainage channels in the same

65 
 
order as the horizontal orientation experimental results while the fourth subsection has

the experimental results of inclined angle experiments in vertical orientation in the order

of 00, 150, 300, 450, and 600.

In the horizontal orientation when the drainage channel is placed at the inlet,

outlet, and both ends, the capture efficiency is about 98 %. For the inclined angle

experiments the 450 media had capture efficiency of 93 % but for the other inclined angle

media the efficiencies were in the 30 to 60% range. The drainage channel experiments in

horizontal orientation had capture efficiencies of 87 % or greater. The capture efficiency

of experiments without any filter but two drainage channels was only 30 %. These results

show that the filter media modified with drainage channels had higher capture

efficiencies than media without drainage channels and higher efficiency than drainage

channels alone. In the vertical orientation the outlet-drainage indicates capture efficiency

of 93%. The improved efficiency with drainage channels may be due to more fiber

surface area being exposed due to the reduced saturation (decreased latter) plus the large

drop may fill the pore forcing the gas flow through the smaller pores.

The capture efficiency of the middle-drainage channel is slightly higher as

compared to the no-drainage channel experiments. It could be because of insufficient

filter bed length on the either side of the drainage channel to cause coalescence as

relatively greater concentration of smaller droplets was observed in downstream particle

size distribution. More likely in the middle of the medium the saturation is normally

lower than at the boundaries and insertion of the drainage channel at this location does

not significantly change the saturation.

66 
 
6.1.2 Pressure drop

The pressure drop of the all the drainage channel incorporated filter media as well

as no drainage channel filter media were recorded. Filter geometries are plotted against

pressure drop as shown in Figure 6.4. This plot is obtained by plotting pressure drop

against filter geometries in similar manner as Figure 6.3. All filter media were

constructed in triplicate and the experimental results show the steady state values

averaged over the three samples. The error bars in the plots indicate one standard

deviation of the three averaged points.

12.00

10.59 10.6010.48
9.79 9.79 10.00
10.00 9.64
9.19

7.87
8.00
Pressure Drop (KPa)

6.00
4.92 4.85
4.61 4.68 4.73 4.53 4.68 4.44
4.11
4.00 3.66 3.80

2.83 2.69

2.00

0.00
H-N H-I H-M H-O H-B . H-0 H-15 H-30 H-40 H-45 H-50 H-60 V-N V-I V-M V-O V-B . V-0 V-15 V-30 V-45 V-60 .

Filter geometry

H-HORIZONTAL ORIENTATION
H-HORIZONTAL ORIENTATION H-0 : Drainage at 00 (three cut) V-VERTICAL ORIENTATION
H-15 : Drainage at 150 (three cut) V-VERTICAL ORIENTATION V-0 : Drainage at 00 (three cut)
H-N : No drainage channel V-N : No drainage channel
H-I : Inlet drainage channel H-30 : Drainage at 300 (three cut) V-15 : Drainage at 150 (three cut)
H-40 : Drainage at 400 (three cut) V-I : Inlet drainage channel V-30 : Drainage at 300 (three cut)
H-M : Middle drainage channel V-M : Middle drainage channel
H-O : Outlet drainage channel H-45 : Drainage at 450 (three cut) V-45 : Drainage at 450 (three cut)
H-50 : Drainage at 500 (three cut) V-O : Outlet drainage channel V-60 : Drainage at 600 (three cut)
H-B : Drainage channel at Both ends V-B : Drainage channel at Both ends
H-60 : Drainage at 600 (three cut)

Figure 6.4 Pressure drop of filter geometries with polypropylene woven drainage

channels.

67 
 
In the horizontal orientation when the drainage channel is placed at the inlet,

outlet, and both ends, the pressure drop of the filter media slightly increases. An increase

in pressure drop usually accompanies an increase in the gas-liquid separation efficiency

of a filter [22]. The inclined angle experiments in horizontal orientation indicate

significantly low pressure drop. These filter geometries indicate half pressure drop as

compared to the no drainage channel filter media. These filter geometries have three

equally spaced woven drainage channels. The drainage channels are porous and open

structures and have 10 times higher air permeability as compared to the glass fiber filter

media which makes the filter geometry very porous structure as compared to the no

drainage channels filter media. These drainage channels are embedded at different

downward angles which aids in draining the oil faster from the filter media and hence

these filter media indicate significantly lower pressure drops.

In the vertical orientation the no drainage channel as well as the outlet drainage

indicate higher pressure drop as compared to the no drainage channel filter media in

horizontal orientation, these could be because of the improved capture efficiency of the

filter. The inclined angle experiments in vertical orientation indicate lower pressure drop

as compared to the no drainage channel filter media in horizontal orientation but these

filters indicate lower capture efficiencies.

6.1.3 Quality factor

Quality factor of the all the drainage channel incorporated filter media as well as

no drainage channel filter media was calculated by using Eq. 6.1 and are plotted against

filter geometry as shown in Figure 6.5. This plot is plotted in similar manner as Figure

6.3.

68 
 
0.90

0.80 0.77

0.70
Quality Factor(KPa-1)

0.60
0.48
0.50 0.45

0.40 0.36 0.36


0.30 0.3 0.28
0.30 0.26
0.23 0.22
0.18 0.19 0.20
0.20 0.17 0.16 0.16
0.13
0.09 0.09
0.10
0.02 0.03
0.00
H-N H-I H-M H-O H-B . H-0 H-15 H-30 H-40 H-45 H-50 H-60 V-N V-I V-M V-O V-B . V-0 V-15 V-30 V-45 V-60
.
Filter geometry

H-HORIZONTAL ORIENTATION H-HORIZONTAL ORIENTATION V-VERTICAL ORIENTATION V-VERTICAL ORIENTATION


H-N : No drainage channel H-0 : Drainage at 00 (three cut) V-N : No drainage channel V-0 : Drainage at 00 (three cut)
H-I : Inlet drainage channel H-15 : Drainage at 150 (three cut) V-I : Inlet drainage channel V-15 : Drainage at 150 (three cut)
H-M : Middle drainage channel H-30 : Drainage at 300 (three cut) V-M : Middle drainage channel V-30 : Drainage at 300 (three cut)
H-O : Outlet drainage channel H-40 : Drainage at 400 (three cut) V-O : Outlet drainage channel V-45 : Drainage at 450 (three cut)
H-B : Drainage channel at Both ends H-45 : Drainage at 450 (three cut) V-B : Drainage channel at Both ends V-60 : Drainage at 600 (three cut)
H-50 : Drainage at 500 (three cut)
H-60 : Drainage at 600 (three cut)

Figure 6.5 Quality factor of filter geometries with polypropylene woven drainage

channels.

For horizontal orientation, inlet-drainage, outlet-drainage and both-end-drainage

indicate improved quality factor as compared to the no drainage channel filter media. The

middle drainage channel filter media in horizontal orientation indicate slightly higher

quality factor as compared to the no drainage channel filter media in horizontal

orientation. The best performance was observed for the 450 downward inclined angle

experiments in horizontal orientation in terms of quality factor, saturation and drainage.

The capture efficiency of the 450 inclined angle samples was not the highest among all

the case studies but the pressure drop of these experiments were significantly lower

compared to the no-drainage, inlet-drainage, middle-drainage and both-end-drainage

channel samples in horizontal and vertical orientation. The reduced pressure drop results

in the 450 inclined angle media having the highest quality factor.

69 
 
Vertical orientation no drainage filter media indicate significantly high quality

factor and lower saturation as compared to the no drainage channel filter media in

horizontal orientation which proves that gravitational force helps in removing the oil

droplets from the filter media.

6.1.4 Saturation

Liquid saturation of the filter medium is the volume fraction of the pores filled by

the liquid phase. It directly affects the pressure drop and local gas velocity within the

medium. Saturation of filter media is calculated by using following formula.

  (6.4)

Average saturations of all the filter geometries as well as no-drainage channel

filter media in horizontal and vertical orientation are plotted in Figure 6.6. This plot is

plotted in similar manner as Figure 6.3. All the filter geometries in the horizontal as well

as vertical orientation indicate lower saturation and higher drainage as compared to the

experimental control. Inlet, outlet and both-end drainage channel filter media in

horizontal orientation indicate higher capture efficiencies, higher drainage and lower

saturation than the experimental control. The 450 downward inclined angle filter

geometry indicate lowest saturation and hence the filter media indicate significantly low

pressure drop. This filter geometry indicates capture efficiency of 93 %. It indicates that

the drainage channels incorporated at 450 downward inclination helps in draining the

bigger coalesced drops faster from the filter media. Hence this filter media has lower

saturation and the filter can remain porous for a longer time with high capture efficiency

which increases filter life. The both end filter media in the horizontal orientation indicate

70 
 
significantly lower saturation and higher quality factor as compared to the experimental

control.

0.4
0.37
0.36
0.35 0.34 0.34 0.34
0.32
0.31
0.3 0.3
0.3 0.29
0.28
0.27 0.27

0.25 0.24 0.24


0.23 0.23
0.22
0.21
Saturation

0.2
0.2
0.17

0.15
0.12

0.1

0.05

0
H-N H-I H-M H-O H-B . H-0 H-15 H-30 H-40 H-45 H-50 H-60 V-N V-I V-M V-O V-B . V-0 V-15 V-30 V-45 V-60 .
Filter Geometry

H-HORIZONTAL ORIENTATION H-HORIZONTAL ORIENTATION V-VERTICAL ORIENTATION


H-0 : Drainage at 00 (three cut) V-VERTICAL ORIENTATION
H-N : No drainage channel V-N : No drainage channel V-0 : Drainage at 00 (three cut)
H-I : Inlet drainage channel H-15 : Drainage at 150 (three cut) V-15 : Drainage at 150 (three cut)
H-30 : Drainage at 300 (three cut) V-I : Inlet drainage channel
H-M : Middle drainage channel V-M : Middle drainage channel V-30 : Drainage at 300 (three cut)
H-O : Outlet drainage channel H-40 : Drainage at 400 (three cut) V-45 : Drainage at 450 (three cut)
H-45 : Drainage at 450 (three cut) V-O : Outlet drainage channel
H-B : Drainage channel at Both ends V-B : Drainage channel at Both ends V-60 : Drainage at 600 (three cut)
H-50 : Drainage at 500 (three cut)
H-60 : Drainage at 600 (three cut)

Figure 6.6 Saturation of filter geometries with polypropylene woven drainage channels.

The no-drainage channel filter media in vertical orientation indicate lower

saturation and higher drainage with very high capture efficiency. While inlet and both-

end drainage channel filter geometries in vertical orientation indicate lower capture

efficiencies, drainage and relatively higher saturation. Outlet drainage in vertical

orientation indicates 93 % capture efficiency but also has higher pressure drop than the

control. These filter geometries indicate low quality factor and high saturation. The

inclined angle filter geometries in vertical orientation indicate lower capture efficiency,

lower drainage but higher saturations which indicate these filter geometries do not drain

71 
 
the droplets effectively from the filter media. Hence the filter media gets clogged with the

droplets faster and hence indicate lower capture efficiencies.

To understand the effect of the change of filter orientation and drainage structures

on the performance of the filter media qualitatively, the filters were dissected. It was

observed that the drainage channel equipped filter media had a more uniform oil

distribution with less oil concentrated at the bottom of the filter media (i.e., the lowest

point in the direction of gravity) while the filters without drainage channels had higher

liquid saturations at the bottom of the media plus the saturation was higher at the inlet

and outlet surfaces compared to the middle of the media (Refer Figure 6.7). Filter media

equipped with three equally spaced drainage channels at different angles had more or less

uniform distribution of liquid with a very small region of high saturation at the bottom.

This is likely the reason for the lower pressure drop of these filter media.

a b

Figure 6.7 Tested filter media (a) no-drainage channel filter media in horizontal

orientation (b) Both-end-drainage filter media in horizontal orientation. The green

colored oil is collected near the bottom edge of the filter medium.

72 
 
When the filter media are oriented for horizontal flow, the liquid drainage must

move through the filter and down the exit surface of the filter before it collects at the

bottom of the filter holder. The presence of this liquid reduces the space available for gas

flow and thus causes an increase in pressure drop. Furthermore, filters with horizontal

flow are often observed to have a small part of the filter at the bottom to be saturated with

liquid, through which no gas can flow. This restriction causes the gas to move through a

reduced area of the filter and hence with a greater face velocity. The capture mechanisms

are sensitive to the face velocity and the capture efficiencies may decrease. The presence

of the drainage channels of the filter medium provide greater pore openings to allow

faster drainage and can reduce some of the flow restriction. The filters with flow oriented

vertically, on the other hand, do not have the problem of liquid filling the pores at the exit

surface as the liquid drains from the filter. When the liquid reaches the exit the surface of

the filter the liquid drips from the filter and the net result is more space available for gas

flow.

The plot in Figure 6.8 shows there is a strong connection between saturation and

quality factor regardless of the filter orientation and drainage channel geometric structure.

The 450 downward inclination filter geometry in horizontal orientation indicate highest

quality factor and lowest saturation. The no-drainage channel filter media in vertical

orientation indicate second best quality factor and hence slightly higher saturation as

compared to the 450 downward inclination filter geometry in horizontal orientation. The

experimental results indicate that performance of the filter media can be enhanced by

reducing the saturation by developing drainage channel incorporated filter geometries.

73 
 
0.90

450
0.80

0.70
Quality Factor (1/kPa)

0.60
No-drainage
0.50

0.40

0.30

0.20

0.10

0.00
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
Saturation

Horizontal Horizontal-angle Vertical Vertical-angle

Figure 6.8 Quality factor versus average saturation of all filter geometries with

polypropylene woven drainage channels.

6.1.4.1 Saturation profile of no-drainage channel filter media

The saturation profile along the filter media bed length for no-drainage channel

filter media was obtained. The schematic of the process is shown in Figure 6.9. The glass

fiber filter media was constructed using the vacuum molding process. The filter media

was characterized to obtain porosity, air permeability and strength. The filter media was

cut in five thin slices of same thickness by using the filter media cutter saw. The mass of

every slice was recorded and these slices were assembled and tested in the coalescence

74 
 
filtration setup. The weight of the filter slices was recorded after the experiment and the

saturation was calculated using Eq. 6.4.

Glass fiber filter Thin slices of filter media Assemble layers

Measure mass

Gas
outlet
Gas
inlet

Measure mass
Filter test Calculate saturation
 

Figure 6.9 Schematic of experimental measurement of saturation profile of no-drainage

filter media in horizontal orientation.

75 
 
Experimental result of saturation profile of three filter media with relatively

similar porosity, air permeability and strength are shown in Figure 6.10. Experimental

measurements show the saturation in a medium of uniform properties is higher near the

inlet and outlet surfaces and lower in the interior of the medium. This is because at the

inlet the drops on the fibers are smaller and tend to move slower than larger drops in the

interior. At the exit surface capillary forces hold the liquid to the fibers and the liquid

drains down the outside of a horizontally oriented filter, causing the saturation to be

higher at the exit surface. Hence when the drainage channels are incorporated at inlet and

outlet, the filter media indicated lower saturation and higher quality factor and drainage.

Hence the filter performance and filter life can be enhanced by incorporating drainage

channels in the coalescing filter media.

0.7

0.6

0.5
Saturation

0.4

0.3

0.2

0.1

0
0.0028 0.0056 0.0084 0.0112 0.014

Length (meter)

Figure 6.10 Saturation profile of no-drainage channel filter media in horizontal

orientation.

76 
 
6.1.5. Performance comparison of filter geometries developed with woven drainage

channels

Filter media performance is quantified in terms of quality factor. The

improvement in quality factor is quantified by the relative quality factor (RQF) given by

  (6.5)

where the control medium is the horizontal no-drainage channel filter media.

Performance comparison in terms of RQF of the no-drainage filter media in

horizontal and vertical orientation is shown in Figure 6.11. When the air flow is in the

direction of gravity the filter media indicate a RQF of 2.67 and lower saturation which

means that the gravitational force helps in removing the oil droplets faster from the filter

medium.

3
2.67
Relative Quality Factor (RQF)

2.5

1.5
1
1

0.5

0
Horizontal Vertical
Filter Geometry (No-drainage)

Figure 6.11 Relative quality factor of no-drainage filter media in horizontal and vertical

orientation.

77 
 
RQF of all the filter geometries and no-drainage channel filter media both in

horizontal orientation and vertical orientation are shown in Figure 6.12. The horizontal

no-drainage channel filter media is the experimental control.

4.50
4.28

4.00

3.50
Relative Quality Factor (RQF)

3.00
2.61
2.46
2.50

1.95 1.95
2.00
1.63 1.63
1.52
1.50 1.42
1.23 1.16
1.09
1.00 1.01
1.00 0.91 0.84 0.89
0.73
0.49 0.50
0.50
0.13 0.16
0.00
H-N H-I H-M H-O H-B . H-0 H-15 H-30 H-40 H-45 H-50 H-60 V-N V-I V-M V-O V-B . V-0 V-15 V-30 V-45 V-60 .
Filter Geometry

H-HORIZONTAL ORIENTATION V-VERTICAL ORIENTATION V-VERTICAL ORIENTATION


H-HORIZONTAL ORIENTATION V-0 : Drainage at 00 (three cut)
H-0 : Drainage at 00 (three cut) V-N : No drainage channel
H-N : No drainage channel V-15 : Drainage at 150 (three cut)
H-15 : Drainage at 150 (three cut) V-I : Inlet drainage channel
H-I : Inlet drainage channel V-30 : Drainage at 300 (three cut)
H-30 : Drainage at 300 (three cut) V-M : Middle drainage channel
H-M : Middle drainage channel V-45 : Drainage at 450 (three cut)
H-40 : Drainage at 400 (three cut) V-O : Outlet drainage channel
H-O : Outlet drainage channel V-60 : Drainage at 600 (three cut)
H-45 : Drainage at 450 (three cut) V-B : Drainage channel at Both ends
H-B : Drainage channel at Both ends
H-50 : Drainage at 500 (three cut)
H-60 : Drainage at 600 (three cut)

Figure 6.12 Relative Quality Factor of filter geometry with polypropylene woven

drainage channels in horizontal and vertical orientation.

RQF of all the filter geometries are plotted in Figure 6.12. This plot is plotted in

similar manner as Figure 6.3. Inlet, outlet, middle and both-end drainage filter geometries

in the horizontal orientation have increased RQF. Filter media equipped with three

equally spaced drainage channels in horizontal orientation at angles of 150, 300, 400, 450,

500, and 600 degrees had increased RQF with the 450 inclined angle media having the

78 
 
greatest RQF and lowest saturation. Outlet-drainage and both-end drainage in vertical

orientation indicate improved RQF. These filter geometries also indicate lower saturation.

The inclined angle experiments in vertical orientation did not show improvement

in RQF.

6.1.6 Factors affecting coalescence filtration

Drainage channel incorporated filter geometries indicate improved performance.

All the parameters which can affect the performance of the filter geometries are listed in

Table 6.1. The highlighted cells are the important parameters of filter media design which

are explored to study their effect on the performance of the geometries.

79 
 
Table 6.1 Factors affecting coalescence filtration.

Parameter Controllable
Fiber Size Not controlled
Surface properties Not controlled
Filter Vacuum molding pressure Controlled/ Measured
Preparation Amount of binder and starch Controlled/ Measured
method
pH of slurry Controlled/ Measured
Filter Surface area Controlled/ Measured
Bed length (thickness) Controlled/ Measured
Porosity Not Controlled/ Measured
Permeability Not Controlled/ Measured
Pore size distribution Not Controlled/ Measured
Strength Not Controlled/ Measured
Surface energy Not Controlled/ Measured
Number of drainage channels Controlled/Measured
Drainage layer Thickness Controlled/ Measured
Woven/Nonwoven structure Controlled
Surface energy Not Controlled/ Measured
Pore opening Not Controlled/ Measured
Filament thickness Not Controlled/ Measured
Porosity Not Controlled/ Measured
Permeability Not Controlled/ Measured
Strength Not Controlled/ Not measured
Experimental Gas flow rate Controlled/ Measured
Parameter Particle size distribution Not Controlled/ Measured
Aerosol generator Not Varied
Humidity Not Controlled/Measured
Temperature Not Controlled/Measured
 

80 
 
6.1.7: Effect of number of drainage channels in inclined angle filter media and upward

inclination

Number of drainage channels in the inclined angle filter media has impact on the

performance of the filter geometry. 450 downward inclination filter geometry with three

equally spaced polypropylene woven drainage channels (500 µm pore opening and 600

µm filament thickness) in horizontal orientation indicated highest quality factor. Filter

geometries with one and two drainage channel at 450 downward inclination were

developed and tested in the horizontal orientation in coalescence filtration setup. The

experimental results in terms of capture efficiency, pressure drop, quality factor and

saturation are shown in Figure 6.13, 6.14, 6.15 and 6.16 respectively.

94 93.28
93
92 91.23
91
Efficiency (%)

90 89.12
89
88
87
86
1 2 3
No of drainage channels

Filter geometry: 45 degree downward inclination

Figure 6.13 Effect of number of channels on capture efficiency of 450 downward

inclination filter geometry in horizontal orientation.

81 
 
12
9.89
10

Pressure drop (kPa) 8


6.23
6
4.53
4

0
1 2 3
No of drainage channels

Filter geometry: 45 degree downward inclination

Figure 6.14 Effect of number of Channels on pressure drop of 450 downward inclination

filter geometry in horizontal orientation.

0.9
0.77
0.8
Quality factor (1/kPa)

0.7 0.61
0.6
0.47
0.5
0.4
0.3
0.2
0.1
0
1 2 3
No of drainage channels
Filter geometry: 45 degree downward inclination

Figure 6.15 Effect of number of channels on quality factor of 450 downward inclination

filter geometry in horizontal orientation.

82 
 
0.18 0.16
0.16
0.14
0.14 0.12
Saturation 0.12
0.1
0.08
0.06
0.04
0.02
0
1 2 3
No of drainage channels

Filter geometry: 45 degree downward inclination

Figure 6.16 Effect of number of channels on saturation of 450 downward inclination filter

geometry in horizontal orientation.

The experimental results indicate that as the number of drainage channels

increases the filter geometry indicate higher capture efficiency, lower pressure drop,

lower saturation hence higher quality factor. The drainage channels are porous and have

lower surface energy as compared to the filter media. As the droplets reach the drainage

channels they bead up and are easily removed from the open drainage channel structure.

Hence as the number of drainage channels increases the porosity of the filter media

increases which helps in lowering the pressure drops. Drainage channel incorporated

filter media indicate higher drainage hence the filter can remain porous for longer time

and can capture more number of droplets. Hence as the number of drainage channels

increases the quality factor and drainage of the media increases while the saturation

decreases. The filter media having three equally spaced woven polypropylene drainage

channels indicate highest quality factor and can have higher filter life.

83 
 
To study the effect of gravitational force on the performance of the filter media,

filter geometries were developed by adding three equally spaced polypropylene woven

drainage channels at 450 upward inclination (Refer Figure 6.17). The experimental results

in terms of capture efficiency, pressure drop, quality factor and saturation are shown in

Figure 6.18, 6.19, 6.20, and 6.21 respectively.

Air flow
450

450
Gravity
450

Figure 6.17 Filter geometry developed with three equally spaced polypropylene woven

drainage channel at 450 upward inclination.

84 
 
90
87.87
88
Efficiency (%) 86
84
82 81.12
80
78
76
No drainage 45 degree upward

Filter Geometry

No Drainage 45 Degree Upward

Figure 6.18 Capture efficiency of three drainage channels at 45 degree upward inclination

and experimental control.

18
15.61
16
Pressure Drop (kPa)

14
12
9.64
10
8
6
4
2
0
No drainage 45 degree upward

Filter Geometry

No Drainage 45 Degree Upward

Figure 6.19 Pressure drop of three drainage channels at 45 degree upward inclination and

experimental control.

85 
 
0.2 0.18
0.18

Quality Factor (1/kPa)


0.16
0.14 0.12
0.12
0.1
0.08
0.06
0.04
0.02
0
No drainage 45 degree upward
Filter Geometry

No Drainage 45 Degree Upward

Figure 6.20 Quality factor of three drainage channels at 45 degree upward inclination and

experimental control.

0.42 0.41
0.41
0.4
Saturation

0.39
0.38 0.37
0.37
0.36
0.35
0.34
No drainage 45 degree upward

Filter Geometry

No Drainage 45 Degree Upward

Figure 6.21 Saturation of three drainage channels at 45 degree upward inclination and

experimental control.

86 
 
Capture efficiency of the filter geometry with three equally spaced drainage

channels at 450 upward inclination is less than the no drainage channel filter media. This

filter geometry also indicates higher pressure drop and higher saturation, hence a lower

quality factor. It could be because the droplets tend to take the path of lower resistance

i.e. they try to move through the drainage channels and the drainage channels in upward

direction are forcing the droplets to move against the gravitational force and hence this

filter geometry indicates lower drainage and higher saturation and hence a lower quality

factor.

6.1.8. Effect of pore size and filament thickness of drainage channel

The pore sizes and filament sizes of the mesh in the drainage channels can affect

the drainage performance. To study the effect of pore opening and thickness of drainage

channel, polypropylene meshes of 210 µm pore opening and 308 µm filament thickness

as well as 1000 µm pore opening and 1024 µm filament thickness were tested in the filter

geometry with the 450 inclined angle media and with drainage channels at both ends with

the media in horizontal orientation. The experimental results are given in Figures 6.22,

6.23, 6.24, and 6.25 respectively. Every data point on the graph is the average of three

filters and error bar indicated one standard deviation.

