You are on page 1of 314

MEMBRANES AND MATERIALS FOR SEPARATION AND

PURIFICATION OF HYDROGEN AND NATURAL GAS

SEYED SAEID HOSSEINI

NATIONAL UNIVERSITY OF SINGAPORE

2009
MEMBRANES AND MATERIALS FOR SEPARATION AND

PURIFICATION OF HYDROGEN AND NATURAL GAS

SEYED SAEID HOSSEINI

(B. Sc., Chemical Engineering, Isfahan University of Technology)

A THESIS SUBMITTED

FOR THE DEGREE OF DOCTOR OF PHILOSOPHY

DEPARTMENT OF CHEMICAL & BIOMOLECULAR

ENGINEERING

NATIONAL UNIVERSITY OF SINGAPORE

2009
ACKNOWLEDEGMENTS

This thesis is the result of years of studies and research works carried out since I joined

the membrane research group at NUS. This provided me the wonderful opportunity of

working with great number of people whose contributions to this accomplishment

deserve special mention. It is my pleasure to take the chance and convey my sincere

gratitude to all of them herein:

In the first place, I am greatly thankful to my supervisor, Professor Tai-Shung Chung, for

his supervision, constructive advices and guidance from the very early stage throughout

the entire parts of this journey. Undoubtedly his constant demand enabled me to gain

extraordinary experiences in diverse aspects of the subject matters covering the areas of

my research. I also would like to extend my gratitude to Dr. Liu Ye for his kind supports

and valuable contributions in various stages of the research works.

Many thanks go in particular to Prof. K (William B. Krantz). I am much indebted to him

as a true teacher, and for his constant encouragements and supports in various aspects. I

really enjoyed every moments of working together and scientific discussions during the

development of EP technique.

I gratefully thank A/Prof. Hidajat, A/Prof. Hong and Dr. Jiang, for their constructive

comments on this thesis. I am thankful that in the midst of all their activity, they accepted

to be members of the examining committee.

i
Special thanks are due to my friends in membrane research group for their kind helps and

supports in countless occasions. Their company over these years made my stay at NUS

joyful and memorable. Dr. Li Yi is especially recognized for his kind mentorship at the

early stage of my research and teaching me valuable skills. I am also thankful to others

namely Ms. Low Bee Ting, Ms. Chng Mei Lin, Dr. Xiao Youchang, Dr. Lim Poh Chong,

Dr. Shao Lu, Ms. Peng Na, Mr. Lau Cher Hon for their valuable contributions and

assistance, especially at the later stage of my studies.

I was extraordinarily fortunate in receiving wonderful supports from Dr. A. Zadhoush,

Dr. S.S. Nouri and Dr. A. Mehrabani during the time I was at Isfahan University of

Technology. I could never have embarked and started all of this without their prior

teachings, trusts and encouragements. Thank you.

I gratefully acknowledge the International Graduate Scholarship (IGS) offered by Agency

for Science, Technology & Research (A*STAR) which provided me wonderful

opportunity pursue my graduate studies at the National University of Singapore (NUS)

through which I could realize abilities and discover other capabilities.

Sincere thanks are due to the Department of Chemical & Biomolecular Engineering

(NUS) and its staff for providing a positive, conducive and professional atmosphere for

research. Special thanks also go to the Institute of Materials, Research & Engineering

(IMRE) for providing me the opportunity of using their excellent facilities and research

environment throughout my studies. I really enjoyed being a research scholar affiliated to

ii
IMRE. I greatly acknowledge funding from NUS, A*STAR and National Research

Foundation (NRF) and other organizations and industrial partners whose financial

supports enabled this work to be successfully completed.

I would like to express my sincere thanks to the management at KEPPEL Offshore &

Marine Technology Centre (KOMtech) for their kind supports and patience which

undoubtedly played an important role in my concurrent progress in both academic and

industrial research.

Where would I be without my family? Words fail me to express my appreciation to my

parents and my sister for their love, dedication, prayers and inseparable supports,

especially in tolerating several years of my absence in pursuit of knowledge.

Last but not least, I must express my special thanks my family members for their love,

patience and endless encouragements and support. Acknowledgements are due to all

those who have assisted me in any way throughout the period of my PhD study.

Finally, I would like to thank everybody who was important and supportive in any

respect to the successful realization of this thesis, and express my apology if I missed

anyone in the list.

iii
TABLE OF CONTENTS

Page
ACKNOWLEDGEMENTS………………………………………………………………..i

TABLE OF CONTENTS………………………………………………………………...iii

SUMMARY…………………………………………………………….…………………x

NOMENCLATURE…………………………………………………………………….xvi

LIST OF TABLES………………………………………………………………………xxi

LIST OF FIGURES……………………………………………………………………xxiv

CHAPTER 1 INTRODUCTION TO GAS SEPARATION MEMBRANES 1

1.1 Historical Developments of Membranes 3

1.2 Industrial Applications of Gas Separation Membrane Technology 6

1.2.1 Hydrogen Separation 11

1.2.2 Air Separation 13

1.2.3 Natural Gas Treatment 14

1.3 Principles of Gas Separation Membranes 15

1.3.1 Membrane Materials 15

1.3.2 Membrane Design and Structures 21

1.3.2.1 Dense Versus Porous Membranes 21

1.3.2.2 Symmetric Versus Asymmetric Membranes 21

1.3.2.3 Dynamic in-situ Membranes 24

1.3.2.4 Liquid Membranes 25

1.3.3 Membrane Module Configurations 25

iv
1.3.3.1 Plate-and-Frame Membranes 26

1.3.3.2 Spiral Wound Modules 27

1.3.3.3 Hollow Fiber Modules 28

1.4 Transport Mechanisms in Gas Separation Membranes 30

1.4.1 Poiseuille Flow 31

1.4.2 Knudsen Diffusion 31

1.4.3 Molecular Diffusion 32

1.4.4 Solution-diffusion model 34

1.4.5 Gas Sorption and Diffusion in Glassy Polymers 34

1.4.6 Free Volume in Glassy Polymers 36

1.5 Research Objectives and Organization of Dissertation 38

1.6 References 43

CHAPTER 2 STRATEGIES AND APPROACHES IN DEVELOPMENT OF GAS

SEPARATION MEMBRANES 52

2.1 Key Limitations of Existing Polymeric Membranes 55

2.1.1 Permeability Versus Permselectivitiy 55

2.1.2 CO 2 -Induced Plasticization and Conditioning 57

2.1.3 Physical Aging 59

2.2 Molecular Design of Polymeric Membranes for Gas Separation 62

2.2.1 Tailoring Pure Polyimide Systems 62

2.2.2 Design of Block Copolymer Systems 70

2.3 Modification of Polymeric Membranes for Gas Separation 71

v
2.3.1 Polymer Blending 72

2.3.2 Hybrid Membranes 74

2.3.3 Thermal Annealing 77

2.3.4 Grafting and Halogenation of Polymer Backbone 79

2.3.5 Cross-linking Modification 81

2.4 Current and Future Developments 88

2.5 References 91

CHAPTER 3 ENHANCED GAS SEPARATION PERFORMANCE OF

NANOCOMPOSITE MEMBRANES USING MGO NANOPARTICLES 109

3.1 Introduction 109

3.2 Experimental 115

3.2.1 Materials 115

3.2.2 Membrane fabrication 116

3.2.3 Post-treatment of membranes 117

3.2.4 Membrane characterizations 117

3.2.5 Gas permeability measurements 118

3.2.6 Measurement of gas sorption 119

3.3 Results and discussion 120

3.3.1 Effect of MgO loading on the gas separation performance of

nanocomposite membranes 120

3.3.2 Effect of heat treatment on the gas separation performance of

nanocomposite membranes 125

vi
3.3.3 Effect of MgO nanoparticles on silver adsorption into the nanocomposite

membranes 127

3.3.4 Effect of silver treatment on the gas separation performance of

nanocomposite membranes 134

3.4 Conclusions 140

3.5 References 141

CHAPTER 4 HYDROGEN SEPARATION AND PURIFICATION IN

MEMBRANES OF MISCIBLE POLYMER BLENDS WITH

INTERPENETRATION NETWORKS 148

4.1 Introduction 148

4.2 Experimental 151

4.2.1 Materials 151

4.2.2 Preparation of dense flat membranes 153

4.2.3 Chemical modification of membranes 153

4.2.4 Characterization techniques 154

4.2.5 Gas permeability measurements 155

4.3 Results and discussion 156

4.3.1 The miscibility and other characteristics of blends 156

4.3.2 Gas separation characteristics of blend membranes 158

4.3.3 Effect of chemical modification on performance of membranes 166

4.4 Conclusions 173

4.5 References 174

vii
CHAPTER 5 CARBON MEMBRANES FROM BLENDS OF PBI AND

POLYIMIDES FOR N 2 /CH 4 AND CO 2 /CH 4 SEPARATION AND HYDROGEN

PURIFICATION 179

5.1 Introduction 179

5.2 Experimental 184

5.2.1 Materials 184

5.2.2 Preparation of polymer precursors 185

5.2.3 Chemical modification of membranes 186

5.2.4 Preparation of carbon molecular sieve membranes 186

5.2.5 Characterization techniques 187

5.2.6 Gas permeability and sorption measurements 188

5.3 Results and discussion 190

5.3.1 The physical and transport properties of polymers and blend precursors190

5.3.2 The effect of pyrolysis protocol on gas separation performance of carbon

membranes 194

5.3.3 The effect of temperature protocol on evolution of carbon molecular sieve

membranes 199

5.3.4 The effect of blend component on gas separation performance of

precursors and derived carbon membranes 201

5.3.5 The effect of blend composition on transport properties of carbon

membranes 206

5.3.6 The effect of chemical modification on properties and gas separation

performance of carbon membranes 215

viii
5.4 Conclusions 220

5.5 References 222

CHAPTER 6 DEVELOPMENT OF NOVEL HIGH-PERFORMANCE GAS

SEPARATION MEMBRANES THROUGH INTEGRATION OF POLYMER

BLENDING TECHNOLOGY AND DUAL-LAYER HOLLOW FIBER SPINNING

PROCESS 230

6.1 Introduction 230

6.2 Experimental 235

6.2.1 Materials 235

6.2.2 Dope formulation and preparation 237

6.2.3 Dual-layer hollow fiber spinning process 238

6.2.4 Silicone rubber coating and chemical modification 240

6.2.5 Membrane characterization and gas permeation tests 240

6.3 Results and discussion 241

6.3.1 Morphological characteristics of dual-layer hollow fiber membranes 241

6.3.2 The effect of air gap distance on the morphology and gas transport

properties of dual-layer hollow fiber membranes 245

6.3.3 The effect of elongational drawing on the morphology and gas transport

properties of dual-layer hollow fiber membranes 251

6.3.4 The effect of outer layer dope flow rate on the morphology and gas

transport properties of dual-layer hollow fiber membranes 253

6.3.5 Anti-plasticization properties of dual-layer hollow fiber membranes 255

ix
6.3.6 The effect of chemical modification on the gas transport properties of

dual-layer hollow fiber membranes 257

6.4 Conclusions 259

6.5 References 261

CHAPTER 7 CONCLUSIONS AND RECOMMENDATIONS 276

7.1 Conclusions 276

7.1.1 Enhanced Gas Separation Performance of Nanocomposite Membranes

Using MgO Nanoparticles 277

7.1.2 Hydrogen Separation and Purification in Membranes of Miscible Polymer

Blends with Interpenetration Networks 279

7.1.3 Carbon Membranes from Blends of PBI and Polyimides for N 2 /CH 4 And

CO 2 /CH 4 Separation and Hydrogen Purification 280

7.1.4 Development of Novel High-performance Gas Separation Membranes

Through Integration of Polymer Blending Technology and Dual-layer Hollow

Fiber Spinning Process 282

7.2 Recommendations for Future Work 283

7.2.1 Nanocomposite membranes 283

7.2.2 Polymeric blend membranes 285

7.2.3 Carbon membranes from polymer blends 287

7.2.4 Dual-layer hollow fiber membranes from polymeric blends 288

x
SUMMARY

Efficient technologies are required in various industries for separation of gas

molecules. Membrane-based gas separation has emerged as one of the fastest growing

technologies, due to the compactness, higher energy efficiency and economic advantages

which can be reaped. However, the inadequacy of the existing membrane materials

hinders the full exploitation of the application opportunities on the industrial scale. The

key challenges are developing membranes that can achieve good gas selectivity without

sacrificing gas permeability and also capable of maintaining the long-term gas separation

performance. Therefore, the purpose of this research study is to identify, explore and

develop viable techniques in order to address these challenges and provide membranes

with enhanced gas separation performance. A comprehensive research report is presented

encompassing various subjects including the development of new nanocomposite

membranes by using MgO nanoparticles with enhanced gas separation performance,

fabrication of membranes with interpenetration networks from miscible polymeric blends

for hydrogen separation and purification, production of carbon molecular sieving

membranes with enhanced performance for diverse gas separation applications and

fabrication of dual-layer hollow fiber membranes for gas separation in the semi-industrial

scale. Various instrumental analyses were employed in the course of this research for

characterization of physical and transport properties of developed membranes. The key

emphases were put on the natural gas treatment and hydrogen separation and purification.

Firstly, gas separation membranes with enhanced performance were developed by

the introduction of nano-size magnesium oxide particles followed by treatment with

xi
silver ions. Nanocomposite membranes were fabricated by incorporating nanoscale

magnesium oxide particles with different loadings into the Matrimid matrix. The addition

of MgO nanoparticles led to an increase in gas permeability of membranes; the highest

permeability was observed for the membranes containing 40 wt% MgO loading.

However, the selectivity of nanocomposite membranes was less than the neat Matrimid.

These changes can be mainly ascribed to the effect of pore dimensions of MgO

nanoparticles which are larger than the size range of gas molecules. Heat treatment at

temperatures below polymer glass transition temperature resulted in densified structure,

while heat treatment above T g resulted in further increase in gas permeability of

membranes through the changes in free volume of membrane. Secondly, nanocomposite

membranes with 20 wt% MgO were modified by silver treatment for stipulated periods of

time. A dipping procedure was adopted for this process which resulted in the

impregnation of silver ions into the nanocomposite membranes. It was found that the

major driving force for penetration of silver ions was offered by MgO nanoparticles and

polymer played the role of a carrier. The adsorption process was confirmed by different

techniques and oxygen sites at the surface of MgO were recognized to be responsible for

this process. The longer immersion time led to more silver adsorption into the membrane.

All the modified membranes exhibited enhanced gas separation performance for selected

gas pairs compared to the neat Matrimid membrane. The best performance was observed

for nanocomposite membranes (with 20 wt% MgO) after a 10-day silver treatment in

which the CO 2 /CH 4 and H 2 /N 2 selectivity increased by 50% and 35%, respectively. To

our best knowledge, this is the first study that combines the synergistic features of

nanocomposite structure and facilitated transport mechanism into a single membrane by

xii
using unique characteristics of MgO for gas separation. The results of this research study

were published in the J. Membr. Sci., 302 (2007) 207-217.

Another research study demonstrates the successful implications of blending

technique cum chemical modification for the fabrication of high performance polymeric

membranes for gas separation applications. The effect of variation in composition on

miscibility and microstructure, gas permeability and selectivity of blend membranes is

investigated. It is found that augmentation in PBI composition results in enhancement in

gas separation performance of membranes which is attributed mainly to the effect of

diffusivity selectivity. Analysis of the microstructure of membranes confirms the

variations in chain packing density, d-spacing and segmental mobility of polymer chains

as a result of blending. Separation performance of membranes is further ameliorated

through chemical modification of blend constituents. Modification of PBI phase with p-

xylenedichloride brings about slight improvements in selectivity performance, especially

for H 2 /CO 2 and H 2 /N 2 . In contrast, the selectivity of membranes is improved

significantly after cross-linking of Matrimid phase with p-xylenediamine. The results

indicate that higher tendency of Matrimid toward cross-linking reaction contributes more

in controlling the transport properties of membranes through diffusion coefficient by

increase in chain packing density and diminishing the excess free volumes. Results

obtained in this study reveal the promising features of developed membranes for gas

separation applications with great potential for hydrogen separation and purification on

industrial scale. To our best knowledge, this is the first study focusing the exploitation of

unique properties of blending high performance polymers of Matrimid and PBI for the

xiii
development of specialty membranes for hydrogen separation and purification. The

results of this research study were published in the Polymer, 49 (2008) 1594-1603. In

addition, a patent application (Ser. No.: 61/109,318) was filed on this subject in the U.S.

Patent Office.

Another research study features investigation on blends of poly(benzimidazole)

(PBI) and various polyimides regarding to their capabilities in formation of homogenous

blends with satisfactory morphology as precursors for carbon molecular sieve

membranes. Attempts are made to unravel the correlations between the pyrolysis process

parameters and separation performance of developed membranes. The results show that

compared to Torlon and P84, Matrimid is a better choice for preparation of blend

precursors in combination with PBI. The findings also suggest that carbon membranes

derived from PBI/Matrimid are congruous candidates for H 2 /CO 2 separation.

Interestingly, attractive performance for separation of other gas pairs was obtained by

tuning the PBI content in membranes. Modification of precursors by chemical cross-

linking prior to carbonization provides membranes with enhanced selectivity suitable for

separation of hydrogen from both nitrogen and carbon dioxide. This research study

accentuates the necessity and importance of adopting a proper strategy in exploiting

synergistic beneficial features of advanced materials (PBI), technology (polymer

blending), process (carbonization) and modification techniques (cross-linking) toward

achieving membranes with desired performance. Developed membranes surpass several

separation performance trade-offs with great potentials for various industrial applications

including CO 2 /CH 4 (α= 203.95), H 2 /CO 2 (α= 33.44) and particularlyN 2 /CH 4 separation

xiv
with unprecedented high permeability (P N2 = 2.78 Barrer) and selectivity (α= 7.99) for

this gas pair. To our best knowledge, this is the highest ever reported selectivity for this

gas pair. The results of this research study were published in the J. Membr. Sci., 328

(2009) 174-185. In addition, a patent application (Ser. No.: 61/103,547) was filed on this

subject in the U.S. Patent Office.

The final aspect of this research study focuses on the exploitation of polymer

blending technology in fabrication of high performance dual-layer hollow fiber

membranes. The functional material for the outer layer of the membranes is a blend of

PBI/Matrimid (50:50wt.%). Attempts are made to unravel the effect of various key

parameters on the morphological evolution, transport properties and separation

performance of the fabricated hollow fiber. The resultant membranes possess desirable

morphological structure, especially at the outer layer. It is important to note that all the

samples show the delamination-free structure at the interface between inner and outer

layers. The results suggest that increasing the air gap distance and spinning with

elongational drawing have improving effects on the gas permeance and selectivity of the

membranes. In contrast, increasing the outer dope flow rates resulted in membranes with

diminished gas permeance and altered separation performance depending on the air gap

distance. Application of silicone rubber coating was found an effective means for tuning

the gas transport and separation performance of the membranes. The findings also

indicate that chemical modification of as-spun hollow fibers, even for a very short period

of time, provides membranes with enhanced H 2 /CO 2 selectivity. Membranes fabricated

in this study essentially exhibit viable potentials for various gas separation applications

xv
including hydrogen and natural gas separation. Especially, fabricated membranes offer

H 2 /CO 2 selectivity values significantly higher than the intrinsic property of the blend

materials and with very good resistance against the CO 2 -induced plasticization.

xvi
NOMENCLATURE

A Effective area of the membrane (cm2)

Ag Silver

b Affinity of the gas molecules to the Langmuir sites

C Total penetrant concentration in the polymer (cm3(STP)/cm3(polymer))

C’ H Langmuir capacity constant

CD Concentration of gas adsorbed at Henry sites (cm3(STP)/cm3(polymer))

CH Concentration of gas adsorbed at Langmuir sites (cm3(STP)/cm3(polymer))

D Penetrant diffusion coefficient (cm2.s-1)

d Kinetic diameter of the gas molecule, inter-segmental spacing between two

polymer chains

D0 Pre-exponential factor for activation energy of penetrant diffusion (cm2.s-1)

D app Apparent diffusion coefficient

DD Henry’s diffusion coefficient (cm2.s-1)

DH Langmuir’s diffusion coefficient (cm2.s-1)

Dk Knudsen diffusion coefficient (m2.s-1)

dp Pore diameter (m)

ED Activation energy of diffusion (kJ.mol-1)

f Polymer dependant parameter

F Ratio of Henry’s diffusion coefficient to Langmuir’s diffusion coefficient

KD Henry’s dissolution constant (cm3(STP)/cm3(polymer).atm)

l Membrane thickness (cm)

Mw Molecular weight

xvii
n Integral number

P Permeability coefficient (1Barrer = 1 x 10-10 (cm3(STP).cm/cm2.s.cmHg)

p Pressure

P0 Pressure of the feed gas in up-stream (psia)

p2 Pressure at the upstream boundary

r Pore radius (m)

R Universal gas constant

S Solubility coefficient (cm3(STP)/cm3(polymer).cmHg)

S app Apparent solubility coefficient

T Absolute temperature (K)

Tb Boiling point (oC)

Tc Gas critical temperature (oC)

Tg Polymer glassy temperature (oC)

V Volume of the down-stream chamber of permeation cell (cm3)

V0 Volume occupied by the chains (cm3/g)

Vsp Specific volume of polymer film (cm3/g)

VW Van der Waals volumes (cm3)

W Mass fractions

α ij Separation factor between molecules i and j

θ Diffraction angle, diffusion time-lag

λ X-ray wavelength

ρ Density (g/cm3)

υ Average molecular velocity (m.s-1)

xviii
Abbreviations

6FDA Hexafluoroisopropylidene-diphthalic anhydride

6FmDA 2,2-bis(3-aminophenyl)hexafluoropropane

6FpDA 2,2-bis(4-aminophenyl)hexafluoropropane

AFM Atomic force microscope

BET Brunauer-Emmett-Teller

BPDA 3,3’,4,4’-biphenyltetracarboxylic dianhydride

CA Cellulose acetate

CDA 3,6-diaminocarbazole

CMS Carbon molecular sieves

CMSM Carbon molecular sieve membranes

CTC Charge transfer complexes

DAF 2,7-diaminofluorene

DAFO 2,7-diaminofluorenon

DMEA N,N-dimethylaminoethyleneamine

DSC Differential scanning calorimetry

ECDA N-ethyl-3,6,-diaminocarbazole

EDX Energy dispersive X-ray

EtOH Ethanol

FFV Fractional free volume

FTIR Fourier Transform Infra-Red

FTM Facilitated transport membranes

xix
GC Gas chromatograph

IPDA Isopropylidenedianiline

Matrimid Poly [3,3’,4,4’-benzophenone tetracarboxylic dianhydride and 5(6)-amino-1-

(4'-aminophenyl-1,3-trimethylindane)]

MDA Methylenedianiline

mDDS 3,3’-diaminodiphenyl sulfone

MgO Magnesium Oxide

MMM Mixed matrix membrane

NMP N-methyl-2-pyrolidone

ODA Oxydianiline

o-Ps Orthopositronium

P84 Copolyimide of 3,3’4,4’-benzophenone tetracarboxylic dianhydride and 80%

methylphenylenediamine + 20% methylenediamine

PALS Positron annihilation lifetime spectroscopy

PAN Polyacrylonitrile

PBI Polybenzimidazole, poly [2,2’-(1,3-phenylene)-5,5’-bibenzimidazole]

PC Polycarbonate

pDDS 4,4’-diaminodiphenyl sulfone

pDDS 4,4’-diaminodiphenyl sulfone

PDMS Polydimethylsiloxane

PEO Poly(ethylene oxide)

PES Polyethersulfone

PI Polyimide

xx
PIM Polymers with intrinsic microporosity

PMDA Pyromellitic dianhydride

PPESK Poly(phthalazinone ether sulfone ketone)

P plas Plasticization pressure

PPO Polyphenyleneoxide

PPTI-1 Phenylene thioether

PSA Pressure swing adsorption

PSf Polysulfone

PSMA Poly(styrene-co-maleric anhydride)

PXDA p-xylenediamine

PXDC p-xylenedichloride

SEM Scanning electron microscope

TGA Thermogravimetric Analysis

UV Ultraviolet

WAXD Wide angle x-ray diffraction

wt. Weight

XPS X-ray photoelectron spectroscopy

XRD X-ray diffractometer

xxi
LIST OF TABLES

Table 1.1. Market development of membrane processes 7

Table 1.2. Future market of membrane gas separation 8

Table 1.3. Industrial applications of gas membranes 11

Table 1.4. Materials for gas separation membranes 17

Table 1.5. Qualitative comparison of various membrane modules 30

Table 2.1. Effect of meta or para linkages on gas separation performance (6FDA-based)

64

Table 2.2. Effect of connector groups on gas separation performance (6FDA-based) 66

Table 2.3. Effect of bulky substituents on gas separation performance (6FDA-based) 67

Table 2.4. Effect of polar functional groups on gas separation performance (6FDA-based)