The results show all of the performances were very similar, indicating that a wider

range in pore sizes should be explored in future work. It also indicates that the drainage

layer performance may not be very sensitive to the mesh size in this pore size range but

may be more sensitive to the surface properties of the mesh such as its contact angle or

surface tension.

87 
 
100
98.23 98.53 97.83
98

Efficiency (%)
96
94 93.38
92.48 92.83
92
90
88
210 μm 500 μm 1000 μm 210 μm 500 μm 1000 μm
(H-B) (H-45)
Filter Geometry
H-HORIZONTAL ORIENTATION

H-B: Drainage channel at Both ends H-45: Drainage at 450 (three cut)

Figure 6.22 Capture efficiency of filter media equipped with woven drainage channel of

different pore openings.

14
11.96
12 10.48 10.38
Pressure Drop (kPa)

10
8 6.47
6 4.68 5.25
4
2
0
210 μm 500 μm 1000 μm 210 μm 500 μm 1000 μm
(H-B) (H-45)
Filter Geometry

H-HORIZONTAL ORIENTATION

H-B: Drainage channel at Both ends H-45: Drainage at 450 (three cut)

Figure 6.23 Pressure drop of filter media equipped with woven drainage channel of

different pore openings.

88 
 
0.9
0.8 0.74 0.77 0.75

Quality Factor (1/kPa)


0.7
0.6
0.5 0.42
0.37 0.4
0.4
0.3
0.2
0.1
0
210 μm 500 μm 1000 μm 210 μm 500 μm 1000 μm
(H-B) (H-45)
Filter Geometry
H-HORIZONTAL ORIENTATION

H-B: Drainage channel at Both ends H-45: Drainage at 450 (three cut)

Figure 6.24 Quality factor of filter media equipped with woven drainage channel of

different pore openings.

0.35
0.29
0.3
0.24
0.25 0.22
Saturation

0.2 0.16
0.15 0.12
0.1
0.1
0.05
0
210 μm 500 μm 1000 μm 210 μm 500 μm 1000 μm
(H-B) (H-45)
Filter Geometry

H-HORIZONTAL ORIENTATION

H-B: Drainage channel at Both ends H-45: Drainage at 450 (three cut)

Figure 6.25 Saturation of filter media equipped with woven drainage channel of different

pore openings.

89 
 
6.1.9 Filer bed length

The filter geometry developed with three equally spaced polypropylene woven

drainage channels at 450 downward inclination filter media were dissected after testing

and the filter media indicated dry fiber patches in region as shown in Figure 6.26. Hence

it seems that the flow is going through the drainage channels. Hence the filter bed length

was reduced to half so that the flow can pass straight through the filter media. The

experimental results in terms of capture efficiency, pressure drop, quality factor and

saturation are given in Figure 6.27, 6.28, 6.29 and 6.30 respectively. The experimental

results of the filter geometry compared to the results of no-drainage channel filter media

of same filter bed length.

Dry fibers

Air flow

Figure 6.26 Flow pattern in the filter geometry with three equally spaced drainage

channels at 450 downward inclination.

90 
 
87.00
86.50 86.28
86.00
85.50
Efficiency (%)
85.00
84.50
84.00 83.66
83.50
83.00
82.50
82.00
H-N H-45
Filter Geometry

H-HORIZONTAL ORIENTATION

H-N: No drainage channel H-45: Drainage at 450 (three cut)

Figure 6.27 Capture efficiency of filter geometries with reduced filter bed length.

8.00
7.00 6.48
6.00
Pressure drop (kPa)

5.00
4.00
2.80
3.00
2.00
1.00
0.00
H-N H-45
Filter Geometry

H-HORIZONTAL ORIENTATION

H-N: No drainage channel H-45: Drainage at 450 (three cut)

Figure 6.28 Pressure drop of filter geometries with reduced filter bed length.

91 
 
0.70
0.60
0.60

Quality factor (1/kPa)


0.50

0.40

0.30 0.28

0.20

0.10

0.00
H-N H-45
Filter Geometry

H-HORIZONTAL ORIENTATION

H-N: No drainage channel H-45: Drainage at 450 (three cut)

Figure 6.29 Quality factor of filter geometries with reduced filter bed length.

0.35
0.29
0.30

0.25
Sauration

0.20

0.15
0.09
0.10

0.05

0.00
H-N H-45
Filter Geometry

H-HORIZONTAL ORIENTATION

H-N: No drainage channel H-45: Drainage at 450 (three cut)

Figure 6.30 Saturation of filter geometries with reduced filter bed length.

Experimental results indicate that when the filter bed length is reduced to half the

capture efficiency of no drainage as well as 450 downward inclination filter geometry

92 
 
decreases. The pressure drop of both the filter geometries also decreases specifically the

450 downward inclination filter geometry indicate significantly low pressure drop and

saturation as compared to the experimental control. Hence this filter geometry has

significantly high quality factor irrespective of its moderate capture efficiency. These

filters indicate very less dry fiber region as compared to the filters which had double the

filter bed length.

6.2 Nonwoven drainage channels

Polypropyelene nonwoven fabric commonly known as polypropylene spunbond is

used as a drainage channel to study the effect of nonwoven drainage channels on the

performance of filter media. Spunbond nonwovens are fabrics produced by depositing

extruded, spun filaments onto a collecting belt in a uniform but random manner followed

by bonding the fibers [84, 85]. The fibers are separated during the web laying process by

air jets or electrostatic charges. Bonding is formed by applying heated rolls or hot needles

to partially melt the polymer and fuse the fibers together [84, 85]. Spunbond fabric is

nonwoven in nature and characterized by its basis weight. Basis weight of spunbond

indicate amount of fibers present in per square meter of fabric and the unit of basis

weight is grams per square meter (gsm). The nonwoven fabric has non uniform size of

pores. The mean and maximum pore sizes of different basis weight polypropylene

spunbond fabric are given in table 6.2 [86]. Polypropylene woven mesh has uniform pore

size of 500 µm and filament thickness of 600 µm while the filament thickness of

polypropylene spunbond is 50 µm. All the filter geometries were developed by

incorporating nonwoven spunbond of 17 gsm basis weight. The effect of basis weight of

drainage channels on the performance of the filter geometries was also studied. The

93 
 
performance of different filter geometries developed by incorporating polypropylene

spunbond is expressed in terms of capture efficiency, pressure drop, quality factor and

saturation.

6.2.1 Capture efficiency

Filter geometries are plotted against capture efficiency as shown in Figure 6.31.

This plot is obtained by plotting the filter geometries versus capture efficiency in similar

manner as Figure 6.3. All filter media were constructed in triplicate and the experimental

results show the steady state values averaged over the three samples. The error bars in

the plots indicate one standard deviation of the three averaged points.

98.29 98.53 98.63 99.48


100.00
93.16 93.03
90.00 87.87 88.14

80.00

70.00

57.79
Efficiency (%)

60.00 56.15
53.30 51.72
48.33
50.00
40.23 41.67
40.00 35.89
32.64 32.22 31.55 32.89
30.00 26.61 25.16

20.00

10.00

0.00
H-N H-I H-M H-O H-B . H-0 H-15 H-30 H-40 H-45 H-50 H-60 V-N V-I V-M V-O V-B . V-0 V-15 V-30 V-45 V-60 .
Filter Geometry

H-HORIZONTAL ORIENTATION H-HORIZONTAL ORIENTATION V-VERTICAL ORIENTATION


V-VERTICAL ORIENTATION
H-N : No drainage channel H-0 : Drainage at 00 (three cut) V-0 : Drainage at 00 (three cut)
V-N : No drainage channel
H-I : Inlet drainage channel H-15 : Drainage at 150 (three cut) V-15 : Drainage at 150 (three cut)
V-I : Inlet drainage channel
H-M : Middle drainage channel H-30 : Drainage at 300 (three cut) V-30 : Drainage at 300 (three cut)
V-M : Middle drainage channel
H-O : Outlet drainage channel H-40 : Drainage at 400 (three cut) V-45 : Drainage at 450 (three cut)
V-O : Outlet drainage channel
H-B : Drainage channel at Both ends H-45 : Drainage at 450 (three cut) V-60 : Drainage at 600 (three cut)
V-B : Drainage channel at Both ends
H-50 : Drainage at 500 (three cut)
H-60 : Drainage at 600 (three cut)

Figure 6.31 Capture efficiency of filter geometries with polypropylene nonwoven

drainage channels.

In the horizontal orientation when the drainage channel is placed at the inlet,

outlet and both-end-drainage the capture efficiency is about 98 %. The capture efficiency

94 
 
of the middle-drainage channel is slightly higher as compared to the no-drainage channel

experiments. For the inclined angle experiments the 450 media had capture efficiency of

93 % but for the other inclined angle media the efficiencies were in the 30 to 60% range.

The capture efficiency of experiments without any filter but two nonwoven drainage

channels was only 27 %. In the vertical orientation the outlet-drainage indicated a capture

efficiency of 93%.

These results show that the filter media modified with nonwoven drainage

channels had higher capture efficiencies than media without drainage channels and higher

efficiency than drainage channels alone. These experiments indicate similar trend as the

filter geometries developed with the woven polypropylene drainage channels.

6.2.2 Pressure drop

Pressure drop of the all the drainage channel incorporated filter media as well as

no-drainage channel filter media were recorded. Filter geometries are plotted against

pressure drop as shown in Figure 6.32. This plot is obtained by plotting pressure drop

against filter geometries in similar manner as Figure 6.3. All filter media were

constructed in triplicate and the experimental results show the steady state values

averaged over the three samples. The error bars in the plots indicate one standard

deviation of the three averaged points.

In the horizontal orientation when the drainage channel is placed at the inlet,

outlet, and both ends, the pressure drop of the filter media slightly increases might be due

to improved capture efficiency of these filters. The inclined angle filter geometries

indicate significantly low pressure drop as compared to the no-drainage channel filter

95 
 
media. These filter geometries have three equally spaced nonwoven drainage channels.

The drainage channels are porous and open structures and have 10 times higher air

permeability as compared to the glass fiber filter media which makes the filter geometry

very porous structure as compared to the no drainage channels filter media. These

drainage channels are embedded at different downward angles which aids in draining the

oil faster from the filter media and hence these filter media indicate significantly lower

pressure drops.

12.00
10.85 10.8810.66
10.36
10.08
10.00 9.64 9.79 9.65
9.36

8.00
Pressure Drop (KPa)

7.11 6.99
6.78 6.89
6.59 6.39
6.33
6.08 6.10
6.00 5.63
5.32 5.12
5.03

4.00

2.00

0.00
H-N H-I H-M H-O H-B . H-0 H-15 H-30 H-40 H-45 H-50 H-60 V-N V-I V-M V-O V-B . V-0 V-15 V-30 V-45 V-60 .

Filter Geometry

V-VERTICAL ORIENTATION
H-HORIZONTAL ORIENTATION H-HORIZONTAL ORIENTATION V-VERTICAL ORIENTATION V-0 : Drainage at 00 (three cut)
H-N : No drainage channel H-0 : Drainage at 00 (three cut) V-N : No drainage channel V-15 : Drainage at 150 (three cut)
H-I : Inlet drainage channel H-15 : Drainage at 150 (three cut) V-I : Inlet drainage channel V-30 : Drainage at 300 (three cut)
H-M : Middle drainage channel H-30 : Drainage at 300 (three cut) V-M : Middle drainage channel V-45 : Drainage at 450 (three cut)
H-O : Outlet drainage channel H-40 : Drainage at 400 (three cut) V-O : Outlet drainage channel V-60 : Drainage at 600 (three cut)
H-B : Drainage channel at Both ends H-45 : Drainage at 450 (three cut) V-B : Drainage channel at Both ends
H-50 : Drainage at 500 (three cut)
H-60 : Drainage at 600 (three cut)

Figure 6.32 Pressure drop of filter geometries with polypropylene nonwoven drainage

channels.

The inclined angle experiments in horizontal orientation developed with

nonwoven drainage channel indicate significantly low pressure drop as compared to the

no-drainage channel filter media. But these filter geometries indicate slightly higher

96 
 
pressure drop as compared to the inclined angle filter geometries of woven drainage

channels.

In the vertical orientation the outlet drainage indicate higher pressure drop as

compared to the no drainage channel filter media in horizontal orientation, it could be

because of the improved capture efficiency of the filter. The inclined angle experiments

in vertical orientation indicate lower pressure drop as compared to the no drainage

channel filter media in horizontal orientation but these filters indicate lower capture

efficiencies. The filter geometries developed with nonwoven drainage channels in

horizontal as well as vertical orientation indicate similar trend as that of the woven

drainage channel filter media. The nonwoven drainage channel incorporated filter

geometries indicate higher saturation and higher pressure drop than the woven drainage

channel incorporated filter geometries.

6.2.3 Quality factor

Quality factor of the all the drainage channel incorporated filter media as well as

no drainage channel filter media are plotted against filter geometry as shown in Figure

6.33. This plot is obtained by plotting quality factor against filter geometries in similar

manner as Figure 6.3.

97 
 
0.90

0.80 0.76

0.70

Quality Factor(1/kPa) 0.60

0.50 0.48
0.42
0.40 0.35 0.35
0.29 0.28 0.26
0.30 0.25
0.22 0.21
0.18 0.19 0.19
0.20 0.15 0.15
0.12 0.13
0.10 0.07 0.05
0.02 0.02
0.00
H-N H-I H-M H-O H-B . H-0 H-15 H-30 H-40 H-45 H-50 H-60 V-N V-I V-M V-O V-B . V-0 V-15 V-30 V-45 V-60 .

Filter Geometry

H-HORIZONTAL ORIENTATION V-VERTICAL ORIENTATION V-VERTICAL ORIENTATION


H-HORIZONTAL ORIENTATION H-0 : Drainage at 00 (three cut) V-0 : Drainage at 00 (three cut)
H-N : No drainage channel V-N : No drainage channel
H-15 : Drainage at 150 (three cut) V-I : Inlet drainage channel V-15 : Drainage at 150 (three cut)
H-I : Inlet drainage channel H-30 : Drainage at 300 (three cut) V-30 : Drainage at 300 (three cut)
H-M : Middle drainage channel V-M : Middle drainage channel
H-40 : Drainage at 400 (three cut) V-O : Outlet drainage channel V-45 : Drainage at 450 (three cut)
H-O : Outlet drainage channel H-45 : Drainage at 450 (three cut) V-60 : Drainage at 600 (three cut)
H-B : Drainage channel at Both ends V-B : Drainage channel at Both ends
H-50 : Drainage at 500 (three cut)
H-60 : Drainage at 600 (three cut)

Figure 6.33 Quality factor of filter geometries with polypropylene nonwoven drainage

channels.

For horizontal orientation, inlet-drainage, outlet-drainage and both-end-drainage

indicate improved quality factor as compared to the no drainage channel filter media. The

middle drainage channel filter media in horizontal orientation indicate slightly higher

quality factor as compared to the no-drainage channel filter media in horizontal

orientation. The best performance was observed for the 450 downward inclined angle

experiments in horizontal orientation in terms of quality factor, saturation and drainage.

The capture efficiency of the 450 inclined angle samples was not the highest among all

the case studies but the pressure drop of these experiments were significantly lower

compared to the no-drainage, inlet-drainage, middle-drainage and both-end-drainage

channel samples in horizontal and vertical orientation. The reduced pressure drop results

98 
 
in the 450 inclined angle media having the highest quality factor. The filter geometries

developed with nonwoven drainage channel indicate slightly lower quality factor as

compared to the respective filter geometries developed by incorporating woven drainage

channels. This is because the nonwoven drainage channel incorporated filter geometries

indicate higher pressure drop and higher saturation corresponding to their woven

counterparts.

6.2.4 Saturation

Average saturation of all the filter geometries in horizontal and vertical

orientation is calculated using Eq. 6.4. Filter geometries are plotted against the saturation

in Figure 6.34 in the similar manner as Figure 6.3.

0.4
0.37 0.37
0.36
0.35 0.35
0.35 0.34
0.33
0.32 0.32
0.3 0.3 0.3 0.3
0.3 0.29
0.28
0.26 0.26
0.25 0.25
0.25 0.24
Saturation

0.2
0.17

0.15 0.14

0.1

0.05

0
H-N H-I H-M H-O H-B . H-0 H-15 H-30 H-40 H-45 H-50 H-60 V-N V-I V-M V-O V-B . V-0 V-15 V-30 V-45 V-60 .
Filter Geometry

H-HORIZONTAL ORIENTATION V-VERTICAL ORIENTATION V-VERTICAL ORIENTATION


H-HORIZONTAL ORIENTATION
H-0 : Drainage at 00 (three cut) V-N : No drainage channel V-0 : Drainage at 00 (three cut)
H-N : No drainage channel
H-15 : Drainage at 150 (three cut) V-I : Inlet drainage channel V-15 : Drainage at 150 (three cut)
H-I : Inlet drainage channel
H-30 : Drainage at 300 (three cut) V-M : Middle drainage channel V-30 : Drainage at 300 (three cut)
H-M : Middle drainage channel
H-40 : Drainage at 400 (three cut) V-O : Outlet drainage channel V-45 : Drainage at 450 (three cut)
H-O : Outlet drainage channel
H-45 : Drainage at 450 (three cut) V-B : Drainage channel at Both ends V-60 : Drainage at 600 (three cut)
H-B : Drainage channel at Both ends
H-50 : Drainage at 500 (three cut)
H-60 : Drainage at 600 (three cut)

Figure 6.34 Saturation of filter geometries with polypropylene nonwoven drainage

channels.

99 
 
All the filter geometries in the horizontal as well as vertical orientation indicate

lower saturation and higher drainage as compared to the experimental control. Inlet,

outlet and both-end drainage channel filter media in horizontal orientation indicate higher

capture efficiencies, higher drainage and lower saturation than the experimental control.

The both-end-drainage channel filter media in the horizontal orientation indicate

significantly lower saturation and higher quality factor as compared to the experimental

control. The 450 downward inclined angle filter geometry indicate lowest saturation and

hence the filter media indicate significantly lower pressure drop. This filter geometry

indicates capture efficiency of 93 %. It indicates that the drainage channels incorporated

at 450 downward inclination help in draining the bigger coalesced drops faster from the

filter media. Hence this filter media has lower saturation and the filter can remain porous

for longer time with high capture efficiency which increases filter life. As shown in

Figure 6.10, in the middle of the medium the saturation is normally lower than at the

boundaries and insertion of the drainage channel at this location does not significantly

change the saturation.

Inlet and both-end drainage channel filter geometries in vertical orientation

indicate lower capture efficiencies, drainage and relatively higher saturation. The inclined

angle filter geometries in vertical orientation also indicate lower capture efficiency, lower

drainage but higher saturations which indicate these filter geometries do not drain the

droplets effectively from the filter media. Hence the filter media gets loaded with the

droplets faster and becomes less porous. So there are less pores remaining in the filter to

capture droplets and hence the filter media indicate lower capture efficiencies.

100 
 
Nonwoven drainage channel incorporated filter geometries indicate higher

saturation, hence higher pressure drop and lower quality factor. The possible reason could

be these drainage channels have smaller pores and a pore size distribution as compared to

the woven drainage channels. Hence some of the bigger drops might not be draining

efficiently due to the small pore sizes of the drainage channels. Hence these filters

indicate slightly higher average saturation as compared to the average saturation of the

filter geometries developed by incorporating woven drainage channels.

Figure 6.35 indicate a strong connection between the quality factor and saturation.

Higher quality factor correspond to lower saturation. The 450 media indicate lowest

saturation and highest quality factor. The second best performance is indicated by the no-

drainage channel filter media in vertical orientation which indicate higher saturation than

the 450 media in horizontal orientation. To understand the effect of drainage structures on

the performance of the filter media qualitatively, these filters were dissected. It was

observed that the drainage channel equipped filter media had a more uniform oil

distribution with less oil concentrated at the bottom of the filter media similar to the filter

geometries developed by incorporating woven drainage channels. The filter geometries

developed with nonwoven drainage channels indicate similar trends in terms of filter

media performance as indicated by the woven drainage channel incorporated filter

geometries. The only difference is these filter geometries indicate higher pressure drop

and higher saturation hence slightly lower quality factor compared to respective woven

drainage channel incorporated filter geometries. These experimental results indicate that

to improve the performance of the filter media the woven drainage channels are effective

than the nonwoven drainage channels because of their uniform and bigger pore sizes.

101 
 
0.80 450

0.70
Quality Factor (1/kPa)
0.60
No-drainage
0.50

0.40

0.30

0.20

0.10

0.00
0 0.05 0.1 0.15 0.2 0.25 0.3 0.35 0.4
Saturation

Horizontal Horizontal-angle Vertical Vertical-angle

Figure 6.35 Quality factor versus average saturation of all filter geometries with

polypropylene nonwoven drainage channels.

6.2.5 Effect of basis weight of drainage channels

Basis weight of the spunbond fabric is related to its pore size. As the basis weight

of the spunbond fabric increases the pore sizes decreases because of more material

addition in the same area. Effect of change in pore sizes of drainage channel is studied by

using different basis weight drainage channels. The different basis weight drainage

channels used are 17, 34, 51 and 68 gsm of spunbond fabric. The basis weight of

polypropylene spunbond fabric and their mean and maximum pore sizes are shown in

table 6.2. Both-end drainage and 450 downward inclination filter geometries were

developed by incorporating the drainage channels of varying basis weight and tested in

horizontal orientation. The performance of the filter geometries is plotted in terms of

102 
 
capture efficiency, pressure drop, quality factor and saturation as shown in Figure 6.36,

6.37, 6.38 and 6.39 respectively.

Table 6.2 Pore sizes of polypropylene spunbond fabric [88].

Sr. No. Basis weigh of polypropylene Pore size (µm)

spunbond fabric (gsm) Mean pore size Maximum pore size

1. 17 135 230

2. 34 110 190

3. 51 60 100

4. 68 30 60

100.00
98.63 98.58 98.50 98.47
98.00

96.00
Efficiency (%)

94.00 93.16
92.11 92.01
92.00 91.18

90.00

88.00

86.00
17 34 51 68 17 34 51 68
(H-B) (H-45)
Filter Geometry

H-HORIZONTAL ORIENTATION

H-B: Drainage channel at Both ends H-45: Drainage at 450 (three cut)

Figure 6.36 Capture efficiency of filter geometries for varying basis weight of

polypropylene spunbond fabric.

103 
 
14.00

12.00 10.98 11.28


10.66 10.76

Pressure Drop (1/kPa) 10.00

8.00
6.89
6.11
6.00 5.32 5.67

4.00

2.00

0.00
17 34 51 68 17 34 51 68
(H-B) (H-45)
Filter Geometry

H-HORIZONTAL ORIENTATION

H-B: Drainage channel at Both ends H-45: Drainage at 450 (three cut)

Figure 6.37 Pressure drop of filter geometries for varying basis weight of polypropylene

spunbond fabric.

0.90

0.80 0.76
0.73 0.71
0.70 0.68
Quality Factor (1/kPa)

0.60

0.50
0.42 0.41 0.41
0.40 0.37

0.30

0.20

0.10

0.00
17 34 51 68 17 34 51 68
(H-B) (H-45)
Filter Geometry
H-HORIZONTAL ORIENTATION

H-B: Drainage channel at Both ends H-45: Drainage at 450 (three cut)

Figure 6.38 Quality factor of filter geometries for varying basis weight of polypropylene

spunbond fabric.

104 
 
0.35 0.33
0.31
0.30 0.29
0.26
0.25
0.21
Saturation

0.20 0.18
0.16
0.15 0.14

0.10

0.05

0.00
17 34 51 68 17 34 51 68
(H-B) (H-45)
Filter Geometry

H-HORIZONTAL ORIENTATION

H-B: Drainage channel at Both ends H-45: Drainage at 450 (three cut)

Figure 6.39 Saturation of filter geometries for varying basis weight of polypropylene

spunbond fabric.

Experimental results indicate that as the basis weight of the drainage channel

increases the capture efficiency of the media decreases. As the basis weight of the

drainage channel increases the drainage channel becomes a tighter structure with small

sizes of pores which makes the filter geometry less effective in draining the oil drops.

Hence the filter media loads up with the liquid droplets and it has fewer pores available to

capture new droplets. Hence the filter geometries indicate lowered capture efficiency. As

the filter media loads with the oil it indicates higher pressure drop, higher saturation and

hence lowered quality factor. Hence a drainage channel needs to be an open and porous

structure for effective drainage and higher filter performance and filter life.

105 
 
6.2.6 Performance comparison of filter geometries developed with woven and nonwoven

drainage channels

The improvement in quality factor is quantified by the relative quality factor

(RQF) and is calculated by using Eq. 6.5. RQF of all the filter geometries developed by

incorporating nonwoven drainage channels and no-drainage channel filter media both in

horizontal orientation and vertical orientation are shown in Figure 6.40. The horizontal

no-drainage channel filter media is the experimental control. This plot is plotted in

similar manner as Figure 6.3.

All of the filter geometries in the horizontal orientation with drainage channels

had increased RQF compared to the control. Filter media equipped with three equally

spaced drainage channels in horizontal orientation at angles of 150, 300, 400, 450, 500, and

600 had increased relative quality factor with the 450 inclined angle media having the

greatest RQF and lowest saturation. Outlet-drainage and both-end drainage in vertical

orientation indicate improved RQF. These filter geometries also indicate lower saturation.

The inclined angle experiments in vertical orientation did not show improvement in

relative quality factor.

106 
 
4.50
4.22

4.00

3.50
Relative Quality Fcator (RQF)
3.00
2.61
2.50 2.28

2.00 1.90 1.90


1.57 1.52
1.50 1.36 1.41
1.19 1.11
1.00 1.01 1.01
1.00 0.83 0.81
0.65 0.73

0.50 0.38
0.28
0.08 0.11
0.00
H-N H-I H-M H-O H-B . H-0 H-15 H-30 H-40 H-45 H-50 H-60 V-N V-I V-M V-O V-B . V-0 V-15 V-30 V-45 V-60 .