68

Table 2.5. CO 2 /CH 4 separation performance of crosslinked dense films

85

Table 2.6. CO 2 /CH 4 separation performance of crosslinked hollow fibers 86

Table 3.1. Properties of magnesium oxide (MgO Plus) nanoparticles. 116

Table 3.2. The ideal selectivity of pure Matrimid and Matrimid-MgO nanocomposite

membranes with different nanoparticle loadings. 124

Table 3.3. The effect of heat treatment and cooling profile on the permeability and

selectivity of Matrimid-MgO nanocomposite membranes. 126

Table 3.4. XPS element binding energies of the neat Matrimid membrane before and after

silver treatment for 2 days. 130

xxii
Table 3.5. The effect of silver treatment on the permeability and selectivity of Matrimid-

MgO nanocomposite membranes. 135

Table 4.1. The permeability and ideal gas selectivity of polymer blend membranes with

various compositions. 160

Table 4.2. Gas diffusion coefficients and solubility coefficients of polyimide, blend

membranes and modified blend membranes. 161

Table 4.3. The effect of blend composition on the physical properties and fraction of free

volume of membranes. 163

Table 4.4. The effect of chemical modification using p-xylenedichloride on gas

permeability and selectivity of blend membranes. 169

Table 4.5. The effect of chemical modification using p-xylenediamine on gas

permeability and selectivity of blend membranes. 169

Table 5.1. The glass transition temperature of polymers and blend precursors measured

by differential scanning calorimetry (DSC). 191

Table 5.2. Gas permeability and ideal selectivity of individual polymers used in this

study. a,b 193

Table 5.3. The effect of pyrolysis temperature on gas permeability and selectivity of

carbon membranes made of PBI/Matrimid blend precursors.a,b 195

Table 5.4. The effect of blend components on gas permeability and ideal selectivity of

polymeric precursors.a,b 202

Table 5.5. The effect of blend components on gas permeability and ideal selectivity of

carbon membranes.a,b,c 204

xxiii
Table 5.6. Gas permeability and ideal selectivity of carbon membranes derived from

individual polymers.a,b,c 204

Table 5.7. The effect of blend composition on gas permeability and selectivity of

resultant carbon membranes.a,b,c 207

Table 5.8. The effect of application of a 5-day cross-linking modification by p-xylene

diamine (PXDA) on the gas transport properties of blend carbon membranes

with various compositions.a,b,c 217

Table 6.1. The spinning parameters and conditions used in preparation of dual-layer

hollow fiber membranes. 239

Table 6.2. Solubility parameters of the solvents, the coagulation agents and the polymers,

coagulation agents and polymers.* 243

Table 6.3. The effect of air gap distance on the gas permeance and separation

performance of dual-layer hollow fiber membranes. 247

Table 6.4. The intrinsic gas transport properties of the dense flat membrane made from

the polymer blend and respective individual polymers.a 250

Table 6.5. The effect of application of elongational drawing on the gas permeance and

separation performance of dual-layer hollow fiber membranes. 251

Table 6.6. The effect of variation in outer dope flow rates on the gas permeance and

separation performance of wet-spun dual-layer hollow fiber membranes. 254

Table 6.7. The effect of chemical cross-linking modification and its period on the gas

permeance and separation factor of dual-layer hollow fiber membranes. 258

xxiv
LIST OF FIGURES

Figure 1.1. Schematic diagram of gas separation process by a membrane 9

Figure 1.2. Trade-off line of oxygen permeability and oxygen/nitrogen selectivity 19

Figure 1.3. Typical membrane morphologies 23

Figure 1.4. Schematic drawing of a plate-and-frame module 27

Figure 1.5. Spiral wound elements and assembly 28

Figure 1.6. Hollow fiber separator assembly 29

Figure 1.7. Schematics of main mechanisms of membrane-based gas separation 31

Figure 1.8. Motion of gas molecules within the cavities through a series of diffusional

jumps 37

Figure 2.1. Robeson’s tradeoff curve for CO 2 /CH 4 gas pair

57

Figure 2.2. Diffusion of free volume and lattice contraction (physical aging) 61

Figure 2.3. General chemical structure of polyimide 63

Figure 2.4. Synthetic strategy for copolymer structure for CO 2 separation

71

Figure 2.5. Schematic of the various gas transport routes through hybrid polymeric

membranes 76

Figure 2.6. Amine modification of polyimides 79

Figure 3.1. Dispersion quality of MgO nanoparticles at the surface of the nanocomposite

membrane with 20wt% MgO. 121

xxv
Figure 3.2. The comparison of gas permeability for pure Matrimid and nanocomposite

membranes with different loadings of MgO nanoparticles. (left: permeability

of helium, hydrogen and carbon dioxide; right: permeability of oxygen,

nitrogen and methane). 122

Figure 3.3. The cross-section morphology of nanocomposite membranes with different

loadings of MgO nanoparticles. (MgO loadings: A: 20wt% B: 30wt% C:

40wt%). Inset: high magnification image of MgO nanoparticles inside the

polymeric matrix. 123

Figure 3.4. The elementary profiles of EDX analysis for the pure Matrimid and

nanocomposite membranes (20wt% MgO loading) after silver treatment for

different periods of time (Red peaks for Mg element, and green peaks for Ag

element). 129

Figure 3.5. XPS Ag 3d spectra of the nanocomposite membranes with 20wt% MgO

nanoparticles after silver treatment for different immersion times. 131

Figure 3.6. The schematic representation and molecular model for the adsorption of Ag

ions onto the oxygen sites at the surface of MgO. (Type of Interaction: Direct:

dotted lines, Secondary: dashed lines) 133

Figure 3.7. CO 2 sorption isotherms for the nanocomposite membranes before and after

silver doping. 138

Figure 3.8. The XRD spectra of the membranes before and after silver treatment with

different profiles. 139

Figure 4.1. Schematic representation of the chemical structure of polymers and hydrogen

bonding interaction between functional groups of Matrimid and PBI. 152

xxvi
Figure 4.2. DSC thermograms of the polymer blend membranes and individual

constituents. (a) 0, (b) 25, (c) 50, (d) 75, (e) 100 wt% PBI composition. 157

Figure 4.3. Composition dependency of Tg and comparison with theoretical values

predicted by Fox equation. 159

Figure 4.4. Analysis and comparison of the status of free volume in the structure of (a)

PBI and (b) Matrimid membranes using molecular simulation. 164

Figure 4.5. The XRD spectra of blends depicting the effect of composition on

microstructure of membranes. 165

Figure 4.6. (a) Proposed mechanism for the chemical cross-linking modification of

Matrimid component of blend using p-xylene diamine. (b) Chemical structure

of p-xylene diamine, (c) possible chain morphology and configuration of p-

xylene diamine cross-linked Matrimid. (cross-linking agents are specified by

circles).xylenediamine. 167

Figure 4.7. (a) Proposed mechanism for the chemical cross-linking modification of PBI

component of blend using p-xylene dichloride. (b) Chemical structure of p-

xylene dichloride, (c) possible chain morphology and configuraion of p-

xylene dichloride cross-linked PBI. (cross-linking agents are specified by

circles). 168

Figure 4.8. The XRD spectra of blend membranes cross-linked with p-xylene dichloride

(PXDC) and p-xylenediamine (PXDA) for different periods of time. 171

Figure 4.9. Molecular simulation analysis of chemical structure of cross-linking agents

(a) p-xylenediamine (b) p-xylenedichloride. 173

xxvii
Figure 5.1. Chemical structure of (a) Polybenzimidazole (PBI) (b) Matrimid® 5218 (c)

Torlon® 4000T and (d) P84. 185

Figure 5.2. Representation of carbonization protocols used in this study with three

distinct final temperatures; Protocol A: 600oC; Protocol B: 700oC; Protocol C:

800oC. 187

Figure 5.3. The elemental analyses of the polymeric precursor and derived carbon

♦ :(
membranes from PBI/Matrimid (50/50 wt. %) at various temperatures.

Carbon, ■ : Oxygen, ▲ : Nitrogen) 197

Figure 5.4. The effect of variations in pyrolysis temperature on microstructure of

PBI/Matrimid (50/50 wt.%) blend carbon membranes evaluated by X-ray

diffraction analysis. 198

Figure 5.5. The results of the thermogravimetric analysis carried out on blend precursor

PBI/Matrimid (50/50 wt. %) applying different temperature protocols. (a)

Protocol C (b) Protocol D (10oC/min). 200

Figure 5.6. The effect of blend component on structural characteristics of blend carbon

membranes evaluated by XRD analysis. (Pyrolysis temperature: 800oC) 205

Figure 5.7. Performance of PBI/Matrimid blend precursors with various compositions

and their corresponding carbon membranes for H 2 /CO 2 separation with

respect to trade-off line. (aData from Ref. [58]) 208

Figure 5.8. Performance of PBI/Matrimid blend precursors with various compositions

and their corresponding carbon membranes for N 2 /CH 4 separation with

respect to trade-off line. (aData from Ref [58]) (Trade-off line was drawn

based on data points collected from Ref. [62]) 209

xxviii
Figure 5.9. Performance of PBI/Matrimid blend precursors with various compositions

and their corresponding carbon membranes for CO 2 /CH 4 separation with

respect to trade-off line. (aData from Ref. [58]) 210

Figure 5.10. Plots of N2 and CH4 sorption isotherms for the carbon membrane derived

from the blend precursor composed of PBI/Matrimid (25/75 wt.%). (Pyrolysis

temperature: 800oC) 212

Figure 5.11. The effect of blend composition on structural characteristics of

PBI/Matrimid blend carbon membranes evaluated by XRD analysis.

(Pyrolysis temperature: 800oC) 213

Figure 5.12. Performance of PBI/Matrimid blend precursors with various compositions

and their corresponding carbon membranes with respect to trade-off line for

various gas pairs. (a) O 2 /N 2 (b) H 2 /N 2 (c) H 2 /CH 4 . (aData from Ref. [58])

214

Figure 5.13. The effect of blend composition on structural characteristics of

PBI/Matrimid blend carbon membranes derived from cross-linked precursors.

(Pyrolysis temperature: 800oC) 219

Figure 6.1. The chemical structure of (a) Polybenzimidazole (PBI) (b) Matrimid® 5218

and (c) UDEL Polysulfone (PSf). 236

Figure 6.2. The trend of changes in dope viscosity as a function of polymer concentration

in the respective solvent. (♦ PBI/DMAc, ■ Matrimid/DMAc, ▲ PSf/NMP, ●

PBI:Matrimid (1:1)/DMAc) 238

xxix
Figure 6.3. Schematics of dual-layer hollow fiber spinning set-up and triple-orifice

spinneret. (a) dope fluid tank and pump; (b) bore fluid tank and pump; (c)

filter; (d) spinneret; (e) coagulation bath and (f) take-up drum. 239

Figure 6.4. SEM images illustrating various morphological aspects of a typical dual-layer

hollow fiber membrane (sample B): (a) membrane cross-section; (b)

membrane sector; (c and d) inner and outer layers; (e) interface region; (f)

outer surface of the outer layer and (g) inner surface of the inner layer. 242

Figure 6.5. The cross-section morphology of dual-layer hollow fiber membranes spun at

different air gap distances. 246

Figure 6.6. The trend of changes in CO 2 permeance as a function of upstream pressure

for silicone rubber coated dual-layer hollow fiber membranes.

256

xxx
CHAPTER ONE

INTRODUCTION TO GAS SEPARATION MEMBRANES

Over the last few decades, membrane science and technology has been facing

revolutionary era in many aspects. During this period, substantial scientific advancements

have been achieved resulting in evolution of numerous applications for membranes.

Among these, one may witness pronounced strides in the use of membranes for

separation of gases. Particularly, an extensive growth can be observed in various energy

sectors where gas separation membranes are widely used. This is perhaps, due to the

substantial rise in demand for various energy resources among which natural gas and

hydrogen are considered as viable and promising candidates, respectively. The

advantageous features of natural gas as an appropriate alternative to the oil have attracted

the interests toward further exploration and development of gas reserves. However, the

beneficial exploitation of natural gas may require overcoming some of the challenges.

One of the key issues which has impeded the widespread exploitation of discovered

natural gas resources is that majority of these fields suffer from contaminations with

species other than the desired product; which is methane. As a result, various gas

separation and purification technologies have been developed to address this issue and

make the development of the gas field feasible. Although the readily established

technologies have been very successful, their application has always been associated with

1
high energy intensity, high operational risk and large footprint. In many cases, these

shortcomings may cause limitations in design and operation of a proper gas separation

unit for specific fields. In this respect, the emergence of membrane technology has

attracted many attentions since its unique features can readily address some of the

challenges associated with conventional technologies.

Despite the presence of various types of membranes with organic or inorganic

roots, more attentions have been toward the development of gas separation membranes

from organic materials. This is mainly due to their advantageous features like availability

of raw materials, ease of processing and possibility for further modification of properties

in order to meet the desired specifications. Over the time, various engineering polymers

such as cellulose acetate, polyimides, polysulfones have undergone extensive studies for

their potentials as high performance membranes for gas separation applications. The

results obtained from these studies have provided good guidelines for material selection

and other important steps required for design and engineering of successful gas

separation membranes. The following section will provide an overview on the historical

development of membranes and how the basic concepts have emerged. Gas separation

membranes and their industrial applications will be introduced. The key features of

membrane technology for gas separation, their advantages and shortcomings compared to

the other contenders will be discussed.

2
1.1 HISTORICAL DEVELOPMENTS OF MEMBRANES

The origin of membranes and the associated phenomena dates back to the

eighteenth century. It was during that time that the basic ideas about membranes were

initiated due to the work of philosopher scientists. The word osmosis was among the

early terms coined by Abbé Nolet (at 1748) to describe the permeation of water across a

diaphragm [1]. It was during early 1830s that Mitchell reported some membrane

phenomena which were very interesting at that point of time [2]. He observed during his

laboratory experiments that natural rubber balloons deflated at different rates when

exposed to different gases. A membrane, in basic terms, is defined as a functional thin

film acting as a barrier between two mediums. Basically, the ability to control the

permeation of chemical species is considered as the essential property associated with

any membrane. For example, considering gas separation membranes, the separation is

typically achieved based on a mechanism that one of the components of the feed gas

permeates the permselective membrane at a much higher rate than the other species. The

driving force for this process is typically the pressure or chemical potential difference

between the feed and permeate sides.

In 1833, Graham reported his observations on the osmosis of an air-carbon

dioxide gas system through a wet animal bladder [3]. Almost two decades later, in 1855,

a systematic series of experiments was conducted by Fick aiming to investigate the gas

transport properties of membranes. These studies were carried out using nitrocellulose

membranes and resulted in development Fick’s First Law which has been of paramount

3
importance as a fundamental formula [4]. In 1866, T. Graham extended the Fick’s

experiments and could quantitatively measure the gas permeation through a membrane.

This experiment resembled the Mitchelle’s but was carried out in vacuum environment

rather than air [5,6]. The results obtained led Graham to propose the “solution-diffusion”

as the governing mechanism in transport of gas molecules through a membrane. In 1879,

S. von Wroblewski defined the permeability coefficient for membranes as the penetrant

flux normalized by pressure and thickness [7]. He also further characterized the

permeability coefficient and demonstrated it as the product of diffusivity and solubility

coefficients. In 1891, Kayser showed the validity of Henry’s law for the absorption of

carbon-dioxide in rubber [8].

All these laboratory experiments for about two centuries assisted in development

of fundamentals and theories of membrane science. This period took about two centuries

in with almost no tangible applications or product from membranes. It eventually turned

into reality and the first industrial and commercial applications of membranes were

emergence at early 20s. It was till only about five decades ago that the intensified

scientific research efforts eventually turned true and a technology based on membranes

was emerged. This breakthrough led to the gain of many attentions for further

development of membranes for various applications including but not limited to

separation of gaseous species.

Although the potentials of membranes for separation of gas mixtures was

identified long time ago, but the majority of the achievements can be traced to only few

4
decades ago. The fundamental efforts in this field took place during 1930s and 1940s

during which membranes from natural polymers were studied to understand the effects of

structure and molecular properties on the gas transport properties [9]. The research

activities were continued in the following decades and many other materials underwent

investigations for their potentials for gas separating membranes. These efforts were along

with the revolutionary development in testing equipments and measurement methods for

analysis and evaluation of membranes. Despite these progressive efforts, the emergence

membranes for practical gas separation applications was still suffering from the lack a

proper technology for fabrication of high performance membranes and modules. The

work by Loeb and Sourirajan in formation of asymmetric membranes from cellulose

acetate for reverse osmosis was a breakthrough which gave a lot promise for viability of

membranes for industrial applications [10]. In this respect, successful extension of

asymmetric structure concept to membranes in the shape of hollow fibers along with the

supplementary process developments for fabrication of defect-free composite membranes

played prominent roles to bring the science of gas separation membranes closer to a

potential technology [11,12].

Interestingly, it was only during the 1980s that gas separation using membranes

eventually turned into reality as a commercially available process. Hydrogen separating

membranes branded as “Prism” (manufactured by Permea) was the first product of its

kind to entre to the market and was coined as the first large industrial application of gas

separation membranes. Since then, membrane-based gas separation has undergone

substantial growth targeting various industrial processes and has become a $200

5
million/year business. The breakthrough that lead to this achievement was emerged based

on pioneering works in the field of reverse osmosis process through which high-flux

anisotropic membranes possessing large-surface-area had been successfully produced.

This development was of great success and the technology was rapidly adopted for

implementation in number of chemical plants. It was only few years later that a new

application was introduced which was the development of cellulose acetate membranes

for removing carbon dioxide from natural gas. It was at about the same time that

membranes were developed for certain air separation applications which could enrich

oxygen or nitrogen. This application has undergone rapid development over the years.

Finally, it can be found in the industry that membranes are also being used for variety of

small but growing applications such as hydrocarbon recovery and dehydration of

compressed air.

1.2 INDUSTRIAL APPLICATIONS OF GAS SEPARATION MEMBRANE

TECHNOLOGY

Separation technology constitutes one of the key components of the industrial

world covering diverse fields including chemical technology. Several separation

technologies have been developed upon the demand in unit operations. Some of them

such as distillation, crystallization, adsorption, absorption solvent extraction, etc are well

matured in certain application [13]. In comparison, membrane technology is considered

as a relatively new comer that has drawn many attentions, especially in recent years.

Membrane-based separation processes have undergone major developments since the

6
early days at about four decades ago. Various types of membrane separation units have

been developed and are in operation now for industrial applications and there are many

interests for further implementations or extension to other sectors.

Membrane is basically a semi-permeable barrier within which the separation is

taken place through the controlled rate of material transport. Membranes can perform

separation between liquid, gaseous or liquid/gas phases. In fact, membrane technology

has been largely in use for separation of liquid/liquid and liquid/solids through processes

like reverse osmosis, ultrafiltration, microfiltration, pervaporation, hemodialysis,

electrodialysis and controlled drug delivery. Table 1.1 provides the market growth for

various membrane processes in the final years of the last century.

Table 1.1. Market development of membrane processes

Sales (US$ Million)


Membrane process Annual growth
1996 2000
Dialysis 1300 2100 10
Microfiltration 600 960 10
Ultrafiltration 350 560 10
Reverse osmosis 250 450 12
Gas exchange 160 260 10
Gas separation 120 250 15
Electrodialysis 80 130 10
Miscellaneous >5 >10 7
Total 2940 4810 >10

The numbers may indicate a relatively small market share for gas separation

7
membranes compared to other membrane applications. However, this might be justified

by taking the young age of gas separation membranes into the considerations.

Nevertheless, with the current intensive research and development programs, particularly

in membrane materials, a bright and promising future is anticipated for gas separation

membranes. This can be highlighted by possessing the largest annual growth of 15%. The

estimated market growth in various industrial applications of gas separation membranes

are shown in Table 1.2 [14].

Table 1.2. Future market of membrane gas separation

Market (US$ Million @ 2000)


Separation
2000 2010 2020
Nitrogen from air 75 100 125
Oxygen from air <1 10 30
Hydrogen 25 60 150
Natural gas
CO 2 30 60 100
NGL <1 20 50
N 2 /H 2 O 0 10 25
Vapor/Nitrogen 10 30 60
Vapor/vapor 0 20 100
Air dehydration/other 15 30 100
Total 155 340 760

It can be found that nitrogen enrichment, CO 2 removal from natural gas and hydrogen

separation and purification constitute the dominant applications of gas separation

membranes. Nevertheless, a huge application cum market growth and expansion is

predicted for the use of membranes in refinery, petrochemical and natural gas industries.

8
Unlike the other essentially liquid phase applications, membranes for gas

separation act as a barrier interposed between two phases and perform a selective

separation through provision of certain degrees of restriction upon targeted species

[15,16]. Number of mechanisms has been identified for gas separation process in

membranes which will be discussed in subsequent sections. Generally in membrane gas

separation processes, the bulk phases are gas mixtures. Gas permeation is a set of

physical phenomena where certain gas components selectively pass through a membrane

where one of the species in the mixture is allowed to be exchanged in preference to

others. Separation of a gas mixture occurs since each type of molecules diffuse at a

different rate through the membrane. One bulk phase is enriched in one of the species

while the other phase is depleted of it [17]. The movement of any species across the

membrane is caused by one or more driving forces. These driving forces arise from a

gradient of chemical potential due to concentration gradient or pressure gradient or both.

An illustrative schematic of a two-phase gas separation is shown in Figure 1.1. Today,

large-scale membrane gas separation systems have found acceptance in many sectors of

industry.

Figure 1.1. Schematic diagram of gas separation process by a membrane


9
Membrane technology compares favorably with other conventional separation

techniques due to its multidisciplinary character of more capital and energy efficient. The

specific features and inherent advantages of membrane separation process can be

described as follows [16,18,19,20,21]:

1. Process simplicity in installation and operational

2. Low capital outlay and energy consumption as often only a compressor is required

without any additional utilities.

3. Compact size, small footprint and modularize capability which offer great savings

in space and weight.

4. System flexibility and easy scalability for various process capacities.

5. Mild operational condition (temperature, pressure, etc) in majority of application

6. Membrane separation can be carried out continuously.

7. Membrane properties can be adjusted “tailor-made” to meet specific requirements

or based on a desired task.

8. Membranes are compatible with other unit operations and can easily be integrated

as part of any complex separation processes.

Membrane gas separation process found its first industrial application by emergence of

PRISM® in early eighties [22]. This was for hydrogen separation and since then various

other applications have attracted attentions and grown rapidly. Table 1.3 lists the

multitude applications of gas separation membranes which will be discussed in the

followings sections.

10
Table 1.3. Industrial applications of gas membranes [23,24]

Gas Separation Application


O 2 /N 2 Oxygen enrixhment, nitrogen (inert gas)
generation
H 2 /Hydrocarbons Refinery hydrocarbon recovery
H 2 /CO Syngas ratio adjustment
H 2 /N 2 Ammonia purge gas
CO 2 /Hydrocarbons Acid gas treatment, enhanced oil recovery,
landfill gas upgrading
H 2 S/Hydrocarbons Sour gas treatment
H 2 O/Hydrocarbons Natural gas dehydration
H 2 /Hydrocarbons Helium separation
H 2 O/Air Air dehydration
He/N 2 Helium recovery
Hydrocarbons/Air Pollution control, hydrocarbon recovery
Hydrocarbons from process streams Organic solvent recovery, monomer
recovery

1.2.1 Hydrogen Separation

Hydrogen is anticipated to play an inevitable role as one of the promising sources

of energy in the future. The major proportions of produced hydrogen are used to supply

the refinery and fertilizer industries. The constant increasing demand for highly purified

hydrogen has resulted in attracting more attentions for extensive research studies on

advancement of the technologies for generation of hydrogen with higher efficiency and

lower production cost.

At present, steam reforming of hydrocarbons is predominantly the established

11
technology for generation of hydrogen and it supplies about half of the world’s demand

for this viable gas. In this process, a stream of natural gas mainly composed of methane

undergoes a water gas shift reaction. The yielded product has to be further processed in a

separation process in order to remove unwanted carbon dioxide and other byproducts

from the mixtures with hydrogen. Currently, commercially available separation processes

for this purpose are pressure swing adsorption, amine absorption and absorption using

aqueous solution of potassium carbonate which are highly energy intensive.

Hydrogen separation is known to be the first large scale application of membranes

for gas separation that took place in an ammonia plant. Survey has shown that separation

of hydrogen from nitrogen, ammonia, methane, coal together with syngas molar ratio

adjustment and other refinery applications totally account for 25% market share of

hydrogen separation by membrane technology [25,26]. Hydrogen separation has been

effectively carried out through polymeric membranes enjoying the advantageous

extremely high diffusion coefficient of hydrogen and its smaller molecular size compared

to other gas molecules [27]. However, the hydrogen separation and purification,

especially in refinery applications has been fraught with some difficulties like poor

reliability, fouling, plasticization, and condensation of hydrocarbon vapors [14]. This

highlights the need for development of new hydrogen-permeable membrane materials

with the capability to operate at high hydrocarbon partial pressures and high

temperatures. Another recommended approach for protection of the membrane is the pre-

treatment of the feed gas to reduce its dew point.

12
1.2.2 Air Separation

Air separation constitutes about 60% of overall membrane-based gas separation as

oxygen and nitrogen are considered as third and fifth largest bulk chemicals produced

worldwide [25]. Membrane based gas separation seems to be extremely attractive for the

enrichment of nitrogen from air which can economically produce nitrogen with purities

of up to 95%. The market share of membrane technology in producing nitrogen is

currently 30% of total production volume and growing. The first membranes used for this

purpose had O 2 /N 2 selectivity of about 4. Membrane modules constitute only about 20%

of the component cost of a nitrogen separation plant and the rest is spent for air

compressor. Thus, besides the improvements in membrane performance, reduction in size

of the feed gas compressor seems to be the key in lowering the nitrogen production costs.

In contrast, the overall efficiency of oxygen production using membranes highly relies on

the improvements in membrane performance. Essentially all approaches for production of

oxygen are based on oxygen-permeation while nitrogen-rejection process. Due to the

partial permeation of nitrogen molecules along with oxygen, pro ones production of pure-

oxygen has been facing difficulties and mostly the product is oxygen-enriched air with

purities in the range of 45-50% [28]. Therefore, due to the large industrial demand for

ultra-pure oxygen, oxygen production via membrane technology has yet to compete with

other rivals like cryogenic distillation and vacuum swing adsorption which can provide

oxygen with purities of about 99.999% and 95%, respectively [29]. Evaluations shows

that, in ideal case, new membrane materials with permeability of 250 (Barrer) and

separation factor of 4-6 would be required to make membrane as an attractive choice for

this application.

13
1.2.3 Natural Gas Treatment

Natural gas has undoubtedly become a viable choice in recent years due to the

high rises in the price of crude. The US production of natural gas has been evaluated to

20 trillion scf/year while the world production is about 50 trillion scf/year. Methane is the

key and most desirable component of natural gas, but it also contains impurities like

ethane, propane, butane, water, carbon dioxide, nitrogen and hydrogen sulfide which

have to be removed to certain levels before delivery to pipelines. Despite the dominant

presence of absorption separation by amines, use of membranes has been accepted as a

promising technology for natural gas treatment [30]. High efficiency is observed in

upgrading of natural gas by CO 2 , H 2 O and H 2 S removal using membranes. However,

membranes usually suffer from plasticization in the presence of trace quantities of

condensable heavy hydrocarbons and CO 2 [31]. Therefore it is an imperative to develop

robust materials which can resist plasticization and help the widespread application of

membranes for natural gas treatment.

For water removal from natural gas or dehydration, membrane technology

competes with glycol absorption process. The advantages of membranes are being more

environmental friendly and energy efficiency. Additionally, membranes can effectively

be used for separation of olefinic and paraffinic hydrocarbons. For example, separation of

ethane/ethylene and propane/propylene are very important processes in petroleum

refining and petrochemical industries as the extracted olefins are the major feedstocks in

production of petrochemical products such as polyethylene, polypropylene and other

copolymers [32,33]. Compared to low temperature distillation which is widely practiced

14
for olefin/paraffin separation, membrane technology enjoys large energy efficiency,

simpler operation, lower power consumption and less capital investment.

1.3 PRINCIPLES OF GAS SEPARATION MEMBRANES

With development of advanced technologies, effectual membrane separation processes

are required to address the immediate market challenges and requirements. It is

formidable that the supplied membranes be economically and technologically viable and

satisfies the gas separation productivity and efficiency. Therefore one may consider and

define the following items as the key components in development of successful gas

separation membranes, viz:

- Membrane material selection

- Membrane preparation and modification

- Membrane characterization and evaluation

- Membrane module design

Nevertheless, fundamental concepts have been always considered as the cornerstone at

anytime during the developments. Following sections aim to address some of the key

issues in this regard.

1.3.1 Membrane Materials

Perhaps this component constitutes the most important part of the membrane-

based gas separation process. It is well known that any specific gas separation application

requires materials with certain performance characteristics. For example it is found that

15
chemical interaction between a membrane material and a gas penetrant strongly affects

the separation efficiency of the membrane process [34]. Basically, the choice of proper

material is based on the specific properties originating from its nature and structural

factors. The key requirements of an effective membrane material may include: (1) high

separation efficiency with reasonable high flux, (2) good chemical resistance (3) good

mechanical stability (4) high thermal stability (5) engineering feasibility (6) satisfactory

manufacturing reproducibility (7) low cost [35,36]. Essentially the capital cost of a

membrane separation is beholden to the total membrane area required to perform the

separation for a specific amount of the gas. In practice, many gas feed streams are

contaminated with chemically reactive organic vapors (aromatic compounds, lubricating

oils and solvents) that potentially has adverse effects on membrane material through

swelling or dissolving. Thus, a membrane must have strong chemical resistance

properties and capable of enduring the environment exposed. Since pressure difference is

the driving force for separation, membrane materials have to maintain the mechanical

rigidity against the stress applied. Similarly, excellent thermal stability is required to

withstand the high operating temperatures. Conclusively, an ideal material should be

easily converted to cost effective membrane structure as well as possessing the intrinsic

characteristics of high permeability, selectivity and durability. The science and

technology for membrane materials has seen rapid developments over the past decades

and new materials with different structures have been tailor-made to perform the specific

separations. These developments have been for both organic and inorganic materials. The

following Table provides a list of successfully developed membrane materials for gas

separation.

16
Table 1.4. Materials for gas separation membranes [37]

Organic (polymeric) materials Inorganic materials


Polysulfone (PSf), Polyethersulfone (PES) Carbon molecular sieves (CMS)
Cellulose acetate (CA) Nanoporous carbon
Polyimide, Polyetherimide Zeolites
Polycarbonate [brominated] (PC) Ultramicroporous amorphous silica
Polyphenyleneoxide (PPO) Palladium alloys
Polymethylpentene Mixed conducting perovskites
Polydimethylsiloxane (PDMS)
Polyvinyltrimethylsilane

Basically, among the two available classes of membrane materials amorphous polymers

have been found to possess good processability, sufficient productivity and selectivity

while being cost-effective. Therefore polymeric membranes are of great promise for

industrial applications. In polymers, gas transport and intrinsic permeation characteristics

are significantly governed by the structure and configuration of macromolecular chains.

Glass transition temperature is one of the key characteristics that determine the state and

subsequently the properties of a polymer. In the glassy state, in which polymer is present

above its T g , the polymer structure is hard and rigid and it becomes soft and flexible

when the conditions change to the rubbery state. In fact, glass transition temperature

denotes the boundary between these two states and the temperature in which thermal

expansion coefficient undergoes drastic changes.

Compared to glassy polymers, generally the gas diffusivity thus the productivity

in rubbery polymers is high. However, this is achieved at the cost of low separation

17
efficiency resulted from the small contribution of diffusivity selectivity. On the other

hand, the low segmental mobility and long relaxation time are of the main characteristics

of glassy polymers [20]. This results in rendering of inherent size and shape selectivity

as well as mechanical stability. Hence, glassy polymers have found to be more promising

for industrial gas separation applications [26].

One of the key measures in development of gas separation membranes is the

“upper bond” or “trade-off” line which has extensively been in use as a measure for

evaluation of the new membranes. This concept was emerged by Robeson [38] and

indicated that polymeric materials empirically lie on or below this line. Each gas pair has

its own specific trade-off line and for instance, the trade-off line for oxygen/nitrogen

separation is illustrated in Figure 1.2. New developments in membrane materials are

targeted for the upper part of this line which indicates the high permeability and

selectivity performance. It is interesting to note that despite the emergence of numerous

polymers in the past years, only about a handful of materials have been utilized in the

industrial scale. The primary choices include polycarbonates, polysulfones, polyesters,

polypyrrolones, and polyimides which the latter constitutes the largest family of polymers

for this application.

Besides to polymeric materials which have been widely in use for many years,

inorganic materials are also receiving attentions owing to their superior separation

properties and a rapid growth in market share for this class of membranes are anticipated

in near future. Inorganic membranes especially molecular sieving materials such as silica,

18
porous glass, crystalline zeolites, microporous beryllium oxide powders and carbon

molecular sieves have been under intensive research studies due to their potential in

surpassing the traditional upper bond of glassy polymers [38,40,41].

Figure 1.2. Trade-off line of oxygen permeability and oxygen/nitrogen selectivity [39].

The position of molecular sieving membranes with respect to trade-off line was depicted

in previous figure. Although among all, zeolites and carbon molecular sieves have

attracted more attentions in development of high performance membranes, nevertheless,

their high production cost, fragility and reproducibility are the main obstacles for large

scale applications.

19
Among the versatile polymeric material, polyimides are a class of organic resins

which exhibit outstanding engineering properties, especially thermal stability. Their

introduction in the 1960s made significant contributions to meet exacting demands of

industry, aerospace, military and nuclear developments. Although polyimide resin

structures were known some years earlier, the commercial developments by DuPont led

in 1961 to the introduction of film (under the tradename of Kapton) and varnishes for

electrical applications. Subsequently moulding powders, laminating resins, foams and

composite materials and especially high performance adhesives became available.

Generally polyimide materials due to their peculiar properties have become

important and are extensively used in variety of technological applications, such as

semiconductor devices, high-temperature adhesives and high-performance composite

materials. Polyimides have exceptional heat resistance, where very high-level mechanical

properties can be sustained at temperatures of 250°C or more. Excellent electrical,

solvent resistance, flame retardance, abrasion resistance, oxidative and radiation

resistance properties have led to a range of critical applications. The possible replacement

of some metal and ceramic components shows the extraordinary advances made in the

developments of organic plastics within a short time span. Industries such as nuclear,

aerospace, electronics, automotive, etc., as well as products such as self-lubricating

pumps, bearings and such, indicate the true value and penetration of polyimides from

exotic engineering areas to those of everyday life.

20
1.3.2 Membrane Design and Structures

1.3.2.1 Dense Versus Porous Membranes

Membranes, according to their structural characteristics, can be divided into two

general categories of dense and porous which can have significant impacts on their

performance as separators and/or reactors. Dense membranes are free from discrete or

well-defined pores or voids. The difference between the two types can conveniently be

determined by the presence of any pore structure under electron microscopy. The

effectiveness of a dense membrane strongly depends on the intrinsic properties of the

material, the species to be separated and their interaction with the membrane. The

microstructure of a porous membrane can vary according to the method of preparation.

Those membranes that possess essentially straight pores across the membrane thickness

are referred to as straight pore or nearly straight pore membranes. The majorities of

porous membranes, however, have interconnected pores with tortuous paths and are

called tortuous pore membranes.

1.3.2.2 Symmetric Versus Asymmetric Membranes

When the separating layer and the bulk support designed for mechanical strength

are indistinguishable and show an integrally homogenous structure and composition in

the direction of the membrane thickness, it is called a symmetric or isotropic membrane.

Since the flow rate through the membrane thickness is inversely proportional to the

membrane thickness, it is very desirable to make the homogenous membrane layer as thin

as possible. However, very thin stand-alone membranes typically do not exhibit

mechanical integrity to withstand the usual handling procedures and processing pressure

21
gradients found in many separation applications. A practical solution to the dilemma has

been the concept of an asymmetric or composite membrane where the thin separating

layer and the open-cell mechanical support structures are distinctly different

[42,43,44,45]. In this anisotropic arrangement, separation of the species in the feed

stream and ideally the majority of the flow resistance (or pressure drop) also take place

primarily in the thin separating layer. The underlying support layer should be

mechanically strong and highly porous that prevent from any considerable contribution to

the transport resistance of the penetrants across the membrane.

A membrane is called asymmetric, if it has a graded pore structure made on the

processing step and frequently from the same material across its thickness. On the other

hand, the membrane is called asymmetric composite, if it is constructed of two or more

distinctively different layers made at different steps. A predominantly thick layer in this

type of membrane provides the necessary mechanical strength to other layers and is

called the support layer or bulk support. Asymmetric composite membranes have the

advantage that the separating layer and the support layer(s) can be tailored with different

materials. Permselectivity and permeability properties of the membrane material are

critically important for the separating layer, while the material of the support layer(s) is

chosen based on its mechanical strength and other considerations such as chemical

resistance. Ideally, the interface between the two layers and the bulk structure of the

support layer must be porous in order to have a minimal gas transport resistance. In

addition, the two layers must be delaimination-free at the interface.

22
Asymmetric composite membranes can have more than two layers in which the

separating layer is superimposed on more than one support layer. In this case, the

intermediate layer serves as a major purpose of regulating the pressure drops across the

membrane by preventing any appreciable penetration of the very fine constituent particles

into the pores of the underlying support layer(s). Typically the intermediate support layer

is also thin. Figure 1.3 illustrates the structural difference among symmetric and

asymmetric membranes.

Figure 1.3. Typical membrane morphologies [45]

Another promising development on asymmetric membranes worthy of mentioning

is dual-layer hollow fiber membranes. Compared to flat membranes, hollow fiber is more

favored configuration due to the following advantages: 1) large membrane area per

volume unit; 2) high gas flux (i.e., high productivity); and 3) good flexibility and easy

23
handling in the module fabrication. Although high performance polymeric materials have

been synthesized recently with the development of material chemistry, it is not

economically attractive due to high price if these materials are applied for the entire

single-layer hollow fibers. Therefore, dual-layer hollow fiber membranes made by a co-

extrusion technology could be a potential solution. Therefore, fabrication of hollow fibers

in the form of dual-layer can bring considerable cost and time savings.

1.3.2.3 Dynamic in-situ Membranes

There is a type of inorganic membranes that are formed in-situ in the application

environment. They are called dynamic membranes which have been studied in a great

deal especially during the 1960s and 1970s. The general concept is to filter a dispersion

containing the suspended inorganic or polymeric colloids through a microporous support

to form a layer of the colloids on the surface of the support. This layer becomes the active

separating layer (membrane).

Over time, this permselective layer is eroded or dissolved and must be replenished

as they are washed away in the retentate. Commonly used materials for the support are

porous stainless steel, carbon or ceramics. Dynamic membranes have been studied for

reverse osmosis applications such as desalination of brackish water, but found to be

difficult to provide consistent performance and the added cost of the consumables makes

the process unattractive economically. Therefore, only limited applications for recovering

polyvinyl alcohol in the field of textile dyeing is commercially practiced.

24
1.3.2.4 Liquid Membranes

There is another type of membrane called liquid membrane where a liquid

complexing or carrier agent supported or immobilized in a rigid solid porous structure

function as the separating transport medium [46,47,48]. The liquid carrier agent

completely occupies the pores of the support matrix and reacts with the permeating

component on the feed side. The complex formed diffuses across the membrane/support

structure and then releases the permeate on the product side and at the same time recovers

the carrier agent which diffuses back to the feed side. Thus permselectivity is

accomplished through the combination of complexing reactions and diffusion.

There is often known as facilitated transport which can be used for gas separation

or coupled transport or separation of metal compounds through ion transport. An example

of the former is some molten salts supported in porous ceramic substrate that are selective

toward oxygen and of the latter is some liquid ion exchange reagent for selectively

transporting copper ions. In this configuration, the composite for the liquid membrane

and its support can be considered to be a special case of dense membranes. Despite their

potential for very high selectivity liquid membranes suffer from physical instability and

are not likely to be a major commercial force in the separation industry.

1.3.3 Membrane Module Configurations

Membrane gas separation system offers great benefit that enables it to be applied

in various industrial processes. Principally, the membrane module is the heart of a

membrane system to determine the yield and efficiency of a separation process. Industrial

25
membrane modules are regularly based on both flat and hollow fiber configurations. The

vast majority of industrial membrane modules are constructed into five basic designs:

plate-and-frame, spiral wound, hollow fiber, tubular and capillary. Flat sheet membranes

are usually contained in the plate-and-frame devices and spiral wound elements, whereas

the modules of hollow fiber, tubular and capillary involve the hollow fiber configuration.

Three most general modules employed in industrial membrane separation processes will

be discussed.

1.3.3.1 Plate-and-Frame Membranes

Plate-and-frame design replicates conventional filtration setup. It is conceptually

simple, which consists of a package of flat sheet membranes. Plate-and-frame modules

are easy to fabricate and the area of membranes are well defined. The flat sheet

membranes, mainly used for experimental purpose to characterize the intrinsic properties

of membrane are stacked together like a multilayer sandwich in a frame. This module

consists of a cylindrical tube, spacer material to separate the membrane envelopes and

rubber gaskets to direct the flow through the module and seal the assembly [44]. The

plate-and-frame package design is illustrated in Figure 1.4. Lowest surface area/unit

separator volume (~100-400m2/m3) is the major drawback of plate-and-frame module.

Consequently, this module configuration is not favorable for gas separation application

and relegated to oxygen enrichment for small scale medical application [49].

26
Figure 1.4. Schematic drawing of a plate-and-frame module [44,50]

1.3.3.2 Spiral Wound Modules

The spiral wound module maintains the simplicity of flat membranes fabrication,

but it is the next logical step from a flat membrane. This element increases the packing

density (membrane surface per module volume) remarkably to 300-1000 m2/m3

compared to plate-and-frame modules. As shown in Figure 1.5, the assembly consists of a

sandwich of flat sheet membranes to form an envelope enclosing a separator/spacer in

them to provide mechanical strength and permeate flow space. The membranes envelope

is wound around a central core of a perforated collecting tube. When a spiral wound

27
module is in operation, the feed gas flows outside the membrane envelope and the

permeate is collected inside and removed through the central collector.

Figure 1.5. Spiral wound elements and assembly

1.3.3.3 Hollow Fiber Modules

The hollow fiber module consists of a large number of hollow fibers assembled

together into a pressure vessel. The membranes are thin hollow tubes with a very small

diameter. The geometric arrangement of module is similar to conventional heat-

exchanger assembly, presented in Figure 1.6. It offers a fairly high packing density with a

maximum membrane surface area per unit volume as high as 10,000 m2/m3 [50,51].

Typically the high pressure feed enters either into shell side or bore side and permeate is

collected from the other side. However, the major disadvantage associated with hollow

fibers is significant pressure drops at bore side when large quantities of gas permeating

the membrane. Therefore, for the gas separation, feed stream must be relatively clean

with preferably shell side feeding to avoid high pressure drops and attaining high

membrane area.

28
(Retentate)

(Permeate)

(Feed)

Figure 1.6. Hollow fiber separator assembly [51]

The selection and application of membrane modules are principally depending on

the economic considerations, separation performance and practicability (ease of cleaning,

ease of maintenance, ease of operation, compactness of the system, scale and the

possibility of membrane replacement). Table 1.5 shows the characteristic of a basic

membrane module. According to this Table, spiral wound and hollow fiber modules are

the two major choices for gas separation application. Spiral wound modules tend to have

a lower permeate pressure, whereas hollow fiber modules exhibit the highest packing

density among all the modules. To overcome the problem of pressure drop in hollow

fibers, the compression and recompression of product gas must often be considered but

they usually result in a significant cost addition for a given process. Universally, the need

to achieve high packing density in a cost-effective manner is probably the most critical

criteria to be considered in a membrane system.

29
Table 1.5. Qualitative comparison of various membrane modules [50]

Plate-and- Spiral Hollow


Tubular Frame wound Capillary fiber
Packing density Low Very high
Investment High Low
Fouling tendency Low Very high
Cleaning Good Poor
Membrane replacement Yes/No Yes No No No

1.4 TRANSPORT MECHANISMS IN GAS SEPARATION MEMBRANES

Gas separation by membranes is a straight forward process concept driven by the

pressure gradient imposed between upstream and downstream. Permeation is a rate

controlled process and separation degree is determined by the selectivity of membrane at

the conditions of separation, including pressure, temperature, flow rate, and membrane

area [52]. A membrane will separate gases only if some components pass through the

membrane more rapidly than others. Depending on the pore sizes in membrane matrix,

there are four fundamental transport mechanisms in gas separation membranes, namely

(1) Poiseuille flow, (2) Knudsen diffusion, (3) Molecular sieving and (4) Solution-

diffusion, as illustrated in Figure 1.7.

From this Figure, it is obvious that there are two major types of membranes

involved in the gas separation. The first is a porous membrane in which the gases are

separated on the basis of their molecular size through the small pores in the membrane

matrix, by Poiseuille flow, Knudsen diffusion or Molecular sieving. Nevertheless, the

vast majority of commercial application is based on nonporous membranes (containing

no holes or pores in the conventional sense) through solution-diffusion mechanism.

30
Poiseuille Flow Knudsen Diffusion Molecular Sieving Solution-diffusion

Figure 1.7. Schematics of main mechanisms of membrane-based gas separation

1.4.1 Poiseuille Flow

The Poiseuille flow also known as viscous flow occurs when the mean pore

diameter is larger than the mean free path of the gas penetrants. The mean free path here

refers to the average distance traversed by a gas molecule between collisions and it is

pressure and temperature dependent. In this condition, membrane contains pores larger

enough to allow convective flow, where gas molecules collide exclusively with each

other and no separation is obtained between the gas components. This type of transport

mechanism is observed for the membranes having the much larger pore sizes than gas

molecules, at pore size, d p >10µm and the flux is proportional to r4.

1.4.2 Knudsen Diffusion

Convective flow will be replaced by Knudsen diffusion in a porous membrane

whose pore sizes are less than the mean free path of the gas molecules [42]. Gas

31
molecules therefore interact with pore walls much more frequently than with one another

and allow lighter molecules to preferentially diffuse through pores to achieve separation.

Knudsen diffusion principally takes place in the membranes with the pore size of 50-100

Å in diameter [27].

Knudsen diffusion is achieved when permeating species flow via the membrane

almost independent of one another. Hence, the Knudsen diffusion coefficient, D k (m2.s-1)

is independent of pressure. For an equimolar feed the permeation rate of Knudsen

diffusion is inversely proportional to the square root of the molecular weight of the

different compounds in the following equation [53]:

𝐷𝐷𝑘𝑘 = 0.667𝑟𝑟𝑟𝑟 = 97𝑟𝑟�𝑀𝑀𝑇𝑇𝑤𝑤 (1.1)

where the average pore radius is given by r (m), υ is the average molecular velocity (m.s-
1
) and T is the operating temperature. Whilst the highest attainable separation factor

between two different gas molecules i and j equals to the square root of the ratio of the

molecular weights of the species, as shown in equation (1.2).

𝑀𝑀
𝛼𝛼𝑖𝑖𝑖𝑖 = � 𝑀𝑀𝑤𝑤𝑤𝑤 (1.2)
𝑤𝑤𝑤𝑤

Consequently, such membranes are not commercially attractive in general for standard

application due to their relatively low selectivity.

1.4.3 Molecular Diffusion

Molecular sieving separation is primarily based on the precise size discrimination

between the gas molecules through the ultramicropores (<7Å in diameter). Molecular

32
sieving membranes become increasingly important in the gas separation especially for

inorganic membranes due to their reported higher productivity and selectivity than

solution-diffusion polymeric membranes [54,55]. Their porous nature has led to high

permeability, while the high selectivity is achieved through effective size and shape

exclusion between the gas species. This happens when the pore diameters are small

enough to allow the permeation of smaller molecules while obstructing the larger

molecules to diffusive through. Even if both molecules enter the pores; the larger one

would experience stronger repulsive forces. This energetically biased selectivity is called

“energetic selectivity”. It is also believed that the more subtle contribution to selectivity

is from “entropic selectivity” [56], in which the rotational freedom of one component is

restricted to a higher degree than that of the other components.

Carbon molecular sieve membranes (CMSMs) and zeolites are the typically

membranes dominated by molecular sieving mechanism and give the high separation

performance. The ratio of the gas molecular size to the micropore diameter controls the

gas permeation rate and separation in molecular sieving materials [57]. For example,

zeolite 4A with a pore size of 3.8 Å has an O 2 /N 2 selectivity of approximately 37 at 35oC

[58].The ideal selectivity of carbon molecular sieve membranes for CO 2 /CH 4 separation

has been reported to be as high as 200 [59]. The pore size of zeolite maybe controlled or

modified by choosing suitable synthesis, de-almunation, and ion-exchange [60]; as for

CMSMs, the pore size can be controlled by choosing polymer precursor, pyrolysis

conditions, pretreatment and post treatment [61].

33
1.4.4 Solution–diffusion Model

The solution–diffusion mechanism is widely employed for explaining gas

transport through dense polymeric membranes [62]. Based on this model, the gas

permeability (P) across the membrane is a product of solubility (S) and diffusivity (D).

P = S×D (1.3)

The ideal gas pair (A/B) permselectivity comprises the solubility (S A/S B ) and diffusivity

(D A /D B ) selectivities, which should complement each other for optimal separation factor.

PA  S A   DA 
α AB = =  ×  (1.4)
PB  S B   DB 

The gas solubility of polymeric membranes is dependent on the (1) condensability of the

penetrant gases, (2) nature of interactions between the polymer and the penetrants and (3)

chain packing density in glassy polymers [63]. For natural gas separation, the solubility

of CO 2 is higher than CH 4 due to the significant difference in the condensability of the

gases (T c, CO2 = 304 K and T c ,CH4 = 191 K). On the other hand, since the gas diffusivity

mainly depends on the kinetic diameter (d KT ) of the penetrants, the diffusivity of CO 2 is

expected to be larger than CH 4 (d KT,CO2 = 3.3Ǻ and d KT,CH4 = 3.8Ǻ) [64]. This explains

the preferential permeation of CO 2, relative to CH 4 for most glassy polymeric

membranes.