Filter Geometry

H-HORIZONTAL ORIENTATION V-VERTICAL ORIENTATION


V-VERTICAL ORIENTATION
H-HORIZONTAL ORIENTATION H-0 : Drainage at 00 (three cut) V-0 : Drainage at 00 (three cut)
V-N : No drainage channel
H-N : No drainage channel H-15 : Drainage at 150 (three cut) V-15 : Drainage at 150 (three cut)
V-I : Inlet drainage channel
H-I : Inlet drainage channel H-30 : Drainage at 300 (three cut) V-30 : Drainage at 300 (three cut)
V-M : Middle drainage channel
H-M : Middle drainage channel H-40 : Drainage at 400 (three cut) V-45 : Drainage at 450 (three cut)
V-O : Outlet drainage channel
H-O : Outlet drainage channel H-45 : Drainage at 450 (three cut) V-60 : Drainage at 600 (three cut)
V-B : Drainage channel at Both ends
H-B : Drainage channel at Both ends H-50 : Drainage at 500 (three cut)
H-60 : Drainage at 600 (three cut)

Figure 6.40 Relative Quality Factor of filter geometry with polypropylene nonwoven

drainage channels in both horizontal and vertical orientation.

Performance comparison of filter geometries developed with woven and

nonwoven drainage channels is indicated in terms of RQF as shown in Figure 6.41. The

nonwoven drainage channel incorporated filter geometries indicate improved quality

factor as compared to the experimental control but indicated lower quality factor as

compared to the respective filter geometries developed by incorporating woven drainage

channels. The possible reason behind this could be the smaller pores and non-uniform

pore size distribution is not efficient as compared to the bigger size pores as well as a

regular structure of the woven drainage channels resulting in similar size pores. Hence

the nonwoven drainage channel incorporated filter geometries indicate higher saturation

hence higher pressure drop and lower quality factor compared to woven drainage channel

107 
 
incorporated filter geometries. Hence these filters will have shorter filter life as compared

to the woven drainage channel incorporated filter geometries. But the nonwoven

spunbond fabric is inexpensive as compared to the woven meshes. Hence developing

filter geometries by incorporating nonwoven drainage channels is a cost effective way of

improving the filter media performance and filter life.

Polypropylene woven drainage channels Polypropylene nonwoven drainage channels


4.50

4.28
4.22
4.00

3.50

3.00
Relative Quality Fcator (RQF)

2.61
2.46

2.50
2.28
1.95

1.95
1.90

1.90

2.00
1.63

1.63
1.57

1.52
1.52
1.42

1.41
1.36

1.50

1.23
1.19

1.16
1.11
1.09
1.03

1.03
1.01
1.00

0.91

0.89
0.84
0.83

0.81
1.00
0.73

0.73
0.65

0.50
0.49
0.38

0.50

0.28
0.16
0.13

0.11
0.08
0.00
H-N H-I H-M H-O H-B H-0 H-15 H-30 H-40 H-45 H-50 H-60 V-N V-I V-M V-O V-B V-0 V-15 V-30 V-45 V-60
Filter Geometry

H-HORIZONTAL ORIENTATION
V-VERTICAL ORIENTATION V-VERTICAL ORIENTATION
H-HORIZONTAL ORIENTATION H-0 : Drainage at 00 (three cut)
V-N : No drainage channel V-0 : Drainage at 00 (three cut)
H-N : No drainage channel H-15 : Drainage at 150 (three cut)
V-I : Inlet drainage channel V-15 : Drainage at 150 (three cut)
H-I : Inlet drainage channel H-30 : Drainage at 300 (three cut)
V-M : Middle drainage channel V-30 : Drainage at 300 (three cut)
H-M : Middle drainage channel H-40 : Drainage at 400 (three cut)
V-O : Outlet drainage channel V-45 : Drainage at 450 (three cut)
H-O : Outlet drainage channel H-45 : Drainage at 450 (three cut)
V-B : Drainage channel at Both ends V-60 : Drainage at 600 (three cut)
H-B : Drainage channel at Both ends H-50 : Drainage at 500 (three cut)
H-60 : Drainage at 600 (three cut)

Figure 6.41 RQF of woven and nonwoven drainage channel incorporated filter

geometries in horizontal and vertical orientation.

6.3 Effect of surface wettability of drainage channels

Surface wettability is vital in separating the oil droplets from the air streams. The

wettability of the drainage channels can be varied by using low, intermediate and high

108 
 
surface energy materials. Polypropylene is intermediate wetting material; hence to study

the effect of surface wettability of drainage channels on the performance of the filter

geometries, low surface energy, Teflon®, and high surface energy nylon, were used as

drainage channels. Polypropylene woven and nonwoven drainage channel experiments

indicated highest quality factor for 500 µm pore openings meshes. Hence nylon and

Teflon® meshes of 500 µm pore openings and 600 µm filament thickness were used as

woven drainage channels. Along with the woven meshes nylon spunbond fabric of 17

grams per meter basis weight and nylon-6 nanofibers electrospun on Teflon® mesh were

used as nonwoven drainage channels. Filter geometries developed by incorporating

polypropylene spunbond fabric of 17 grams per meter basis weight indicated highest

quality factor among other basis weight experiments. Hence the nylon spunbond fabric of

17 grams per meter basis weight was used as drainage channels. It has larger pores as

compared to glass fiber filter media. The mean pore size of the 17 grams per meter square

basis weight nylon spunbond fabric is 140 µm while the maximum size of the pore is 220

µm [86]. The nylon electrospun nanofiber on Teflon® mesh has smaller pore opening as

compared to the glass fiber filter media.

Glass is hydrophilic and more oleophilic than nylon, polypropylene and Teflon®.

Nylon is hydrophilic and oleophilic while polypropylene is weak hydrophobic and

oleophobic and Teflon® is hydrophobic and oleophobic. Surface energy of the drainage

channels were evaluated by measuring the Sullube® 32 oil’s contact angle on drainage

channels. Figure 6.42 and 6.43 indicate the contact angle images on nylon,

polypropylene and Teflon® plane surfaces as well as drainage channels. Nylon,

polypropylene and Teflon® were melted on glass slides and the oil contact angle was

109 
 
measured on the plane surface. The oil contact angle was also measured directly on the

drainage channel porous surfaces. The oil contact angle of nylon, polypropylene and

Teflon® plane surfaces were 40, 460 and 950 respectively. The oil contact angle on the

polypropylene and Teflon® woven drainage channels were 760 and 1250. The oil spreads

on the nylon mesh hence it was very difficult to measure the contact angle on nylon

mesh. The contact angle on the mesh was increased as compared to the contact angle on

the plane surface it could be because of formation of the composite surface of air and

polymer material. It can be compared to the Cassie-Baxter case. The Cassie-Baxter case

explains the increase in contact angle due to surface roughness and formation of a

composite surface which is usually made up by trapping the air underneath the roughness

or features of the surface [87].

a b

Figure 6.42 Sullube® 32 oil contact angle on plane surfaces (a) Nylon, (b) Polypropylene

and (c) Teflon®.

110 
 
a b

Figure 6.43 Sullube 32® oil contact angle on woven surfaces (a) Polypropylene and (b)

Teflon®.

Performance of all the filter geometries developed with the woven and nonwoven

drainage channels of varying surface energy are expressed in terms of capture efficiency,

pressure drop, quality factor and saturation.

6.3.1 Capture efficiency

All filter media were constructed in triplicate and the experimental results show

the steady state values averaged over the three samples. The error bars in the plots

indicate one standard deviation of the three averaged points. Inlet, outlet, both end and

450 downward inclination in horizontal orientation case studies indicated improved

performance among all the filter geometries developed by incorporating woven and

nonwoven polypropylene drainage channels. Hence these filter geometries were

developed by incorporating woven and nonwoven drainage channels of varying surface

wettability. Drainage channel material is plotted against capture efficiency, pressure drop,

quality factor and saturation as shown in Figures 6.44, 6.45, 6.46 and 6.47 respectively.

The plots are divided into nine sections. The first section, marked “No”, is glass

microfiber media only and does not have a drainage channel, and serves as the
111 
 
experimental control. The second and third sections report results for woven meshes of

nylon and Teflon® respectively. The fourth and fifth sections are nylon + Teflon® and

Teflon® + nylon where both nylon and Teflon® woven mesh of 500 µm pore opening are

used as drainage channel with nylon and Teflon® facing the flow respectively. The sixth

and seventh sections are nylon (spundbond) + Teflon® and Teflon® + nylon (spunbond).

These sections show results for nylon spunbond and Teflon® mesh both used as drainage

channels in which nylon and Teflon® facing the flow respectively. The eighth and ninth

sections are nylon (NF) +Teflon® and Teflon® + nylon (NF) respectively. These sections

indicate the performance of filter media equipped with nylon nanofibers electrospun on

Teflon® woven mesh of 500 µm pore opening, used as drainage channels with the nylon

nanofibers facing the incoming flow or the Teflon® facing the incoming flow

respectively. All the drainage channels are 600 µm thick. All the sections on the plot

except the first section have the filter geometries plotted in the order of “Inlet-Drainage”,

“Outlet-Drainage”, “Both-End-Drainage” and 450 downward inclination with three

equally spaced drainage channels respectively.

112 
 
99.82

99.78
99.75

99.67
99.53
99.57

99.51
99.32

99.24

99.16

99.17
99.07
99.02

99.11
98.82

98.59
98.53

98.38

98.31
98.28
98.23

98.23
97.62
96.48
96.48

95.12
94.24

93.45

93.45
93.23

92.98
92.84
100.00

87.87
90.00

80.00

Efficiency (%) 70.00

60.00

50.00

40.00

30.00

20.00

10.00

0.00

Drainage Channel Material

No Drainage Inlet Drainage Outlet Drainage Both End Drainage 45 Degree Three Cut

Figure 6.44 Capture efficiency of filter geometries developed with varying

surface wettability of drainage channels.

The experimental results indicate that all of the filters with drainage channels had

higher capture efficiency when compared to the microfiber media with no-drainage

channel. It may be that action of the drainage channels to remove liquid from the

medium exposes more fiber for capture of incoming droplets. In the Inlet, Outlet, and

Both End Drainage geometries the capture efficiencies all fall in the 96 to 99 % range,

and within each geometry the performances are within the one-standard deviation error

bars of the data points. The capture efficiencies of the 450 inclined drainage channels

were lower than the other geometries, with capture efficiencies of 93 to 95%, except for

the Teflon® drainage channel that had 99% efficient capture.

113 
 
6.3.2 Pressure drop

Pressure drop of all the filter geometries were recorded and are plotted against the

drainage channel material as shown in Figure 6.45. This plot is plotted in similar manner

as Figure 6.44.

15.93
18.00

14.61
14.56
14.15
16.00

13.56

13.41
13.27
12.81
12.23
12.15
14.00
11.51
11.34

11.25
10.92
Pressure Drop (kPa)

10.59

10.41

10.23
10.07
12.00
9.51
9.89

9.86

9.81
9.64

9.38
9.37
9.27
8.48

10.00

8.14

7.81
6.93

8.00

5.67

5.14
6.00
3.53

4.00
2.00
0.00

Drainage Channel Material

No Drainage Inlet Drainage Outlet Drainage Both End Drainage 45 Degree Three Cut

Figure 6.45 Pressure drop of filter geometries developed with varying surface wettability

of drainage channels.

The pressure drop data show more variability between the different geometries.

For most geometries the Inlet, Outlet, and Both End Drainage designs had higher

pressure drops than the control. The Teflon® and Teflon® + nylon drainage channels had

slightly lower pressure drops than the control. More noteworthy is the pressure drops for

the 450 inclined angle designs are much lower than the control except for the nylon (NF)

114 
 
+ Teflon® and the Teflon® + nylon (NF) drainage channels. The exception may be due to

the small pore sizes associated with the nanofiber layer. The mean and maximum pore

size of the nylon (NF) + Teflon® are 0.5 µm and 5 µm respectively. The general trend is

the drainage channels with larger pores and lower surface energy than the control

medium results in lower pressure drop. The lowest pressure drop was obtained for the

450 inclined drainage channels made of Teflon® woven mesh.

6.3.3 Quality factor

The quality factor of all the geometries are plotted against drainage channel

material as shown in Figure 6.46. This plot is plotted in similar manner as Figure 6.44.

This figure shows all drainage channel incorporated filter geometries indicate increased

quality factor compared to the control. Low surface energy Teflon® woven mesh

incorporated filter media indicate higher quality factor for respective geometries as

compared to the high surface energy nylon or intermediate surface energy polypropylene

incorporated filter geometries. The woven drainage channel incorporated filter

geometries have higher quality factor than the nonwoven drainage channel incorporated

filter geometries. As the pore size of the drainage channel decreases the quality factor

decreases. Hence the nylon (NF) + Teflon® and Teflon® + nylon (NF) drainage channel

incorporated filter geometries indicate low quality factor as compared to the big pore

opening woven and nonwoven drainage channels. The 450 filter geometry indicated best

performance in terms of quality factor and saturation in all the experiments but for high

surface energy and smaller pore opening the 450 filter geometry’s performance is

comparable to the other filter geometries (Refer Figure 6.46). One configuration in

particular stands out among all of the others, the Teflon 450 inclined drainage channels.

115 
 
The Teflon 450 inclined drainage channels had a significantly higher quality factor due to

the high capture efficiency and the low pressure drop. This geometry is developed by

incorporating low surface energy and big pore opening drainage channels.

1.80

1.60
1.60

1.40

1.20
Quality Factor (1/kPa)

1.00
0.71
0.80

0.54
0.52
0.49

0.60

0.47
0.43

0.42
0.41
0.40

0.40

0.40
0.39

0.39
0.38

0.38

0.38

0.37

0.37
0.36

0.36

0.35

0.35
0.35

0.34

0.34

0.34
0.33

0.32
0.30

0.30
0.28
0.40
0.18

0.20

0.00

Drainage Channel Material

No Drainage Inlet Drainage Outlet Drainage Both End Drainage 45 Degree Three Cut

Figure 6.46 Quality factor of filter geometries developed with varying surface wettability

of drainage channels.

These results show that the filter media performance can be enhanced by

incorporating woven low surface energy drainage channels. These drainage channels are

porous and open structures and have low surface energy that enhances liquid drainage

from the filter.

116 
 
6.3.4 Saturation

The average saturation of different filter geometries were calculated by using Eq. 6.4 and

plotted against drainage channel material as shown in Figure 6.47. This plot is plotted in

similar manner as Figure 6.44.


0.37

0.4

0.34
0.34

0.32
0.32
0.35
0.305

0.31
0.31

0.31

0.29
0.29
0.3

0.3

0.3

0.3
0.28

0.28
0.27
0.27

0.27
0.3

0.26
0.25
0.25

0.24
0.25
Saturation

0.2

0.2
0.18

0.18

0.18

0.17
0.2

0.16
0.12
0.15
0.09

0.1

0.05

Drainage Channel Material

No Drainage Inlet Drainage Outlet Drainage Both End Drainage 45 Degree Three Cut

Figure 6.47 Saturation of filter geometries developed with varying surface wettability of

drainage channels.

The saturation data show the saturation levels in all of the filter geometries were

lower than the control. The least saturations occurred for the 450 inclined drainage

channel geometries in each configuration. Comparison of the saturations with the

pressure drops show that in general lower saturation may contribute to a lower pressure

drop but other factors including the drainage channel geometry, surface energies, and

fiber structures are also important.

117 
 
6.3.5 Performance comparison of filter geometries

The improvement in quality factor is quantified by the RQF and is calculated by

using Equation 6.5. RQF of all the filter geometries developed by incorporating woven

and nonwoven drainage channels of varying wettability and no-drainage channel filter

media in horizontal orientation are shown in Figure 6.48. This plot is plotted in similar

manner as Figure 6.44. The horizontal no-drainage channel filter media is the

experimental control. All the filter geometries indicate higher RQF as compared to the

experimental control. The woven mesh incorporated filter geometries indicate higher

RQF as compared to the filter geometries developed with nonwoven drainage channels.

The 450 downward inclination filter geometry indicate higher RQF as compared to other

filter geometries. The Teflon® woven mesh incorporated drainage channels indicate

higher quality factor as compared to the respective filter geometries developed with the

nylon drainage channels. The nylon (NF) + Teflon® indicate lowest quality factor. Nylon

is olephilic and nylon (NF) + Teflon® has smallest pore sizes and being olephilic in

nature the drainage channel has tendency to hold the oil. Hence these filter geometries

indicate higher saturation and lower quality factor. Hence these filters will have lower

filter life as compared to the other filter geometries.

118 
 
10
H-N H-I H-O H-B H-45

8.89
9

8
Relative Quality Factor (RQF)
7

3.94
4

3.00
2.88
2.72

2.61
2.39
3

2.33
2.28
2.22

2.22

2.22
2.17

2.17
2.11

2.11

2.11

2.11

2.06
2.00

2.00

1.94

1.94
1.93

1.89

1.89

1.89
1.83

1.78
1.67

1.67
1.56
2
1

Drainage Channel Material

H-HORIZONTAL ORIENTATION

H-N : No drainage channel H-I : Inlet drainage channel H-O : Outlet drainage channel H-B : Drainage channel at Both ends H-45 : Drainage at 450 (three cut)

Figure 6.48 RQF of filter geometries developed with woven and nonwoven drainage

channels of varying surface wettability.

Experimental results indicate that the filter media equipped with three equally

spaced drainage channels at 450 downward inclination performs better than other inclined

angle experiments. The filter media is 6 cm filter disc with 1.4 cm thickness. For the

inclined angle filter geometries the drainage channels are incorporated in the filter to

keep the filter geometry symmetric and height or the distance between the drainage

channels is decided to get the effect of drainage channels over the entire media height.

119 
 
0.02
0.018
0.016
Height (Meter) 0.014
0.012
0.01
0.008
0.006
0.004
0.002
0
0 10 20 30 40 50 60 70
Angle (Degree)

Figure 6.49 Height between drainage channels for different angles of the inclined angle

experiments.

The height or distance between the drainage channels is plotted against the angle

in Figure 6.49. Figure 6.49 indicate that the height is maximum for the 450 filter

geometry. Based on the current filter geometry designs of all the inclined angle case

studies, it is difficult to explain why the 450 works better than the other angles. Distance

between the drainage channels contributes to the performance of the filter geometry.

Hence distance between the drainage channels needs to be optimized which will also

allow adding more than three drainage channels to develop new filter geometries. The

addition of more than three drainage channels as well as optimizing the distance between

the drainage channels will help in understanding why 450 filter geometry works better

than other inclined angle filter geometries.

120 
 
The current size of the filter media puts restrictions in cutting the filter geometry

at certain distances due to the sharp curved surface. In future, the filter media should be

developed in bigger size which will give better chance to optimize the distance between

the drainage channels.

6.4 Composite filter media

Composite filter media is made up of glass fibers of varying fiber diameter as

discussed in Chapter III. Performance of the composite filter media is qualified in terms

of capture efficiency, pressure drop, quality factor and saturation as shown in Figure

6.50, 6.51, 6.52 and 6.53 respectively.

100.00 94.06
90.00 86.69
80.00 75.90
70.00
Efficiency (%)

60.00
50.00
40.00
30.00
20.00
10.00
0.00
0.006 0.013 0.018
Filter bed length (meter)

Figure 6.50 Capture efficiency of composite filter media.

121 
 
45.00
40.00 37.48
35.00
Pressure Drop (kPa)

29.82
30.00 27.51
25.00
20.00
15.00
10.00
5.00
0.00
0.006 0.013 0.018
Filter bed length (meter)

Figure 6.51 Pressure drop of composite filter media.

0.16
0.14 0.13
0.12
Quality Factor (1/kPa)

0.12
0.10
0.08
0.06 0.05

0.04
0.02
0.00
0.006 0.013 0.018
Filter bed length (meter)

Figure 6.52 Quality factor of composite filter media.

122 
 
0.45 0.42
0.40
0.35
0.35 0.31
0.30
Saturation

0.25
0.20
0.15
0.10
0.05
0.00
0.006 0.013 0.018
Filter bed length (meter)

Figure 6.53 Saturation of composite filter media.

Experimental results indicate that as the filter bed length of the filter media

increases the capture efficiency, pressure drop and saturation increases. Hence the quality

factor of the filter media decreases. Even the glass fibers exhibit different surface

wettability because of their significantly different way synthesis. To study local

saturation qualitatively the filters were dissected and the filter media indicated that the

fibers of fiber diameter 6 µm being most oleophilic in nature indicated the highest

saturation while the 3 µm fiber diameter fibers indicated less amount of oil saturation as

compared to the 6 µm fiber diameter while the 10 µm and 39 µm indicated no oil

saturation. The filter media was designed with 3 µm fibers facing the air flow followed

by 6 µm, 10 µm and 39 µm fibers. The oil saturation was observed in 3 µm and 6 µm

fiber diameter area of the filter while the 10 µm and 39 µm indicated no oil saturation

hence all the filters indicated significantly high pressure drop as the droplets were not

able to move to the exit boundary and drain from the filter. Hence the filter geometry
123 
 
needs to have optimum wettability along with the optimum pore sizes to have efficient oil

drainage from the filter media and to obtain higher performance and higher filter life.

124 
 
CHAPTER VII

NANOFIBER AUGMENTED FILTER GEOMETRIES

7.1 Introduction

Drainage channel incorporated micro glass fiber filter media indicated improved

performance. Filter geometry with three equally spaced drainage channels at 450

downward inclination in horizontal orientation indicated the best performance among all

the filter geometries. These filter geometries indicate significantly low pressure drops

compared to the experimental control with moderate capture efficiencies except the

Teflon® woven drainage channel incorporated filter geometries which indicate 99.75 %

capture efficiency. The capture efficiencies of the nylon and polypropylene woven and

nonwoven drainage channel incorporated filter geometries at 450 downward inclination in

horizontal orientation are in the range of 92-95 % and can be enhanced by adding

nanofibers. By augmenting micro glass fiber media with nanofibers, a wide range of

micron and sub-micron (0.3 to 0.8 micron) droplets can be effectively captured [81]. The

glass fibers, being olephilic and high surface energy material, aids in coalescence hence

high surface energy nanofibers needs to be incorporated in glass fiber media to enhance

the coalescence. Hence nylon nanofibers of varying fiber diameter were incorporated to

enhance the performance of the filter media and the effect of fiber diameter of nanofibers

was studied on the performance of drainage channel incorporated nanofiber augmented

glass fiber filter media. The nanofiber augmented glass fiber filter media making process

125 
 
in discussed in Chapter III. The presence of the nanofibers increases the surface area

available for capture of the droplets but because the diameters are so small the gas phase

“slips” around the fibers and the corresponding increase in pressure drop is not as large as

when micron size fibers are used. The net effect is an increased quality factor with the

use of nanofibers [88-89].

7.2 Electrospinning

Electrospinning is a simple process that uses an electric field to make polymer

nanofibers of submicron and nanometer sizes. In electrospinning, solid fibers of small

diameters are produced as a result of stretching of the electrified jet [90]. Generally, the

suspended drop starts to stretch and form a Taylor cone as a result of an electric field

being applied to the solution [91-93]. A droplet at the tip of the syringe needle, when

electrified, experiences two types of electrostatic forces- the electrostatic repulsion

between the surface charges and the Coulombic force by external electric field [92]. The

distortion of this solution drop is caused by the balance of the repulsive forces induced on

the drop due to the charge distribution and the surface tension of the liquid. When the

voltage reaches a critical value, the electric force overcomes the surface tension and

viscoelastic forces of the deformed drop in the suspended polymer solution formed at the

tip of the syringe [93, 94]. Then a jet can be produced. The electrified jet undergoes

stretching and bending, eventually forming into a continuous thread. After the jet travels

through the air, the polymer fibers are produced by the evaporation of the solvent and are

collected at an electrically grounded target. The surface morphology of the electrospun

fiber is affected by many parameters, such as the polymer concentration, applied voltage,

spinning distance, air friction, gravity, temperature, and ambient parameters [91-99]. In

126 
 
electrospinning, the spinning of fibers is achieved largely by the charges placed on the jet

by the applied field, which move with the jet.

7.2.1 Electrospinning setup

The schematic and lab scale setup of electrospinning process is shown in Figure

7.1 and 7.2 respectively.

The electrospinning setup mainly consists of a syringe, a high voltage power

supply and a collector. The polymer is usually dissolved in suitable solvent. Nanofibers

in the range of 10 to 2000 nm diameter can be achieved by choosing the appropriate

polymer solvent system [89, 100]. The syringe is filled with the polymer solution. A

syringe pump is used to pump the solution at a constant flow rate. The polymer solution

is electrified by applying a high voltage, usually in the range of 1-30 kV. The fibers are

deposited onto a collector which is grounded. The diameter of the deposited fiber can be

controlled through electrospinning parameters like solution viscosity, applied voltage,

distance between the syringe needle tip and collector surface and many others [92].

127 
 
 

Figure 7.1 Schematic of single jet electrospinning.

Teflon® tubing connecting


syringe and needle
Needle
Syringe
High voltage terminal
Syringe pump

Grounded collector
High voltage supply

Figure 7.2 Lab scale set up of single jet electrospinning.