1.4.5 Gas Sorption and Diffusion in Glassy Polymers

Several models have been established to explain the sorption of gas molecules in

glassy polymeric membranes but the use of a model based on the dual-mode sorption

theory has been more prevalent [65]. The dual-mode sorption model (Eq. (1.5)) suggests

the presence of Henry and Langmuir sorption sites in glassy polymers [31,32,33]. The
34
solubility coefficient is the ratio of total penetrant concentration to pressure, given by Eq.

(1.6).

C ' H bp
C = CD + CH = k D p + (1.5)
1 + bp

C c' b
S= = kD + H (1.6)
p 1 + bp

C is the total gas concentration, p is the pressure and k D represents the Henry law

constant. C D and C H are the gas concentrations at the Henry and Langmuir sites,

respectively. C’ H is the Langmuir capacity constant which varies with polymer type and

structure. b is associated with the affinity of the gas molecules to the Langmuir sites.

Paul and Koros proposed the dual-mobility model to account for the difference in

inherent gas diffusion of the Langmuir and Henry sites [67,68,69]. This model is

presented by equation 1.5 where D D and D H are the diffusion coefficients in the Henry

and Langmuir sites, respectively. F is the ratio of D H to D D and K is (c’ H b/k D ). p 2 is the

pressure at the upstream boundary.

 FK 
P = k D 1 +  DD (1.7)
 1 + bp 2 

From a classical approach, the gas diffusion coefficient, D can be expressed in the form

of an Arrhenius equation shown by equation 1.8.

 − ED 
D = D0 exp  (1.8)
 RT 

Here, D 0 is the pre-exponential factor, R is the gas constant, T is the temperature in

Kelvin and E D is the activation energy for diffusion. Based on another model proposed

35
by Brandt, E D is related to polymer dependent parameters c and f as well as the kinetic

diameter (d) of the gas penetrant which is presented in equation 1.9 [70].

E D = cd 2 − f (1.9)

According to this model, polymers that have high diffusivity selectivity should have high

values of c, which implies high polymer chain stiffness. The term f / c is related to the

average distance between polymer chains.

1.4.6 Free Volume in Glassy Polymers

The presence of free volume in polymers is attributed to the non-equilibrium state

of the polymeric matrix induced by the limited mobility and long relaxation times of

entangled polymer chains [71,72]. For polyimides with flexible backbones, the formation

of irregular voids is due to the minimum barriers of each rotational bonds occurring at

irregular intervals [73,74]. The way in which the gas molecules transverse across the

membranes is described by a series of diffusional jumps through the temporary channels

created by the constantly vibrating polymer chains as shown in Figure 1.8. Therefore, the

size and distribution of free volume, polymer chain mobility and penetrant size govern

the overall gas transport. The free volume entities of polymers are frequently expressed

by the fractional free volume (FFV) [75,76,77,78,79]. The FFV model provides a rough

estimate for gas transport in polymeric membrane but fails to account completely for

experimentally determined results, especially for highly condensable gases [80]. The

discrepancy in the physical interpretation of the parameters in the FFV model created

additional problem for the general acceptance of this model [80,81].

36
1

A
1

B 4 4
1

5 5
1
1 C
2 B
A 2
C
3
3

Position of molecules at one instant Position of molecules at the next instant

Figure 1.8. Motion of gas molecules within the cavities through a series of

diffusional jumps

Free volume elements are characterized by the wide angle x-ray diffraction

(WAXD), which provides the average d-spacing of polymers [81]. Recent investigations

using molecular modelling reveals that the d-spaces are better reflected by the x-ray

scattering associated with 2θ values between 20o and 28o [73]. In a recent work by Miyata

et al., the correlation between the dielectric constant of polyimide and FFV was

established [82]. The positron annihilation lifetime spectroscopy (PALS) is also

employed for free volume analysis [83-87]. A correlation between the measured lifetime

of orthopositronium (o-Ps) and the size of the void volume is used to determine the

average hole size. The reported cavity radii of polyimide free volume usually fall in the

range of 2.5 to 4.5 Ǻ [ 79,86-88] which are comparable to d-spaces measured by WAXD.

These analytical techniques provide valuable information which can be utilized for

optimizing the gas separation performance of polyimide membranes. However, the

characterization is not conducted in-situ and thus, the results are not truly reflective of the

37
actual operating conditions. It should be noted that the interactions between the polymer

chains and the gas molecules affects the chain mobility and the free volume of the

polymeric membranes. Therefore, novel analytical methods capable of examining

polymeric membranes that are subjected to a certain gas pressure are necessary. In

addition, characterization techniques with higher sensitivity and accuracy, simpler

operating procedures and lower costs are required.

1.5 RESEARCH OBJECTIVES AND ORGANIZATION OF DISSERTATION

This research study comprised of four main aspects including the development of

new nanocomposite membranes by using MgO nanoparticles with enhanced gas

separation performance, fabrication of membranes with interpenetration networks from

miscible polymeric blends for hydrogen separation and purification, production of carbon

molecular sieving membranes with enhanced performance for diverse gas separation

applications and fabrication of dual-layer hollow fiber membranes for gas separation in

the semi-industrial scale.

The key impetus for this research is underachievement of conventional polymeric

membranes in surpassing the trade-off between productivity and selectivity. The main

objective of this research is to identify, explore and develop viable techniques in order to

address this challenge and provide membranes with enhanced gas separation

performance. In every aspect of this research study the effect of various variables of

interests will be discussed. For the nanocomposite membranes, the effect of incorporation

38
of nanoscale magnesium oxide particles at different loading was examined. Number of

post-treatment approaches was employed in order to optimize the separation performance

of the membranes. On the other hand, blending technology was exploited for fabrication

of polymeric and carbon molecular sieve membranes. The properties of these membranes

were finely tuned through proper adjustment of parameters and application of suitable

modification techniques. The potentials of polymeric blend membranes were further

extended by fabrication in form of hollow fiber membranes. In order to achieve the above

objectives, the scopes of this study are drawn as follows:

1. Development of new nanocomposite membranes by using MgO nanoparticles

with enhanced gas separation performance

2. Fabrication of membranes with interpenetration networks from miscible

polymeric blends for hydrogen separation and purification

3. Production of carbon molecular sieving membranes with enhanced

performance for diverse gas separation applications

4. Fabrication of novel dual-layer hollow fiber membranes from polymeric blends

for gas separation

This dissertation is organized and structured into seven chapters. Chapter One is

an introductory chapter of this dissertation. It gives an introduction to gas separation

membranes, encompassing the historical development of membranes, industrial

applications and principles of gas separation membrane technology. The discussion

includes membrane materials, structure and design of membranes and membrane module

configurations. In addition, the governing transport mechanisms in gas separation

39
membranes are described. The research objective and outline of this dissertation are also

presented in this chapter.

The information of strategies and approaches in development of membranes for

gas separation are given in Chapter Two. This chapter begins with the description of the

key limitations of existing polymeric membranes. Approaches for molecular design and

tailoring the properties of membranes materials are introduced. Various key modification

techniques including polymer blending, hybridation, thermal annealing, grafting and

halogentaion of polymer backbone and chemical cross-linking are reviewed. Finally, the

current and future developments in each area are presented.

Chapter Three describes the development of gas separation membranes with

enhanced performance by introduction of nanosize magnesium oxide particles followed

by treatment with silver ions. The effects of particle loading on the gas permeability and

separation performance of the membranes are investigated. It is shown that the highest

permeability was observed for the membranes containing 40 wt% MgO loading but with

diminished selectivity compared to the neat Matrimid membrane. The effect of applying

various modification techniques including heat treatment at various temperatures and also

silver ion treatment on the structural and transport properties of the membranes are

discussed.

Chapter Four demonstrates the successful implications of blending technique cum

chemical modification for the fabrication of high performance polymeric membranes for

40
gas separation applications. The effect of variation in composition on miscibility and

microstructure, gas permeability and selectivity of blend membranes is investigated. The

results of the analysis of the microstructure of membranes in terms of their chain packing

density, d-spacing and segmental mobility of polymer chains are provided. Separation

performance of membranes is further ameliorated through chemical modification of blend

constituents. The results indicate that higher tendency of Matrimid toward cross-linking

reaction contributes more in controlling the transport properties of membranes through

diffusion coefficient by increase in chain packing density and diminishing the excess free

volumes. The promising features of developed membranes for various gas separation

applications are presented.

Chapter Five introduces the development of membranes composed of

poly(benzimidazole) (PBI) and various polyimides and the study regarding to their

capabilities in formation of homogenous blends with satisfactory morphology for

fabrication carbon molecular sieve membranes. Attempts are made to unravel the

correlations between the pyrolysis process parameters and separation performance of

developed membranes. Modification of precursors by chemical cross-linking prior to

carbonization and its effects on the separation performance of the membranes is

discussed. This research study accentuates the necessity and importance of adopting a

proper strategy in exploiting synergistic beneficial features of advanced materials (PBI),

technology (polymer blending), process (carbonization) and modification techniques

(cross-linking) toward achieving membranes with desired performance.

41
Chapter Six introduces the development of novel high performance gas separation

membranes through integrating of the concepts of polymer blending and dual-layer

hollow fiber spinning process. Attempts are made to unravel the effect of various key

parameters on the morphological evolution, transport properties and separation

performance of the fabricated hollow fiber. The effects of various parameters such as air

gap distance, application of elongational drawing, outer dope flow on the gas permeance

and selectivity of the membranes are discussed. Silicone rubber coating and chemical

modification are used as potential tools for modification of the membranes and tuning

their gas transport and separation performance.

The general experimental approaches and methodologies, along with the materials

involved are provided in their respective research study. This includes the details of

membrane preparation, characterization, modification, treatment and other miscellaneous

tests.

General conclusions drawn from this research study are summarized in Chapter

Seven. Inclusive in this final chapter are some recommendations and suggestions for

future research related to this work.

42
1.6 REFERENCES

[1] J.K. Mitchell, On the Penetrativeness of Fluids, J. Med. Sci., 13 (1830) 36.

[2] J.K. Mitchell, On the Penetration of Fluids, Am. J. Med. Sci., 13 (1833) 100.

[3] T. Graham, On the Law of the Diffusion of Gases, Phil. Mag., 2 (1833)

pp.175,269,351.

[4] A. Fick, Ueber Diffusion, Ann. Phys. Chem. 94 (1855) 59.

[5] T. Graham, On the Absorption and Dialytic Separation of Gases by Colloid Septa.

Part I. Action of A Septum of Caoutchouc, Phil. Mag., 32 (1866) 401.

[6] T. Graham, On the Absorption and Dialytic Separation of Gases by Colloid Septa.

Part II. Action of Metallic Septa at a Red Heat, Phil. Mag., 32 (1866) 503.

[7] S. von Wroblewski, Ueber die Natur der Absorption der Gase, Ann. Phys. Chem. 8

(1879) 29.

[8] H. Kayser, Wied. Ann. Phys., 43 (1891) 544.

[9] D.R. Paul and Y.P. Yampol’skii, Polymeric Gas Separation Membranes, CRC, Boca

Raton, 1994, p.1.

[10] S. Loeb, S. Sourirajan, Sea water demineralization by means of an osmotic

membrane, Adv. Chem. Ser., 38 (1962) 117.

[11] J.M.S. Henis, and M.K. Tripodi, Multicomponent membranes for Gas Separations,

US Patent 4,230,463, Monsanto Company, 1980.

[12] J.M.S. Henis, and M.K. Tripodi, Composite Hollow Fiber membranes for Gas

Separation: the Resistance Model Approach, J. Membr. Sci., 8 (1981) 233.

[13] Mohammadi A.T., A review of the applications of membrane separation technology

43
in natural gas treatment, Sep. Sci. Technol., 34 (1999) 2095.

[14] Baker, R.W. Future directions of membrane gas separation technology, Ind. Eng.

Chem. Res. 41 (2002) 1393.

[15] Koros W.J., Fleming G.K., Membrane-based gas separation, J. Membr. Sci., 81

(1993) 1.

[16] Mulder M., Basic principles of membrane technology. Dordrecht: Kluwer Academic

Publishers, 1996

[17] Geankoplis C.J., Membrane separation processes. In: Transport processes and unit

operations, pp 745-799, Prentice Hall, Inc.,1995.

[18] Paul D.R., Yampol’skii Y.P. (Eds.), Polymeric gas separation membranes, CRC

Press, Boca Raton, 1994

[19] Berry R.I. Membrane separate gas, Chem. Eng. 88 (1981) 63.

[20] Stern S.A., polymers for gas separation: the next decade, J. Membr. Sci. 94 (1994) 1.

[21] Spillman R.W., Sherwin M.B., Gas separation membranes: the first decade.

Chemtech, 20 (1990) 378.

[22] Strathmann H., Membrane separation processes, J. Membr. Sci., 9 (1981) 121.

[23] Spillman R.W. Economics of gas separation membranes, Chem. Eng. Prog. 85

(1989) 41.

[24] Noble R.D., Stern S.A., Membrane Separation Technology., Principles and

applications, Amsterdam: Elsevier Science, 1995.

[25] Crull A. Applications of membrane technology, processings of the 1998 membrane

technology/separation planning conference, pp1-16, 1998 .

[26] Zolandz R.R., Fleming G.K., Gas permeation applications, In: Membrane

44
Handbook, Ed. By Ho W.S.W. and Sircar K.K., New York: Chapman & Hall, 1992.

[27] Koros W.J., Gas separation. In: Membrane Separation Systems: Recent

developments and future directions, Ed. by Baker R.W. Cussler E.L., Eykamp W.,

Koros W.J., Riley R.L., Strathmann H., pp. 189-241, U.S.A.: Noyes Data

Corporation, 1991.

[28] Koros W.J., Mahajan R., Pushing the limits on possibilities for large scale gas

separation: which strategies?, J. Membr. Sci. 175 (2000) 181.

[13] Puri P.S., Gas separation membranes-current status, Chem. Ind. Digest, 9 (1996) 98.

[30] Kesting R.E., Fritzsche A.K., Polymeric gas separation membranes, New York: John

Wiley & Sons, 1993.

[31] Djoekita, G. Characterization and analysis of asymmetric hollow fiber membranes

for natural gas purification in the presence of hydrocarbons, M.Sc. Thesis, University

of Texas at Austin, 2000.

[32] Kirk-Othmer, Encyclopedia of chemical technology, Vol.19, New York: John Wiley

& Sons, 1982.

[33] Rosenberg L., Chong V., Propylene in the twenty-first century, Business intelligence

program, SRI consulting, 1997.

[34] Aoki, T. Macromolecular design of permselective membranes, Prog. Polym. Sci., 24

(1999) 951.

[35] Scott K., Membrane separation technology, Oxford: scientific & technical

information, 1990.

[36] Strathmann H., Membrane separation processes:current relevance and future

opportunities, AIChE Journal, 47 (2001) 1077

45
[37] Nunes S.P., Peinemann K.-V., Membrane technology in the chemical industry,

Weinheim: Wiley-VCH, 2001.

[38] Brek D.W. Zeolite molecular sieves: structure, chemistry and use. Florida: Robert E.

Kreiger Publishing Company.1973

[39] Singh A., Koros W.J., Significance of Entropic Selectivity for Advanced Gas

Separation Membranes, Ind. Eng. Chem. Res. 1996, 35, 1231-1234.

[40] Centeno T.A., Fuertes A.B., Carbon molecular sieve gas separation membranas base

don Poly(vinylidene chloride-co-vinyl chloride) carbon, 38, pp 1067-1073, 2000

[41] Saufi S.M., Ismail A.F., Fabrication of carbon membranes for gas separation-

Review, Carbon 42 (2004) 241.

[42] W.J. Koros, G.K. Fleming, Membrane-based gas separation, J. Membr. Sci., 83

(1993) 1.

[43] J.M.S. Henis, M.K. Tripodi, Composite hollow fiber membranes for gas separation:

The resistance model approach, j. Membr. Sci., 8 (1981) 233.

[44] D.R. Paul, Y.P. Yampol’skii (Eds.), Polymeric gas separation membranes, CRC

press, Boca Raton, 1994.

[45] T.S. Chung, A review of microporous composite polymeric membrane technology

for air-separation, Polym. And Polym. Comp., 4 (1996) 269.

[46] M. Ashraf Chuadry, S. Ahmad, M.T. Malik, Supported liquid membrane technique

applicability for removal of chromium from tannery wastes, Waste Manage., 17

(1997) 211.

[47] J. de Gyves, E. Rodriguez desan Miguel, Metal ion separation by supported liquid

membrane, Ind. Eng. Chem. Res., 38 (1999) 2182.

46
[48] F.J. Algucil, M.A. Villegas, Liquid membranes and the treatment of metal bearing

wastewaters, Rev. metal. Madrid, 38 (2002) 45.

[49] R.E. Kesting, A.K. Fritzsche, Polymeric gas separation membranes, John Wiley &

Sons, New York, 1993.

[50] M. Mulder, Basic principles of membrane technology, Dordrecht: Kluwer Academic

Publishers, 1996.

[51] C.J. Geankoplis, Membrane separation processes: In transport processes and unit

operations, Prentice-Hall, Inc., American, (1995) 745.

[52] R.W. Spillman, M.B. Sherwin, Gas separation membranes: The first decade,

Chemtech, 20 (1990) 378.

[53] R.J.R. Ulhum, K. Keizer, A.J. Burggraff, Gas and surface diffusion in modified

gamma alumina systems, J. Membr. Sci., 46 (1989) 225.

[54] J. Koresh, A Soffer, The carbon molecular sieve membranes: General properties and

the permeability of methane/hydrogen mixtures, Sep. Sci. Technol., 22 (1987) 973.

[55] J.D. Way, D.L. Roberts, Hollow fiber inorganic membranes for gas separations, Sep.

Sci. Technol., 27 (1991) 29.

[56] A. Singh, W.J. Koros, Significance of entropic selectivity for advanced gas

separation membranes , Ind. Eng. Chem. Res., 35 (1996) 1231.

[57] A.B. Fuertes, D.M. Nevskaia, T.A. Centeno, Carbon composite membranes from

Matrimid and Kapton polyimides for gas separation, Micro. Meso. Mat., 33 (1999)

115.

[58] D.M. Ruthven, R.I. Derrah, Diffusion of monatomic and diatomic gases in 4A and

5A zeolites, J. Chem. Soc., Faraday Trans., 71 (1975)2031.

47
[59] K. Steel, Carbon membranes for challenging gas separations, Ph.D. Dissertation,

The University of Texas at Austin, 2000.

[60] J.C. Jensen, Advanced zeolite science and applications, Amsterdam, Elsevier, New

York, 1994.

[61] P.S. Tin, Membrane materials and fabrication for gas separation, Ph.D. Thesis,

National University of Singapore, 2004.

[62] Koros WJ, Fleming GK, Jordan SM, Kim TH, Hoehn HH. Polymeric membrane

materials for solution–diffusion based permeation separations. Prog Polym Sci

1988;13:339–401.

[63] Bhide BD, Voskericyan A, Stern SA. Hybrid processes for the removal of acid gases

from natural gas. J Membr Sci 1998;140:27–49.

[64] Cecopieri-Gomez ML, Palacios-Alquisira J, Dominguz JM. On the limits of gas

separation in CO2/CH4, N2/CH4 and CO2/N2 binary mixtures using polyimide

membranes. J Membr Sci 2007;293:53–65.

[65] YampolskiiY, PinnauI, FreemanBD.Materials science ofmembranes for gas and

vapor separation. Hoboken, NJ: Wiley-Interscience; 2006.

[66] Sanders ES. Penetrant-induced plasticization and gas permeation in glassy polymers.

J Membr Sci 1988;37:63–80.

[67] KorosWJ, Paul DR, Rocha AA. Carbon dioxide sorption and transport in

polycarbonate. J Polym Sci Polym Phys Ed 1976;14:687–702.

[68] Paul DR, KorosWJ. Effect of partially immobilizing sorption on permeability and

the diffusion time lag. J Polym Sci Polym Phys Ed 1976;14:675–85.

48
[69] Wang R, Cao C, Chung TS. A critical reviewon diffusivity and characterization of

6FDA-6FpDA polyimide membranes for gas separation. J Membr Sci 2002;198:259–

71.

[70] Brandt WW. Model calculation of the temperature dependence of small molecule

diffusion in high polymers. J Phys Chem 1959;63:1080–5.

[71] Prausnitz JM, Lichtenthaler RN, Azevedo ED. Molecular thermodynamics of fluid-

phase equilibrium. New York: Prentice-Hall; 1986.

[72] Koros WJ, Hellums MW. Transport properties. In: Kroschwitz JI, editor.

Encyclopedia of polymer science. 2nd ed. New York: Wiley-Interscience; 1989.

[73] Jacobson SH. Molecular modeling studies of polymeric gas separation and barrier

materials: structure and transport mechanisms. Polym Adv Technol 1994;5:724–32.

[74] Herchel M, Hofmann D, Pullumbi P. Molecular modeling of small-molecule

permeation in polyimides and its correlation to free-volume distribution.

Macromolecules 2004;37:201–14.

[75] Van Krevelen DW. Properties of polymers-their correlation with chemical structure;

their numerical estimation and prediction from additive group contributions.

Amsterdam: Elsevier; 1990.

[76] Bondi A. Physical properties of molecular crystals, liquids and glasses. New York:

Wiley; 1968.

[77] FujitaH. Diffusion in polymer-diluent systems. Fortschr Hochpolym Forsch

1961;3:1–47.

[78] Lee WM. Selection of barrier materials from molecular structure. Polym Eng Sci

1980;20:65–9.

49
[79] Alentiev AY, Yampolskii YP. Meares equation and the role of cohesion energy

density in diffusion in polymers. J Membr Sci 2002;206:291–306.

[80] Park JY, Paul DR. Correlation and prediction of gas permeability in glassy polymer

membrane materials via a modified free volume based group contribution method. J

Membr Sci 1997;125:23–39.

[81] Boehme RF, Cargill GS. X-ray scattering measurements demonstrating in-plane

anisotropy in Kapton polyimide films. In: Mittal KL, editor. Polyimides. New York:

Plenum Press; 1984.

[82] Miyata S, Sato S,Nagai K,Nakagawa T, Kuda K. Relationship between gas transport

properties and fractional free volume determined from dielectric constant in

polyimide films containing the hexafluoroisopropylidene group. J Appl Polym Sci

2008;107:3933–44.

[83] Shrader DM, Jean YC. Positron and positronium chemistry. Amsterdam: Elsevier;

1988.

[84] Victor JG, Torkelson JM. On measuring the distribution of local free volume in

glassy polymers by photochromic and fluorescence techniques. Macromolecules

1987;20:2241–50.

[85] Bondar VI, Freeman BD, Yampolski YP. Sorption of gases and vapors in an

amorphous glassy perfluorodioxole copolymer. Macromolecules 1999;32:6163–71.

[86] Tanaka K, Kawi T, Kita H, Okamoto K, Ito Y. Correlation between gas diffusion

coefficient and positron annihilation lifetime in polymers with rigid polymer chains.

Macromolecules 2000;33:5513–7.

50
[87] Nagel A, Gu1nther-Schade K, Fritsch D, Strunskus T, Faupel F. Free volume and

transport properties in highly selective polymer membranes. Macromolecules

2002;35:2071–7.

[88] Alfrey T, Goldfinger G,Mark H. The apparent second-order transition point of

polystyrene. J Appl Phys 1943;14:700–5.

51
CHAPTER TWO

STRATEGIES AND APPROACHES IN DEVELOPMENT OF GAS

SEPARATION MEMBRANES

The worldwide demand for the use of natural gas as a cleaner and more efficient

fuel is constantly rising. In addition, natural gas (methane) is considered as a principle

feedstock for the chemical industry. The global consumption of natural gas is projected to

increase from 95 trillion cubic feet in 2003 to 182 trillion cubic feet in 2030 [1]. Since

raw natural gas from various origins is different in composition, the potential growth in

the demand for methane implies the need for separation technologies with outstanding

efficacy. Although methane constitutes the key component of natural gas, it may also

contain considerable amounts of impurities including water, carbon dioxide, nitrogen,

hydrogen sulfide and other hydrocarbons. One of the major steps in natural gas treatment

involves the removal of acid gases (CO 2 and H 2 S) before it is delivered to the pipelines

or stored into portable cylinders as compressed natural gas. The necessity for natural gas

sweetening is to (1) increase the fuel heating value, (2) decrease the volume of gas to be

transported in pipelines and cylinders, (3) reduce pipeline corrosion within the gas

distribution network and (4) prevent atmospheric pollution [2].

Currently, natural gas treatment is one of the principal industrial gas separation

processes and the market for new purification systems may exceed $5 billion per year [3].
52
Absorption of acid gases in basic solvents (such as amine or hot aqueous potassium

carbonate solutions), pressure swing adsorption and membrane technology are examples

of natural gas purification technologies [4]. Despite the full maturity of the gas absorption

processes, they basically suffer fromthe need for solvent regeneration process, difficulty

to be justified for small gas fields, large footprint for offshore applications and lack of

robustness against fluctuations of feed composition. Similarly, pressure swing adsorption

(PSA) is highly capital intensive and requires significant energy consumption. In

comparison, the advantages of membrane technology include potentially higher energy

efficiency, ease of scale-up, better utilization of space and environmentally friendly [5].