128 
 
7.2.2 NanospiderTM

NanospiderTM is a patented, needle-less, high voltage, free liquid surface

electrospinning process enabling the production of nanofibers on an industrial scale

[101]. NanospiderTM technology allows the production of nanofibers from polymers

dissolved in water, acids or bipolar solvents as well as from melted polymers and is

suitable for the production of organic and inorganic fibers of differing diameters from

nanometer to micrometer [102]. This technology is modular and versatile in nature and

can be easily adapted to optimize specific properties of the produced nanofibers [102].

Nanospider’sTM free liquid surface electrospinning allows higher fiber packing density

and thus an increased productivity as well as better fiber homogeneity and more

consistent web morphology.

7.2.2.1 Principle of NanospiderTM technology

The principle of NanospiderTM is based on the discovery that it is possible to

create multiple Taylor Cones from a thin layer of polymer solution [103]. So, unlike other

methods, NanospiderTM uses a cylinder, not nozzles or capillaries, to form the fibers. The

cylinder is partially immersed in the polymer solution, and, as it rotates, a defined amount

of the polymer solution is carried to the top part of the cylinder where Taylor Cones are

formed and the multiple jets immerging from multiple Taylor Cones produces the

nanofibers. The Taylor Cones are formed next to each other, throughout the entire length

of the cylinder, resulting in the high production capacity of Nanospider’sTM spinning

head. The streams or jets of polymer solutions become solid nanofibers as the solvent

evaporates and the nanofibers are collected on the opposite collecting electrode. It is free

129 
 
liquid surface electrospinning that allows the electric field to optimize the distance

between Taylor cones. NanospiderTM technology allows optimizing the solution

parameters (conductivity, temperature, surface tension, etc.) as well as the equipment

parameters (voltage, electrode distance, etc.). Advantages of this technology are top

quality of nanofibers and homogeneity of fiber web, ability to control the fiber diameters,

its mechanical simplicity and high throughput. Lab scale NanospiderTM equipment is

shown in Figure 7.3. The wire electrode used to electrospin nylon nanofibers is shown in

Figure 7.4. The wire electrodes have equally spaced 12 stainless steel wires of diameter

0.0002 meter. This electrode can be used to make nanofibers which needs acids as

solvents. The solution bath in which the electrode can be placed is made up of

polypropylene hence a solvent which dissolves polypropylene cannot be used as well as

polypropylene nanofibers cannot be produced using the machine.

130 
 
Top electrode
Distance between the electrodes
Bottom electrode immersed in solution bath

Electrode rotation control


Voltage control

Figure 7.3 Lab scale NanospiderTM module.

Solution bath

Wire electrode

Figure 7.4 Wire electrode and solution bath.

131 
 
7.3 Production of nylon nanofibers

Nylon-6 (Sigma-Aldrich) was dissolved in formic acid (Sigma-Aldrich) in various

concentrations by weight and the polymer solution was electrospun using the single jet

electrospinning as well as NanospiderTM. The various solution concentration and the

electrospinning parameters are given in Table 7.1. 6 %, 8%, and 12% by weight solutions

were electrospun using the NanospiderTM. The rotation speed of the spinning head or the

electrode is inversely proportional to the fiber size distribution. All the nanofibers were

fabricated at same and high electrode rotation speed to get narrow fiber size distribution.

20 wt% solution is too viscous to make fibers using the NanospiderTM, hence it was

electrospun using the single jet electrospinning.

Table 7.1 Electrospinning process parameters for producing nylon-6 nanofibers.

Method of Polymer Distance Flow rate Voltage Electrode

fabrication concentration between (milliliters/minute) (kV) rotation

(wt %) electrodes speed

(cm) (Hz)

Single jet 20 20 0.002 20 ---

electrospinning

NanospiderTM 12 20 --- 55 20

8 20 --- 65 20

6 20 --- 80 20

132 
 
a b

c d

Figure 7.5 SEM images of nylon nanofibers (a) 6 wt% (b) 8 wt% (c) 12 wt% (d) 20 wt%.

The fiber diameters of the nanofibers were measured using ImageJ software from

the Scanning Electron Microscope (SEM) images. The SEM images of the fibers are

given in Figure 7.5. Figure 7.5 indicate that as the concentration of the polymer increases

the fiber size increases. 6 wt%, 8 wt%, 12 wt%, and 20 wt% nylon solution can produce

100 nanometer (nm), 150 nm, 300 nm and 600 nm size nylon fibers. The nanofiber

augmented micro glass fiber filter media were developed by adding 100, 300 and 600

nanometer size nanofibers in the glass fibers proportionally by keeping the area ratio of

nanofibers to micro glass fiber as one. The amount of nanofiber added to the glass fiber

133 
 
filter is measured as the ratio of surface area of nanofibers per surface area of

microfibers. All of the nanofiber augmented filter media have an area of 1.0. The SEM

image of nanofiber augmented glass fiber filter media is shown in Figure 7.6. Figure 7.6

indicate a filter media can be made by mixing nylon nanofibers and glass fibers.

Figure 7.6 SEM image of nanofiber augmented glass fiber filter media.

7.4 Experimental results of nanofiber augmented filter media modified with woven

drainage channels of varying surface wettability

7.4.1 Capture efficiency

The coalescence filtration set up is described in Chapter V. All of the filter media

were tested in the coalescence filtration setup. All the experiments were carried out in

134 
 
horizontal flow orientation. The experimental results for the filter geometries developed

by inserting nylon and polypropylene woven mesh of 500 µm pore opening and 600 µm

filament thickness as well as “No drainage” filter media in horizontal orientation are

expressed in terms of capture efficiency, pressure drop, quality factor and saturation.

Filter geometry is plotted against capture efficiency as shown in Figure 7.7. All filter

media were constructed in triplicate and the experimental results show the steady state

values averaged over the three samples. The error bars in the plots indicate one standard

deviation of the three averaged points.

No Drainage 45 Degree Nylon 45 Degree Polypropylene


99.79
100.00 99.44

98.00 97.05
96.55
96.00 95.45 95.44
94.60
94.00 92.84 93.28 93.02
92.23
Efficiency (%)

92.00

90.00
87.87
88.00

86.00

84.00

82.00

80.00
Glassfiber 600 nm Nylon NF 300 nm Nylon NF 100 nm Nylon NF
Filter geometry

Glassfiber : No nanofibers
600 nm Nylon NF : Glassfiber + Nylon nanofiber (Fiber diamter: 600 nanometer)
300 nm Nylon NF : Glassfiber + Nylon nanofiber (Fiber diamter: 300 nanometer)
100 nm Nylon NF : Glassfiber + Nylon nanofiber (Fiber diamter: 100 nanometer)

Figure 7.7 Capture efficiency of nanofiber augmented glass fiber filter media eqipped

with nylon and polypropylene woven drainage channels.

135 
 
The plot is divided in four sections. The first section is only glass fiber filter

media without any nanofibers. Second, third and fourth sections are glass fiber filter

media augmented with nanofibers of fiber diameter 600 nm, 300 nm and 100 nm

respectively. In all the section the filter geometries are plotted in the order of no-drainage,

450 downward inclination with three equally spaced nylon woven drainage channels and

450 downward inclination with three equally spaced polypropylene woven drainage

channels respectively. The no drainage channel glass fiber filter media is the

experimental control.

The no-drainage channel filter media augmented with nanofibers indicate a

capture efficiency of 90-96 % which is improved from 87% for a no nanofiber filter

media. Experimental results indicate that all the filters augmented with oleophilic nylon

nanofibers indicate improved capture efficiency as compared to the experimental control.

As the fiber diameter of the nanofiber decreases, the capture efficiency of the filters

increases, with the 100 nanometer nylon nanofiber incorporated filter media indicating

the highest capture efficiency among all no drainage channel incorporated filter media.

The nylon and polypropylene woven mesh incorporated at 450 downward inclination

filter geometries indicated improved capture efficiency compared to the experimental

control as well as no drainage channel nanofiber augmented filter media. The filter

geometries with polypropylene woven drainage channels at 450 indicate higher capture

efficiency than the nylon woven drainage channel incorporated filter geometries. As the

fiber diameter of the nanofibers decreases, the capture efficiency of the both the nylon

and polypropylene incorporated filter geometries increases. 100 nm nylon nanofiber

augmented filter media equipped with woven polypropylene drainage channels at 450

136 
 
downward inclination indicated highest capture efficiency and the same filter geometry

augmented with nylon woven drainage channels indicated second best capture efficiency.

7.4.2 Pressure drop

Pressure drop of the all the drainage channel incorporated filter media as well as

no drainage channel filter media were recorded. Filter geometries are plotted against

pressure drop as shown in Figure 7.8. This plot is obtained by plotting pressure drop

against filter geometries in similar manner as Figure 7.7. All filter media were

constructed in triplicate and the experimental results show the steady state values

averaged over the three samples. The error bars in the plots indicate one standard

deviation of the three averaged points.

No Drainage 45 Degree Nylon 45 Degree Polyproylene


25.00

20.67
20.00

16.34
Pressure Drop (kPa)

15.00

11.52
10.24
9.64
10.00 8.58 8.49
6.93 7.43 7.09
6.47
4.68
5.00

0.00
Glassfiber 600 nm Nylon NF 300 nm Nylon NF 100 nm Nylon NF
Filter geometry

Glassfiber : No nanofibers
600 nm Nylon NF : Glassfiber + Nylon nanofiber (Fiber diamter: 600 nanometer)
300 nm Nylon NF : Glassfiber + Nylon nanofiber (Fiber diamter: 300 nanometer)
100 nm Nylon NF : Glassfiber + Nylon nanofiber (Fiber diamter: 100 nanometer)

Figure 7.8 Pressure drop of nanofiber augmented glass fiber filter media equipped with

nylon and polypropylene woven drainage channels.

137 
 
All the filter media augmented with nanofibers indicate higher pressure drop than

the experimental control. An increase in pressure drop usually accompanies an increase

in the gas-liquid separation efficiency of a filter [22]. As the fiber size of nylon

nanofibers increases the pressure drop of the filter media increases. The 450 inclined

angle experiments with nanofiber augmented filter media indicated lower pressure drop

as compared to the experimental control. The polypropylene incorporated filter

geometries indicate lower pressure drop and saturation compared to nylon incorporated

filter geometries. But as compared to the no nanofiber filter media with drainage channel

these filter geometries had higher pressure drop.

Polypropylene drainage channels incorporated filter geometries indicated lower

pressure drop as compared to the nylon drainage channel incorporated filter geometries.

Experimental results indicated that as the surface energy of drainage channels decreases

the pressure drop of the filter media decreases. The intermediate surface energy drainage

channels bead up the oil droplets and these drainage channels are very porous and open

structures hence the beaded oil droplets drains very easily. As the surface energy of the

drainage channels increases the droplets spread on the drainage channels which might

block the pores of the drainage channel, hence these filters are less efficient in draining

the oil droplets from the filter geometries.

600 nm size nanofiber augmented filter geometries with drainage channels at 450

indicate lower pressure drop than the 600 nm nanofiber augmented 450 filter geometries.

For nanometer-scale fibers, the effect of slip flow at the fiber surface is taken into

account. This is because the scale of the fiber becomes small enough that the molecular

movements of the air molecules are significant in relation to the size of the fibers and the

138 
 
flow field. The Knudsen number is used to describe the importance of the molecular

movements of air molecules at the fiber surface to the overall flow field. The Knudsen

number can be written as

(7.1)

where “λ” is the gas mean free path (the dimension of the non-continuous nature of the

molecules), and Rf is the mean radius of the fibers. Slip flow generally needs to be

considered when Kn > 0.1. Slip flow definitely needs to be considered when Kn is around

0.25 [11]. For air at standard conditions, the mean free path is 0.066 µm; therefore, for

fibers with diameters smaller than 0.5 µm, slip flow must be considered. In slip flow, the

air velocity at the fiber surface is assumed to be non-zero. Due to the slip at the fiber

surface, drag force on a fiber is smaller than that in the case of non-slip flow, which

translates into lower pressure drop [88]. Hence glass fiber filter media augmented with

100 nm and 300 nm nanofibers indicate moderate increase in pressure drop while the

filter media developed by augmenting 600 nm nanofiber indicate significantly high

pressure drops.

7.4.3 Quality factor

Quality factor of the all the drainage channel incorporated filter media as well as

no drainage channel filter media are calculated using Equation 6.1 and plotted against

filter geometry as shown in Figure 7.9.

139 
 
No Drainage 45 Degree Nylon 45 Degree Polypropylene

1.40
1.24
1.20

1.00
Quality Factor (1/kPa)

0.82
0.77 0.78
0.80
0.69

0.60

0.40 0.41 0.43


0.40 0.37

0.23
0.18 0.19
0.20

0.00
Glassfiber 600 nm Nylon NF 300 nm Nylon NF 100 nm Nylon NF
Filter geometry

Glassfiber : No nanofibers
600 nm Nylon NF : Glassfiber + Nylon nanofiber (Fiber diamter: 600 nanometer)
300 nm Nylon NF : Glassfiber + Nylon nanofiber (Fiber diamter: 300 nanometer)
100 nm Nylon NF : Glassfiber + Nylon nanofiber (Fiber diamter: 100 nanometer)

Figure 7.9 Quality factor of nanofiber augmented glass fiber filter media equipped with

nylon and polypropylene woven drainage channels.

This plot is obtained by plotting quality factor against filter geometries in similar

manner as Figure 7.1. Nanofiber incorporated no drainage channel filter media indicate

improved performance compared to the experimental control. As the fiber size of the

nanofiber decreases the quality factor of the filter media increases. The smaller nanofiber

incorporated filter media indicate significantly higher capture efficiency as compared to

the experimental control but, because of the slip flow condition, the pressure drop of the

filter media increases moderately. The 450 nanofiber augmented filter geometry indicated

significantly higher quality factor as compared to experimental control. As compared to

140 
 
the nylon the polypropylene woven drainage channel incorporated filter media indicate

significantly high quality factor.

7.4.4 Saturation

Saturation of the filter geometries and the no drainage channel filter media are

calculated using Eq. 6.4 and plotted in Figure 7.10. This plot is obtained by plotting

saturation against filter geometries in similar manner as Figure 7.1.

No Drainage 45 Degree Nylon 45 Degree Polypropylene

0.45
0.41
0.40 0.39
0.38
0.37
0.35

0.30
Saturation

0.25
0.22
0.21
0.20 0.19
0.18
0.15
0.15 0.14
0.13
0.12
0.10

0.05

0.00
Glassfiber 600 nm Nylon NF 300 nm Nylon NF 100 nm Nylon NF
Filter geometry

Glassfiber : No nanofibers
600 nm Nylon NF : Glassfiber + Nylon nanofiber (Fiber diamter: 600 nanometer)
300 nm Nylon NF : Glassfiber + Nylon nanofiber (Fiber diamter: 300 nanometer)
100 nm Nylon NF : Glassfiber + Nylon nanofiber (Fiber diamter: 100 nanometer)

Figure 7.10 Saturation of nanofiber augmented glass fiber filter media incorporated with

nylon and polypropylene woven drainage channels.

Nanofiber augmented no drainage channel filter geometries indicate higher

saturation and lower drainage as compared to the experimental control. It is because

141 
 
nylon nanofibers are oleophilic in nature and augmenting the filter media with nanofibers

indicated higher capture efficiency. As the fiber size of nanofiber increases the saturation

of the filter media increases. The same trend is observed for the pressure drop. The nylon

drainage channel incorporated nanofiber augmented filter media indicate significantly

lower saturation as compared to the experimental control.

The polypropylene incorporated filter geometries indicated significantly lower

saturation as compared to the nylon incorporated drainage channels as well as no

drainage channel filter media. The polypropylene incorporated filter media indicate lower

pressure drop as compared to the nylon as well as no drainage channel filter media. To

understand the effect of drainage structures on the performance of the nanofiber

augmented filter media qualitatively, the filters were dissected. The filters without

drainage channels had higher liquid saturations at the bottom of the media plus the

saturation was higher at the inlet and outlet surfaces compared to the middle of the media.

Filter media equipped with three equally spaced drainage channels at 450 downward

angle had more or less uniform distribution of liquid with a very small region of high

saturation at the bottom. This is likely the reason for the lower pressure drop of these

filter media. The higher saturation of the nanofiber augmented filter could be the

oleophilic nature of nanofibers. These filters also indicated higher capture efficiencies

and hence higher pressure drop as compared to the no nanofiber filter media. The plot in

Figure 7.11 shows there is a strong connection between saturation and quality factor

regardless of the drainage channel geometric structure. All the filter geometries equipped

with intermediate surface energy polypropylene incorporated filter geometries indicate

higher quality factor and lower saturation.

142 
 
1.40
450 Polypropylene (100 nm)
1.20

Quality Factor (1/kPa) 1.00

0.80

0.60

0.40

0.20

0.00
0.00 0.05 0.10 0.15 0.20 0.25 0.30 0.35 0.40 0.45

Saturation

No Drainage 45 Degree Nylon 45 Degree Polypropylene

Figure 7.11 Quality factor versus average saturation of all nanofiber augmented filter

geometries with nylon and polypropylene woven drainage channels.

7.4.5 Performance comparison of nanofiber augmented filter geometries

The improvement in quality factor is quantified by the relative quality

factor (RQF) and calculated by using Eq. 6.5 and plotted in Figure 7.12. This plot is

obtained in similar manner as the Figure 7.1. All of the nanofiber augmented no drainage

filter geometries indicated higher RQF as has compared to the experimental control. The

nylon and polypropylene woven drainage channel incorporated nanofiber augmented

drainage channels indicated higher RQF as compared to the experimental control. As the

nanofiber diameter decreases the RQF increases due to the “slip flow” condition. The

polypropylene woven drainage channel incorporated 100 nm nylon nanofiber augmented

filter geometries indicated the highest quality factor.

143 
 
No Drainage 45 Degree Nylon 45 Degree Polypropylene
8.00

6.86
7.00
Relative Quality Factor (RQF)
6.00

5.00 4.55
4.20 4.33
4.00 3.83

3.00
2.22 2.30 2.37
2.07
2.00
1.28
1.00 1.06
1.00

0.00
Glassfiber 600 nm Nylon NF 300 nm Nylon NF 100 nm Nylon NF
Filter Geometry

Glassfiber : No nanofibers
600 nm Nylon NF : Glassfiber + Nylon nanofiber (Fiber diamter: 600 nanometer)
300 nm Nylon NF : Glassfiber + Nylon nanofiber (Fiber diamter: 300 nanometer)
100 nm Nylon NF : Glassfiber + Nylon nanofiber (Fiber diamter: 100 nanometer)

Figure 7.12 Relative Quality Factor of nanofiber augmented filter geometry with nylon

and polypropylene woven drainage channels.

144 
 
CHAPTER VIII

MODELING OF DRAINAGE CHANNEL INCORPORATED FIBROUS FILTER

MEDIA

This chapter focuses on multiphase and volume averaging theories. The

multiphase equations discussed in this chapter are used to model flow through drainage

channel incorporated glass fiber filter media. Pressure and velocity profiles are generated

in the drainage channel incorporated media. Physical space of the media had drainage

channels incorporated at different angles. To smooth the irregularities at the boundaries

of media and drainage channel the physical space is converted to a logical space for the

calculations. The Jacobian transformation is used to transform physical space into logical

space and the pressure profile calculations are done is logical space. The velocity profiles

are developed by using the Darcy’s law calculated from the pressure profile and the

logical space is transformed back to physical space by inverse Jacobian transformation.

This model allows understanding the pressure and velocity profiles of pure air (no

droplets) for the drainage channel incorporated filter media. The pressure and velocity

profile affects the pressure drop of the filter media and ultimately affects the quality

factor of the filter media. The ideal coalescing filter model should account for gas, liquid

flow, capture of droplets and liquid drainage. In the current model the capture of droplets

on fibers, migration of droplets and drainage is not incorporated. We do not have an

adequate model for the movement of captured droplets and drainage of droplets in

145 
 
drainage channel incorporated filter media. Hence this model is limited to determine the

gas flow vectors in an unloaded filter i.e. no capture of droplets. A model which can

predict the velocity and pressure profiles of the drainage channel incorporated filter

media including the capture of droplets, migration and drainage of droplets will be able to

predict the overall performance of a filter. This model is the first step in predicting the

drainage channel incorporated filter media’s performance. Future model including the

capture of droplets, migration and drainage of droplets should build on this model in

cylindrical coordinates for better comparison with the experimental results.

8.1 Multiphase theory in porous media

Multiphase flow deals with the flow of mixture of various phases [104].

Multiphase continuum theory is used to describe a multiphase flow by considering that

the local measurement of a dispersed phase in a multiphase flow is representative of the

averaged effect of all processes at microscale [105]. Volume averaging is a technique

used to develop multiphase continuum equations. In older theories, the interfaces

between phases were considered to be discontinuous in the system which resulted in

inaccurate representation of multiphase systems [106, 107]. It is important that these

interface regions at the microscale be accounted for while developing theories to describe

a system [108]. Knowing the importance of interfaces, the newer multiphase theories

started including interfacial properties for a better representation of a multiphase system.

8.2 Volume average theory

The purpose of volume averaging is to formulate a theoretical structure based on

average values which are equivalent to those that are detected within the field and time

146 
 
scales of a local measurement in a dispersed multiphase material. The balance is

formulated in terms of a phase average, average of a gradient, and average of a time

derivative.

The definition of a locally averaged property must correspond to the value that is

measured locally and must reflect the average of the total amount of the property in the

bulk phases and the interface region. It is implied that the local measurement cannot

isolate the effect of the interface region from that of the bulk phase or detect conditions

within a bulk phase.

The volume averaged for any property for the  phase can be written as follows

[109, 110].


ρ φ + ∇ ⋅ ρα φα v α + ∇ ⋅ i α − ρα f α − ρα gα + ( E α + I α + S α + G α ) = 0 (8.1)
∂t α α

The excess terms are constrained by

∑ ( E α + I α + S α + Gα ) = 0 (8.2)
α

The existing theories for interfacial phenomena are inadequate for describing the

dynamic interactions that occur at an interface. A dynamic model is needed for scaling

up to the continuum scale. The jump balance for a two dimensional surface is developed

by subtracting the balances for the adjoining phases from the balance on a material

volume, containing the surface [111]. The jump balance approach for property for

singular surfaces can be written as follows.

[ρ1φ1 (w − v1 ) − i1 ] ⋅ n12 = [ρ 2φ2 (w − v 2 ) − i 2 ] ⋅ n12 (8.3)

147 
 
Where the velocity of phase 1 and 2 are v1 and v 2 respectively and w is the velocity of

the interface. and are the density of phase 1 and 2 respectively.

8.3 Conservation equations

The dispersed multiphase volume averaging theory provides the starting equations

for a continuum analysis [105]. These equations were simplified for modeling the

drainage channel incorporated clean filter media. The continuum theory provides

balances for the conservation of mass, momentum, mechanical energy, thermal energy,

total energy, chemical species, and entropy for each phase present in the system [106].

For evaluation of pressure drop and velocity profile of a clean filter media in sections

only the mass and momentum balances for the gas phase are needed. Mass and

momentum balances were used to model the interior of media and drainage channels

while at the interface of media and drainage channel the mass and momentum jump

balances were used.

8.3.1 Mass and momentum balaces

The heterogeneous scale gas phase mass balance can be written as follows.

∇.   (E mG + G mG )
G
v 0 (8.4)

Where   is the accumulation of mass per unit volume

∇. v
G
   is the gas phase convection per unit volume

The mass balance has a constraint on its excess terms where

∑ (EG
G
m + GmG = 0 ) (8.5)

148 
 
The superscript G in Equations 8.4 and 8.5 means the quantity represents the gas phase.

The gas phase momentum balance can be written as follows.

∂ G G G
∂t
( ) ( ) (
ε ρ v + ∇ ⋅ ε G ρ G v G v G + ∇ ⋅ ε G t G − ε G ρ G g + E GM + I GM + S GM + G GM = 0) (8.6)

where
∂t
(
∂ G G G
)
ε ρ v is the rate of increase of gas phase momentum per volume,

(
∇⋅ ε GρGv v
G G
) is the rate of gas phase momentum gain by convection per unit volume,

∇ ⋅ε G t
G
is the rate of gas phase momentum gain by shear and normal stresses per unit

volume, and ε G ρ G g is the gravity force on the gas phase per unit volume.

G G G G
The interphase transfer excess terms E M , I M   ,  S M   and  G M represent different

G
mechanisms by which momentum is exchanged between the phases. The term, E M ,

is the rate of momentum transfer due to phase change at the interface between the phases.

G
The term, I M , is the rate of momentum transfer due to drag per unit volume between the

G
phases, S M is the rate of momentum transfer due to slip across a 3-D interface region

when 3 or more phases are in contact (i.e., due to capillary forces) per unit volume, and

G
G M is the momentum transfer due to heterogeneous reaction at the interface between the

phases per unit volume.

In most processes of flows through porous media the dominant forces are the

stresses and drag forces. The stress term is the sum of the normal stress (i.e. pressure)

and the shear (deviatoric) stress term

149 
 
t = Pδ + τ
G G

(8.7)

The interphase exchange of momentum due to mass transfer or chemical reaction

G G
are assumed to be insignificant; hence the terms E M and G M are dropped from Equation

(8.6). In flows through porous media, as long as the pore sizes are significantly smaller

than the macrodimensions of the medium, then the wall effects are negligible and the

shear term is dropped from the momentum balance. Also the inertial terms are negligible

compared to the pressure and drag terms. The momentum balance reduces to

ε G ∇P + F G − ε G ρ G g = 0 (8.8)

where

G G G
F = IM +SM (8.9)

represents the effective drag force momentum transfer between the phases. The

constraint on excess terms for the momentum balance requires momentum transfers

across the interface between the phases sum to zero

∑ (F ) = 0
G
(8.10)
G

8.3.2 Mass and momentum jump balances

To develop the pressure and velocity profiles in the drainage channel incorporated filter

media, the mass and momentum jump balances must be applied at the boundaries of the

filter media and drainage channels.

A schematic of a drainage channel incorporated media and the interface between

media and drainage channel is shown in Figure 8.1.

150 
 
θ
Interface Interface
Region I Region II
θ

VI VII
θ
Momentum

Figure 8.1 Interface of media and drainage channel.

The mass jump balance analogous to Eq. (8.3) applied between the multiphase

regions is written as


( G ⎤
( ⎡
))
G G G ⎤
⎢∑ ε ρ w − v ⎥ ⋅ n12 = ⎢∑ ε ρ w − v ⎥ ⋅ n12  
G G
( ( ))       (8.11)
⎣G ⎦1 ⎣G ⎦2

Similarly, the momentum jump balance is


(G G G G
(
G G⎤
⎢∑ ε ρ v w − v − ∑ ε t ⎥ ⋅ n12 ))
⎣G G ⎦1
(8.12)

( G G
( G⎤
= ⎢∑ ε G ρ G v w − v − ∑ ε G t ⎥ ⋅ n12 ))
⎣G G ⎦2

8.4 ASSUMPTIONS

The volume averaged continuum equations are used to model the coalescing filter

performance. The complete sets of equations are complicated and cannot be solved

151 
 
analytically. The following assumptions simplify the equations to obtain a tractable

solution.

• The process is isothermal.

• The flow is one dimensional (only in the X direction).

• Steady state.

• Filter media is incompressible. Solid phase is rigid and stationary.

• The fibers do not move within the matrix (held stationary by the binder).

• Binder material (i.e. glue) accounts for only a small fraction of the filter

volume and can be neglected.

• The porosity of a new unused media and drainage channels (no oil present)

are uniform and constant. Alternatively, the fiber phase volume fraction of

media and drainage channel are uniform and constant.

• The permeability of a glass fiber media as well as drainage channel is uniform

and anisotropic.

• No chemical reaction.

• No mass transfer.

• Inertial (convection) terms are neglected.

• No property transfer due to slip.

• Viscosity and density of air are constant. No temperature effects.

• Neglect wall stress.

• Interface between the media and drainage channel is stationary.

152 
 
8.5 Permeability of media and Darcy’s law

If the porosity and/or permeability vary in position, the Darcy’s law at the local

scale is used and defines a local permeability as given in equation 8.13.

k
q=− ∇P (8.13)
μ

Hassanizadeh and Gray [109] introduce the constitutive relation


G
F = R⋅ v −v ( G S
) (8.14)

S
where the stationary solid phase velocity, v is zero. R is a second order tensor
resistance function that is characteristic of the material properties and the flow
conditions. Combining equations (8.14) and (8.8) gives the expression for the local
volume averaged velocity as
−1
v = − R ε G ⋅ ∇P
G
(8.15)

in which R −1 is the inverse of the resistance function tensor.

Darcy’s law may be written at the local scale by substituting the definition for the volume
averaged velocity ε G v = q and letting the permeability vary in position and direction to
G

have the form

G k ⋅ ∇P
v =− (8.16)
ε Gμ

By inspection it can be seen that the resistance function and the local permeability are

related by

−1 2
k = R εG μ (8.17)

The significance of writing the permeability or resistance function in local form is

that it allows us to account for variations in the magnitude of the permeability (or

153 
 
resistance function) over position. The local permeability depends upon several factors

including the local porosity and the effective wetted surface area [112]. A number of

references provide correlations relating local permeability or resistance functions to

surface area, porosity, particle size, and velocity [113, 114].

The significance of the tensorial nature of the local permeability or the resistance

function is that it provides a way to account for a directional dependence of the direction

to flow. For an isotropic medium it should not matter what direction is the flow and what

direction is the coordinate origin. An anisotropic permeability is characterized by

different permeabilities in each coordinate flow direction. Anisotropic permeability is

applicable to two-dimensional systems shown in Figure 8.2. The Y-direction

permeability, kY, differs from the X-direction permeability, kX, in the 2-D geometry.

kX

kY

Figure 8.2 Anisotropic permeability.

When a material has anisotropic behavior, in a 2-D X-Y planar geometry (i.e., it

has symmetry and uniformity in Z direction) then the permeability tensor can be written

as follows,

154 
 
11 12 (8.18)
k=
21 22

where the 1, 2 indices represent the, X, Y coordinate directions, in any orthogonal

coordinate system. A symmetric tensor has equal valued off-diagonal components, such

that k12 = k 21 . Normally, anisotropic porous media are assumed to be orthotropic, [109,

115, 116] meaning that there is symmetry in the permeability tensor with respect to two

orthogonal planes [117]. With an orthotropic material it is possible to rotate the

coordinate system to a primed coordinate system , , where the coordinate axes are

aligned with a principal axis of the medium and the off-diagonal components are zero.

When aligned with the principle axes, the permeability tensor is

0 (8.19)
k' = 11
0 22

Figure 8.3 indicates the principle and system axes for the drainage channel

incorporated media. The system axes are X, Y, and Z while the principle axes are X’, Y’,

and Z’. If the material is orthotropic then the principle axes are perpendicular (i.e., the

direction at which the permeability tensor component values are maximum and minimum

are perpendicular) and the coordinate system may be rotated such that the coordinate

directions coincide with the principle axes. In this situation the off diagonal components

are zero. The principle axes of the media coincide with the system axes but for the

drainage channels the principle axes are tilted as indicated by angle θ.

155 
 
Y θ
X’
θ
θ
X
θ

θ
Z
Y’
Z’

Figure 8.3 Principle and system axes for drainage channel incorporated filter media.

The and for the glass fiber media as well as drainage channel were obtained

by using the using the Frazier® Differential Air Permeability Measuring Instrument and

analyzing the data using Darcy’s law. The k’22 and k’11 experimental values for glass

fiber media were 4.47E-10 m2 and 2.24E-10 m2 while k’22 and k’11 experimental values

for drainage channels were 14.24E-9 m2 and 3.389E-9 m2. The k’22 value for the drainage

channel is a good starting value, but the true value is questionable because in order to

make the measurement multiple sheets of the drainage channel materials were stacked

together. This parameter should be fitted between a 3-D numerical model and

experimental measurements of filter samples constructed with drainage channels to

obtain a more reliable value. This is left for future work because the current model is

2-D. The 2-D model only gives accurate results for the disk-shaped filter when the filter

properties are uniform (i.e. symmetric) in the Y-Z plane which only occurs when there

156 
 
are no drainage channels; otherwise the computer program must be rewritten in 3

dimensions which is beyond the scope of this particular work.

For the drainage channels, the principle axes are tilted relative to the filter

geometry coordinate origin by angle θ is shown in Figure 8.3. Hence the permeability

tensor in X, Y coordinates of the filter geometry for drainage channel is written as

follows.

(8.20)

Applying coordinate transformations from the primed to unprimed coordinates [117,118].

    (8.21)

      (8.22)

    (8.23)

The permeability tensor for the drainage channel has   , hence it is symmetric.

The drainage channels and the media have different porosities also. The

experimental porosity of the media was 0.96 using the pycnometer while the

experimental porosity of drainage channel was obtained gravimetrically by using a

specific gravity bottle. The porosity of the drainage channel is 0.986. Local variations in

medium permeability affect the observed pressure drop. The effective permeability

observed at the macroscale can be correlated with the local variations for models.

157 
 
8.6. Simplified conservation equations

8.6.1 Mass and momentum balances

The assumptions reduce the mass balance equation (Equation 8.4) as follows.

(
∇⋅ εGv
G
)= 0 (8.24)

Equation 8.24 in 2-D rectangular coordinates can be written as follows.

0 (8.25)

Combing Darcy’s law Eq.8.16, with Eq. (8.24) gives

⎛ 1 ⎞
∇ ⋅ ⎜ k ∇P ⎟ = 0 (8.26)
⎜ μ ⎟
⎝ ⎠

which can be expanded as

k : ∇∇P − ∇P.(∇.k ) = 0 (8.27)

In regions where k is not a function of position then the 2nd term is zero. In a

system with a jump within the system volume, the second term in Eq. 8.27 becomes

infinite at the jump. Hence a filter with drainage channels must be modeled in separate

regions and the jump balance must be applied across the jump. Within each region Eq.

8.27 reduces to

k : ∇∇P = 0 (8.28)

Eq. 8.28 in 2-D rectangular coordinates can be written as

0 (8.29)

We can define the dimensionless pressure as where ΔP is the pressure drop


across the filter. Equation 8.29 becomes

0 (8.30)

158 
 
8.6.2 Mass and momentum jump balances

The mass jump balance in Eq. 8.11 reduces to

(ε I
G
−I − II
)
υ G − ε II G υ G ⋅ n12 = 0
(8.31)

Hence, separated into vector components, Eq. 8.31 becomes

i   j 0

Both of the bracketed terms are independent. Hence the y-component can be written as

(8.32)

Combining this with Darcy’s law Eq. 8.16, and the definitions of the dimensionless

pressure, we get

    (8.33)

which is used to determine the values F at the region boundaries between the medium

and the drainage channels. Equations 8.30 and 8.33 are solved in discretized form over

the volume of the filter medium to determine the pressure profile. The pressure data are

applied with Darcy’s law to determine the local velocities. The momentum jump balance,

Eq. 8.12 with Eq. 8.7, the convection terms and the shear terms are negligible compared

to the pressure terms. Hence the momentum jump balance reduces to

⎡ ⎤ ⎡ ⎤
⎢∑ ε P ∂ ⎥ ⋅ n12 = ⎢∑ ε P ∂ ⎥ ⋅ n12
G G
(8.34)
⎣G ⎦1 ⎣G ⎦2

which ultimately reduces to

  (8.35)

at the jump interfaces. This proves the pressure at the boundary between the regions

when approached from region I equals the pressure at the boundary when approached

from region II.

159 
 
8.7 Boundary conditions

The top and bottom boundaries of the filter are impermeable hence the 0 at

both the B.Cs and hence 0 at these boundaries. The dimensionless pressure at the

outlet is zero and the dimensionless pressure at the inlet is unity.

8.8 Grid generation

Drainage channels are incorporated at different angles into the filter bed length.

Hence the boundaries between the regions do not parallel rectangular grid lines if a

rectangular grid system is used to model the geometry. To get the grid lines to coincide

with the boundaries, the solution grid must be deformed. Knupp and Steinberg [116]

describe how the grid transformations are obtained using the Jacobian. The Jacobian is

used to convert from X, Y coordinate system to u, v coordinate system by the following

expression.

∂( X , Y )
∫∫ φ ( X , Y )dXdY = ∫∫ φ [ f (u, v ), g (u, v )] ∂(U ,V ) dudv
A G
(8.36)

The Jacobian matrix j can be defined as follows.

(8.37)

The Jacobian determinant, also known as Jacobian, is the determinant of the

()
Jacobian matrix i.e. J = det j . The Jacobian matrix need not be square. If the

dimensions of the physical objects equal the dimensions of the physical space i.e. the

object is not a curve or non-flat surface, then the Jacobian is square. The schematic of the

grid transformation is shown in Figure 8.4.

160 
 
η  Z
Logical space Physical space

Transformation

Inverse
Transformation

ξ X

Figure 8.4 Physical and logical space.

The Jacobean of the transformation is required to be non-zero, and consequently

the transformation has an inverse. Grids are first chosen in logical space and then mapped

to physical space.

8.8.1 Goals of grid generation

1. Avoid a folded transformation (to preserve physical and math properties).

2. Produce smooth grids where spacing varies smoothly and angles do not become

too small (to minimize the error).

3. PDEs representing the physical phenomena such as fluid flow or heat transfer are

known as hosted equations. If the solution of the hosted equations varies rapidly

in some part of the physical region, choose a finer grid in that part of the region

(to reduce error in the math solution).

161 
 
8.8.2 Grid mapping

1. Each point in physical space is the image of a unique point in logical space.

2. Each point in logical space maps to a unique point in physical space.

3. The corners of the logical space must map with to the corners of the physical

object. Coordinate functions X (ξ ) must be continuous and have continuous

derivative in the interior of Uk and on the boundary of Uk.

4. Use grid to solve the PDEs in irregular region.

5. PDE must be transformed to the logical space.

6. The hosted equation is invariant under general nonsingular coordinate

transformation i.e. the physics is preserved.

Figure 8.5 shows a section of the media in physical as well as logical space. The

top and the bottom sections of the media in physical space are trapezoids except for the 00

angle media. The section between the drainage channels are parallelograms. For the 00

angle media the media has rectangular shaped four sections separated by equally spaced

three drainage channels. The logical space will have same dimension of the physical

space.

162 
 
,   ,

Physical space
Inlet , ,
Y
Outlet

,   ,
, ,
Y0
X0 X X1

Logical space
η 


0  ξ  1

Figure 8.5 Physical and logical space conversion.

The ,   values are the physical locations of the corners of the physical

trapezoidal region. To convert from the logical space coordinates , to the physical

space coordinates (X, Y), the following relations are used

, 1 1 ,   1   ,   1 ,   , (8.38)

, 1 1 ,   1   ,   1 ,   , (8.39)

163 
 
The Jacobian of the bilinear map is computed as follows.

1 , , 1 , , (8.40)

1 , 1 , , , (8.41)

1 , , 1 , , (8.42)

1 , 1 , , , (8.43)

PDES are written in discretized form by finite differencing. The pressure profile is

calculated at local points by using a nine point solver and the Gauss Seidel iterative

approach is used to solve the pressure profile over all of the points in the grid space. The

nine point solver uses values at grid points surrounding each (i ,j) point in the logical grid

space, as indicated in Figure 8.6.

1, 1   , 1 1, 1

1,   , 1,

1, 1   , 1 1, 1

Figure 8.6 Nine point-solver.

164 
 
Using the nine point-solver, the value at point (i, j) is calculated from the

surrounding 8 points. The tridiagonal matrix is applied to calculate the pressures at local

points. A FORTRAN program published by Knupp and Stteinberg [116] gives a code for

solving the nine-point solver. This code is used in the FORTRAN program (Appendix B)

to calculate the pressure and velocity profile for the filter geometries described above.

The boundary condition at an impermeable surface (at the top and bottom surfaces

of the filter) requires that the gradient in the pressure at the surface be zero, 0.

A first order approximation to this boundary condition is applied, where

, , (8.44)

Eq. 8.30 is used to calculate the pressure profile in the media and drainage

channel region and Eq. 8.33 is used calculate pressure profile at the interface of the media

and drainage channel. To smoothen the irregularities at the boundary of media and

drainage channel, the physical space is transferred to logical space by using the Jacobian

transformation. In section I to VII Eq. 8.30 is used and at all the media and drainage

channel interface Eq. 8.33 is used to calculate the pressure profile. The schematic of

physical and logical space for the drainage channel incorporated filter media and the

equations used to calculate the pressure profile for the media and drainage channel and at

the interface of the media and drainage channel is shown in Figure 8.7.

165 
 
Interface Interface
B.C.   0  
VII
VII
θ
VI
VI
V

θ V
IV
B. C. F = 1 B.C. F = 0
IV
III

θ III
II

II
I
I

B.C.   0  

Physical space Logical space

Region I, II, III, IV, V, VI, VII: Eq. 8.30


Interfaces between regions: Eq. 8.33

Figure 8.7 Physical and logical space of drainage channel incorporated filter media and

equations used to calculate the pressure profile in the logical space.

8.9 Algorithm

A flow chart for the mathematical algorithm of the FORTRAN code is given in

Figure 8.8. This algorithm shows the key features of the FORTRAN code. The pressure

166 
 
profile is repeatedly calculated until the pressure values converge to values that change

less than a error tolerance value between successive iterations.

START

Define media properties and variables

Generate geometry in physical space

Transformation of physical space to logical space using Jacobian transformation

Initial guess for pressure profile using linear approximation

Calculate pressure profile using the discretized PDE (Gauss Seidel method)

If
No
convergence
obtained

Yes
Generate velocity profile using the pressure profile based on Darcy’s law

Print velocities

END

Figure 8.8 Algorithm.

167 
 
8.10 Results

8.10.1 Optimizing the grid size

For the analytical solution to be independent of grid size, the optimum grid size

was determined that minimizes error in the calculations. For glass fiber isotropic media

the grid size was varied by keeping the grid spacing the same, ∆ ∆ . The volumetric

flow rate, given by

      (8.45)

is calculated using trapezoidal rule to evaluate the integral over the exit surface. Here W

and H are the width and height of the media respectively (assuming a square face area

filter media). is the porosity of the media. vX is the velocity in X direction.

The flow rate was calculated for a special case solution of an isotropic medium

for which an analytical solution is available. The analytical volumetric flow rate, i.e.

QAnal, was obtained by solving the macro scale Darcy’s law for the special case. The

QAnal was obtained from the Frazier test Analyzer. The QAnal can only be obtained for an

isotropic media by solving the macro scale Darcy’s law using the Frazier® Differential

Pressure Air Permeability Measuring Instrument. The experimental procedure is similar

which is described in Chapter III. The model calculated, flow rate QCalc for the same filter

size, geometry, properties, and pressure drop as the analytical solution. The difference

between the two flow rates defines the error in the calculation. The error is plotted in

Figure 8.9 as a function of the grid size. The error is calculated by using following

formula

168 
 
  (8.46)

At grid sizes less than 0.0005 meter, corresponding to the filter depth divided into 30

grids and higher, for the special case i.e. the isotropic media, the QCalc = 1.5735531E-3

m3/s while QAnal = 1.57E-3 m3/s. The QCalc and QAnal are determined for the same ∆ and

permeability. The volumetric flow rate of the coalescence filtration experiment for dry

air (without droplets) is 1.57E-3 m3/s. Hence the isotropic filter media model is in a very

good agreement with the experimental results of Darcy’s law (Frazier® Differential

Pressure Air Permeability Measuring Instrument test) and the coalescence filtration test.

The calculations show 30 grid points or larger provide a small enough grid spacing for

reasonably accurate calculations.

1.20E+02

1.00E+02

8.00E+01

6.00E+01
Error

4.00E+01

2.00E+01

0.00E+00
0 0.0005 0.001 0.0015 0.002 0.0025 0.003

-2.00E+01
Grid size (meter)

Figure 8.9 Error analysis at different grid sizes.

169 
 
The current model is optimized for the grid size for comparing to the analytical

solution of special case i.e. the isotropic filter media. But the pressure and velocity

profiles are developed for the anisotropic media. Analytical solution is not available for

the anisotropic media. Hence in future the grid size needs to be optimized by converging

the QCalc for the anisotropic media.

8.10.2 Velocity profile of isotropic filter media without drainage channel

The velocity profiles were calculated in the FORTRAN Code by evaluating

Darcy’s Law using 1st and 2nd order discretized approximations of the pressure gradients.

The velocity vectors were plotted on the grid by using a vector plot code [119]. The

vector plot code generates two plots, a velocity magnitude plot and a velocity direction

plot. The velocity magnitude plot gives information about the magnitude and direction of

the velocity along the filter bed length. In the velocity magnitude plot, the lengths of the

arrows indicate the magnitude of the velocity and the directions of arrows indicate the

direction of velocity. The velocity direction plot only shows the velocity directions, hence

the arrow in the velocity direction plot are of similar length.

170 
 
0.06 m

0.06 m

0.014 m 0.014 m
(b)
(a)

Figure 8.10 (a) Velocity magnitude of isotroipc media without drainage channel (b)

velocity direction of isotroipc media without drainage channel.

The velocity magnitude and direction plot of the isotropic glass fiber filter media

without drainage channels are shown in Figure 8.10. The velocity magnitude plot for the

isotropic filter media indicate that the velocities are of similar magnitude and the

direction plot indicate that the velocity vectors are all parallel to each other and all

velocities are exiting from the media. The velocity magnitude and direction plot indicates

that in the isotropic filter media a uniform flow field is generated. In this particular test

171 
 
case the media and drainage channels had the same isotropic permeabililties. As

expected the velocities are nearly uniform and have y-components that are very small.

8.10.3 Velocity profile of anisotropic drainage channel incorporated filter media

The velocity magnitude and direction plot of drainage channel incorporated filter

media of 0.06 m height and 0.014 m thickness at 450 downward inclination are shown in

Figure 8.11 (a) and (b) respectively. In both figures, the drainage channels are boxed for

better visualization. Both the figures show the velocity magnitude and directions by blue

and red arrows for the drainage channels and media respectively. The length of the arrow

indicates the magnitude and the direction of arrow indicate velocity direction. It is

evident that the magnitude of velocity flowing through the drainage channel is a lot

bigger than the magnitude of the velocity flowing through the filer media. Figure 8.12

shows the closer view of the two different sections of media at the media and drainage

channels interface. The velocity in the drainage channel is indicated by blue arrows while

the magnitude of velocity in the filter media is shown with the red dots. The red dots are

actually red arrows but because the magnitude of velocity in that region is very small

they are seen as dots. The direction plot shows that all the velocities at the top and

bottom boundary are straight pointing towards exit which indicates that the upper and

lower boundaries of the media are impermeable to flow because the filter is enclosed in

the filter holder.

172 
 
0.06 m
0.06 m

0.014 m 0.014 m
(a) (b)

Figure 8.11 (a) Magnitude and (b) direction of velocity profile in drainage channel

incorporated anisotropic filter media with an angle of 45 deg. The drainage channels are

marked inside of the rectangular boxes.

173 
 
Figure 8.12 Close-up view of Figure 8.10 (a). Velocity magnitude of drainage channel

near the drainage channel and filter media interface.

8.10.4 Parametric study

8.10.4.1 Varying the angle of drainage channel

Filter geometries were developed by incorporating drainage channels at 00, 150,

300, 450 and 600 downward inclination. The model calculations were also obtained for a

filter medium having three equally spaced drainage channels of 500 µm thickness. The

model predicted volumetric flow rates are reported in table 8.1. The volumetric flow rates

were obtained by keeping all other model parameters constant. As the angle of drainage

channel increases the model predicted volumetric flow rate decreases.

174 
 
Table 8.1 Volumetric flow rates of filter geometry with drainage channels at different

angles

Sr. No Angle of drainage channel Volumetric flow rate (m3/s)

1. 00 3.31E-02

2. 150 3.07E-02

3. 300 2.93E-02

4. 450 2.35E-02

5. 600 9.31E-03

8.10.4.2 Changing the thickness of drainage channel

The thickness of drainage channel was varied from 500 µm to 1000 µm and 7200 µm by

keeping all other model parameters constant. The model predicted flow rate of varying

drainage channel thickness are given in Table 8.2. As the thickness of the drainage

channel is increased the volumetric flow rate increases. The drainage channel are porous

than the media hence when the thickness of the drainage channel increases the volumetric

flow rate increases as expected. According to Darcy’s law if these flow rates were the

same as the QAnal, then the pressure drop of the media will be significantly lower as

compared to the isotropic no drainage channel filter media.

175 
 
Table 8.2 Volumetric flow rates of filter geometry with drainage channels of varying

thickness

Sr. No. Drainage channel thickness (meter) Volumetric flow rate (m3/s)

1. 0.005 2.35E-2

2. 0.001 3.61E-2

3. 0.0072 7.44E-2

8.10.4.3 Exchanging porosity and permeability of media and drainage channel

When porosity and permeability of the drainage channels and media were

interchanged by keeping all other model parameters constant, for the 450 filter geometry

the model predicted volumetric flow rate is 0.137531 m3/s. In this geometry media is

more porous and 10 times permeable than the drainage channels and media size is

comparable to the size of entire filter geometry hence the volumetric flow rate is

significantly bigger as compared to the volumetric flow rate of the actual 450 filter media

geometry. 

For this geometry, the velocity magnitude and direction plot at the interface of

drainage channel and media are shown in Figure 8.13 and 8.14 respectively. Figure 8.15

indicate velocity magnitude and direction for the entire filter geometry. Figure 8.13

indicate that when the drainage channels are less porous and permeable than the media

then the velocity magnitudes in the media and drainage channel are comparable. Figure

8.14 indicates that when the drainage channels are less porous and permeable than the

media then all the velocities are diverted towards the low resistance drainage channels.

176 
 
Figure 8.13 Velocity magnitude near the drainage channels when the drainage channel

porosity and permeability is less than the filter media

Figure 8.14 Velocity direction near the drainage channels when the drainage channel

porosity and permeability is less than the filter media.

177 
 
6 cm

6 cm

1.4 cm 1.4 cm

(a) (b)

Figure 8.15 Velocity (a) magnitude and (b) directions in drainage channel and media

when porosity and permeability of filter media is higher than drainage channels.

178 
 
The velocity profiles are useful to help to visualize how the gas moves through

the medium at the start of the filtration process and gives an indication of where liquid

may collect fastest (i.e. at larger velocities). The model results indicate that the velocity

magnitude is higher in the drainage channel region as compared to the filter media which

will promote drainage. The velocity profile of the media has effect on the performance of

the filter. Hence it is important to study the velocity profile of the filter media. When the

droplet capture, migration and drainage will be accounted in the model the model will

able to predict the performance of filter media better. But to find out why the 450 media

in horizontal orientation works better more work is needed.

Gas flow in Gas flow out

Drainage

Figure 8.16 Drop motion through drainage channel and media.

In the parametric study of the model, when the drainage channels have lower

porosity and permeability than the filter media the velocity direction plot indicate that the

179 
 
velocity is moving towards the media. In this case study, the media offer lower resistance

to flow compared to the drainage channels, hence the velocities are going towards media.

Even though we consider the filter media has higher porosity and permeability than the

drainage channel, the pore sizes of the filter media are smaller than the drainage channels.