In particular, from the economic point of view, membrane technology may appear more

advantageous for small-to-medium scale separations and non-stringent product purity

requirements. It is worthwhile to note that the current trend is the development of hybrid

membrane/absorption processes capable of eliminating both CO 2 and H 2 S impurities

from natural gas [6]. This is due to the fact that membrane separation technology alone is

not yet well proven for H 2 S removal [7].

The gas transport through membranes has been studied since the 1950s but the

immaturity of the fabrication technology to produce high flux membranes has retarded

their use in industry. The application of membrane technology for CO 2 removal from

natural gas emerged in the 1980s after several major breakthroughs. One of the principal

revolutions was the preparation of high flux asymmetric Loeb-Sourirajan membranes

with selective layer thicknesses of less than 0.1µm [8]. However, the formation of defects

during membrane fabrication restrained the direct application of these membranes for gas

53
separations. Henis and Tripodi’s invention of applying a thin layer of silicone rubber

coating on the membrane surface provided a solution to overcome this limitation [9]. The

emergence of new module designs like spiral-wound and hollow fibers led to the

fabrication of smaller membrane configurations with higher surface-to-volume ratios

[10].

In addition to the fabrication methods and module designs, works have also been

in progress to develop membrane materials with enhanced properties and separation

performance in process environments [11]. For membranebased natural gas purification,

it is highly desirable to explore materials which can selectively remove CO 2 from the

mixtures, thereby maintaining CH 4 at or near feed pressure and avoiding costly gas

recompression. There are many criteria for the selection of membrane materials and the

most important ones are high gas permeation flux and gas pair selectivity [8]. Other

factors may include the mechanical and thermal stability, manufacturing reproducibility,

tolerance to contaminants and economical aspects. A number of reviews examining the

various materials for gas separation membranes have been published over the last two

decades [3,6,8,12–19]. Among the studied materials, it was noted that polyimides (PI)

which exhibit high permselectivity for various gas pairs (e.g. CO 2 /CH 4 and O 2 /N 2 ), high

chemical resistance, thermal stability and mechanical strength have attracted much

attention. This is well supported by the significant increase in the number of publications

related to polyimide membranes over the years. Du Pont Co. (USA) and Ube Industries

(Japan) have been pioneers in the commercial application of polyimides in separation

processes [21]. The industrial application of polyimide membranes by DuPont began in

54
1962 with separation of helium from natural gas which was subsequently applied to the

separation of CO 2 from various gas streams.

In view of the potential use of polyimides as materials for fabricating high

performance gas separation membranes, an understanding of the physical and chemical

properties in relation to the gas separation performance must be established. This is due

to the fact that the intrinsic gas transport properties of polyimide membranes have been

found to be linked to their chemical structures. The interdependencies have been well

summarized in several articles [22-27]. Specific tailoring of the polyimide molecular

structure can be regarded as a viable approach in obtaining the desired gas separation

properties of the resultant membranes. Alternatively, exploration of modification

techniques may open up additional opportunities for better exploitation of the existing

polyimides.

2.1 KEY LIMITATIONS OF EXISTING POLYMERIC MEMBRANES

2.1.1 Permeability Versus Permselectivitiy

It is desirable to obtain gas separation membranes with both high permeability

and permselectivity. In the early 90s, Robeson gathered the permeability and

permselectivity data for various binary gas pairs from the literature and discovered the

existence of upper bound curves indicating a general tradeoff between permeability and

permselectivity [28]. The empirical correlation of the Robeson upper bound is

represented by equation 2.1.

55
ln α AB = ln β AB − λ AB ln PA (2.1)

In a subsequent study by Freeman, a more comprehensive physical interpretation of the

upper bound was conducted and the proposed correlation is given by equation 2.2 [29].

For glassy polymers, a and b are constants equal to 0.64 and 11.5 cm2/s, respectively and

f is a polymer dependent parameter [30-31].

 d  2    S   d  2  
1 − a 
ln α AB = −   − 1 ln PA + ln A  −  B  − 1 b −
B
f 
 − ln S A (2.2)
 d A     S B   d A    RT   

A comparison of equations 2.1 and 2.2 indicates that the slope of the upper bound, λ AB is

dependent only on the gas kinetic diameters (d A and d B ) while the intercept of the upper

bound is a strong function of the nature of the gaseous penetrants and the polymer.

Hence, for a given gas pair, the slope of the upper bound remains constant. This implies

that a performance surpassing the upper bound is achievable by placing greater emphasis

on increasing the solubility selectivity, inter-chain spacing and chain stiffness. Recently,

Robeson utilized a larger data set than before to re-examine the upper bound [32].

Despite the continuous efforts to develop novel membrane materials, the upper bound

line for CO 2 /CH 4 exhibited only minor shifts [32]. For our current study, a thorough

literature survey was conducted to obtain the CO 2 /CH 4 separation performance of

aromatic polyimide membranes. The collected data are summarized and compared with

the upper bound in Figure 2.1. It is evident that only a minority of the aromatic

polyimides fall close to or above the trade-off line.

56
1000

Present upper bound (2008)


CO2/CH4 permselectivity

Prior upper bound (1991)


100

10

1
0.1 1 10 100 1000
CO2 permeability

Figure 2.1. Robeson’s tradeoff curve for CO 2 /CH 4 gas pair

2.1.2 CO 2 - Induced Plasticization and Conditioning

Penetrant-induced plasticization and conditioning are two aspects which need to

be considered in the fabrication of high performance glassy polymeric membranes. These

phenomena are found to be more severe in natural gas separation which involves highly

condensable gases (e.g. CO 2 , H 2 S, H 2 O) and other heavy hydrocarbons [33,34-40].

These undesirable effects are present in polyimide membranes, despite their good

chemical integrity compared to other polymers. In the following discussion, the gist of

plasticization and conditioning are presented to better understand their impacts on the

separation performance of polyimide membranes for natural gas purification.

57
Plasticization is a repertoire of pressure dependent phenomena caused by the

dissolution of certain components within the polymer matrix which disrupts chain

packing and enhances inter-segmental mobility [33,34]. At relatively low CO 2 pressures,

the gas permeability of a polyimide membrane decreases with increasing pressure. This

behavior is attributed to the gradual saturation of microvoids in the membrane (i.e.

reduced solubility coefficient) which in turn reduces the permeability. Further increase in

pressure results in an upward inflection in gas permeability. The pressure at which the

corresponding gas permeability exhibits a minimum is known as the plasticization

pressure (P plas). The pressure-permeability relationship is inadequate for determining the

incipient point of plasticization since the phenomenon occurs considerably below P plas

[41]. Therefore, another diagnosis to determine the beginning of plasticization involves

examining the absolute heat of sorption as a function of pressure. The onset of

plasticization corresponds to the point at which the change in the heat of sorption with

pressure transforms from a negative value to a positive one [41].

For CO 2 /CH 4 separation using glassy polymeric membranes, CO 2 -induced

plasticization accelerates the permeation of methane, thus deteriorating the CO 2 /CH 4

permselectivity. CO 2 -induced plasticization leads to pressure, temperature, time and

thickness dependence in the gas transport properties of glassy polymeric membranes [41,

42-47]. The temperature dependence of CO 2 -induced plasticization is investigated by

Duthie et al [47]. The variations in CO 2 permeability as a function of CO 2 fugacity at

different operating temperatures reveal that the plasticization phenomenon is more

pronounced at lower temperatures [47]. Plasticization results in non-constant CO 2

58
permeability. For example, the relative increase in the gas permeability of 6FDA-MDA

film is larger than 40% after the lapse of several hours and it does not reach a steady-state

value within the experimentation time [45]. The thickness-dependence of CO 2 -induced

plasticization was clearly illustrated using 6FDA-durene polyimide in a study by Zhou et

al [46]. An accelerated plasticization behavior is observed in ultra-thin dense polyimide

films, attributed to the weak micro-mechanical properties and severe swelling effects

caused by CO 2 .

Another phenomenon closely associated with CO 2 -induced plasticization is the

effects of conditioning. Conditioning is related to the alterations in the properties of

glassy polymers upon exposure to highly soluble gases [48]. CO 2 sorption in polymer can

severely and in most cases permanently change the morphology and transport properties

of the membrane even after the gas molecules are fully desorbed [48]. The resultant

conditioned membranes exhibit irreversible volume dilation and new steady-state gas

transport properties. CO 2 conditioning leads to simultaneous increment in gas diffusivity

and solubility which is dependent on the extent of volume expansion. Polyimides show

significant volume dilation and increment in gas permeability [42]. Since the effects of

CO 2 -induced plasticization and conditioning significantly deteriorate the separation

performance of polyimide membranes, any effort toward its suppression or even

elimination would have great beneficial impacts.

2.1.3 Physical Aging

59
It is well demonstrated that the non-equilibrium state of glassy polymers is mainly

due to the thermal cooling process across the glass transition temperature [49-51]. Hence,

there is a tendency for polymeric chains to undergo “physical aging” which is a gradual

molecular rearrangement process toward attaining an equilibrium state. One consequence

of aging is the densification of the polymer matrix with a corresponding decrease in the

free volume. There abound several numbers of research studies and publications on this

subject of aging.

Two mechanisms have been proposed to account for the collapse of free volume

during the physical aging process. Alfrey proposed that some of the free volumes in the

membrane structure exist temporarily and can be lost in the course of time via diffusion

toward the membrane surface [52]. This phenomenon is shown in Figure 2.2. It should be

highlighted that the rate of free volume diffusion is a function of membrane thickness; it

occurs in an accelerated pace for thinner membranes [48,50, 53-54]. Subsequently, Curro

and co-workers and Hirai and Eyring proposed that lattice contraction (Figure 2.2.) also

contributes to the overall reduction in the free volume [55]. Lattice contraction is not

thickness dependent. The rate of aging is strongly associated with the polymer type

[48,53-54,56-59]. For a polymer with poor chain packing efficiency and high FFV

values, the decline in free volume resulting from physical aging is more significant than

that of a polymer with lower FFV. Kim and co-workers investigated the aging

phenomenon in homo- and co-polyimides, and faster aging rate was observed for

polymers with higher free volume [58-59].

60
Time = 0
z

Glassy polymer matrix y


forming the membrane x

Free volume within the Membrane thickness, l


matrix – Langmuir sites

Time > 0
3-dimensional diffusion
of free volume from the
interior to the surface
For simplicity, only Lattice
diffusion in the z contraction
direction is shown in the radial
direction
Time >>> 0

Reduction in free volume


and densification of the Reduction in membrane
polymer matrix thickness, l - Δl

Figure 2.2. Diffusion of free volume and lattice contraction (physical aging).

One critical consequence caused by physical aging is that the gas separation

performance of these glassy polymeric membranes becomes time dependent. As the

aging time increases, the decrease in free volume reduces the gas permeability and

possibly a corresponding increase in gas pair permselectivity. This introduces uncertainty

in the gas separation performance in the long run which acts as a barrier to the

commercialization of polyimide membranes for industrial gas separation. Therefore,

overcoming the problem of physical aging in polyimide membranes represents a big

challenge for membrane researchers.

61
2.2 MOLECULAR DESIGN OF POLYMERIC MEMBRANES FOR GAS

SEPARATION

Polyimides exhibit excellent gas separation performance and good

physiochemical properties. Therefore, polyimide membranes have been extensively

explored for their potential applications in natural gas purification. Molecular tailoring of

polyimides is one approach that is utilized in the search for better membrane materials for

CO 2 /CH 4 separation. The molecular design of polyimides can be broadly classified into

two categories. The first encompasses purely polyimide system whereby the chemical

constituents and configurations of the dianhydrides and/or diamines used for polyimide

synthesis are varied in a systematic manner. The other category is the chemical fusion

(copolymerization) of polyimides with polymers from other classes.

2.2.1 Tailoring Pure Polyimide Systems

Numerous dianhydrides and diamines are available and their combinations give

rise to various polyimides which have the potential for use in natural gas separation.

Among them, aromatic polyimides containing hexafluoroisopropylidene-diphthalic

anhydride (6FDA) are of particular interest [60-63]. The CF 3 groups in 6FDA

considerably increase chain stiffness on the basis of steric bulk hindrances and effectively

reduce chain packing. Moreover, 6FDA-based polyimides exhibit good balance between

high CO 2 permeability and CO 2 /CH 4 selectivity [60-63]. In view of the advantages of

6FDA-based polyimides, this section aims to provide further insights on the relationship

between their polymer structure and the membrane separation performance. The general

62
structure of a polyimide is shown in Figure 2.3., where R 1 and R 2 refer to the dianhydride

and diamine moieties, respectively. Generally, polymer chain rigidity determines the

permselectivity while inter-chain spacing and chain mobility governs the permeability. In

the molecular design of polyimides, the main factors affecting the gas transport properties

are (1) spatial linkage configurations, (2) type of bridging groups, and (3) bulkiness and

polarity of pendant groups constituting the polymer.

O O

C C

N R1 N R2

C C

O O

Figure 2.3. General chemical structure of polyimide

For spatial linkage configurations, unsymmetric meta-linked polymers exhibit

better chain packing efficiency and restricted rotational freedom compared to their

symmetric para-linked isomers [64-68]. Therefore, the meta-isomers show lower FFV

and gas permeability but higher gas pair selectivity [64-66]. A comparison of 6FDA-

based polyimides containing different types of linkages is presented in Table 2.1. [64-66].

63
Table 2.1. Effect of meta or para linkages on gas separation performance (6FDA-based)
PCO 2 S CO 2 DCO 2
PCO 2 S CO 2 DCO 2
Name and chemical structure of diamine PCH 4 S CH 4 DCH 4 FFV Ref.
4,4’ – oxydianiline (4,4’ODA) 16.7 4.1 3.1 49 3.4 14 0.165 [64]

3,4’ – oxydianiline (3,4’ODA) 6.11 3.5 1.3 49 4.3 14 0.162 [64]


O

2,2-bis(4-aminophenyl)hexafluoropropane (6FpDA) 63.9 4.7 10.3 39.9 3.3 12.2 0.190 [65]
CF3
C
CF3

2,2-bis(3-aminophenyl)hexafluoropropane (6FmDA) 5.1 2.9 1.3 63.8 3.6 17.7 0.175 [65]
CF3
C
CF3

4,4’-diaminodiphenyl sulfone (pDDS) 16 5.6 2.1 47 3.8 12 0.149 [66]


O
S
O

3,3’-diaminodiphenyl sulfone (mDDS) 2.3 4.3 0.4 74 3.8 18 0.146 [66]


O
S
O

Units for P, S and D are barrer (1 barrer = 1 x 10-10 cm3 (STP)-cm/cm2 s cmHg), cm3 (STP)/cm3 atm and 10-8 cm2 /s respectively

64
An exception to the general trend between 6FDA-4,4’ODA and 6FDA-3,4’ODA

polyimides has been observed [64]. Despite the presence of meta linkages, there is no

change in CO 2 /CH 4 selectivity because the effect is diluted by the presence of other para

linkages. A comprehensive discussion on the dilution effect for polysulfones, similarly

applicable to polyimides, has been investigated by Aitken et al [69].

The nature of the bridging groups present in the polymer backbone influences gas

separation performance. Table 2.2. lists several 6FDA-based polyimides with different

bridging groups in the diamine moieties [65-66,70]. Bridging groups with low energy

barriers (e.g. O and S linkages) facilitate chain motions and results in higher gas

permeability while the incorporation of bulky linkage groups (e.g. CH 2 and O=S=O)

lowers gas permeability. With reference to Table 2.2., the order of increasing CO 2 /CH 4

selectivity of the polyimides (PPTI-1 < pDDS < 4,4’ODA) corresponds to the order of

increasing electronegativity of the bridging groups (S < O=S=O < O) [65-66,70]. Higher

electronegativity leads to stronger chain-to-chain interactions and enhanced selectivity.

The presence of bulky pendant groups increases inter-chain spacing and reduces

the packing efficiency of the polymer chains. Due to these bulky appendages (e.g. CH 3 ,

CF 3 and C 2 H 5 ), the FFV and gas permeability increases [64-66]. The effect of substituent

size on the gas separation performance of polyimides is shown in Table 2.3. [64-66]. The

polarity of the substituent groups affects chain-to-chain interactions which subsequently

influences chain rigidity and packing efficiency. With reference to Table 2.4., the greater

polarity of C=O groups in 6FDA-DAFO accounts for its higher permselectivity and

65
Table 2.2. Effect of connector groups on gas separation performance (6FDA-based)

PCO 2 S CO 2 DCO 2
PCO 2 S CO 2 DCO 2
Name and chemical structure of diamine PCH 4 S CH 4 DCH 4 FFV Ref.
4,4’ – oxydianiline (4,4’ODA) 23 4.89 3.58 60.5 3.70 16.3 0.164 [65]

methylenedianiline (MDA) 19.3 3.96 3.70 44.9 3.36 13.4 0.160 [65]
H
C
H

phenylene thioether (PPTI-1) 23.1 4.96 3.54 34.8 3.84 3.54 0.161 [70]

4,4’-diaminodiphenyl sulfone (pDDS) 16 5.55 0.41 47 3.8 18 0.149 [67]


O
S
O

Units for P, S and D are barrer (1 barrer = 1 x 10-10 cm3 (STP)-cm/cm2 s cmHg), cm3 (STP)/cm3 atm and 10-8 cm2 /s respectively

66
bulky substituents on gas separation performance (6FDA-based)
TableTable
3 Effect
2.3.ofEffect
bulkyofsubstituents on gas separation performance (6FDA-based)

CO 2
PPCO CO 2 CO 2
2 SSCO 2 DDCO 2
P
PCO 2 CO 2 DCO 2
D
Name CO 2 SS CO 2 CO 2 PPCH 4 SS CH 4 DDCH 4 FFV Ref.
Nameand
andchemical
chemicalstructure
structure of
ofdiamine
diamine CH 4 CH 4 CH 4 FFV Ref.
Methylenedianiline
Methylenedianiline(MDA)
(MDA) 19.3
19.3 3.96
3.96 3.70
3.70 44.9
44.9 3.36
3.36 13.4
13.4 0.160 [64]
0.160 [85]
HH
CC
HH

Isopropylidenedianiline
Isopropylidenedianiline(IPDA)
(IPDA) 30 4.24 5.38 42.9
42.9 3.53
3.53 12.1
12.1 0.168 [64]
0.168 [85]
CH
CH3 3
CC
CH
CH3 3

2,2-bis(4-aminophenyl)hexafluoropropane
2,2-bis(4-aminophenyl)hexafluoropropane (6FpDA)
(6FpDA) 63.9 4.7 10.3 39.9
39.9 3.3
3.3 12.2
12.2 0.190 [64]
0.190 [85]
CF
CF3 3
CC
CF
CF3 3

3,6-diaminocarbazole (CDA)
3,6-diaminocarbazole(CDA) 17.1 5.09 2.6 57
57 3.0
3.0 19
19 0.154
0.154 [87]
[66]
HH
NN

N-ethyl-3,6,-diaminocarbazole (ECDA)
N-ethyl-3,6,-diaminocarbazole (ECDA) 33.8 4.79 5.3 33
33 2.8
2.8 12
12 0.161
0.161 [87]
[66]
2H
CC2H 55
NN

Units
UnitsforforP,P,S Sand
andDDare
arebarrer
barrer(1(1barrer
barrer==11xx10
10-10-10cm
cm3 3(STP)-cm/cm
(STP)-cm/cm22sscmHg),
cmHg), cm
cm33 (STP)/cm
(STP)/cm33 atm
atm and 10-8
and 10 -8 cm2 /s respectively
cm2 /s respectively

67
Table 2.4. Effect of polar functional groups on gas separation performance (6FDA-based)

PCO 2 S CO 2 DCO 2
PCO 2 S CO 2 DCO 2
Name and chemical structure of diamine PCH 4 S CH 4 DCH 4 FFV Ref.
2,7-diaminofluorene (DAF) 11.3 4.64 1.8 54 3.2 1.8 0.155 [71]
H H
C

2,7-diaminofluorenon (DAFO) 7.68 4.48 1.3 60 3.2 1.3 0.154 [71]


O
C

Units for P, S and D are barrer (1 barrer = 1 x 10-10 cm3 (STP)-cm/cm2 s cmHg), cm3 (STP)/cm3 atm and 10-8 cm2 /s respectively

68
lower permeability over 6FDA-DAF [71]. Side appendages with stronger polarity such as

sulfonic groups were substituted on the polyimide main chains to alter the gas transport

properties [64,72-73]. Based on studies related to non-6FDA based sulfonated polyimide

membranes, it was shown that the permeability of and selectivity for CO 2 increase in the

presence of water vapor. This is attributed to the strong interactions between the polymer

chains and CO 2 upon water uptake which facilitates the transport of CO 2 .

In a nutshell, the polymer chain rigidity can be enhanced by integrating (1) meta

linkages, (2) polar side groups, and (3) bridging groups with high energy barrier. Since

the problems of plasticization and physical aging rooted from the flexibility and the non-

equilibrium state of the polymer chains, increasing polymer chain rigidity may reduce the

severity of these undesirable phenomenon. The inter-chain spacing can be increased by

incorporating bulky pendant or bridging groups and designing polyimides with para

linkages. The increase in inter-chain spacing should be coupled with backbone stiffness

in order to achieve both high permeability and selectivity. This strategy is only effective

for certain inter-chain spacings whereby penetrant diffusion is still governed by polymer

segmental motions.

Another feasible approach involves the use of polymers with intrinsic

microporosity (PIMs) which was first disclosed by McKeown’s group [74]. The PIMs

consist of random rod-like structures which generates intrinsic cavities within the

membrane. They exhibit behavior analogous to conventional molecular sieves, but with

69
greater solubility. Due to the intrinsic rotation of the linkages in polyimides structure,

PIMs for this particular class of polymers have not been discovered.

2.2.2 Design of Block Copolymer Systems

One approach to increase CO 2 solubility is to introduce functional groups that

increase CO 2 affinity with the resultant polymer. This is realizable via the synthesis of a

block co-polymer system comprising of desirable segments from different polymers. A

good example is a copolymer consisting of a hard block from a well-packed and rigid

glassy polymer and a soft block based on flexible rubbery polymer as shown in Figure

2.4. [75]. The glassy polymer segments provide the main structural framework and better

thermal resistance while the rubbery segments govern the transportation of gas

molecules.

The CO 2 removal performance of a series of poly(ethylene oxide)

(PEO)/polyimide have been investigated [76-79]. In the work by Okamoto and co-

workers, the PEO/polyimide block copolymers exhibited both high CO 2 permeabilities

and selectivities, possibly due to the improved CO 2 solubility in the PEO segments [76-

77]. Smaihi et al. investigated a series of poly(imide siloxane) copolymers and concluded

that the nature of the coupling agent and the proportion of siloxane significantly affects

the gas permeabilities [79]. Similarly, Hibshman et al. studied a series of poly(imide

siloxane) copolymers and in most cases an enhancement in CO 2 permeability was

achieved but at the expense of CO 2 /CH 4 selectivity [80]. The gas transport properties of a

co-poly(imide pyrrolone), 6FDA-TAB/DAM was studied and the extremely rigid

70
polypyrrolone structure executes molecular sieving function [81]. Most of these co-

poly(imide pyrrolone) membranes exhibited CO 2 /CH 4 separation performances that fall

near to or above the upper bound.

At one instant: (e.g. time = 0)

Hard segments (glassy Soft segments (rubbery


polymer) polymer)

At the next instant: (e.g. time >0)

Configurations of the hard segments Soft segments are flexible


remain the same. Provides the and hence configurations
structural framework change with time

Figure 2.4. Synthetic strategy for copolymer structure for CO 2 separation

2.3 MODIFICATION OF POLYMERIC MEMBRANES FOR GAS

SEPARATION

The synthesis of new materials generally involves risk, high costs and

considerable time. Hence, modification of the existing membrane materials is another


71
viable approach, capable of creating new prospects aimed at overcoming the present

limitations for better separation performance.

2.3.1 Polymer Blending

Polymer blending can be considered as a preferred route of modification in the

light of its simplicity, reproducibility, processability and low development cost.

Moreover, the blending of polymers may synergistically combine the advantages of

different materials. Polymer blends are generally classified as miscible and immiscible.

Miscible polymer blends are the ones desirable for membrane fabrication since

homogeneity is essential to produce membranes with uniform performance while

retaining physical properties.

As mentioned previously, some membranes made of polyimide materials are

seriously susceptible to highly soluble penetrants such as CO 2 . Taking into account the

complimentary features of polyimides and other polymers that possess anti-plasticization

characteristics, it is reasonable to examine the compatibility of these polymers to form

blend membranes with improved properties. Researchers have found that polyimides can

be blended on a molecular level with a number of polymers such as polycarbonate [82],

polysulfone [83], polyethersulfone [84-86], polyaniline [87-88] and polybenzimidazole

[89]. In CO 2 permeation experiments, Kapantaidakis et al. [83] showed that blending

Matrimid with polysulfone (a polymer with significantly higher plasticization pressure)

resulted in anti-plasticization effects. Bos et al. [90] extended this approach to mixed gas

experiments and demonstrated similar improved plasticization behavior, but at the

72
expense of gas permeation and separation performance. In their work, they revealed

another polymer blend system comprising of Matrimid and P84 (a copolyimide) which

can stabilize the membrane against plasticization while simultaneously improving the gas

separation performance [90].