There will be some resistance for the big oil droplets to move from larger pores to smaller

pores. But they will tend to drain from the filter media through the drainage channels

because of the bigger pores of the drainage channels will provide an easier path for the

drops. The small droplets might get carried to the media with the air flow where they will

coalesce which will improve the rate of coalescence (Refer Figure 8.16). The drainage

channels are incorporated at downward angles hence the droplet draining from the

drainage channels will also get an added advantage of the gravitational force. Hence a

media with larger pores than drainage channels needs to be tested in future. A

microscopic study of drainage channel incorporated filter media needs to be done to

observe the droplet motion in the media and drainage channels. The microscopic results

will be helpful for incorporating the droplets capture, droplet migration and drainage

mechanisms in the mathematical model.

The model also needs to be modified to account for wettability, pore sizes,

geometry effects along with the droplet capture, droplet migration and drainage. Current

model indicates the velocity profile of the dry air. The velocity profile of the air flow has

effect on the droplet capture, migration and drainage of droplets hence it is important to

know the velocity profile in the filter media. But in future the model needs to account for

the droplet capture, droplet migration and drainage in cylindrical coordinates for better

comparison with the experimental results.

180 
 
CHAPTER IX

CONCLUSIONS

This dissertation focuses on improving performance and drainage of coalescing

filters. The objectives of the research work are as follows.

a. Improve performance and drainage of coalescing filters by utilizing the

gravitational force.

b. Develop different filter geometries by incorporating woven and nonwoven

drainage channels. Develop filter geometries by using glass fiber filter media

of similar porosity, permeability and strength.

c. Study the parameters affecting on the improvement of the drainage channel

incorporated filter media and improve the filter media design.

d. Improve the performance of the filter media by developing nanofiber

augmented filter geometries and study the performance of nanofiber

augmented filter geometries experimentally.

e. Develop a composite filter media design. The composite filter media will be

made up of different fiber sizes and it will indicate higher performance than a

filter media made with one size fibers. Evaluate the performance of composite

filter media experimentally.

181 
 
f. Develop a mathematical model of the drainage channel incorporated filter

geometry to predict pressure and velocity profile with no oil drops present at

steady state.

The glass fiber filter media with similar porosity, permeability and strength were

made by adding Megasol® S50 binder by using the vacuum molding process. All The

filter geometries were developed by adding the drainage channels at various positions

along the filter bed length. The middle-drainage and the inclined angle geometries were

developed by cutting the filter discs at different positions and the drainage channel were

incorporated to develop the filter geometries. All filter geometries were tested in

coalescence filtration setup in horizontal and vertical flow orientation to study the effect

of gravitational force on the performance of the filter media. The filter media

performance was expressed in terms of quality factor and saturation. Drainage channel

incorporated filter geometries indicated improved quality factor and drainage. Hence

factors affecting the performance of the filter geometries were experimentally evaluated.

A mathematical model was developed to predict the pressure and velocity profile of the

45 degree filter geometry and was compared with other inclined angle geometries in

horizontal flow orientation.

The results of each work are documented in respective chapters. Several

conclusions are drawn based on those results which are summarized as follows.

• Experimental results indicate that when the air flow is in the direction of

gravitational force the no-drainage channel glass fiber filter media had higher

182 
 
quality factor and lower saturation. Experimental results indicate that gravitational

force helps in enhancing the drainage and performance of filter media.

• No-drainage channel filter media had higher saturation at the inlet and outlet and

low saturation in the middle of the filter media. Hence when drainage channels

are incorporated at the inlet and outlet as well as both-end of the filter media the

filter geometry had higher quality factor and lower saturation.

• Filter geometries developed with three equally spaced drainage channels at 150,

300, 400, 450, 500 and 600 in horizontal flow orientation indicated improved

performance. The Filter geometries equipped with three equally spaced drainage

channels at 450 downward inclination in horizontal flow orientation indicated the

highest quality factor and lowest saturation among all of the filter geometries.

• Filter geometries developed with polypropylene woven drainage channel of 500

µm pore opening and 600 µm filament thickness had best performance among

other polypropylene woven drainage channel filter media that were tested.

• Filter geometries developed with nonwoven drainage channels indicated

improved performance as compared to no-drainage channel filter media.

• Filter geometries developed with polypropylene woven drainage channels of 500

µm pore opening and 600 µm filament thickness indicate better performance than

filter geometries developed with polypropylene spunbond fabric used as drainage

channels. The experimental results showed that regular spaced porous woven

drainage channel are effective in draining the oil droplets from the filter media

compared to the nonuniform pore opening nonwoven drainage channels.

183 
 
• The pore sizes of the drainage channels affect significantly the performance of the

filter geometries. As the basis weight of the spunbond fabric increases the pore

sizes decreases. Hence the basis weight of the drainage channels indicates an

effect on the performance of the filter geometries. As the pore size of the drainage

channel decreases the pressure drop and saturation increases and the quality factor

decreases. Hence the low basis weight drainage channel incorporated filter media

performed better than higher basis weight drainage channels incorporated filter

media. Hence the drainage channel needs to be open structure for draining the

coalesced drops faster from the filter media.

• Surface energy of the drainage channels affects the performance of the filter

geometries significantly. The low surface energy Teflon® incorporated drainage

channels had higher quality factor and lower saturation as compared to the

respective filter geometries developed by incorporating intermediate surface

energy polypropylene. The intermediate surface energy polypropylene

incorporated drainage channels had higher quality factor and drainage and lower

saturation as compared to the respective filter geometries developed by

incorporating high surface energy nylon. Experimental results indicate that low

surface energy drainage channels are effective in draining the oil from the filter

media. It could be the oil droplets beads up on the surface of the low surface

energy drainage channels and due to the porous and open structure of drainage

channels the droplets drains very effectively while for the high surface energy

drainage channels the oil spreads on the drainage channel possibly blocking the

pores of the drainage channels.

184 
 
• Nanofiber incorporated glass fiber filter media had improved quality factor. The

fiber diameter of the nanofiber affect the performance of the nanofiber augmented

filter geometries. Nanofiber incorporated filter geometries had higher capture

efficiency, but as the diameter of the nanofiber decreases the pressure drop

increases moderately due to the “slip flow” condition. Hence as the nanofiber

diameter decreases the quality factor of the filter media increases. Nanofiber

augmented 450 downward inclination filter geometries had significantly higher

quality factor and lower saturation than the no-drainage channel glass fiber as

well as no-drainage channel nanofiber augmented filter media. The 450 downward

inclination filter geometries developed with intermediate surface energy

polypropylene had higher quality factor and lower saturation as compared to the

high surface energy nylon incorporated 450 downward inclination filter

geometries. Experimental results indicate that as the surface energy of the

drainage channel decreases the performance of the nanofiber augmented filter

geometries increases.

• 450 downward inclination filter geometry indicated highest quality factor and

drainage among all the filter geometries developed. Polypropylene woven

drainage channel incorporated 450 geometry indicated relative quality factor of

4.20 while the same geometry with Teflon® woven drainage channels indicated

relative quality factor of 8.89. The 100 nm nanofiber augmented 450 filter

geometry with polypropylene woven drainage channels indicated relative quality

factor of 6.86 while the same geometry with nylon woven drainage channels

indicated a relative quality factor of 3.83.

185 
 
• Composite filter media indicated significantly high pressure drop. The fibers were

layered by changing wettability from high to low along the filter bed length.

olepophilic fibers being at the inlet accumulated the oil droplets. Less oleophilic

fibers were layered after these fibers. Due to the less olephilic nature of the fibers

in the exit side of the filter the oil droplets remained accumulated in the filter

instead of migrating towards the exit boundary and draining from the filter. Hence

the filter media had very less drainage and significantly higher pressure drop.

• 450 downward inclination with three equally spaced drainage channel filter

geometry was modeled using the volume averaging continuum theory. Pressure

and velocity profiles of the drainage channel incorporated filter media were

developed for air flow. The velocity profile of the filter geometry indicates that

the velocity magnitude in the drainage channel is significantly high as compared

to the filter media. Because the drainage channels have lower flow resistance to

air. Hence the pressure drop of 450 downward inclination filter geometry are

significantly lower than the no-drainage channel filter media.

• The 450 filter geometry: The 450 filter geometry in horizontal orientation works

best among all the filter geometries that were developed. The filter geometry

indicates high capture efficiency and significantly low average saturation. The

low average saturation of the media results in significantly low pressure drop

which improves the quality factor of the filter media. We do not fully understand

why the 450 filter geometry works so much better than the other inclined angle

geometries. The performance of the filter media is certainly a function of filter

size, filter geometry, pore sizes, saturation and wettability. In general, but not in

186 
 
every case when the saturation is low, pressure drop is low. Low saturation of the

filter geometry is due to liquid drain faster through the drainage channels. Further

work needs to be done to explain improved performance of the 450 filter geometry

than other filter geometries by studying filter size, filter geometry, spacing

between drainage channels, pore sizes and wettability effects.

187 
 
CHAPTER X

FUTURE WORK

1. Develop filter geometries by incorporating more than three drainage channels at

inclined angles and evaluate the filter media performance. The thickness of the

filter media sections or height between the drainage channels needs to be

optimized to study its effect on the performance of the filter geometries as well as

to further enhance the filter geometry performance.

2. Perform microscopic study to observe drop motion in drainage channel

incorporated filter media when the permeability of drainage channels is higher

than media as well as when the permeability of media is higher than drainage

channels. Filter geometries developed with filter media having high porosity and

permeability but smaller pore sizes as compared to the drainage channels needs to

be tested.

3. Cartridge filters and conical filters are commonly used in industry. Along with

disc shaped media the cartridge shaped and conical shaped filters can be used to

develop new filter geometries.

4. Wettability of drainage channel has proven effect on the performance of the filter

media. Hence filter geometries can be developed by incorporating different

surface energy material drainage channels. Filter can also be made by using fibers

188 
 
of varying surface energy and fiber sizes and the gradient in surface energy of

filter and drainage channel can be studied.

5. Composite filter media with fibers of oleophilic and olephobic nature can be

developed to enhance the performance and drainage of the filter media.

6. The current capture model needs to be optimized for the grid size of anisotropic

media. It also needs to be modified to account for wettability, pore sizes of media

and drainage channel, geometry effects, and size of the filter. Along with these

parameters the model needs to account the capture of droplets, droplet migration

and drainage to predict the overall performance of the media in cylindrical

coordinates for better comparison with the experimental results.

189 
 
BIBLIOGRAPHY

1. Rebours, A., Boulaud, D., Renoux, A. (1992). Production of monodispersed


aerosol by evaporation and condensation of vapor under control process. Journal
of Aerosol Science, 23 (Suppl. 1), 189.

2. Schaber, K. J., Körber, O., Ofenloch, R., Ehrig, P. D. (2002). Aerosol formation
in gas–liquid contact devices—nucleation, growth and particle dynamics.
Chemical Engineering Science, 57 (20), 4345.

3. Gina, L. J., Raynor P. C., Schumann, R. L. (2003). Selecting Fiber Materials to


Improve Mist Filters. Aerosol Science, 34, 1481.

4. Thornburg, J. (2000). Size Distribution of Mist Generated During Metal


Machining. Applied Occupational and Environmental Hygiene, 15(8), 618.

5. Woskie, S. R., Smith, T. J., Hallock, M. F., Hammond, S. K., Rosenthal, F.,
Eisen, E. (1994). Size-Selective Pulmonary Dose Indices for Metal Working Fluid
Aerosols in Machining and Grinding Operations in the Automobile
Manufacturing Industry. American Industrial Hygiene Association Journal, 55,
20.

6. Gillespie, T., Rideal, E. (1955). On the adhesion of drops and particles on impact
at solid surfaces. Journal of Colloid and Interface Science, 11(10), 281.

7. Retrieved 22 October 2009, from http://www.parker.com

8. Piacitelli, G. M. (2001). Metalworking Fluid Exposures in Small Machine Shops:


An overview. American Industrial Hygiene Association Journal, 62(3), 356.

9. Leith, D., Raynor, P. C., Boundy, M. G., Cooper, S. J. (1996). Performance of


Industrial Equipment to Collect Coolant Mist. American Industrial Hygiene
Association Journal, 57, 1142.

10. Leith, D., Leith, F. A., Boundy, M. G. (1996). Laboratory measurements of oil
mist concentrations using filters and an electrostatic precipitator. American
Industrial Hygiene Association Journal, 57 (12), 1137.

11. Barhate R.S., Ramakrishna S. (2007). Nanofibrous filtering media: Filtration


problems and solutions from tiny materials, Journal of Membrane Science 296, 1.

190 
 
12. Hei, W. A., Yacher, M., Deye, G. J., Spencer, A. B. (2000). Mist control at a
machining center. Part 1: mist characterization. American Industrial Hygiene
Association Journal, 61, 275.

13. http://www.parker.com/literature/Racor/7678%20(CCV%20Technical%20Brochu
re).pdf accessed on 26 April 2010.

14. Jandos, E., Lebrun, M., Brezezinski, C., Canizares, S. C. (2007). Filtration
Technologies in the Automotive Industry. Multifunctional barriers for flexible
structure, Berlin: Springer, 269.

15. Gunter, K. L., Sutherland, J. W. (1999). An experimental investigation into the


effects of process conditions on the mass concentration of cutting fluid mist in
turning. Journal of Cleaner Production, 7, 341.

16. Mehta, K., Chase, G. G., Liu, J., Rangarajan, S. (1999). Development of novel
filter media for coalescence and compressed air filtration. Advances in Filtration
and Separation Technology, 13(B), 933.

17. Fochtman, E. J., (1982). Coalescence of water from oil: Single fiber studies,
World Congress III, 654.

18. Jodi, W., Chase, G. G. (1982). Water-based polymer binders for coalescing filters,
Advances in Filtration and Separation Technology, American Filtration and
Separations Society, Vol 10, 40.

19. Read, B. (1998). Introduction to coalescence and separation, Advances in


filtration and separation technology, American Filtration and Separations Society,
Vol 12, 449.

20. Sherony, D. E., Kintner, R. C., Wasan, D. T. (1978). Coalescence of Secondary


Emulsions in Fibrous Bed. Surface Colloid Science, 10, 99.

21. Viraraghavan, T., Mathavan, G. N., Scoular G. N., Kurucz J. R. L. (1995)


Coalescence/filtration of water-in-oil emulsions, advances in filtration and
separation technology, American Filtration and Separations Society, Vol 9, 747.

22. Contal, P., Simao, J., Thomas, D., Frising, T. S., Callé, J. C. (2004). Clogging of
fiber filters by submicron droplets. Phenomena and influence of operating
conditions. Journal of Aerosol Science, 35 (2), 245.

23. Moldavsky, L., Fichman, M., Gutfinger, C. (2006). Enhancing the performance
of fibrous filters by means of acoustic waves. Aerosol Science 37, 528.

191 
 
24. Charvet, A., Gonthier, Y., Bernis, A., Gonze E. (2008). Filtration of liquid
aerosols with a horizontal fibrous filter, Chemical engineering research and
design, 86, 569.

25. Miller, D., Koslow, E. E., Williamson, K. M. (Jul 26, 1988). Coalescing filter for
removal of liquid aerosols from gaseous streams, US Patent 4759782.

26. Juvinall, R. A., Kessin, R. W., Steindler, M. J. (June 1978). Sand-bed filtration of
aerosol: a review of published information of their use in industrial and atomic
energy facilities, ANL-7683, Argonne National Laboratory, Argonne, IL.

27. Wang, J., Kim, S. C., Pui, D. Y. H. (2008). Figure of merit of composite filters
with micrometer and nanometer fibers, Aerosol Science and Technology, 42, 722.

28. Liu B. Y. H., Rubow, K. L. (June 1990). Efficiency, pressure drop and figure of
merit of high efficiency fibrous and membrane filter media, 5th World filtration
Congress, 112, Paris, France.

29. Brown, R.C. (1993). Air Filtration: An Integrated Approach to the Theory and
Applications of Fibrous Filters, Pergamon Press, New York.

30. Davies, C. N. (1973). Air Filtration. Academic Press Inc, London.

31. Davies, C. N. (1952). The separation of airborne dust and particles, Proceedings
of the Institute of Mechanical Engineers, 18, 185.

32. Stenhouse, J. I. T., Trottier, R. (1991). The Loading of Fibrous Filters with
Submicron Particles. Journal of Aerosol Science, 22 (Suppl. 1), 777.

33. Hunter, G. S., Chalmers, A. G. (Jul 14, 1992). Oil coalescing filter and filtering
process, US Patent 5129923.

34. Hunter, G. S. (July 16, 2002). Coalescing filters, US Patent 6419721 B1.

35. Hunter, A. G. (October 11, 1995). Filter for purification of gas, European patent
618835 B1.

36. Hunter, A. G. (October 14, 1999). Improvements in coalescing filters, WO 51319.

37. Spencer, D. M. (December 4, 2008). The improvements in coalescing filters, WO


146045 A1.

38. Waltl, H. G. A. (October 23, 2008). Improvements in coalescing filters, WO


125885 A2.

192 
 
39. Chokdeepanich, S. (2002). Coalescence filtration and drainage design. Ph.D.
dissertation, The University of Akron, Akron, Ohio, USA.

40. Bharadwaj, R., Patel, A., Chokdeepanich, S., Chase, G.G. (2008). Oriented fiber
filter media. Journal of Engineered Fibers and Fabrics, Special Issue 2008 –
Filtration, 29.

41. Chase, G.G., Beniwal, V., Venkataraman, C. (2000). Measurement of uni-axial


fiber angle in non-woven fibrous media. Chemical Engineering Science, 55 (12),
2151.

42. Andan S., Hariharan S. I., Chase G. G. (May 2008). Modeling of saturation in
coalescence filtration. American Filtration and Separations Society, 22th annual
conference, Valley Forge, PA, USA.

43. S. Andan, S. (2010). Modeling of drainage in coalescing filters. PhD Dissertation,


The University of Akron, Akron, Ohio, USA.

44. Andan, S., Hariharan, S. I., Chase, G. G. (2008). Continuum model evaluation of
the effect of saturation on coalescence filtration, Separation Science and
Technology, 43, 1955.

45. Richardson, J. G., Kerver, J. L., Hafford, J. A., Osaba, J. (1952). Laboratory
measurements of relative permeability, AIME transactions, 195.

46. Bitten, J. F., Fochtman, E. G. (1971). Water distribution in Fibrous bed


coalescers. Journal of Colloid and Interface Science, 37(2), 312.

47. Dawar, S., Chase, G. G. (2010). Correlations for transverse motion of liquid drops
on fibers. Separation and Purification Technology, 72, 282.

48. Dawar, S. Chase, G. G. (2008). Drag correlation for axial motion of drops on
fibers. Separation and Purification Technology 60, 6.

49. Mullins, B. J., Braddock, A., Roger, D. and Agranovski, I. E. (2004). Particle
Capture Processes and Evaporation on a Microscopic Scale in Wet Filters. Journal
of Colloid and Interface Science, 279, 213.

50. Mead-Hunter, R., Mullins, B. J., Kampa, D., Kasper, G. (2009). Predicting
saturation in coalescing filters-capillary based model. American Filtration and
Separations Society Conference, Bloomington, Minnesota, USA.

51. http://www.engr.utk.edu/mse/Textiles/Nanofiber%20Nonwovens.htm accessed on


28 May 2010.

193 
 
52. Kosmider K., Scott J. (July/August 2002). Polymeric nanofibers exhibit an
enhanced air filtration performance. Filtration + Separation. Featured article,
Vol.39 (6), 20.

53. Shin, C., Chase, G. G. (2006). Separation of liquid drops from air by glass fiber
filters augmented with polystyrene nanofibers. Journal of Dispersion Science and
Technology, 27, 5.

54. Dharmanolla S. (2007). A computer program for filter media design optimization.
Master’s thesis, The University of Akron, Akron, Ohio, USA.

55. Rangarajan S. (2001). Nanofibers in coalescer filter media. Master’s thesis, The
University of Akron, Akron, Ohio, USA.

56. Srinivasan, P., Chase, G. G. (2005). Steady State Filter Media Performance
Modeling with and without Nanofibers. American Filtration and Separations
Society: 18th Annual Conference, Atlanta, Georgia, USA.

57. George, J. (March 2007). Nanofiber manufacturing processes for filtration media.
American Filtration & Separation Society Annual Conference, Orlando, Florida,
USA.

58. Gopal R., Kaur, S., Maa, Z., Chanc, C., Ramakrishna, S., Matsuura, T. (2006).
Electrospun nanofibrous filtration membrane. Journal of Membrane Science 281,
581.

59. Yoon, K., Hsiao, B. S., Chu, B. (2008). Functional nanofibers for environmental
applications. Journal of Materials Chemistry, 18, 5326.

60. Wang, H., Fu, G, Li, X. (2009). Functional Polymeric Nanofibers from
Electrospinning. Recent Patents on Nanotechnology, 3, 21.

61. Zhang, S., Shim, W. S., Kim, S. J. (2009). Design of ultra-fine nonwovens via
electrospinning of Nylon 6: spinning parameters and filtration efficiency,
materials and design. Korean Journal of Chemical Engineering, 30, 3659.

62. Kim, G. T., Ahn, Y. C., Le, J. K. (2008). Characteristics of Nylon 6 nanofilter for
removing ultra fine particles. Korean Journal of Chemical Engineering, 25(2),
368.

63. Bitten, J. F., Fochtman, E. G. (1971). Water Distribution in Fibrous Bed


Coalescers. Journal of Colloid and Interface Science, 37(2), 312.

64. Happel J. (1959). Viscous flow relative to arrays of cylinders, American Institute
of Chemical Engineers, Journal 5 (2), 174.

194 
 
65. Kuwabara, S. (1959). The forces experienced by randomly distributed parallel
circular cylinders of spheres in a viscous flow at small Reynolds number. Journal
of the Physical Society of Japan, 14 (4), 527.

66. Hasimoto, H. (1959). On the periodic fundamental solutions of the Stokes


equations and their application to viscous flow past a cubic array of spheres.
Journal of Fluid Mechanics, 5 (A335), 355.

67. Spielman, L., Goren, S. L. (1968). Model for predicting pressure drop and
filtration efficiency in fibrous media. Environmental Science and Technology, 2,
279.

68. Krisch, A. A., Fuchs, N. A. (1967). The fluid flow in a system of parallel
cylinders perpendicular to the flow direction at small Reynolds numbers, Journal
of Fluid Mechanics, 22, 1251.

69. Davies, C. N. (1952). The separation of airborne dust and particle. Proceedings of
Institute of Mechanical Engineers, B1 185, London, United Kingdom.

70. Jackson G. W., James D. F. (1984). The permeability of fibrous porous media,
Canadian Journal of Chemical Engineering 64 (3), 364.

71. Jaganathan S., Vahedi Tafreshi H., Pourdeyhimi B. (2008). On the pressure drop
prediction of filter media composed of fibers with bimodal diameter distributions,
Powder Technology 181, 89.

72. Frising T., Thomas D., Contal1 P., Bé mer D., Leclerc D. (2003). Influence of
filter fiber size distribution on filter efficiency calculations, Trans I ChemE, Vol
81, Part A, 1179.

73. Brown R. C. (1984). A many-fiber model for airflow through a fibrous filter,
Journal of Aerosol Science Vol. 15, 583.

74. Walsh, D., Stenhouse, I. (1996). Experimental Studies of Electrically Active


Fibrous Filter Loading. Particle & Particle Systems Characterization, 13(1), 47.

75. Vasudevan G. (2005). Modeling and testing of the transient phase of coalescence
filtration PhD Dissertation, The University of Akron, Akron, Ohio, USA.

76. Vasudevan, G., Shin, C.G., Raber, B., Suthar A., Chase, G.G. (2002). Modeling
the startup stage of coalescence filtration, Fluid Particle Separation Journal, 14(3),
169.

195 
 
77. Vasudevan, G., S.I. Hariharan, S. I., Chase, G. G. (2005). Modeling the loading
stage in coalescence filtration, Journal of Porous Media, 8(3), 299.

78. Retrieved on 1 June 2010, from http://www.lubrizonl.com.

79. http://www.wesbond.com/westar+.htm accessed on 10 May 2010.

80. http://www.wesbond.com/megasol_s50.htm accessed on 12 April 2010.

81. Srinivasan, P. (2005). Nanofiber incorporated glass fiber filter media, The
University of Akron, Akron, Ohio, USA.

82. Jodi W. (1999). Permeability study on laboratory scale water-based coalescing


filter, Master’s thesis, The University of Akron, Akron, Ohio.

83. http://www.drgair.com/Sullube-MSDS.pdf accessed on 10 May 2010.

84. http://www2.dupont.com/Separation_Solutions/en_US/tech_info/spunbonded/spu
nbound_poly.html accessed on 24 May 2010.

85. http://www.ktnonwoven.com/ accessed on 24 May 2010.

86. http://www.spunfab.com/products/index.html accessed on 25 May 2010.

87. Cassie, A.B. D., Baxter, S. (1944). Wettability of Porous Surfaces. Transactions
of Faraday Society, 40, 546.

88. Chase, G. G. (April 2007). Improved microfiber filter performance by


augmentation with nanofibers, Society of Automotive Engineers, Detroit,
Michigan, USA.

196 
 
89. Davies C.N. (1948). Fibrous filters for dust and smoke, Proceedings of 9th
International Medical Congress, London, United Kingdom.

90. Taylor, G. I. (1964). Disintegration of water drops in an electric field.


Proceedings of Royal Society of London, London, United Kingdom.

91. Yarin, A. L., Koombhongse, S., Reneker, D. H. (2001). Taylor Cone and Jetting
from Liquid droplets in Electrospinning of Nanofibers, Physics of Fluids, 90 (9).

92. Reneker D. H., Yarin A.L., Fong H., Koombhongse S. (2000). Bending Instability
of electrically charged liquid jets of polymer solutions in electrospinning, Journal
of Applied Physics, 87, 4531.

93. Doshi, J., Reneker, D. H. (1995). Electrospinning process and application of


electrospun fibers. Journal of Electrostatics, 35, 151.

94. Renekar D. H., Yarin, A. L., Hao, F., Koombhongse. S. (2000). Bending
instability of electrically charged liquid jets of polymer solutions in
electrospinning. Journal of Applied Physics, 87, 4531.

95. Deitzwl, J. M., Kleinmeyer, J. D., Harris D., Becktan, N. C. (2001). The effect of
processing variables on the morphology of electrospun nanofibers and textiles.
Polymer, 42, 261.

96. Ahn, Y. C., Park, S. K., Kim, G. T., Hwang, Y. J., Lee, C. G., Shin H. S., Lee, J.
K. (2006). Development of high efficiency nanofilters made of nanofibers.
Current Applied Physics, 6 (6), 1030.

97. Pedicini A., Farris R. J. (2004). Thermally induced color change in electrospun
fiber mats. Journal of Polymer Science Part B: Polymer Physics, 42(5), 752.

197 
 
98. Deitze, J. M., Kosik, W., McKnight, S. H., Beak-Tan N. C., Desimone J. M.,
Crette S. (2002). Electrospinning of polymer nanofibers with specific surface
chemistry. Polymer, 43(3), 1025.

99. Dai, H., Gong, J., Kim, H. Y., Lee, D. R. (2002). A novel method for preparing
ultra-fine alumina-borate oxide fibres via an electrospinning technique,
Nanotechnology, 13, 674.

100. Fong, H., Reneker, D. H. (1999). Elastomeric nanofibers of styrene-butadiene-


styrene triblock copolymer, Journal of Polymer Science B: Polymer Physics 37,
3488.

101. http://www.nanopeutics.net/nanospider.html accessed on 31 May 2010.

102. http://www.elmarco.cz/technology/nanospider%3Csup%3Etm%3Csup%3E-
technology/ accessed on 31 May 2010.

103. Graham, K., Ouyang, M., Raether, T. Grafe, B. McDonald T., Knauf, P. (April
2002). Polymeric nanofibers in air filtration applications, 15th Annual Technical
Conference & Expo of the American Filtration & Separations Society, Galveston,
Texas, USA.

104. Bachmat, Y., Bear, J. (1972). Macroscopic Modeling of Transport Phenomena in


Porous Media. Transport in Porous Media, 1, 213.

105. Hassanizadeh S. M., Gray, W. G. (1980). General Conservation Equations for


Multi-phase Systems: 3. Constitutive Theory for Porous Media Flow, Advances in
Water Resources, 3 (1), 25.

106. Bear, J. (1972). Dynamics of Fluids in Porous Media, American Elsevier


Publishing Company, New York.

198 
 
107. Gray, W. G., Hassanizadeh, S. M. (1998). Macroscale Continuum Mechanics for
Multiphase Porous Media flow Including Phases, Interfaces, Common Lines and
Common Points, Advances in Water Resources, 21, 261.

108. Willis, M. S., Tosun, I., Choo, W., Chase, G. G., Desai, F. (1991). A Dispersed
Multiphase Theory and its Application to Filtration. In M. Y. Corapcioglu (Ed.),
Advances in porous media, Elsevier, Amsterdam.

109. Hassanizadeh, M., Gray, W. G. (1979). General Conservation Equations for


Multi-phase Systems: 1 Averaging Procedure, Advances in Water Resources, 2,
(3), 131.

110. Hassanizadeh, M. Gray, W. G. (1979). General Conservation Equations for Multi-


phase Systems: 2, Mass, Momentum, Energy, and Entropy Equations. Advances
in Water Resources, 2, (4), 191.

111. Whitaker, S. (1969). Advances in theory of fluid motion in porous media,


Industrial and Engineering Chemistry, 61(12), 14.

112. Dullien, F. A. L. (1992). Porous Media: Fluid Transport and Pore Structure,
Academic Press, San Diego.

113. Coulson, J. M., Richardson, J. F., Backhurst, J. R., Harker, J. H. (1991). Chemical
Engineering, Volume 2, Particle Technology and Separation Processes,
Pergamon, Oxford.

114. Guin, J. A., Kessler, D. P., Greenkorn, R. A. (1971). The permeability tensor for
anisotropic non-uniform porous media, Chemical Engineering Science, 26, 1475.

115. Slattery, J. C. (1981). Momentum, Energy, and Mass Transfer in Continua,


Krieger, Huntington.

199 
 
116. Knupp, P., Steinberg, S. (1993). Fundamentals of grid generation, CRC Press,
Bocaraton.

117. Spiegel, M. R. (1959). Vector Analysis, Schaum’s Outline Series, McGraw-Hill,


New York.

118. Eringen, A. C. (1980). Mechanics of Continua, Krieger, Huntington.

119. http://people.sc.fsu.edu/~burkardt/f_src/vector_plot/vector_plot.html accessed on


16 June, 2010.
 

200 
 
APPENDIX A
EXPERIMENTAL RESULTS OF DRAINAGE CHANNEL INCORPORATED FILTER
GEOMETRIES BY USING CARBOSET 560 BINDER

Raw experimental data


The filter properties in terms of permeability, porosity and strength as well as the
filter performance in terms of capture efficiency, pressure drop, quality factor and
saturation is given below. The data for no-drainage, inlet-drainage, outlet-drainage in
horizontal orientation and the no drainage channel experiments in the vertical orientation
are shown in table A.1, A.2, A.3, and A.4 respectively.

Table A.1: No-drainage channel media in horizontal orientation

Permeability Porosity Strength Efficiency Pressure QF Saturation


(m2) (%) drop (1/kPa)
(kPa)

1.65E-10 0.96 15 86.861 9.24 0.17 0.37


1.28E-10 0.96 15 84.170 9.61 0.15 0.36
1.68E-10 0.96 16 88.368 9.76 0.16 0.37

Table A.2: Inlet-drainage channel media in horizontal orientation

Permeability Porosity Strength Efficiency Pressure QF Saturation


(m2) (%) drop (1/kPa)
(kPa)

1.07E-10 0.96 15 97.163 10.60 0.35 0.27


1.79E-10 0.96 16 96.917 10.62 0.36 0.26
2.13E-10 0.96 15 97.623 10.61 0.33 0.27

201 
 
Table A.3: Outlet-channel drainage media in horizontal orientation

Permeability Porosity Strength Efficiency Pressure QF Saturation


(m2) (%) drop (kPa) (1/kPa)

3.81E-10 0.95 16 98.112 10.65 0.36 0.26


1.35E-10 0.96 17 95.516 10.66 0.35 0.26
1.06E-10 0.95 16 94.390 10.63 0.35 0.27

Table A.4: No-drainage channel media in vertical orientation

Permeability Porosity Strength Efficiency Pressure QF Saturation


(m2) (%) drop (1/kPa)
(kPa)

1.49E-10 0.96 17 98.933 10.51 0.45 0.24


1.01E-10 0.96 15 98.859 10.53 0.44 0.25
1.17E-10 0.96 15 98.669 10.48 0.44 0.24

202 
 
APPENDIX B
FORTRAN CODE FOR DRAINAGE CHANNEL INCORPORATED FILTER MEDIA
 

! PROGRAM THREE DRAINAGE CHANNELS


! George G Chase and Shagufta U Patel

! THIS PROGRAM SOLVES THE 2D MASS AND MOMENTUM BALANCES FOR


! PERMEABLE FLOW THROUGH A POROUS MEDIUM WITH THREE DRAINAGE
! CHANNELS
! IMPERMEABLE
! SURFACE
! X07, Y07 +**********+ X17, Y17 THE DIAGRAM SHOWS THE
! +...................+ LOCATION OF THE CORNER
! +...................+ POINTS OF EACH REGION
! +...................+ OUTLET REGIONS: MEDIUM, CHANNEL
! X06, Y06 ++.MEDIUM.+ PRESSURE
! +.+.............+ SURFACE
! X05, Y05 +.+..+...........+ POUT
! +...+..+........+ MEDIUM => REGION 7
! INLET +.....+..+......+
! PRESSURE +.......+..+....+
! SURFACE +.........+..+..+
! PIN +...........+..++ X16,Y16
! X04, Y04 ++............+.+ CHANNEL => REGION 6
! +..+............+ X15,Y15
! X03, Y03 +.+..+..........+
! +...+..+........+
! +.....+..+......+ MEDIUM => REGION 5
! +.......+..+....+
! +.........+..+..+
! X02, Y02 ++..........+..++ X14, Y14
! +.+...........+.+ CHANNEL => REGION 4
! X01, Y01 ++..+...........+ X13, Y13
! +..+..+.........+
! +....+..+.......+ MEDIUM => REGION 3
! +......+..+.....+
! +........+..+...+
! +..........+..+.+ X12,Y12
! +MEDIUM.+. CHANNEL => REGION 2
! +...............++ X11, Y11

203 
 
! +................+
! +................+ MEDIUM => REGION 1
! X00, Y00 +********+ X10,Y10
! IMPERMEABLE SURFACE

! THE PROGRAM IS BASED UPON THE PROGRAMS DESCRIBED IN


! P KNUPP, S STEINBERG, FUNDAMENTALS OF GRID GENERATION,
! CRC PRESS, BOCA RATON, 1993.
! RUN THIS PROGRAM TO CALCULATE THE PRESSURE FIELD AND THE VELOCITY
! PROFILES. RUN VECTOR_PLOT TO CREATE FILES THAT SHOW THE VELOCITY
VECTORS.
! THE PROGRAM SOLVES THE MASS AND MOMENTUM BALANCES IN THE FORM
!
! MASS del(por v) = 0 por=porosity, v=pore avg velocity
!
! MOMENTUM por v = PERMX K del P / vis DARCY'S LAW
! k=dimensionless permeability = perm/permx
! P=pressure
! |K11 K12| vis = viscosity
! K= |K21 K22| K=DIMENSIONLESS anisotropic permeability
! F=P/PDROP dimensionless pressure
!
! COMBINED EQUATIONS
! del( k del f)=g g=forcing function (g=0 in this application)
!
! THE ANISOTROPIC PERMEABILITY IS PROGRAMMED AS A FUNCTION OF
POSITION
! WHERE
! K11=ALP(X,Y), K12=K21=BET(X,Y), K22=GAM(X,Y)
!
! THE COMPUTER CODE SOLVES THE COMBINED EQUATION IN THE
FORM
! L(F) = g (where g=0)
! WHERE
! L(F) = (alp * f_x)_x f_x is df/dx, etc.
! + (bet * f_x)_y F IS THE DIMENSIONLESS PRESSURE
! + (bet * f_y)_x
! + (gam * f_y)_y
!
! THE VELOCITY PROFILE IS DETERMINED FROM THE PRESSURE FIELD
! BY APPLYING DARCY'S LAW

! UNITS
!PRESSURE IS IN Pa BUT F = DIMENSIONLESS PRESSURE, F = P/PRESSURE DROP
! X, Y LENGTHS IN m
! VISCOSITY IN kg/m/s
! PERMEABILITY IN m^2 ALP, BET, GAM = DIMENSIONLESS PERMEABILITY
! = PERM/PERMX
! VELOCITY IN m/s
!*****************************************************************************
! MAIN PROGRAM

204 
 
! TO SIMPLIFY APPLYING THIS PROGRAM, THE MAIN PROGRAM IS USED
! SET THE GEOMETRY, AND THEN TO CALL THE SUBROUTINES THAT
! CALCULATE PRESSURE AND VELOCITY
!*******************************************************
PARAMETER(MMAX=280,NMAX=280,NREG=7) !SET THE MAXIMUM GRID
SIZE AND ! NO.OF REGIONS
IMPLICIT INTEGER (I-N)
IMPLICIT DOUBLE PRECISION (A-H,O-Z)
! VARIABLES
! NUMERICAL SOLUTION: F (PRESSURE), VX,VY (VELOCITY COMPONENTS)
DIMENSION
F(0:MMAX,0:NMAX),VX(0:MMAX,0:NMAX),VY(0:MMAX,0:NMAX)
DIMENSION G(0:MMAX,0:NMAX) !FORCING FUNCTION
! THE GRID
DIMENSION X(0:MMAX,0:NMAX),Y(0:MMAX,0:NMAX),XC(0:1),YC(0:1,0:NREG)
DIMENSION N(0:NREG)
! NREG,M,DXI,DET(NREG)
! TRANSFORMED VARIABLES
DIMENSION ALPHAT(0:MMAX,0:NMAX), BETHAT(0:MMAX,0:NMAX),
GAMHAT(0:MMAX,0:NMAX), GHAT(0:MMAX,0:NMAX)
! STENCILS
DIMENSION STN_C(0:MMAX,0:NMAX), STN_W(0:MMAX,0:NMAX),
STN_E(0:MMAX,0:NMAX)
DIMENSION STN_S(0:MMAX,0:NMAX),STN_N(0:MMAX,0:NMAX)
! OTHER VARIABLES
DIMENSION POR(NREG),DET(NREG)
! NCNTMAX,TOL, OMG, PDROP
integer alpha
! SET THE GRID SPACE AND GEOMETRY
N(0)=0
N(1)=40 !UPPER BOUND FOR REGION 1
N(2)=80 ! REGION 2
N(3)=120 ! REGION 3
N(4)=160 ! REGION 4
N(5)=200 ! REGION 5
N(6)=240 ! REGION 6
N(7)=280 ! REGION 7 N(NREG)<NMAX
M=40 !WIDTH GRID SIZE, M<= MMAX
!XY POSITION OF CORNER POINTS IN METERS
XC(0)=0.0D0 ! X POSITIONS OF CORNER POINTS
XC(1)=0.014D0 ! ALL OTHER XC POINTS ARE SAME AS X00 OR X10

!READ ANGLE OF THE DRAINAGE CHANNEL


write(6,*)"Choose angle of the drainage channel=> 0, 15, 30 ,45 or 60"
read(5,*)alpha

if(alpha.eq.0)then

YC(0,0)=0.0D0 !Y POSITIONS OF CORNER POINTS


YC(0,1)=0.01475D0
YC(0,2)=0.01525D0

205 
 
YC(0,3)=0.02975D0
YC(0,4)=0.03025D0
YC(0,5)=0.04475D0
YC(0,6)=0.04525D0
YC(0,7)=0.06D0

YC(1,0)=Y(0,0)
YC(1,1)=0.01475D0
YC(1,2)=0.01525D0
YC(1,3)=0.02975D0
YC(1,4)=0.03025D0
YC(1,5)=0.04475D0
YC(1,6)=0.04525D0
YC(1,7)=0.060D0

PDROP=2150.0D0 ! PRESSURE DROP, Pa


Angle=atan((XC(1)-XC(0))/(YC(0,1)-YC(1,1))) !Angle in radians

else if(alpha.eq.15)then

YC(0,0)=0.0D0 !Y POSITIONS OF CORNER POINTS


YC(0,1)=0.014D0
YC(0,2)=0.0145D0
YC(0,3)=0.03162D0
YC(0,4)=0.03212D0
YC(0,5)=0.04925D0
YC(0,6)=0.04975D0
YC(0,7)=0.06D0

YC(1,0)=Y(0,0)
YC(1,1)=0.01025D0
YC(1,2)=0.01075D0
YC(1,3)=0.02787D0
YC(1,4)=0.02837D0
YC(1,5)=0.0455D0
YC(1,6)=0.0460D0
YC(1,7)=0.060D0

PDROP=2150.0D0 ! PRESSURE DROP, Pa


Angle=atan((XC(1)-XC(0))/(YC(0,1)-YC(1,1))) !Angle in radians

else if(alpha.eq.30)then

YC(0,0)=0.0D0 !Y POSITIONS OF CORNER POINTS


YC(0,1)=0.01835D0
YC(0,2)=0.01885D0
YC(0,3)=0.0338D0
YC(0,4)=0.0343D0
YC(0,5)=0.04925D0
YC(0,6)=0.04975D0
YC(0,7)=0.06D0

206 
 
YC(1,0)=Y(0,0)
YC(1,1)=0.01025D0
YC(1,2)=0.01075D0
YC(1,3)=0.0257D0
YC(1,4)=0.0262D0
YC(1,5)=0.04115D0
YC(1,6)=0.04165D0
YC(1,7)=0.060D0

PDROP=2150.0D0 ! PRESSURE DROP, Pa


Angle=(atan((XC(1)-XC(0))/(YC(0,1)-YC(1,1)))) !Angle in radians

else if(alpha.eq.60)then

YC(0,0)=0.0D0 !Y POSITIONS OF CORNER POINTS


YC(0,1)=0.0293D0
YC(0,2)=0.0298D0
YC(0,3)=0.0419D0
YC(0,4)=0.0424D0
YC(0,5)=0.0545D0
YC(0,6)=0.055D0
YC(0,7)=0.06D0

YC(1,0)=Y(0,0)
YC(1,1)=0.005D0
YC(1,2)=0.0055D0
YC(1,3)=0.0176D0
YC(1,4)=0.0181D0
YC(1,5)=0.0302D0
YC(1,6)=0.0307D0
YC(1,7)=0.060D0

PDROP=2150.0D0 ! PRESSURE DROP, Pa


Angle=(atan((XC(1)-XC(0))/(YC(0,1)-YC(1,1)))) !Angle in radians

else if(alpha.eq.45)then

YC(0,0)=0.0D0 !Y POSITIONS OF CORNER POINTS


YC(0,1)=0.01925D0
YC(0,2)=0.01975D0
YC(0,3)=0.03675D0
YC(0,4)=0.03725D0
YC(0,5)=0.05425D0
YC(0,6)=0.05475D0
YC(0,7)=0.06D0

YC(1,0)=Y(0,0)
YC(1,1)=0.00525D0
YC(1,2)=0.00575D0
YC(1,3)=0.02275D0

207 
 
YC(1,4)=0.02325D0
YC(1,5)=0.04025D0
YC(1,6)=0.04075D0
YC(1,7)=0.060D0

PDROP=2150.0D0 ! PRESSURE DROP, Pa


Angle=(atan((XC(1)-XC(0))/(YC(0,1)-YC(1,1)))) !Angle in radians

end if

! CALCULATE THE GRID SPACING


XM=M
DXI=1.0D0/XM
DO K = 1,NREG
XN=N(K)-N(K-1)
DET(K)=1.0D0/XN
ENDDO
!
! SOLVER PARAMETERS
NCNTMAX=5000000 ! MAXIMUM NUMBER OF ITERATIONS IN THE
CALCULATIONS
TOL = 1.D-12 ! ERROR TOLERANCE FOR CONVERGENCE OF POINTS IN
ITERATIONS
OMG = 1.2D0 ! RELAXATION PARAMETER TO SPEED UP CONVERGENCE
!
! CREATE THE GRID FOR EACH REGION
DO K=1,NREG
CALL
GRID_GEN(M,MMAX,N,NMAX,NREG,X,Y,DXI,DET,XC,YC,K)

ENDDO
!
! SET THE FORCING FUNCTION, G
DO J=0,N(NREG)
DO I=0,M
G(I,J)=0.0D0
ENDDO
ENDDO
!
! MEDIUM MATERIAL AND OPERATING CONDIIONS

POR(1)=0.96D0 ! POROSITY OF EACH REGION


POR(2)=0.9864D0
POR(3)=0.96D0
POR(4)=0.9864D0
POR(5)=0.96D0
POR(6)=0.9864D0
POR(7)=0.96D0
VIS=0.00001D0 ! VISCOSITY, Kg/m/s, AIR
PERMX=3.389D-09 ! PERMEABILITY SCALE FACTOR TO MAKE
! DIMENSIONLESS PERMEABILITY ABOUT 0 - 1

208 
 
write(6,*)"Calculating..."

! SET UP FIXED BOUDARY CONDITIONS


DO J=0,N(NREG)
F(0,J)=1.0D0 !INLET DIMENSIONLESS PRESSURE
F(M,J)=0.0D0 !OUTLET DIMENSIONLESS PRESSURE
ENDDO
!
! INITIAL GUESS FOR F INTERIOR POINTS
DO I=1,M-1
DO J=0,N(NREG) !ASSUME A LINEAR APPROXIMATION
D=I ! DISTANCE=D, SLOPE=(FMJ-F0J)/1.0
D=D*DXI ! INTERCEPT = F0J
F(I,J)=(F(M,J)-F(0,J))*D+F(0,J)
ENDDO
ENDDO
!
! CALCULATE THE TRANSFORMED COEFFICIENTS
DO K=1,NREG
CALL TRANS(M,MMAX,N,NMAX,NREG,X,Y,G,DXI,DET, &
ALPHAT,BETHAT,GAMHAT,GHAT,K,ANGLE)
ENDDO

! LOAD THE STENCILS


DO K=1,NREG
call stencils(m, mmax, n, nmax,NREG,K,&
alphat, bethat, gamhat,det,dxi,&
stn_c,&
stn_w, stn_e, stn_s, stn_n,&
stn_ne, stn_nw, stn_se, stn_sw)
ENDDO
!
! SOLVE THE LINEAR SYSTEM OF EQUATIONS
call sor_9pt(m, mmax, n, nmax, NREG, ncntmax, tol, omg,&
stn_c,&
stn_w, stn_e, stn_s, stn_n,&
stn_ne, stn_nw, stn_se, stn_sw, ghat, f,X,Y,&
ANGLE)

!
! CALCULATE THE VELOCITY PROFILES
call velocity(m,mmax,n,nmax,NREG,f,x,y,&
det,dxi,PDROP,POR,PERMX,VIS,ANGLE)
!
! OUTPUT RESULTS
call results(m,mmax,n,nmax,NREG,f)
!
end
!
! ********************************************************************

209 
 
function alp(N,NREG,J,ANGLE) !external function
! Compute a coefficient of the differential equation.
! ALP = K11 = DIMENSIONLESS PERMEABILITY COMPONENT
! Input:
! The point:
double precision XK1, XK2, THETA, ANGLE
integer N(0:NREG)
! Output:
! The coefficients.
double precision alp
! Internal:

if (J.le.N(1)) then
THETA=0.0d0
XK2=0.132D0
XK1=0.066D0
else if(J.le.N(2)) then
THETA=ANGLE
XK1=1.0D0
XK2=1.4D0
else if(J.le.N(3)) then
THETA=0.0d0
XK2=0.132D0
XK1=0.066D0
else if(J.le.N(4)) then
THETA=ANGLE
XK1=1.0D0
XK2=1.4D0
else if(J.le.N(5)) then
THETA=0.0d0
XK2=0.132D0
XK1=0.066D0
else if(J.le.N(6)) then
THETA=ANGLE
XK1=1.0D0
XK2=1.4D0
else if(J.le.N(7)) then
THETA=0.0d0
XK2=0.132D0
XK1=0.066D0

end if
alp=(cos(THETA)**2)*XK1+(sin(THETA)**2)*XK2
return
end
!
! ******************************************************************
function bet(N,NREG,J,ANGLE) !external function
! Compute a coefficient of the differential equation.
! BET = K12=K21 = DIMENSIONLESS PERMEABILITY COMPONENTS
! Input:

210 
 
! The point:
double precision XK1, XK2, THETA, ANGLE
integer N(0:NREG)
! Output:
! The coefficient
double precision bet
! Internal:
!
if (J.le.N(1)) then
THETA=0.0d0
XK2=0.132D0
XK1=0.066D0
else if(J.le.N(2)) then
THETA=ANGLE
XK1=1.0D0
XK2=1.4D0
else if(J.le.N(3)) then
THETA=0.0d0
XK2=0.132D0
XK1=0.066D0
else if(J.le.N(4)) then
THETA=ANGLE
XK1=1.0D0
XK2=1.4D0
else if(J.le.N(5)) then
THETA=0.0d0
XK2=0.132D0
XK1=0.066D0
else if(J.le.N(6)) then
THETA=ANGLE
XK1=1.0D0
XK2=1.4D0
else if(J.le.N(7)) then
THETA=0.0d0
XK2=0.132D0
XK1=0.066D0
end if
bet=(cos(THETA)*sin(THETA)*(XK2-XK1))
return
end
!
!
******************************************************************************
*
function gam(N,NREG,J,ANGLE) !external function
! GAM = K22 DIMENSIONLESS PERMEABILITY COMPONENT
!
! Input:
! The point:
double precision XK1, XK2, THETA, ANGLE
integer N(0:NREG)

211 
 
! Output:
! The coefficient.
double precision gam
! Internal:
!
if (J.le.N(1)) then
THETA=0.0d0
XK2=0.132D0
XK1=0.066D0
else if(J.le.N(2)) then
THETA=ANGLE
XK1=1.0D0
XK2=1.4D0
else if(J.le.N(3)) then
THETA=0.0d0
XK2=0.132D0
XK1=0.066D0
else if(J.le.N(4)) then
THETA=ANGLE
XK1=1.0D0
XK2=1.4D0
else if(J.le.N(5)) then
THETA=0.0d0
XK2=0.132D0
XK1=0.066D0
else if(J.le.N(6)) then
THETA=ANGLE
XK1=1.0D0
XK2=1.4D0
else if(J.le.N(7)) then
THETA=0.0d0
XK2=0.132D0
XK1=0.066D0
end if
gam=(sin(THETA)**2)*XK1+(cos(THETA)**2)*XK2
return
end
!
! **************************************************************************
SUBROUTINE GRID_GEN(M,MMAX,N,NMAX,NREG,X,Y,DXI,DET,XC,YC,K)
!
! GENERATES XY GRID LOCATIONS MAPPED TO IJ OF LOGICAL SPACE
!
IMPLICIT INTEGER (I-N)
IMPLICIT DOUBLE PRECISION (A-H,O-Z)
DIMENSION X(0:MMAX,0:NMAX),Y(0:MMAX,0:NMAX)
DIMENSION DET(NREG),XC(0:1),YC(0:1,0:NREG),N(0:NREG)
!
X00=XC(0) !SET CORNER VALUES (POSITION IN METERS)
X01=XC(0)
X10=XC(1)

212 
 
X11=XC(1)
Y00=YC(0,K-1)
Y01=YC(0,K)
Y10=YC(1,K-1)
Y11=YC(1,K)
!
!SET INTERIOR VALUES

DO J=N(K-1),N(K)
ET=J-N(K-1)
ET=ET*DET(K)
DO I=0,M
XI=I
XI=XI*DXI
X(I,J)=(1.0D0-XI)*(1.0D0-ET)*X00+(1.0D0-XI)*ET*X01 &
+XI*(1.0D0-ET)*X10 +XI*ET*X11
Y(I,J)=(1.0D0-XI)*(1.0D0-ET)*Y00+(1.0D0-XI)*ET*Y01 &
+XI*(1.0D0-ET)*Y10 +XI*ET*Y11
ENDDO
ENDDO
RETURN
END
!
! ************************************************************************
SUBROUTINE SOR_9PT(M,MMAX,N,NMAX,NREG,NCNTMAX,TOL,OMG,&
stn_c,&
stn_w, stn_e, stn_s, stn_n,&
stn_ne, stn_nw, stn_se, stn_sw, g, f, X,Y,&
ANGLE)
!
! iterates on a 9-point stencil in one independent variable, which arises from an equation of the
form:
! stn_w*f(i-1,j) + stn_e*f(i+1,j) + stn_s*f(i,j-1) + stn_n*f(i,j+1) +
! stn_ne*f(i+1,j+1) + stn_nw*f(i-1,j+1) + stn_se*f(i+1,j-1) +
! stn_sw*f(i-1,j-1) + stn_c*f(i,j) = g
!
integer m, mmax, n(0:NREG), nmax, i, j,K,L,LSTOP,NREG
integer ncntmax, ncnt,kk,jnear
double precision tol, omg
double precision xerr, sum, ftmp, error
double precision ANGLE
!
double precision stn_c(0:mmax,0:nmax)
double precision stn_w(0:mmax,0:nmax), stn_e(0:mmax,0:nmax)
double precision stn_s(0:mmax,0:nmax), stn_n(0:mmax,0:nmax)
double precision stn_ne(0:mmax,0:nmax), stn_nw(0:mmax,0:nmax)
double precision stn_se(0:mmax,0:nmax), stn_sw(0:mmax,0:nmax)
double precision g(0:mmax,0:nmax), f(0:mmax,0:nmax)
DOUBLE PRECISION X(0:MMAX,0:NMAX),Y(0:MMAX,0:NMAX)
DOUBLE PRECISION A,B,FTMP1,FTMP2,X1,Y1,X2,Y2,DX1,DX2,DY1,DY2,D
!