One strategy to make miscible polymer blends is to select polymers which have

strong interactions. Su et al. successfully prepared polyimide/polyaniline blends using the

polyamic acid as a “polymeric” dopant [87]. Gas permeability studies showed that 50/50

polyaniline/polyimide blends provided higher CO 2 permeability compared to the

respective homopolymers. This blend membrane exhibited a corresponding CO 2 /CH 4

separation factor of 85. The separation factors for the polyimide and polyaniline are 59

and 180, respectively [87]. In a recent study, Hosseini et al. carefully selected aromatic

polybenzimidazole (PBI) as the blending material to improve the properties of Matrimid

[89]. It was found that incorporating various amounts of PBI into the Matrimid matrix

can boost the CO 2 /CH 4 separation performance but at the expense of permeability. The

CO 2 /CH 4 selectivity was doubled for the blend comprising 75% of PBI content. PBI is a

high T g glassy polymer that possesses both donor and acceptor hydrogen-bonding sites

[91-93], which are capable of participating in specific interactions [94-95]. The good

processability of Matrimid can compensate for the poor processability of PBI by

controlling a low PBI content in blends. In addition, Matrimid contains electron donating

carbonyl groups which can interact with certain electron accepting groups in PBI and

form strong interactions, as highlighted by Musto et al. [93].

73
Another approach to ensure the miscibility of two polymers involves using a

compatibilzer agent within the membrane matrix. Ekiner and Simmons have proposed

that the addition of the compatibilizer, poly(styrene-co-maleric anhydride) (PSMA)

prevents phase separation within the membranes made of poly(phenylene oxide) and

Nylon 6. The resultant membrane showed good separation performance above the

Robeson’s upper bound for O 2 /N 2 [96]. This approach may be used to fabricate

polyimide blend membranes for CO 2 /CH 4 separation.

Although the promising features of polymer blending were demonstrated, it seems

that the shortage of right pair candidates from the currently available materials is the

limiting factor for further materialization of this technique for gas separation membranes.

In our opinion, the adoption of an approach for the systematic design of suitable pairs is

the most efficient way in revitalizing the true benefits of polymer blending. Based on this,

we foresee good prospects in further utilization of polymer blending of polyimide for

fabrication of high performance membranes.

2.3.2 Hybrid Membranes

Hybrid membrane consists of an organic polymer as the continuous phase and

nano-particles as the dispersed phase. The function of the fillers is to systematically

manipulate the molecular packing of polymer chains, thereby enhancing the separation

properties of polyimide membranes. Typically, the hybrid effects resulting from the

combination of organic and inorganic phases are more prominent when employing fillers

that have smaller dimensions and hence larger surface area. The preparation of nano-

74
composite polyimide membranes can be broadly classified into two categories based on

the state of the fillers before membrane formation. The first type involves the direct

addition of the nano-fillers to a polymer solution. For this direct method, the mixing of

the two phases is accomplished by means of mechanical stirring or ultra-sonication to

reduce particle agglomeration. Thereafter, the organic-inorganic mixture is used for

membrane fabrication. The second approach, also known as the in-situ method entails the

use of nano-particle precursors. For in-situ preparation of nano-composite membranes,

the precursors of the fillers are added to the polymer solution and the nano-particles are

formed during the process of membrane formation. In this approach, the nano-particles

precursors generally have good interactions with the polymer, thereby enabling good

dispersion of the precursor and the resultant fillers [97].

As shown in Figure 2.5., four mechanisms can be proposed to describe how nano-

size fillers may tailor the gas transport mechanism through membranes. Based on the

nature and functional role, the fillers can be classified as: (1) non-porous inert, (2) non-

porous activated, (3) high affinity with polymers, and (4) porous nanoparticles.

Some examples of non-porous inert nanoparticles are fume silica [98], C60 [99]

and polymeric particles [100]. Due to the incompatibility or weak interaction between

these inert particles and the polymer chains, the presence of the fillers disrupts chain

packing and creates interfacial voids between the two phases. The enhanced free volume

in the resultant membrane makes the solubility selectivity as the enhancing factor in the

overall gas separation performance. This is especially effective for the separation of CO 2

75
and CH 4 where the characteristic properties of CO 2 (e.g. condensability and the

quadrupolar nature) can also play a part through enhancing the surface flow [101].

Case 1 Case 2 Case 3 Case 4

Surface diffusion

High
Activated Porous
Inner Filler Affinity
Filler Filler
Filler

Void Rigid

Figure 2.5. Schematic of the various gas transport routes through hybrid

polymeric membranes

Activated carbon particles [102] and ion exchanged zeolites [103] belong to the

group of non-porous activated nanoparticles. Generally, these particles show preferential

sorption for CO 2 by creating a surface flux along the particles. Due to the enhanced CO 2

sorption, the CO 2 /CH 4 permselectivity increases.

Another group of fillers is comprised of non-porous activated nanoparticles such

as metal [104], metal oxides [105] and surface modified zeolites [106]. The high affinity

of these particles with the polymers hinders the mobility of the polymer chains which are

in close contact with the particles. In this case, the presence of the fillers within the

76
polymeric matrix serves a similar purpose as chemical cross-linking since both

approaches reduce the mobility of the polymeric chains. This so called “rigidification

effect” is found to improve the selectivity of membrane.

In the last category, porous nano-size particles such as zoelites [107-108] and

carbon molecular sieves [109] generally integrates the high diffusivity selectivity of

inorganic molecular sieves and good processability of polymeric material. Depending on

the nature of polymer and fillers, the gas separation performance of the resultant hybrid

membranes can be tailored. Detailed discussion can be found in another review related to

mixed matrix membranes [17].

2.3.3 Thermal Annealing

The formation of the glassy polymeric structure from an equilibrium liquid or

rubbery state is a time-dependent relaxation process. It is also well known that the

structure of glassy polymers depends upon their thermal history. The local structure of a

glassy material is expected to change as the conditions of glass formation vary. However,

microvoids or frozen free volume at the glassy state might be retained by glass formation

from the liquid state. In particular, more microvoids or larger specific volume might be

formed if the glass formation is more rapid. Therefore, quenching can be adopted to

enhance microvoid formation [38, 110].

On the other side, the glassy state of amorphous polymer tends to relax towards

an equilibrium state. The relaxation time from the glassy state to the equilibrium super

77
cooled liquid state depends markedly on the temperature of the material. The relaxation

time is very long when the glassy polymers are kept at a temperature far below their T g .

However, the relaxation of the glassy state to the equilibrium state is relatively fast near

the glass transition temperature. Sub-T g annealing, where the sample is annealed near,

but below T g , can provide a state much closer to equilibrium. This translates to a

decrease in the frozen free volume or microvoids of glassy polymers which in turn can

potentially reduce the physical aging effects seem in polyimide membranes for natural

gas separation. Using dense aromatic polyimide 6FDA-m-DDS films, Kawakami et al.

[111] observed that heat-treatment at temperatures well below the T g densified the films.

Both density and glass transition temperature were found to increase with higher thermal

treatment temperatures. In addition, lower gas permeabilities were observed when

compared to untreated film. Similar phenomena have been reported by Krol et al. [112],

Chung et al. [113], Zhou and Koros [114], and Qiao et al. [115]. Polyimides contain

alternating electron donating and electron accepting groups which are able to form the

charge transfer complex. The electron donating molecules correspond to the six-

membered ring (benzene ring) and the electron accepting molecules correspond to the

five-membered ring (imide ring). When the rings approach to the proximity of each other,

π-electron interactions occur, forming a complex which restricts the polymer chain

mobility. The formation of such complexes by thermal curing was proven by

fluorescence and UV-Vis spectroscopy [111-112, 114-115]. As a result of thermal curing,

higher chain packing density, smaller free volumes and slightly higher T g values were

obtained. The suppression of CO 2 -induced plasticization was observed as another

consequence of thermal curing.

78
2.3.4 Grafting and Halogenation of Polymer Backbone

Modification of the pure polyimide membrane by means of chemical grafting

which involves the reaction of suitable reagents with the carbonyl groups via covalent

bonding can be achieved [116]. It is expected that such chemical modification can

substantially alter the gas transport properties of the membranes. For example, Liu et al.

reported the amidation of polyimides by immersing the dense films in a 10% (w/v) N,N-

dimethylaminoethyleneamine (DMEA) dissolved in hexane [117]. This chemical

modification involves the opening of some imide rings to form ortho-diamide, as shown

in Figure 2.6.

NH
R

NH F3C CF3 O
O C C C
NH NH R1
C C O
F3C CF3 O O
O NH
C C C H2N-R-NH2
R1 N R
H2 N N R2
C NH
C
O O F3C CF3
O
6FDA moiety C C C O
N
R: N,N-dimethylaminoethyleneamine [137] C C O
O
HN R NH2
p-xylenediamine [157-161]
ethylenediamine [162]
1,3-cyclohexanebis(methylamine) [163]
PAMAM dendrimers [164-166]

Figure 2.6. Amine modification of polyimides

It was found that amidation significantly reduces the gas diffusion coefficients,

probably due to stronger intersegmental hydrogen bonding between the newly formed
79
amide groups and space filling effects of DMEA. However, the effect of amidation on

CO 2 /CH 4 selectivity depends on the chemical structure of the polyimide; for polyimide

with high side group density (e.g. 6FDA-Durene), the space filling by DMEA reduces the

gas solubility which in turn reduces CO 2 /CH 4 gas selectivity. For polyimide with low

side group density (e.g. 6FDA-durene/mPDA), diffusion dominates the overall gas

transport; the denser packing of intermolecular chains increases the diffusion selectivity,

leading to better CO 2 /CH 4 discrimination. Beside amines, other reagents like hydrogen

cyanide and phosphorus ylides can similarly be grafted to polyimides for better CO 2 /CH 4

separation performance [116].

Alternatively, the benzene rings of aromatic polyimides can be used as functional

sites for side group grafting via electrophilic aromatic substitution reaction. However,

due to the strongly deactivating effects of neighboring carbonyl groups, only the benzene

rings of the diamine unit could be grafted. Bromine atoms were first introduced into

polyimides by Okamoto et al. [128] and Liou et al. [129] using the free-radical

bromination of alkyl groups and the polymerization of halogen-containing monomers,

respectively. Subsequently, Guiver et al. [130] introduced the direct bromination of

Matrimid by electrophilic bromination. The special characteristics of brominated

polyimides are the improved solubility by reduction in chain packing and increased

reactivity.

The fluorination of polyimide films with fluorine, fluorine-helium, fluorine-

nitrogen and fluorine-oxygen gas mixtures has also been reported [131]. However, the

80
degree of fluorination is limited to the immediate contactable surface. A fluorine-treated

film is comprised of fluorinated and untreated layers which are separated by a very thin

layer in which the main chemical transformations proceeds. The major changes occurring

during fluorination of polyimides are: 1) replacement of hydrogen atoms by fluorine, 2)

saturation of double bonds by fluorine, 3) disruption of at least one C-N bond in imide

ring and 4) significant increase in the polymer density. Although the He/CH 4 separation

performance of the fluorinated Matrimid surpasses the trade-off line but the potential

application of this technique for CO 2 /CH 4 separation has remained unexplored.

2.3.5 Cross-linking Modification

The extent of cross-linking and distribution of the resultant network can

profoundly affect the physical properties of polymeric membranes. Generally, it is

believed that cross-linking can be used to 1) increase the membrane stability in the

presence of aggressive feed components, 2) suppress CO 2 -induced plasticization, and 3)

achieve better gas separation properties. Cross-linking modifications of polyimides can

be induced by several methods, such as ultraviolet (UV) and ion beam radiations, thermal

treatment and chemical modification.

Kita at el applied UV radiation to induce cross-linking reactions in

benzophenone-containing polyimides [132]. Depending on the duration of the irradiation,

a considerable improvement in the gas pair selectivity could be achieved but with greatly

reduced gas permeability, presumably due to densification of membrane structure and

reduced mobility of polymer chains. Won et al. and Hu et al. investigated the effects of

81
surface modification of polyimide membranes by ion beam [133-134]. They concluded

that ion beam irradiation could provide a useful mean for improving selectivity for gas

separation membranes. Despite the potential use of ion irradiation in enhancing gas

separation pair permselectivity, the high cost and difficulty in implementing uniform

irradiation on large sheets of flat membranes and hollow fibers limit the industrial

application of this technique.

Cross-linking through heat treatment has been studied extensively on polyimide

membranes. It was shown that thermal annealing at elevated temperatures for short

periods of time can significantly induced cross-linking in 6FDA polyimide [113] and

Matrimid films [135]. Polyimides containing acetylene groups have also been

investigated and cross-linking occurs via a Diels-Alder type reaction under heat treatment

[136-137]. The cross-linking modification improved gas selectivity with an acceptable

decrease in gas permeability and significantly enhanced CO 2 plasticization resistance.

Another approach for improving the resistance of polyimide membranes involves

the formation of semi-interpenetrating polymer networks, obtained by heating a mixture

of polyimides with monomers or oligomers containing reactive groups with the ability to

crosslink (e.g., acetylene end groups). Rezac et al. [138] blended 6FDA-IPDA with a

diacetylene containing oligomer (6FDA-DIA) and cross-linking was achieved by heating

the membranes to 340oC. The crosslinked polyimides rendered better

permeability/permselectivity results. Semi-interpenetrating networks derived from blends

82
of Matrimid and the oligomer containing acetylene groups, Thermid FA-700 also showed

good plasticization resistance [139].

Koros, Paul and their co-workers have synthesized carboxylic containing

polyimide membranes which can be cross-linked by either a transesterification reaction

or reaction with a metal ion [140-144]. The effect of cross-linking the polyimide depends

on the chemical structure of the diol cross-linking agents and the specific annealing

temperatures. Unlike other cross-linking modifications which typically decrease the

permeability of the resultant polymeric membranes, cross-linking of carboxylic

containing polyimides with ethylene glycol can lead to an increase in the permeability

without sacrificing selectivity. It should be noted that transesterification occurs only at

elevated temperatures (above 200oC). The increase in the free volume may be due to the

coupling of thermal annealing and the removal of pendant diol groups during the

transesterification reaction. The aging phenomenon of these cross-linked membranes

(thermally treated at 140 oC) has been investigated by Kim et al. [144] and it was found

that the polyimide membranes suffered more from physical aging after cross-linking.

This is perhaps attributed to the incomplete transesterification reaction at a low

temperature of 140oC. The cross-linking reaction may proceeds continuously during the

gas permeation testing which accounts for the decline in gas permeability with time. This

problem can be rectified by performing the cross-linking reaction at elevated

temperatures. Nevertheless, the application of high temperature treatments to asymmetric

membranes alters the membrane morphology and causes densification of the membrane

structure. It should be noted that crosslinking of the bulk polyimide matrix reduces the

83
polymer processability during membrane fabrication and increases the brittleness of the

resultant membranes. Therefore, the preferred cross-linking method should have the

following characteristics: (1) can be performed at room temperature and (2)

functionalizes only the thin selective layer of asymmetric membrane which reduces

substructure resistance and brittleness of the support layer.

Another chemical modification method has been proposed to overcome the

difficulties of post-treatment processes. Hayes devised a new way of cross-linking by

immersing polyimide in a diamino compound solution [118,145]. This approach by

Hayes was first employed for increasing O 2 /N 2 selectivity which was further explored by

Chung’s group. Chung and his co-workers have contributed a number of works related to

the use of diamines and multi-amines to cross-link polyimides [119-127,146-147]. The

results obtained from these works are summarized in Tables 2.5. and 2.6. A typical

mechanism for such cross-linking has been shown in Figure 2.6. Liu et al., Cao et al. and

Tin et al. used p-xylenediamine as a cross-linking agent to modify different polyimide

hollow fiber membranes at ambient temperature [119-122]. It is proven that the

chemically modified membrane does not plasticize easily since the cross-linked structure

prevents the material from swelling in the presence of plasticizing agents. Generally,

increasing the immersion time increases the degree of crosslinking which significantly

reduces polymer swelling and chain mobility. Hence, enhancement in the CO 2 /CH 4

permselectivity may be achieved. Compared to the ethylene glycol approach, the diamine

and multi-amine approaches are more advantageous especially for the modification of

hollow fibers because of easy operation and localized chemical modification at the outer

84
Table 2.5. CO2/CH4 separation performance of crosslinked dense films

Polyimide Temperature Pressure Crosslinking agent and immersion time PCO2 αCO2/CH4 Reference
(oC) (atm) (Barrer)
Matrimid® 35 10 No modification 6.5 34 [123]
5218
10 (w/v)% p-xylenediamine in methanol, 24 h 7.4 36
6FDA-durene 35 10 No modification 600 15 [124]*
10 (w/v)% ethylenediamine in methanol, 2 min 100 28
10 (w/v)% 1,3-cyclohexanebis(methylamine) 55 19 [125]
in methanol, 30 min
10 (w/v)% G1 diaminobutane dendrimer in 430 22 [127]*
methanol, 5 min
5 (w/v)% G0 polyamidoamine dendrimer in 20 47 [126,146]
methanol, 10000 min
5 (w/v)% G1 polyamidoamine dendrimer in 300 30 [146]*
methanol, 10000 min
5 (w/v)% G2 polyamidoamine dendrimer in 370 29
methanol, 10000 min
* Values are approximated from plots
P: permeability
Barrer = 1×10−10 cm3 (STP)-cm/cm2 s cmHg = 7.5005 x 10-18 m2s-1Pa-1

85
Table 2.6. CO2/CH4 separation performance of crosslinked hollow fibers

Polyimide Temperature(o Pressure Crosslinking agent and (P/l)CO2 αCO2/CH4 Reference


C) (psi) immersion time (GPU)

6FDA-durene/mPDA 23 CO2: 100 psi No modification 70.4 44 [121]


(50:50) (selective layer) CH4: 200
5% (w/v) p-xylenediamine in 28.3 101
– PES (inner layer) psi
methanol, 5 min
6FDA-2,6-DAT 25 Mixed gas No modification 55 60 [120]
– single layer CO2 (40%)
1% (w/v) p-xylenediamine 19 63
and CH4
in methanol, 3 min
(60%), total
pressure: 1% (w/v) m-xylenediamine 18 59
200 psi in methanol, 3 min
6FDA-ODA/NDA 25 CO2: 50 psi 5% (w/w) p-xylenediamine in 30 58 [122]
(50:50) – single layer CH4: 200 methanol, 24 h
psi
(P/l): Permeance
1 GPU = 1x10−6 cm3 (STP)/cm2 s cmHg

86
selective skin without damaging the entire membrane substructure.

Chemical modifications of 6FDA-durene dense membranes were performed using

a linear aliphatic diamine, ethylenediamine (EDA) [124] and a bulky diamine, 1,3-

cyclohexanebis(methylamine) by Shao et al [125]. Low temperature thermal annealing

(100~150oC) was utilized to facilitate the amidation reaction. The coupling effects of

diamine-induced cross-linking and thermal annealing accelerates the formation of charge

transfer complexes (CTCs) which densifies the polyimide membrane structure. As a

result, the gas transport properties of the modified polyimides generally show reduced

permeability and enhanced selectivity upon cross-linking and thermal annealing. CO 2

plasticization tests on the resultant membranes indicate that the coupling effects of cross-

linking and thermal annealing can also improve the plasticization resistance of 6FDA-

durene membranes from around 300 psia to more than 720 psia, which may be attributed

to the effective formation of CTCs.

Chung, Xiao and their co-workers modified 6FDA-polyimide membranes using

G0, G1 and G2 PAMAM dendrimers which have multi-functional groups [126-127,146-

147]. Compared with the simulated results, the morphology and conformation of the

grafted PAMAM dendrimers on the polymer surfaces are disk-shaped molecular clusters

which were further proven by AFM. In fact, immersion time and different dendrimer

generation are two important factors in determining the characteristics of modified

polyimide films. A longer immersion time yields higher dendrimer loading and

equilibrium may be reached after modifying for at least 24 hours. Due to their big

molecular size, the cross-linking modification was believed to occur mainly on the
87
membrane surface. The modified polyimide films also exhibited better gas separation

performance, especially for CO 2 /CH 4 because of the affinity between PAMAM

dendrimer and CO 2 . In another study, the effects of temperature on the multi-amine

modification have been investigated [147]. Thermal treatment at temperatures below 150
o
C facilitated the amidization reaction between the polyimide and dendrimers and

enhanced the degree of cross-linking by the dendrimers. On the other hand, high

temperature treatment (>250oC) destroyed the cross-linked structures of the modified

polyimide via the decomposition of the dendrimer molecules and the degradation of the

polyimide backbone.

Although the amino-modification of polyimide membranes enhances CO 2 /CH 4

permselectivity, the long term operational stability of the membranes remains unclear.

One drawback of this approach is the reversible reaction between the amino-compounds

and the imide groups at high temperatures which has been verified by Powell et al. [148-

149]. Nevertheless, the modified membranes can be employed for CO 2 /CH 4 separation at

relatively mild operating temperatures. In addition, since the amidation reaction is carried

out at room temperature, the reaction may be incomplete which result in time-dependent

gas separation performance of the resultant membranes. Moderate temperature (~150oC)

treatment has been approved to facilitate the amidation reaction [125,147]. More studies

on the long-term stability of the modified membranes are required.

2.4 CURRENT AND FUTURE DEVELOPMENTS

88
An ideally structured polyimide suitable for fabricating high performance

CO 2 /CH 4 separation membranes should possess the characteristics of backbone stiffness

coupled with large inter-chain spacing. The basis of this claim is that diffusivity

selectivity plays a dominant role in the separation of CO 2 /CH 4 . It is believed that these

characteristics will result in polyimide membranes with remarkable CO 2 permeability

and CO 2 /CH 4 permselectivity. Current studies on the molecular design of polyimides

tend to accomplish this via employing 6FDA because the CF3 groups considerably

increase the stiffness of the chain on the basis of steric bulk hindrances and effectively

reduce the chain packing. However, there is no significant improvement in gas separation

performance beyond the upper bound. Due to the intrinsic rotation of the linkages in

polyimide structure, it seems difficult to increase the inter-chain spacing without losing

backbone stiffness. In view of the current state of novel polyimide development for

CO 2 /CH 4 separation, researchers should perhaps divert their attention towards enhancing

the solubility and solubility selectivity in the design of new polyimides. Possible

strategies to achieve this include the incorporation of functional groups into capable of

increasing the affinity of CO 2 to the polymer and creating more sorption sites for CO 2 .

In consideration of the higher stiffness of inorganic materials, many researchers

attempt to transform polyimide to carbon state by high temperature pyrolysis [150-153].

In this way, the gas transport switches from solution-diffusion to primarily molecular

sieving. As a result, the gas separation performance surpasses the upper bound. Other

than the high cost of carbon membrane, another critical problem is its brittleness which

89
makes its processing difficult. Detailed information about carbon membrane derived from

polyimides could be found in another review [154].

Recently, Islam et al. proposed a low-temperature pyrolysis method for

fabricating flexible organic-inorganic intermediate membranes by the cleavage of SO 3 H

groups in the sulfonated polyimides [155]. The resultant membranes possess attractive

gas separation performance with high degree of flexibility. As an extension of this idea,

Park et al. suggested the systematic tailoring of an unusual microstructure via thermally-

driven “molecular rearrangement” process [67]. In this process, the single bond linkages

of the polyimide are changed to stiff and rigid chain elements (e.g. benzoxazole-

phenylene ring or benzithiazole-phenylene ring). Additionally, the indiscriminant

torsional rotation increases the efficiency of free volume formation. Consequently, the

gas separation performance of such membranes can surpass the upper bound.

Nonetheless, it is important to keep in mind that this approach similarly requires thermal

treatment at elevated temperatures which may limit its application to asymmetric

membranes.

In addition to good gas separation performance, the durability and the lifetime of

the polyimide membranes are also important. Aging and plasticization phenomena create

undesirable time dependency of the gas separation performance for polyimide

membranes. In order to improve the stability of membrane, thermal annealing, polymer

blending and cross-linking can be employed. However, for successful commercial

application of membrane-based natural gas treatment using polyimides, hollow fiber

90
processing and the overall costs must be adequately tuned. Therefore, the discussed

methods are still a distance away from the requirements. The focus of polyimide

membrane research in the near future may remain on new polyimide molecular

architectures and seeking better cross-linking agents which can work under low

temperature and only at the surface layer to obtain desirable membranes for CO 2 /CH 4

separation. Significant research efforts must be tilted to asymmetric flat and hollow fiber

configurations rather than dense flat membranes in order to have real and immediate

impact on scale up and commercialization.

2.5 REFERENCES

[1] Energy information administration,USDepartment of Energy. International energy

outlook 2006.

[2] Bhide BD, Voskericyan A, Stern SA. Hybrid processes for the removal of acid gases

from natural gas. J Membr Sci 1998;140:27–49.

[3] Baker RW. Future directions of membrane gas separation technology. Ind Eng Chem

Res 2002;41:1393–411.

[4] Kohl AL, Nielson R. Gas purification. Houston, TX: Gulf Publishing; 1997.

[5] Koros WJ, Fleming GK. Membrane-based gas separation. J Membr Sci 1993;83:1–

80.

[6] Mc. Kee RL, Changela MK, Reading GJ. Carbon dioxide removal: membrane plus

amine. Hydrocarb Process 1991;70:63–5.

91
[7] Merkel TC, Toy LG. Comparison of hydrogen sulfide transport properties in

fluorinated and nonfluorinated polymers. Macromolecules 2006;39:7591–600.

[8] Paul DR, Yampolskii YP. Polymeric gas separation membranes. Boca Raton: CRC;

1994.

[9] Henis JM, Tripodi MK. The developing technology of gas separating membranes.

Science 1983;220:11–7.

[10] HoWSW, Sirkar KK. Membrane handbook. NewYork: Van Nostrand Reinhold;

1992.

[11] Koros WJ, Mahajan R. Pushing the limits on possibilities for large scale gas

separation: which strategies? J Membr Sci 2000;175:181–96.

[12] Koros WJ. Gas separation membranes: needs for combined materials science and

processing approaches. Macromol Symp 2002;188:13–22.

[13] Maier G. Gas separation with polymer membranes. Angew Chem Int Ed

1998;37:2960–74.

[14] Stern SA. Polymers for gas separations: the next decade. J Membr Sci 1994;94:1–65.

[15] FreemanBD, Pinnau I. Polymericmembranes for gas and vapor separations:

chemistry and materials science. In: ACS symposium series, vol. 733. 1999.

[16] Caro J, Noack M, Kolsch P, Schafer R. Zeolite membranes-state of their

development and perspective. Micropor Mesopor Mater 2000;38:3–24.

[17] Chung TS, Jiang L, Li Y, Kulprathipanja S. Mixed matrix membranes (MMMs)

comprising organic polymers with dispersed inorganic fillers for gas separation.

Prog Polym Sci 2007;32: 483–507.

92
[18] Saufi SM, Ismail AF. Fabrication of carbon membranes for gas separation—a

review. Carbon 2004;42:241–59.

[19] Uemiya S. State-of-the-art of supported metal membranes for gas separation. Sep

Purif Methods 1999;28:51–85.

[20] SciFinder Scholar 2007. CAS a division of the American Chemical Society.

[21] Ohya H, Kudryavtsev VV, Semenova SI. Polyimide membranes: applications,

fabrications and properties. Tokyo: Gordon and Breach Publishers; 1996.

[22] Kim TH. Gas sorption and permeation in a series of aromatic polyimides. Ph.D.

Dissertation, Univ Texas: Austin, 1988.

[23] Kim TH, Koros WJ, Husk GR, O’Brien KC. Relationship between gas separation

properties and chemical structure in a series of aromatic polyimides. J Membr Sci

1988;37:45–62.

[24] Hirayama Y, Yoshinaga T, Kusuki Y, Ninomiya K, Sakakibara T, Tamari T.

Relation of gas permeability with structure of aromatic polyimides I. J Membr Sci

1996;111:169–82.

[25] Akira S, Tsukasa M, Masatoshi M, Kenichi I. Relationships between the chemical

structures and the solubility, diffusivity, and permselectivity of propylene and

propane in 6FDA-based polyimides. J Polym Sci Part B: Polym Phys

2000;38:2525–36.

[26] Ayala D, Lozano AE, Abajo JD, Perez CG, Campa JG, Peinemann KV, et al. Gas

separation properties of aromatic polyimides. JMembr Sci 2003;215:61–73.

93
[27] Wang YC, Huang SH, Hu CC, Li CL, Lee KR, Liaw DJ, et al. Sorption and

transport properties of gases in aromatic polyimide membrane. J Membr Sci

2005;248:15–25.

[28] Robeson LM. Correlation of separation factor versus permeability for polymeric

membrane. J Membr Sci 1991;62:165–85.

[29] Freeman BD. Basis of permeability/selectivity tradeoff relations in polymeric gas

separation membranes. Macromolecules 1999;32:375–80.

[30] Van Amerongen AJ. The permeability of different rubbers to gases and its relation to

diffusivity and solubility. J Appl Phys 1946;17:972–85.

[31] Barrer RM. Permeability in relation to viscosity and structure of rubber. Trans

Faraday Soc 1942;38:322–30.

[32] Robeson LM. The upper bound revisited. J Membr Sci 2008;320:390–400.

[33] Sanders ES. Penetrant-induced plasticization and gas permeation in glassy polymers.

J Membr Sci 1988;37:63–80.

[34] Chiou JS, Barlow JW, Paul DR. Plasticization of glassy polymers by CO2. J Appl

Polym Sci 1985;30:2633–42.

[35] Wissinger RG, Paulaitis ME. Glass transitions in polymer/CO2 mixtures at elevated

pressures. J Polym Sci Polym Phys Ed 1991;29:631–3.

[36] Handa YP, Lampron S, O’Neill ML. On the plasticization of poly(2,6-dimethyl

phenylene oxide) by CO2. J Polym Sci Polym Phys Ed 1994;32:2549–53.

[37] Wang WCV, Kramer EJ, Sachse WG. Effects of high pressure CO2 on the glass

transition temperature and mechanical properties of polystyrene. J Polym Sci Polym

Phys Ed 1982;20:1371–84.

94
[38] Fried JR, Liu HC, Zhang C. Effect of sorbed carbon dioxide on the dynamic

mechanical properties of glassy polymers. J Polym Sci: Polym Lett 1989;27:385–

92.

[39] Chow TS. Molecular interpretation of the glass transition temperature of polymer

diluent systems. Macromolecules 1980;13:362–4.

[40] Wessling M, Borneman Z, Boomgaard TVD, Smolders CA. Carbon dioxide foaming

of glassy polymers. J Appl Polym Sci 1994;53:1497–512.

[41] Chung TS, Cao C, Wang R. Pressure and temperature dependence of the gas-

transport properties of dense poly[2,6-toluene-2,2-bis(3, 4dicarboxylphenyl)

hexafluoropropane diimide] membranes. J Polym Sci Part B: Polym Phys

2004;42:354–64.

[42] Coleman MR, Koros WJ. Conditioning of fluorine-containing polyimides. 2. Effect

of conditioning protocol at 8% volume dilation on gas-transport properties.

Macromolecules 1999;32:3106–13.

[43] Kanehashi S, Nakagawa T, Nagai K, Duthie X, Kentish S, Stevens G. Effects of

carbon dioxide-induced plasticization on the gas transport properties of glassy

polyimide membranes. J Membr Sci 2007;298:147–55.

[44] Wessling M, Huisman I, Boomgaard TVD, Smolders CA. Timedependent

permeation of carbon dioxide through a polyimide membrane above the

plasticization pressure. J Appl Polym Sci 1995;58:1959–66.

[45] Chiou JS, Paul DR. Effects of carbon dioxide exposure on gas transport properties of

glassy polymers. J Membr Sci 1987;32: 195–205.

95
[46] Zhou C, Chung TS, Wang R, Liu Y, Goh SH. The accelerated CO2 plasticization of

ultra-thin polyimide films and the effect of surface chemical cross-linking on

plasticization and physical aging. J Membr Sci 2003;225:125–34.

[47] Duthie X, Kentish S, Powell C, Nagai K, Qiao GG, Stevens G. Operating

temperature effects on the plasticization of polyimide gas separation membranes. J

Membr Sci 2007;294:40–9.

[48] Wonders AG, Paul DR. Effects of CO2 exposure history on sorption and transport in

polycarbonate. J Membr Sci 1979;5:63–75.

[49] Prausnitz JM, Lichtenthaler RN, Azevedo ED. Molecular thermodynamics of fluid-

phase equilibrium. New York: Prentice-Hall; 1986.

[50] Koros WJ, Hellums MW. Transport properties. In: Kroschwitz JI, editor.

Encyclopedia of polymer science. 2nd ed. New York: Wiley-Interscience; 1989.

[51] Park JY, Paul DR. Correlation and prediction of gas permeability in glassy polymer

membrane materials via a modified free volume based group contribution method. J

Membr Sci 1997;125: 23–39.

[52] Alfrey T, Goldfinger G,Mark H. The apparent second-order transition point of

polystyrene. J Appl Phys 1943;14:700–5.

[53] Pfromm PH, Koros WJ. Accelerated physical ageing of thin glassy polymer films:

evidence from gas transport measurements. Polymer 1995;36:2379–87.

[54] Zhou C, Chung TS,Wang R, Goh SH. A Governing equation for physical aging of

thick and thin fluoropolyimide films. J Appl Polym Sci 2004;92:1758–64.

[55] Curro JG, Lagasse RR, Simba R. Diffusion model for volume recovery in glasses.

Macromolecules 1982;15:1621–6.

96
[56] Lin WH, Chung TS. The physical aging phenomenon of 6FDA-durene polyimide

hollow fiber membranes. J Polym Sci Part B: Polym Phys 2000;38:765–75.

[57] Chung TS, Kafchinski ER. Aging phenomenon of 6FDApolyimide/polyacrylonitrile

composite hollow fiber. J Appl Polym Sci 1996;59:77–82.

[58] Kim JH, Koros WJ, Paul DR. Effects of CO2 exposure and physical aging on the gas

permeability of thin 6FDA-based polyimide membranes. Part 1. Without

crosslinking. J Membr Sci 2006;282: 21–31.

[59] Kim JH, KorosWJ, Paul DR. Physical aging of thin 6FDA-based polyimide

membranes containing carboxyl acid groups. Part I. Transport properties. Polymer

2006;47:3049–103.

[60] Stern SA, Mi Y, Yamamoto H. Structure/permeability relationships of polyimide

membranes. Applications to the separation of gasmixtures. J Polym Sci Part B:

Polym Phys 1989;27:1887–909.

[61] Cheng SX, Chung TS, Wang R, Vora RH. Gas-sorption properties of 6FDA-

durene/1,4-phenylenediamine (pPDA) and 6FDA-durene/1,3-phenylenediamine

(mPDA) copolyimides. J Appl Polym Sci 2003;90:2187–93.

[62] Lin WH, Vora RH, Chung TS. Gas transport properties of 6FDAdurene/ 1,4-

phenylenediamine (pPDA) copolyimides. J Polym Sci Part B: Polym Phys

2000;38:2703–13.

[63] Liu SL, Wang R, Chung TS, Chng ML, Liu Y, Vora RH. Effect of diamine

composition on the gas transport properties in 6FDAdurene/3,3_-diaminodiphenyl

sulfone copolyimides. J Membr Sci 2002;202:165–76.

97
[64] Tanaka K, Kita H, Okano M, Okamoto KI. Permeability and permselectivity of

gases in fluorinated and non-fluorinated polyimides. Polymer 1992;33:585–92.

[65] Coleman MR, Koros WJ. Isomeric polyimides based on fluorinated dianhydrides

and diamines for gas separation applications. JMembr Sci 1990;50:285–97.

[66] Kawakami H, Anzai J, Nagaoka S. Gas transport properties of soluble aromatic

polyimides with sulfone diamine moieties. J Appl Polym Sci 1995;57:789–95.

[67] Park HB, Jung CH, Lee YM, Hill AJ, Pas SJ, Mudie ST, et al. Polymers with

cavities tunes for fast selective transport of small molecules and ions. Science

2007;318:254–8.

[68] KorosWJ, MaddenW. Transport properties. In: Mark HF, Kroschwitz JI, editors.

Encyclopedia of polymer science and technology, vol. 12. 3rd ed. John Wiley &

Sons; 2004. p. 291–381.

[69] Aitken CL, Koros WJ, Paul DR. Effect of structural symmetry on gas transport

properties of polysulfones. Macromolecules 1992;25:3424–34.

[70] Xu ZK, Böhning M, Schultze JD, Li GT, Springer J, Glatz FP, et al. Gas transport

properties of poly(phenylene thioether imide)s. Polymer 1997;38:1573–80.

[71] Tanaka K, Osada Y, KitaH, Okamoto KI. Gas permeability and permselectivity of

polyimides with large aromatic rings. J Polym Sci Part B: Polym Phys

1995;33:1907–15.

[72] Piroux F, Espuche E, Mercier R. The effects of humidity on gas transport properties

of sulfonated copolyimides. J Membr Sci 2004;232:115–22.

[73] Tanaka K, Islam MN, Kido M, KitaH, Okamoto K. Gas permeation and separation

properties of sulfonated polyimidemembranes. Polymer 2006;47:4370–7.

98
[74] McKeown NB, Budd PM. Polymers of intrinsic microporosity (PIMs): organic

materials for membrane separations, heterogeneous catalysis and hydrogen storage.

Chem Soc Rev 2006;35:675–83.

[75] Powell CE, Qiao GG. Polymeric CO2/N2 gas separation membranes for the capture

of carbon dioxide from power plant flue gases. J Membr Sci 2006;279:1–49.

[76] Okamoto K, Umeo N, Okamyo S, Tanaka K, Kita H. Selective permeation of carbon

dioxide over nitrogen through polyethyleneoxidecontaining polyimide membranes.

Chem Lett 1993;2:225–8.

[77] Yoshino M, Ito K, Kita H, Okamoto KI. Effects of hard-segment polymers on

CO2/N2 gas-separation properties of poly(ethylene oxide)-segmented copolymers. J

Polym Sci Part B: Polym Phys 2002;38:1707–15.

[78] Maya EM, Mu˜noz DM, de la Campa JG, de Abajo J, Lozano AE. Thermal effect on

polyethyleneoxide-containing copolyimide membranes for CO2/N2 separation.

Desalination 2006;199: 188–90.

[79] Smaihi M, Schrotter JC, Lesimple C, Prevost I, Guizard C. Gas separation properties

of hybrid imide–siloxane copolymers with various silica contents. J Membr Sci

1999;16:157–70.

[80] Hibshman C, Cornelius CJ, Marand E. The gas separation effects of annealing

polyimide–organosilicate hybrid membranes. J Membr Sci 2003;211:25–40.

[81] Burns RL, Koros WJ. Structure-property relationships for poly(pyrrolone-imide) gas

separationmembranes. Macromolecules 2003;36:2374–81.

[82] Coleman MR, Kohn R, Koros WJ. Gas separation applications of miscible blends of

isomeric polyimides. J Appl Polym Sci 1993;50:1059–64.

99
[83] Kapantaidakis GC, Kaldis SP, Dabou XS, Sakellaropoulos GP. Gas permeation

through PSF-PI miscible blend membranes. J Membr Sci 1996;110:239–47.

[84] Grobelny J, Rice DM, Karasz FR, MacKnight WJ. High resolution solid state 13C

nuclear magnetic resonance study of poly(ether sulphone)/polyimide blends. Poly

Commun 1990;31: 86–9.

[85] Liang K, Grebowicz J, Valles E, Karasz FE, MacKnight WJ. Thermal and

rheological properties of miscible polyethersulfone/polyimide blends. J Polym Sci

Part B: Polym Phys 1992;30:465–76.

[86] Cha YJ,KimET,AhnTK, Choe S.Mechanical and morphological phase behavior in

miscible polyethersulfone and polyimide blends. Polym J 1994;26:1227–35.

[87] Su TM, Ball IJ, Conklin JA, Huang S, Larson RK, Nguyen SL, Lew BM, Kaner RB,

Polyaniline/polyimide blends for pervaporation and gas separation studies.

Synthetic Met 1997;84:801–2.

[88] Chatzidaki EK, Favvas EP, Papageorgiou SK, Kanellopoulos NK, Theophilou NV.

New polyimide–polyaniline hollow fibers: synthesis, characterization and behavior

in gas separation. Eur Polym J 2007;43:5010–6.

[89] Hosseini SS, Teoh MM, Chung TS. Hydrogen separation and purification in

membranes of miscible polymer blends with interpenetration networks. Polymer

2008;49:1594–603.

[90] Bos A, Punt I, Strathmann H,Wessling M. Suppression of gas separation membrane

plasticization by homogeneous polymer blending. AIChE J 2001;47:1088–93.

100
[91] Guerra G, Choe S, Williams DJ, Karasz FE, MacKnight WJ. Fourier transform

infrared spectroscopy of some miscible polybenzimidazole/polyimide blends.

Macromolecules 1988;21:231–4.

[92] Chung TS, Guo WF, Liu Y. Enhanced Matrimid membranes for pervaporation by

homogenous blends with polybenzimidazole (PBI). J Membr Sci 2006;271:221–31.

[93] Musto P, Karasz FE, MacKnight WJ. Hydrogen bonding in

polybenzimidazole/polyimide blends: a spectroscopic study. Macromolecules

1991;24:4762–9.

[94] Jaffe M, Chen P, Choe EW, Chung TS, Makhija S. High performance polymer

blends. Adv Polym Sci 1994;117:297–327.

[95] Chung TS. A critical reviewof polybenzimidazoles: historical development and

future R & D. J Macromol Sci RevMacromol Chem Phys C 1997;37:277–301.

[96] Ekiner M, Simmons JW. Novel separation membrane made from blends of

polyimide with polyamide or polyimide-amide polymers. US20060156920, 2006.

[97] Cong H, Radosz M, Towler BF, Shen Y. Polymer-inorganic nanocompostie

membranes for gas separation. Sep Purif Technol 2007;55:281–91.

[98] Moaddeb M, Koros WJ. Gas transport properties of thin polymeric membranes in

the presence of silicon dioxide particles. JMembr Sci 1997;125:143–63.

[99] Chung TS, Chan SS, Wang R, Lu Z, He C. Characterization of permeability and

sorption in Matrimid/C60 mixed matrix membranes. J Membr Sci 2003;211:91–9.

[100] Xu Z, Xiao L, Wang J, Springer J. Gas separation properties of PMDA/ODA

polyimide membranes filling with polymeric nanoparticles. J Membr Sci

2002;202:27–34.

101
[101] Merkel TC, Freeman BD, Spontak RJ, He Z, Pinnau I. Ultrapermeable, reverse-

selective nanocomposite membranes. Science 2002;296:519–22.

[102] Anson M, Marchese J, Garis E, Ochoa N, Pagliero C. ABS copolymeractivated

carbon mixed matrix membrane for CO2/CH4 separation. J Membr Sci

2004;243:19–28.

[103] Li Y, Chung TS, Kulprathipanja S. Novel Ag+-zeolite/polymer mixed matrix

membranes with a high CO2/CH4 selectivity. AIChE J 2007;53:610–6.

[104] Yoda S, Hasegawa A, Suda H. Preparation of a platinum and palladium/ polyimide

nanocomposite film as a precursor of metal-doped carbon molecular sieve

membrane via supercritical impregnation. Chem Mater 2004;16:2363–8.

[105] Hu Q, Marand E, Dhingra S, Fritsch D, Wen J, Wilkes G. Poly(amide-imide)/TiO2

nano-composite gas separation membranes: fabrication and characterization. J

Membr Sci 1997;135: 65–79.

[106] Li Y, Guan HM, Chung TS, Kulprathipanja S. Effects of novel silane modification

of zeolite surface on polymer chain rigidification and partial pore blockage in

polyethersulfone (PES)—zeolite A mixed matrix membranes. J Membr Sci

2006;275:17–28.

[107] Kulprathipanja S, Neuzil RW, Li NN. Separation of fluids by means of mixed

matrix membranes. US patent 4,740,219, 1988.

[108] Mahajan R, Koros WJ. Mixed matrix membrane materials with glassy polymers.

Part 1. Polym Eng Sci 2002;42:1420–31.

[109] Vu DQ, KorosWJ, Miller SJ. Mixed matrix membranes using carbon molecular

sieves. I. Preparation and experimental results. J Membr Sci 2003;211:311–34.

102
[110] Fuhrman C, Nutt M, Vichtovonga K, Coleman MR. Effect of thermal hysteresis on

the gas permeation properties of 6FDA-based polyimides. J Appl Polym Sci

2004;91:1174–82.

[111] Kawakami H, Mikawa M, Nagaoka S. Gas transport properties in thermally cured

aromatic polyimide membranes. J Membr Sci 1996;118:223–30.

[112] Krol JJ, Boerrigter M, Koops GH. Polyimide hollow fiber gas separation

membranes: preparation and the suppression of plasticization in propane/propylene

environments. J Membr Sci 2001;184:275–86.

[113] Chung TS, Ren JZ, Wang R, Li DF, Liu Y, Pramoda KP, et al. Development of

asymmetric 6FDA-2,6DAT hollow fiber membranes for CO2/CH4 separation. Part

2. Suppression of plasticization. J Membr Sci 2003;214:57–69.

[114] Zhou FB, KorosWJ. Study of thermal annealing on Matrimid® fiber performance

in pervaporation of acetic acid and water mixtures. Polymer 2006;47:280–8.

[115] Qiao X, Chung TS, Pramoda KP. Fabrication and characterization of BTDA-

TDI/MDI (P84) copolyimide membranes for the pervaporation dehydration of

isopropanol. J Membr Sci 2005;264: 176–89.

[116] Carey FA. Organic chemistry. New York: McGraw Hill; 2006.

[117] Liu Y, Chng ML, Chung TS, Wang R. Effects of amidation on gas permeation

properties of polyimide membranes. J Membr Sci 2003;214:83–92.

[118] Hayes RA. Amine-modified polyimide membranes. US patent 4,981,497, 1991.

[119] Liu Y, Wang R, Chung TS. Chemical cross-linking modification of polyimide

membranes for gas separation. J Membr Sci 2001;189:231–9.

103
[120] Cao C, Chung TS, Liu Y,Wang R, Pramoda KP. Chemical cross-linking

modification of 6FDA-2, 6-DAT hollow fibermembranes for natural gas separation.

J Membr Sci 2003;216:257–68.

[121] Liu Y, Chung TS, Wang R, Li DF, Chng ML. Chemical crosslinking modification

of polyimide/poly(ether sulfone) dual-layer hollow-fiber membranes for gas

separation. Ind Eng Chem Res 2003;42:1190–5.

[122] Ren J, Wang R, Chung TS, Li DF, Liu Y. The effects of chemical modifications on

morphology and performance of 6FDA-ODA/NDA hollow fiber membranes for

CO2/CH4 separation. J Membr Sci 2003;222:133–47.

[123] Tin PS, Chung TS,Wang R, Liu Y, Liu SL, Pramoda KP. Effects of crosslinking

modification on gas separation performance of Matrimid membranes. J Membr Sci

2003;225:77–90.

[124] Shao L, Chung TS, Goh SH, Pramoda KP. Polyimide modification by a linear

aliphatic diamine to enhance transport performance and plasticization resistance. J

Membr Sci 2005;256:46–56.

[125] Shao L, Chung TS, Goh SH, Pramoda KP. The effects of 1,3-cyclohexane

bis(methylamine) modification on gas transport and plasticization resistance of

polyimide membranes. J Membr Sci 2005;267:78–89.

[126] Chung TS, Chng ML, Pramoda KP, Xiao Y. PAMAM dendrimer induced cross-

linking modification of polyimide membranes. Langmuir 2004;20:2966–9.

[127] Shao L,ChungTS,GohSH, PramodaKP. Transportproperties of crosslinked

polyimide membranes induced by different generations of diaminobutane (DAB)

dendrimers. J Membr Sci 2004;238:153–63.

104
[128] Okamoto KI, Ijyuin T, Fujiwara S, Wang H. Synthesis and characterization of

polyimides with pendant phosphonate ester groups. Polym J 1998;30:492–8.

[129] Liou GS, Wang JSB, Tseng ST, Tsiang RCC. New organo-soluble aromatic

polyimides based on 3,3_,5,5_-tetrabromo-2,2-bis[4-(3,4dicarboxyphenoxy)

phenyl] propane dianhydride and aromatic diamines. J Polym Sci Part A: Polym

Chem 1999;37:1673–80.

[130] Guiver MD, Robertson GP, Dai Y, Bilodeau F, Kang YS, Lee KJ, et al. Structural

characterization and gas-transport properties of brominated Matrimid polyimide. J

Polym Sci Part A: Polym Chem 2003;40:4193–204.

[131] Kharitonov AP, Mokvin YL, Syrtsova DA, Starov VM, Teplyakov VV. Direct

fluorination of the polyimide Matrimid 5218: the formation kinetic and

physicochemical properties of the fluorinated layers. J Appl Polym Sci 2004;92:6–

17.

[132] Kita H, Inada T, Tanaka K, Okamoto K. Effect of photocrosslinking on

permeability and permselectivity of gases through benzophenonecontaining

polyimide. J Membr Sci 1994;87:139–47.

[133] Won J, Kim MH, Kang YS, Park HC, Kim UY, Choi SC, et al. Surface

modification of polyimide and polysulfone membranes by ion beam for gas

separation. J Appl Polym Sci 2000;75:1554–60.

[134] Hu L, Xu X, Coleman MR. Impact of H+ ion beam irradiation on Matrimid. II.

Evolution in gas transport properties. J Appl PolymSci 2007;103:1670–80.

[135] Bos A, Pünt IGM,Wessling M, Strathmann H. Plasticization-resistant glassy

polyimide membranes for CO2/CH4 separation. Sep Purif Technol 1998;14:27–39.

105
[136] Takeichi T, Ogura S, Takayama Y. Soluble polyimides that contain curable internal

acetylene groups in the backbone. J Polym Chem Part A: Polym Chem

1994;32:579–85.

[137] Xiao Y, Chung TS, Guan HM, GuiverMD. Synthesis, cross-linking and

carbonization of co-polyimides containing internal acetylene units for gas

separation. J Membr Sci 2007;302:254–64.