213 
 
ncnt = 0
1 ncnt = ncnt+1 !iteration count
xerr = 0.
!
! SET IMPERMEABLE LOWER BOUNDARY AT Y(0,J)
DO I=1,M-1
F(I,0)=F(I,1)
ENDDO
!
DO K=1,NREG ! SOLVE ONE REGION AT A TIME
! CALCULATE INTERIOR POINTS
DO J=N(K-1)+1,N(K)-1
DO I=1,M-1
sum = stn_w(i,j)*f(i-1,j)+&
stn_e(i,j)*f(i+1,j)+&
stn_s(i,j)*f(i,j-1)+&
stn_n(i,j)*f(i,j+1) +&
stn_ne(i,j)*f(i+1,j+1)+&
stn_nw(i,j)*f(i-1,j+1)+&
stn_se(i,j)*f(i+1,j-1)+&
stn_sw(i,j)*f(i-1,j-1)
!
ftmp = ( g(i,j) - sum )/stn_c(i,j)
!
ftmp = f(i,j) + omg*( ftmp - f(i,j) )
error=abs( (ftmp-f(i,j)))
xerr = max( xerr, error)
f(i,j) = ftmp
!
ENDDO
ENDDO
! SET UPPER BOUNDARY OF K REGION
IF(K.EQ.NREG)THEN
! SET IMPERMEABLE UPPER BOUNDARY AT Y(NMAX,J)
DO I=1,M-1
F(I,NMAX)=F(I,NMAX-1)
ENDDO
ELSE ! SET MASS JUMP BAL FOR INTERMEDIATE BOUNDARY
J=N(K)
DO I=1,M-1
DX1=X(I,J)-X(I-1,J)
DY1=Y(I,J)-Y(I,J-1)
DX2=X(I+1,J)-X(I,J)
DY2=Y(I,J+1)-Y(I,J)
!FIND FTMP1
LSTOP=N(K-1)
DO L=J,N(K-1),-1
IF(Y(I-1,L+1).GE.Y(I,J).AND.&
Y(I-1,L).LE.Y(I,J))LSTOP=L
ENDDO
D=(Y(I-1,LSTOP+1)-Y(I,J))/(Y(I-1,LSTOP+1)-Y(I-1,LSTOP))

214 
 
FTMP1=D*F(I-1,LSTOP)+(1.0D0-D)*F(I-1,LSTOP+1)
!FIND FTMP2
LSTOP=N(K+1)
DO L=J,N(K+1)
IF(Y(I+1,L+1).GE.Y(I,J).AND.&
Y(I+1,L).LE.Y(I,J))LSTOP=L
ENDDO
D=(Y(I+1,LSTOP+1)-Y(I,J))/(Y(I+1,LSTOP+1)-
Y(I+1,LSTOP))
FTMP2=D*F(I+1,LSTOP)+(1.0D0-D)*F(I+1,LSTOP+1)
! CALCULATE BOUNDARY PRESSURE
!
A=((BET(N,NREG,J-1,ANGLE)/DX1&
+GAM(N,NREG,J-1,ANGLE)/DY1&
+BET(N,NREG,J+1,ANGLE)/DX2&
+GAM(N,NREG,J+1,ANGLE)/DY2))
B=((BET(N,NREG,J-1,ANGLE)*FTMP1/DX1&
+GAM(N,NREG,J-1,ANGLE)*F(I,J-1)/DY1&
+BET(N,NREG,J+1,ANGLE)*FTMP2/DX2&
+GAM(N,NREG,J+1,ANGLE)*F(I,J+1)/DY2))
F(I,J)=B/A
ENDDO
ENDIF
ENDDO
if ( (xerr .gt. tol) .and. (ncnt .lt. ncntmax) ) go to 1
!
if (xerr .gt. tol) then
print*, ""
write(6,*) ' sor tol not satisfied after ', ncntmax
pause
else
print*, ""
write(6,*) ' sor tol satisfied after ', ncnt
pause
end if
!
return
end
!
! *************************************************************************
subroutine stencils(m, mmax, n, nmax,NREG,K,&
alphat, bethat, gamhat,det,dxi,&
stn_c,&
stn_w, stn_e, stn_s, stn_n,&
stn_ne, stn_nw, stn_se, stn_sw)
!
! Initializes the stencils.
! Input:
! Array and problem sizes and indices.
integer m, mmax, N(0:NREG), nmax, i, j,NREG,K
! The transformed coefficients:

215 
 
double precision alphat(0:mmax,0:nmax)
double precision bethat(0:mmax,0:nmax)
double precision gamhat(0:mmax,0:nmax)
! Output:
! The stencils:
double precision stn_c(0:mmax,0:nmax)
double precision stn_w(0:mmax,0:nmax), stn_e(0:mmax,0:nmax)
double precision stn_s(0:mmax,0:nmax), stn_n(0:mmax,0:nmax)
double precision stn_ne(0:mmax,0:nmax), stn_nw(0:mmax,0:nmax)
double precision stn_se(0:mmax,0:nmax), stn_sw(0:mmax,0:nmax)
! Internal variables.
double precision dxi, det(NREG)
!
!
DO J = N(K-1)+1, N(K)-1
DO I = 1, M-1
stn_w(i,j) = alphat(i,j)/(dxi**2)
stn_e(i,j) = alphat(i+1,j)/(dxi**2)
stn_s(i,j) = gamhat(i,j)/(det(K)**2)
stn_n(i,j) = gamhat(i,j+1)/(det(K)**2)
stn_ne(i,j) = bethat(i+1,j+1)/(2.0d0*dxi*det(K))
stn_nw(i,j) = - bethat(i,j+1)/(2.0d0*dxi*det(K))
stn_se(i,j) = - bethat(i+1,j)/(2.0d0*dxi*det(K))
stn_sw(i,j) = bethat(i,j)/(2.0d0*dxi*det(K))
stn_c(i,j) = - ( stn_w(i,j) + stn_e(i,j) + &
stn_s(i,j) + stn_n(i,j) + &
stn_ne(i,j) + stn_nw(i,j) + &
stn_se(i,j) + stn_sw(i,j) )
ENDDO
ENDDO
!
RETURN
END
!
! *********************************************************************
SUBROUTINE TRANS(M,MMAX,N,NMAX,NREG,X,Y,G,DXI,DET,&
ALPHAT,BETHAT,GAMHAT,GHAT,K,&
ANGLE)
!
! Compute the transfomred coefficients.
!
! Input:
! Array size, problem size, and indices.
integer m, mmax, n(0:NREG), nmax, i, j,K,NREG
! The transformation.
double precision x(0:mmax,0:nmax), y(0:mmax,0:nmax)
! The coefficient definitions. external alp, bet, gam
double precision alp, bet, gam
double precision ANGLE
! The right-hand side.
double precision g(0:mmax,0:nmax)

216 
 
! Output:
! The transformed coefficients:
double precision alphat(0:mmax,0:nmax)
double precision bethat(0:mmax,0:nmax)
double precision gamhat(0:mmax,0:nmax)
! The forcing term:
double precision ghat(0:mmax,0:nmax)
!
double precision dxi, det(NREG), x1, x2, y1, y2, jac
!
! Compute the alpha hat coefficient for the interior points.
! Here i <-- (i-1/2).

DO J=N(K-1)+1,N(K)-1
DO I=1,M
x1 = ( + x(i,j) - x(i-1,j) )/dxi
x2 = ( + x(i,j+1) - x(i,j-1) + x(i-1,j+1)&
- x(i-1,j-1) )/(4.0d0*det(K))
y1 = ( + y(i,j) - y(i-1,j) )/dxi
y2 = ( + y(i,j+1) - y(i,j-1) + y(i-1,j+1)&
- y(i-1,j-1) )/(4.0d0*det(K))
jac = x1 * y2 - y1 * x2

alphat(i,j) = (alp(N,NREG,J,ANGLE)*y2*y2&
- 2.0d0*bet(N,NREG,J,ANGLE)*x2*y2&
+ gam(N,NREG,J,ANGLE)*x2*x2)/jac
ENDDO
ENDDO
!
! Compute the gamma hat coefficient.
! Here j <-- (j-1/2).

DO J=N(K-1)+1,N(K)
DO I=1,M
x1 = ( + x(i+1,j) - x(i-1,j) + x(i+1,j-1)&
- x(i-1,j-1) )/(4.0d0*dxi)
x2 = ( + x(i,j) - x(i,j-1))/det(K)
y1 = ( + y(i+1,j) - y(i-1,j) + y(i+1,j-1)&
- y(i-1,j-1) )/(4.0d0*dxi)
y2 = ( + y(i,j) - y(i,j-1) )/det(K)
jac = x1 * y2 - y1 * x2

gamhat(i,j)=(alp(N,NREG,J,ANGLE)*y1*y1&
- 2.0d0*bet(N,NREG,J,ANGLE)*x1*y1&
+ gam(N,NREG,J,ANGLE)*x1*x1)/jac

ENDDO
ENDDO
!
! Compute the beta hat coefficient.
! Here i <-- (i-1/2), j <-- (j-1/2).

217 
 
DO J=N(K)+1,N(K)
DO I=1,M

x1 = ( + x(i,j) + x(i,j-1) - x(i-1,j) - x(i-1,j-1))/(2.0d0*dxi)


x2 = ( + x(i,j) - x(i,j-1) + x(i-1,j) - x(i-1,j-1))/(2.0d0*det(K))
y1 = ( + y(i,j) + y(i,j-1) - y(i-1,j) - y(i-1,j-1))/(2.0d0*dxi)
y2 = ( + y(i,j) - y(i,j-1) + y(i-1,j) - y(i-1,j-1))/(2.0d0*det(K))
jac = x1 * y2 - y1 * x2

bethat(i,j) = - (alp(N,NREG,J,ANGLE)*y1*y2&
- bet(N,NREG,J,ANGLE)*(x1*y2+x2*y1)&
+ gam(N,NREG,J,ANGLE)*x1*x2)/jac
ENDDO
ENDDO
!
! Compute the g hat coefficient.

DO J=N(K-1)+1,N(K)-1
DO I=1,M-1
x1 = (x(i+1,j)-x(i-1,j))/(2.0d0*dxi)
x2 = (x(i,j+1)-x(i,j-1))/(2.0d0*det(K))
y1 = (y(i+1,j)-y(i-1,j))/(2.0d0*dxi)
y2 = (y(i,j+1)-y(i,j-1))/(2.0d0*det(K))
jac = x1 * y2 - y1 * x2
ghat(i,j) = jac*g(i,j)

ENDDO
ENDDO
RETURN
END
!
! ***************************************************************************
subroutine results(m,mmax,n,nmax,NREG,f)
integer m, mmax, N(0:NREG),NREG, nmax, i, j
double precision f(0:mmax,0:nmax)
!
OPEN(UNIT=11, FILE='Pressure.OUT',STATUS='UNKNOWN')
! initially, just write results to screen
do i=0,m
do j=0,N(NREG)
write(11,*)i,f(i,j)
write(6,*)i,j,f(i,j)

enddo
enddo
pause
return
end
!**********************************************************************

218 
 
subroutine velocity
(m,mmax,n,nmax,NREG,f,x,y,det,dxi,PDROP,POR,PERMX,VIS,ANGLE)
IMPLICIT DOUBLE PRECISION (A-H,O-Z)
IMPLICIT INTEGER (I-N)
DIMENSION f(0:mmax,0:nmax)
DIMENSION x(0:mmax,0:nmax), y(0:mmax,0:nmax)
DIMENSION vx(0:mmax,0:nmax), vy(0:mmax,0:nmax)
DIMENSION det(NREG),POR(NREG),N(0:NREG)
DIMENSION q(0:mmax,0:NREG)
double precision ANGLE
!
Z=0.0d0 !FLOW RATE AT THE EXIT BOUNDRY, INITIALLY SET TO ZERO

! NEEDS MODIFIED FOR REGIONS, BE SURE TO INCLUDE DIFFERENT POROSITIES


FOR EACH REGION
! POR V = (K11 FX + K22 FY) PDROP PERMX /POR / VIS
!
! ZERO ALL POINTS
DO J=0,N(NREG)
DO I = 0,M
VX(I,J)=0.0
VY(I,J)=0.0
ENDDO
ENDDO

! CALCULATE INTERIOR POINTS OF EACH REGION


! BOUNDARIES BETWEEN REGIONS ARE NOT CALCULATED
DO K=1,NREG
DO J=N(K-1)+1,N(K)-1
DO I = 1,m-1
ip=i+1
im=i-1
jp=j+1
jm=j-1
sx=2.0d0
se=2.0d0
call fyfx(x,y,f,fx,fy,jac,ip,i,j,im,&
jp,jm,sx,se,mmax,nmax,dxi,det,K,NREG)
! Calculate velocities
vx(i,j)=-(alp(N,NREG,J,ANGLE)*fx+bet(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)
vy(i,j)=-(bet(N,NREG,J,ANGLE)*fx+gam(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)

ENDDO
ENDDO
ENDDO
! EXTERNAL BOUNDARIES
! TOP
K=NREG
J=N(NREG)

219 
 
do i=1,m-1
ip=i+1
im=i-1
jp=1
jm=0
sx=2.0d0
se=1.0d0
call fyfx(x,y,f,fx,fy,jac,ip,i,j,im,&
jp,jm,sx,se,mmax,nmax,dxi,det,K,NREG)
! Calculate velocities
vx(i,j)=-(alp(N,NREG,J,ANGLE)*fx+bet(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)
vy(i,j)=-(bet(N,NREG,J,ANGLE)*fx+gam(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)
enddo
! BOTTOM
J=0
K=1
do i=1,m-1
ip=i+1
im=i-1
jp=n(K)
jm=n(K)-1
sx=2.0d0
se=1.0d0
call fyfx(x,y,f,fx,fy,jac,ip,i,j,im,&
jp,jm,sx,se,mmax,nmax,dxi,det,K,NREG)
! Calculate velocities
vx(i,j)=-(alp(N,NREG,J,ANGLE)*fx+bet(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)
vy(i,j)=-(bet(N,NREG,J,ANGLE)*fx+gam(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)
enddo
! INLET AND OUTLET
! i=0 AND i=m
DO K=1,NREG
I=0
DO J=1,N(K)-1
ip=1
im=0
jp=j+1
jm=j-1
sx=1.0d0
se=2.0d0
call fyfx(x,y,f,fx,fy,jac,ip,i,j,im,&
jp,jm,sx,se,mmax,nmax,dxi,det,K,NREG)
! Calculate velocities

vx(i,j)=-(alp(N,NREG,J,ANGLE)*fx+bet(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)

220 
 
vy(i,j)=-(bet(N,NREG,J,ANGLE)*fx+gam(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)
ENDDO
i=m
do j=1,N(NREG)-1
ip=m
im=m-1
jp=j+1
jm=j-1
sx=1.0d0
se=2.0d0
call fyfx(x,y,f,fx,fy,jac,ip,i,j,im,&
jp,jm,sx,se,mmax,nmax,dxi,det,K,NREG)
! Calculate velocities

vx(i,j)=-(alp(N,NREG,J,ANGLE)*fx+bet(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)
vy(i,j)=-(bet(N,NREG,J,ANGLE)*fx+gam(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)
ENDDO
ENDDO
! CORNERS
i=0
j=0
K=1
ip=1
im=0
jp=1
jm=0
sx=1.0d0
se=1.0d0
call fyfx(x,y,f,fx,fy,jac,ip,i,j,im,&
jp,jm,sx,se,mmax,nmax,dxi,det,K,NREG)
! Calculate velocities
vx(i,j)=-(alp(N,NREG,J,ANGLE)*fx+bet(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)
vy(i,j)=-(bet(N,NREG,J,ANGLE)*fx+gam(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)
I=0
J=N(NREG)
K=NREG
ip=1
im=0
jp=n(K)
jm=n(K)-1
sx=1.0d0
se=1.0d0
call fyfx(x,y,f,fx,fy,jac,ip,i,j,im,&
jp,jm,sx,se,mmax,nmax,dxi,det,K,NREG)
! Calculate velocities
vx(i,j)=-(alp(N,NREG,J,ANGLE)*fx+bet(N,NREG,J,ANGLE)*fy)&

221 
 
*PERMX*PDROP/(POR(K)*VIS)
vy(i,j)=-(bet(N,NREG,J,ANGLE)*fx+gam(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)
i=m
j=0
K=1
ip=m
im=m-1
jp=1
jm=0
sx=1.0d0
se=1.0d0
call fyfx(x,y,f,fx,fy,jac,ip,i,j,im,&
jp,jm,sx,se,mmax,nmax,dxi,det,K,NREG)
! ! Calculate velocities
vx(i,j)=-(alp(N,NREG,J,ANGLE)*fx+bet(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)
vy(i,j)=-(bet(N,NREG,J,ANGLE)*fx+gam(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)
I=M
J=N(NREG)
K=NREG
ip=m
im=m-1
jp=n(K)
jm=n(K)-1
sx=1.0d0
se=1.0d0
call fyfx(x,y,f,fx,fy,jac,ip,i,j,im,&
jp,jm,sx,se,mmax,nmax,dxi,det,K,NREG)
! Calculate velocities
vx(i,j)=-(alp(N,NREG,J,ANGLE)*fx+bet(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)
vy(i,j)=-(bet(N,NREG,J,ANGLE)*fx+gam(N,NREG,J,ANGLE)*fy)&
*PERMX*PDROP/(POR(K)*VIS)
!
! write results to file
OPEN(UNIT=7, FILE='xy.OUT',STATUS='UNKNOWN') !Grid file
OPEN(UNIT=8, FILE='uv.OUT',STATUS='UNKNOWN') !Velocities vx, vy
do i=0,m
do j=0,N(NREG)
write(7,*)x(i,j),y(i,j)
write(8,*)vx(i,j),vy(i,j)
enddo
enddo

! Calculate Flow rate at the exit boundary


OPEN(UNIT=10, FILE='Q.OUT',STATUS='UNKNOWN')
w=0.0471D0
!
! !Calculate flow rate at bottom corner point

222 
 
q(m,0)=POR(1)*vx(m,0)*((y(m,1)-y(m,0))/2)*w
!Calculate flow rate at exit boundry
do j=1,N(NREG)-1
if (j.lt.N(1))then
K=1
q(m,j)=POR(K)*vx(m,j)*(((y(m,j)-y(m,j-1))/2)+((y(m,j+1)-
y(m,j))/2))*w
else if (j.lt.N(2))then
K=2
q(m,j)=POR(K)*vx(m,j)*(((y(m,j)-y(m,j-1))/2)+((y(m,j+1)-
y(m,j))/2))*w
else if (j.lt.N(3))then
K=3
q(m,j)=POR(K)*vx(m,j)*(((y(m,j)-y(m,j-1))/2)+((y(m,j+1)-
y(m,j))/2))*w
else if (j.lt.N(4))then
K=4
q(m,j)=POR(K)*vx(m,j)*(((y(m,j)-y(m,j-1))/2)+((y(m,j+1)-
y(m,j))/2))*w
else if (j.lt.N(5))then
K=5
q(m,j)=POR(K)*vx(m,j)*(((y(m,j)-y(m,j-1))/2)+((y(m,j+1)-
y(m,j))/2))*w
else if (j.lt.N(6))then
K=6
q(m,j)=POR(K)*vx(m,j)*(((y(m,j)-y(m,j-1))/2)+((y(m,j+1)-
y(m,j))/2))*w
else if (j.lt.N(7))then
K=7
q(m,j)=POR(K)*vx(m,j)*(((y(m,j)-y(m,j-1))/2)+((y(m,j+1)-
y(m,j))/2))*w
end if

enddo
!Calculate flow rate at upper corner
q(m,N(NREG))=POR(7)*vx(m,N(NREG))*((y(m,N(NREG))-y(m,(N(NREG)-
1)))/2)*w
! !Total flow rate at the exit boundry
do j=0, N(NREG)
Z=Z+q(m,j)
write(10,*)q(m,j)
enddo
write(10,*) "Flow rate and grid width are"
write(10,*)Z, m
return
end
!
! ***********************************************************************
!
SUBROUTINE fyfx(x,y,f,fx,fy,jac,ip,i,j,im,jp,jm,sx,se,mmax,nmax,dxi,det,K,NREG)
double precision f(0:mmax,0:nmax)

223 
 
double precision x(0:mmax,0:nmax), y(0:mmax,0:nmax)
double precision jac,x1,x2,y1,y2,fx,fy,sx,se,dxi,det(NREG)
integer i,j,ip,im,jp,jm,nmax,mmax,NREG,K
x1 = (x(ip,j)-x(im,j))/(sx*dxi)
x2 = (x(i,jp)-x(i,jm))/(se*det(K))
y1 = (y(ip,j)-y(im,j))/(sx*dxi)
y2 = (y(i,jp)-y(i,jm))/(se*det(K))
jac = x1 * y2 - y1 * x2 !jacobian
!
if(abs(jac).lt.1.0d-10)then !rectangular square grid
fx=(f(ip,j)-f(im,j))/dxi
fy=(f(i,jp)-f(i,jm))/det(K)
else
!
fx=( f(ip,j)*(y(ip,jp)-y(ip,jm))&
- f(im,j)*(y(im,jp)-y(im,jm))&
- f(i,jp)*(y(ip,jp)-y(im,jp))&
+ f(i,jm)*(y(ip,jm)-y(im,jm)))&
/sx/se/jac/dxi/det(K)
fy=( -f(ip,j)*(x(ip,jp)-x(ip,jm))&
+ f(im,j)*(x(im,jp)-x(im,jm))&
+ f(i,jp)*(x(ip,jp)-x(im,jp))&
- f(i,jm)*(x(ip,jm)-x(im,jm)))&
/sx/se/jac/dxi/det(K)
Endif
return
end
!*********************************************************

224 
 
APPENDIX C

NOMENCLATURE

= Density of α phase

vα = Velocity of α phase

ε = Porosity of filter medium

= Gas volume fraction

= Gas phase density

g= Gravity

iα = Flux of property φ of α phase

G
EM = Rate of momentum transfer due to phase change

G
IM = Rate of momentum transfer due to drag per unit volume between the

phases

G
SM = Rate of momentum transfer due to slip per unit volume

G
GM = Momentum transfer due to heterogeneous reaction at the interface between

the phases per unit volume

225 
 
τG= Shear stress for gas phase

w = Velocity of the interface

v1 = Velocity of phase 1

v2 = Velocity of phase 2

G
v = Local volume averaged velocity

∇P = Pressure drop

µ= Viscosity of gas

k= Air permeability

226 
 

You might also like