[138] Rezac ME, Sorensen ET, Beckham HW. Transport properties of cross-linkable

polyimide blends. J Membr Sci 1997;136: 249–59.

[139] Punt IGM, Bos A, Wessling M, Strathmann H. Suppression of CO2-plasticization

by semi-interpenetrating polymer network formation. J Polym Sci Part B: Polym

Phys 1998;36:1547–56.

[140] Hillock AMW, Koros WJ. Cross-linkable polyimide membrane for natural gas

purification and carbon dioxide plasticization reduction. Macromolecules

2007;40:583–7.

[141] Staudt-Bickel C, Koros WJ. Improvement of CO2/CH4 separation characteristic of

polyimides by chemical crosslinking. J Membr Sci 1999;155:145–54.

[142] Wind JD, Staudt-Bickel C, Paul DR, KorosWJ. The effect of crosslinking

chemistry on CO2 plasticization of polyimide gas separation membranes. Ind Eng

Chem Res 2002;4:6139–48.

[143] Wind JD, Staudt-Bickel C, Paul DR, Koros WJ. Solid-sate covalent cross-linking

of polyimide membranes for carbon dioxide plasticization reduction.

Macromolecules 2003;36:1882–8.

106
[144] Kim JH, Koros WJ, Paul DR. Effects of CO2 exposure and physical aging on the

gas permeability of thin 6FDA-based polyimide membranes Part 2. With

crosslinking. J Membr Sci 2006;282: 32–43.

[145] Hayes RA. Polyimide gas separation membranes. US patent 4,717,393, 1988.

[146] Xiao Y, Chung TS, Chng ML. Surface characterization, modification chemistry

and separation performance of polyimide and PAMAM dendrimer composites.

Langmuir 2004;20:8230–8.

[147] Xiao Y, Shao L, Chung TS, Schiraldi DA. The effects of thermal treatments and

dendrimer chemical structures on the properties of highly surface cross-linking

polyimide films. Ind Eng Chem Res 2005;44:3059–67.

[148] Powell CE, Duthie XJ, Kentish SE. Reversible diamine cross-linking of polyimide

membranes. J Membr Sci 2007;29:199–209.

[149] Duthie XJ, Kentish SE, Powell CE, Qiao GG, Nagai K, Stevens GW. Plasticization

suppression in grafted polyimide epoxy network membranes. Ind Eng Chem Res

2007;46: 8183–92.

[150] Park HB, Kim YK, Lee JM, Lee YM. Relationship between chemical structure of

aromatic polyimides and gas permeation properties of their carbon molecular sieve

membranes. J Membr Sci 2004;229:117–27.

[151] Steel KM, Koros WJ. An investigation of the effects of pyrolysis parameters on gas

properties of carbon materials. Carbon 2005;43:1843–56.

[152] Xiao Y, Chung TS, Chng ML, Tamai S, Yamaguchi A. Structure & properties

relationships for aromatic polyimides and their derived carbon membranes:

experimental and simulation approaches. J Phys Chem B 2005;109:18741–8.

107
[153] Shao L, Chung TS, PramodaKP. The evolution of physicochemical and transport

properties of 6FDA-durene toward carbon membranes; from polymer, intermediate

to carbon. Micropor Mesopor Mater 2005;84:59–68.

[154] Tin PS, Xiao Y, Chung TS. Polyimide-carbonized membranes for gas separation:

structural, composition, and morphological control of precursors. Sep Purif Rev

2006;35:285–318.

[155] Islam MN, Zhou W, Honda T, Tanaka K, Kita H, Okamoto K. Preparation and gas

separation performance of flexible pyrolytic membranes by low-

temperaturepyrolysis of sulfonatedpolyimides. J Membr Sci 2005;26:17–26.

108
CHAPTER THREE

ENHANCED GAS SEPARATION PERFORMANCE OF NANOCOMPOSITE

MEMBRANES USING MGO NANOPARTICLES

3.1 INTRODUCTION

After the emergence of commercial membranes for separation, this technology

has yet been firmly struggling to establish its position as the preponderant choice for

various applications. During this time span, membrane technology has been the subject of

numerous investigations in various aspects which has led to the better performance and

broader range of applications. After successful achievements in this field, nowadays,

membranes have found their proper candidates in various types of liquid and gas based

separations. Low operation cost, relatively small footprint, low energy requirement and

consortion with environmental regulations are of the major benefits of membranes over

their conventional counterpart technologies. Membranes have been successfully

developed for separation of gas molecules and they are widely used in chemical and

petrochemical plants. Separation of hydrogen from its mixtures with nitrogen or

hydrocarbons, nitrogen purification and the foremost CO 2 removal from natural gas are

recognized as major established and high value industrial applications of membranes.

109
Different types of materials have been used in the membrane fabrication, where

polymers have attracted more attentions. This is mainly due to the low production cost

and ease of processing, especially for making them in the form of hollow fibers which

offer membranes with higher surface area per unit volume. However, conventional

polymeric membranes generally suffer from the trade-off limitation between productivity

and selectivity which yet remains as one of the most important challenges for further

developments in this field [1]. Different methodologies have been adopted by researchers

in the architecture andmodification of polymers as well as fabrication technology in order

to overcome the shortcomings of polymeric membranes.

Nanocomposite membranes (also known as mixed matrix membranes) and

facilitated transport membranes (FTM) are two distinguished and promising generations

of gas separation membranes which have attracted considerable amount of attention for

their potential use in gas separation and purification.

The microstructure of a nanocomposite membrane is commonly composed of an

inorganic material, in the form of micro- or nanoscale particles, as the discrete phase, and

a polymeric material which forms the continuous or matrix phase. The successful features

of nanocomposite membranes are mainly beholden to the synergistic combination of easy

processability of polymers and superior gas separation factor of inorganic materials. In

addition they may offer enhanced physical, thermal and mechanical properties for

aggressive and adverse environments [2,3]. Some studies also indicated that the particles

110
added into the polymer matrix might have a hope of stabilizing the polymer membrane

against change in permselectivity with temperature [4].

Several inorganic materials have been proposed as good candidates for the

fabrication of nanocomposite membranes. These materials are generally categorized into

two major classes of porous and nonporous fillers. Each type of these materials offer

distinguished transport mechanisms to the membrane.

Zeolites are probably the frontier family of molecular sieves and have been

widely used in both rubbery and glassy polymers for gas separation membranes [5-11].

Zeolite filled membranes from rubbery polymers showed significantly improved

performance for separation of oxygen/nitrogen and carbon dioxide/methane compared to

their pristine polymeric membranes [6,7]. However the use of zeolites with the glassy

polymers has confronted some obstacles specially in controlling the polymer-particle

interface [10-14]. Studies on zeolite-based nanocomposite membranes using glassy

polymers have revealed the pivotal roles of polymer-zeolite affinity at interface on the

resultant gas separation performance; therefore, applying some modification techniques

has resulted in progress in this area [9-11,15,16].

Carbon molecular sieves (CMS) are considered another candidate which exhibit

excellent intrinsic performance for gas separation processes. This micro to nano-size

materials are generally obtained through the pyrolysis of polymeric precursors. CMS

materials can be used as membranes directly or after milling be introduced into a

111
polymeric matrix to form nanocomposite membranes. It is well documented the

polymeric precursor plays a crucial role in the performance of CMS and extensive

researches have been carried out on the preparation of CMS from both rubbery and glassy

polymers using different techniques [17-23].

On the other hand, metal oxides nanoparticles have also important materials due

to their potential applications for membrane-based separation purposes. Moaddeb and

Koros [24] prepared composite membranes by introduction of silicon dioxide particles

into the matrix of six high performance polymers. They found that presence of silica

improves the gas separation performance of the selective layer particularly for oxygen

and nitrogen. This was mainly attributed to the enhanced rigidity of polymer matrix due

to the adsorption of polymer chains to the surface of silica.

Hu et al. [4] developed nanocomposite membranes based on fluorinated

poly(amide-imide) polymer matrix and TiO 2 component via sol-gel technique. These

composite membranes similarly showed higher permselectivities for selected gas pairs

when compared to the neat polymeric counterpart membrane. They attributed the results

to the denser and more rigid structure of the composite membranes which detected by the

increase in apparent activation energy of permeation.

Very recently, Lin et al. [25] introduced the use of MgO nanoparticles in the form

of nanocomposite membranes for CO 2 /H 2 separation. In the optimum condition, they

achieved significantly improved CO 2 permeability while maintaining selectivity by

112
loading around 32wt% of magnesium oxide nanoparticles into the crosslinked

poly(ethylene glycol) diacrylate matrix. The enhanced CO 2 permeability was ascribed to

the effect of nanoparticles on gas diffusivity.

Similar to the nanocomposite membranes, FTMs have also shown promising

results for gas separation. The structure of FTM is designed in a way that some particular

carriers inside the membrane may have the special affinity toward the target gas

molecules and this interaction controls the rate of transport. The majority of the carrier

candidates introduced by the researchers are in the class of metal ions. Generally, the

application of ions for membrane-based gas separation can be categorized into four major

types of immobilized liquid membranes, ion exchange membranes, polymer/metal ion

blends and modified polymer membranes [26-29]. Limitations associated with

immobilized membranes concerning operating conditions [27,28] led to the emergence of

ion exchange membranes. Way et al. prepared ion exchange membranes containing

organic amine counter-ions which showed enhanced performance for CO 2 removal from

its mixture with methane [30].

The application of silver ions has been extensively studied as a promising carrier

for the fabrication of FTMs. It is well documented that the presence of some special

metal ions like Ag+ facilitates the transport of olefins. This is ascribed to the ability of

silver ions to form reversible π-complexation with double-bond gas molecules, leading to

a subsequent solubility enhancement. Therefore, this mechanism has been successfully

exploited as facilitated transport membranes for olefin/paraffin separation [31,32]. As a

113
result, selection of suitable carrier with high affinity toward the desired molecules, in

conjunction with the adoption of appropriate strategy for design and fabrication of

membranes are of the principal criteria to wield the advantages of facilitated transport for

gas separation.

The presence of affinity and interaction between MgO surface and some gas

species like carbon dioxide [33] provides a great motivation to investigate the potential

application of this material as carriers in FTM structure for separation of gas molecules.

In addition, research results have shown that MgO surface has a good adsorption affinity

toward transition metals which is highly advantageous for alteration of surface properties.

Klabunde et al. [34] demonstrated magnesium oxide nanocrystals containing a monolayer

of a transition metal oxide on the surface have greatly enhanced efficiency for the

adsorption of chlorocarbons, hydrogen sulfide, and sulfur dioxide. Therefore, it seems to

be a novel idea that one combines the superior performance of nanocomposite and

facilitated transport membranes together in a single membrane. The objective of this

research study is to develop high performance nanocomposite membranes for hydrogen

separation and CO 2 removal from natural gas by combining the features of inorganic and

organic materials. Nanocomposite membranes were fabricated by a solution-casting

method with different loadings of MgO nanoparticles. Matrimid, a commercially

available glassy polymer with high T g was selected as the continuous phase for all

nanocomposite membranes. Matrimid possesses high gas permeability and selectivity as

well as high thermal stability and solvent resistance properties [35-37]. Firstly, the

characteristics of both MgO and its derived membranes were evaluated by various

114
techniques. The effects of MgO loading on structure-property relationship and gas

separation performance of these membranes were examined. Secondly, these membranes

were post-treated with silver nitrate solution for stipulated periods of time. The effects of

silver treatment and immersion time on gas separation performance of Matrimid-MgO-

Ag+ nanocomposite membranes were investigated. To our best knowledge, this is the first

study that combines the synergistic features of nanocomposite structure and facilitated

transport mechanism into one membrane by using unique characteristics of MgO for gas

separation.

3.2 EXPERIMENTAL

3.2.1 Materials

The polyimide (Matrimid5218) was purchased from Vantico Inc., (Luxemburg).

The solvents including N-methyl-2-pyrolidone (NMP) and methanol were supplied by

Merck and were used without further purification. The magnesium oxide nanoparticles

(NanoActive MgO Plus) were purchased from Nanoscale Materials Inc. (Manhattan,

Kansas). The physical properties of magnesium oxide nanoparticles are shown in Table

3.1. Silver nitrate solid with a reagent grade was supplied by Merck. Both polymer

powders and MgO nanoparticles were dried overnight at 120oC under vacuum prior to

use.

115
Table 3.1. Properties of magnesium oxide (MgO Plus) nanoparticlesa

Property Value
Specific surface area (BET) > 600 m2/g
Crystallite size < 4nm
Average pore diameter 30 Å
Total pore volume > 0.4 cc/g
Mg content (based on metal) > 99.2 %
True density 2.4 g/cc
aData obtained from the manufacturer’s MSDS.

3.2.2 Membrane fabrication

Since MgO is very susceptible to water adsorption and agglomeration, the

nanoparticles were firstly dried under vacuum for overnight and then stipulated amounts

of MgO nanoparticles were sifted on a 20μm-size test sieve (Retsch®, Germany) to

remove any coarse agglomerates. The fine particles were then dispersed in NMP under

continuous stirring for about an hour followed by sonication for 5 min to break any

aggregation and improve the dispersion quality. The polymer powders were then

gradually added up to 20wt% concentration and the stirring was continued for at least one

day to achieve a homogenous solution. The high polymer concentration was selected in

order to prevent the sedimentation of MgO nanopaticles during the casting. Matrimid-

MgO nanocomposite membranes were fabricated via solution-casting at high

temperatures followed by natural cooling. The technique adopted here was similar to the

one previously developed in our group [15]. The maximum loading of MgO was set to

40wt% due to the handling problems and brittleness of MMMs at higher loadings.

116
3.2.3 Post-treatment of membranes

Nanocomposite membranes containing 20wt% MgO underwent various heat

treatment steps. Small pieces of samples were heat treated in a precise furnace under

vacuum condition at different temperature profiles of below and above the glass

transition temperature of the polymer matrix (Matrimid T g =318˚C). The effect of cooling

profile was also investigated by applying two distinguished procedures. For the course of

natural cooling, the samples were allowed to reach the room temperature by turning off

the heating source. During this step, the samples remained under vacuum. In contrast,

other counterpart samples were quenched inside the liquid nitrogen immediately after

removing from the furnace. They were further dried under vacuum before permeation

tests.

Silver treatment was carried out through a dip coating technique. Firstly, silver

nitrate grains were dissolved in methanol. A saturated solution was obtained after stirring

for almost one hour. A piece of membrane was cut and immersed in this solution for a

certain period of time. Afterwards, silver treated membranes were washed in fresh

methanol to remove the residual and unused silver ions. Membranes were dried at room

temperature followed by vacuum drying at 150oC overnight to eliminate any trace of

methanol. Finally, membranes were cooled naturally and were ready for characterization

and gas permeation tests.

3.2.4 Membrane characterizations

117
The morphology of membranes was examined by scanning electron microscope

(SEM). The electron micrographs were obtained on a JEOL JSM-5600LV and JSM-

6700F. Both Energy dispersive X-ray (EDX) and X-ray photoelectron spectroscopy

(XPS) techniques were employed for elemental analysis of membranes. EDX analysis

was carried out using Oxford Instruments Inca EDX system equipped with SEM

machine. XPS data were acquired on a Kratos AXIS-HSi spectrometer with built-in

charge neutralization system. A dual anode Al/Mg X-ray source was operated at 300W

(15 kV, 20 mA).

The structural properties and inter-chain spacing of membranes were analyzed

qualitatively on an X’Pert PRO-MRD wide-angle X-ray diffractometer (XRD) from

PANalytical. The measurements were carried out by using Cu Kα radiation with a

wavelength of λ = 1.54Å at room temperature. The scan range was from 5 to 85


˚ with a

step increment of 0.04̊/s. Average d-spacing was determined based on the Bragg’s law,

viz.

nλ = 2d sin θ (3.1)

where n is an integral number (1, 2, 3, ...), λ denotes the X-ray wavelength, d stands for

the intersegmental spacing between two polymer chains and θ indicates the diffraction

angle.

3.2.5 Gas permeability measurements

The gas permeation properties of membranes were determined by a variable-

pressure constant-volume method. Detailed experimental design and procedures can be

118
found elsewhere [38]. The thickness of membranes used for permeability measurements

was about 50μm. The rate of pressure increase (dp/dt) at steady state was used for the

calculation of gas permeability with the following Eq. (3.2):

273.15 × 1010 Vl  dp 
P=   (3.2)
760 AT (( P0 × 76) / 14.7)  dt 

where P is the gas permeability of a membrane in Barrer (1Barrer = 1×10-10cm3 (STP)-

cm/cm2.sec.cmHg), V is the volume of the down-stream chamber (cm3), A refers to

effective area of the membrane (cm2), l is the membrane thickness (cm), T is the

operating temperature (K) and finally pressure of the feed gas in up-stream is given by P 0

in psia.

The pure gas permeability was obtained in a sequence of He, H 2 , O 2 , N 2 , CH 4

and CO 2 at 35oC for each nanocomposite membrane. The H 2 permeability test was

conducted at 3.5atm, while the testing pressure for other gases was 10atm. The ideal

selectivity of a membrane for gas A to gas B was evaluated as follows:

PA
α A/ B = (3.3)
PB

3.2.6 Measurement of gas sorption

CO 2 gas sorption tests were conducted for nanocomposite membrane films using

a Cahn D200 microbalance sorption cell at 35 oC over a pressure range of 0-15 atm.

Approximately 100 mg membrane films were placed on the sample pan. The system was

evacuated for 24 h prior to testing. The gas at a specific pressure was fed into the system

and the sample started to sorb the gas until the equilibrium was achieved. From the

119
weight gain, the amount of gas dissolved in the membrane was calculated after counting

the buoyancy correction.

3.3 RESULTS AND DISCUSSION

3.3.1 Effect of MgO loading on the gas separation performance of nanocomposite

membranes

Matrimid-MgO nanocomposite membranes with MgO loadings from 20wt% to

40wt% were fabricated according the procedure mentioned in the previous section. For

comparison, one piece of neat Matrimid membrane was fabricated in the form of dense

flat following the same procedure. Figure 3.1. shows a uniform distribution quality of

MgO nanoparticles inside the polymer matrix analyzed by EDX. This figure also

confirms that the dope viscosity has been high enough to prevent the particles from

sedimentation during the membrane formation.

Figure 3.2. exhibits the gas permeability of neat Matrimid and nanocomposite

membranes with different MgO loadings. It can be seen that the permeability of

nanocomposite membranes is higher than that of neat Matrimid membrane. This figure

also shows a monotonous increase in permeability with increasing the MgO content for

all the gases. The average increment in gas permeability is about 75% which is obtained

for the membrane with 40wt% MgO loading. Clearly, the enhancement in gas

permeability is mainly due to the incorporation of highly porous MgO nanoparticles

which are substituted some portions of the dense structure of polymer chains in the

membrane structure.

120
Figure 3.1. Dispersion quality of MgO nanoparticles at the surface of the nanocomposite

membrane with 20wt% MgO.

In principle, the increase in permeability of nanocomposite membranes may be

attributed to the presence of microvoids at the polymer-particles interface and/or the

inherent transport properties of porous particles if encompassing pores larger than the

kinetic diameter of gases. The cross-section morphology of nanocomposite membranes

with different MgO loadings is shown in Figure 3.3. The morphological analysis of the

membranes reveals that particles are intimately intercalated inside the polymer matrix

and no macrovoids or defect could be observed at the polymer-particle interface. We

believe, adoption of appropriate casting procedure, controlling of temperature and phase

inversion process, and final cooling profile play important roles in achieving successful

interfacial adhesion between polymer chains and particles.

121
50 4

45 Pure Matrimid Pure Matrimid


3.5
20% MgO Loading 20% MgO Loading
40
30% MgO Loading 3 30% MgO Loading
35 40% MgO Loading 40% MgO Loading
30
2.5

25 2

20
1.5
15

Permeability (Barrer)
Permeability (Barrer)
1
10
0.5
5

0 0
Helium Hydrogen Carbon dioxide Oxygen Nitrogen Methane
Test gas Test gas

Figure 3.2. The comparison of gas permeability for pure Matrimid and nanocomposite membranes with different

loadings of MgO nanoparticles. (left: permeability of helium, hydrogen and carbon dioxide; right: permeability of

oxygen, nitrogen and methane)

122
A B C
Figure 3.3. The cross-section morphology of nanocomposite membranes with different loadings of MgO

nanoparticles. (MgO loadings: A: 20wt% B: 30wt% C: 40wt%). Inset: high magnification image of MgO

nanoparticles inside the polymeric matrix.

123
Therefore, it can be concluded that the permeability increment observed for

nanocomposite membranes are resulted from the contribution of highly porous structure

of magnesium oxide nanoparticles in which all the gas molecules, regardless of size, can

pass through the particles without any resistance against their transport. The separation

performance of membranes was calculated for selected gas pairs. The ideal selectivity

results of the as-fabricated MgO containing nanocomposite membranes are listed in

Table 3.2. These data denote a general ebb in selectivity performance of nanocomposite

membranes compared to the neat Matrimid membrane. This corroborates our reasoning

for the gas permeability results and reveals that the pore size (30 Å) of MgO

nanoparticles is much larger than the kinetic diameter of gas molecules. As a result, the

porous structure of MgO nanoparticles cannot selectively distinguish the gas molecules

based on their size difference, thus the overall diffusivity selectivity of membranes is

diminished.

Table 3.2. The ideal selectivity of pure Matrimid and Matrimid-MgO nanocomposite

membranes with different nanoparticle loadings.

Membrane name Ideal Selectivity


H2/N2 He/N2 CO2/CH4 O2/N2 H2/CO2
Matrimid®5218 97.0 93.1 33.3 7.14 3.88
Matrimid+MgO(20%) 89.5 86.0 29.8 6.31 3.89
Matrimid+MgO(30%) 89.3 85.5 26.9 6. 28 3.64
Matrimid+MgO(40%) 90.1 86.1 26.4 6.50 4.60

124
3.3.2 Effect of heat treatment on the gas separation performance of nanocomposite

membranes

Membranes containing 20wt% MgO were used to investigate the effect of heat

treatment on the transport and separation performance of nanocomposite membranes. The

results of the gas permeability and selectivity of heat treated membranes are listed in

Table 3.3. Two distinct behaviors are detected with respect to the treatment temperature.

A considerable reduction in gas permeability values are observed for the membranes heat

treated at 240˚C. However, the permeability was significantly increased for the samples

heat treated at 350˚C. In comparison with the results obtained from untreated membranes,

it seems that heat treatment at the temperature below the polymer T g has resulted in the

densification of the polymeric matrix; thus diminishing free volumes and bringing about

more resistance for the gas transport across the membrane. On the other hand, heat

treatment above the T g of polymer provides more freedom in the mobility of polymer

chains which results in the increase in fraction of free volume in the membrane structure.

Furthermore, it was observed that the sample with immediate quenching exhibited even

further increased permeability. From the structural point of view, this can be attributed to

the fact that natural cooling provides polymer chains sufficient time for rearrangements

and position themselves more ordered structure. Therefore, this process annihilates some

portion of the generated excess free volumes. However, in the sample with quench

profile, generated free volumes instantly become frozen at the incipient point of cooling.

The results are well in agreement with the study of Tsujita [39] on the behavior of glassy

polymers toward thermal conditioning. It is interesting to note that the selectivity

performance of the membranes with different treatment profiles remained unchanged in

125
Table 3.3. The effect of heat treatment and cooling profile on the permeability and selectivity of Matrimid-MgO nanocomposite
b
Membrane: Permeability (Barrer) Ideal Selectivity
Matrimid-MgO(20%)
Heat treatment Cooling
condition profile He H2 O2 N2 CH4 CO2 H2/N2 He/N2 CO2/ CH4 O2/N2
Untreated - 31.0 32.2 2.27 0.360 0.278 8.27 89.5 86.0 29.8 6.31
Heat treated at
Natural 25.3 27.0 2.06 0.319 0.245 7.36 84.6 79.3 30.0 6.46
240˚C for 12 hrs
Heat treated at
Natural 37.6 41.2 3.22 0.503 0.389 10.78 81.9 74.8 27.7 6.40
350˚C for 1 hr
Heat treated at
Quench 41.1 44.7 3.31 0.520 0.406 11.64 86.0 79.0 28.7 6.37
350˚C for 0.5 hr
1 Barrer = 1*10-10 cm3(STP) cm/cm2 s cm Hg

Testing Temperature 35oC

126
the same range. This may imply that the intimate interface of polymer-particle has been

in its perfect initial condition.

3.3.3 Effect of MgO nanoparticles on silver adsorption into the nanocomposite

membranes

Membranes containing 20wt% MgO loading were immersed inside the silver

nitrate solution for stipulated period of time. One piece of neat Matrimid membrane was

also treated with silver ions for comparison. Different techniques were used to analyze

the characteristics and structural changes of membranes after silver treatment. Visual

observation indicated that all nanocomposite membranes were discolored from yellowish

to brownish after silver treatment. The discoloration was only observed for the

membranes containing MgO nanoparticles while the color of neat Matrimid membrane

remained intact. Taking into account the natural color of silver element, the observed

discoloration suggests that some interactions have been taken place between membrane

and silver ions. The uniformity of the brownish color throughout the membrane structure

may be a sign of successful penetration of silver ions into the nanocomposite membranes.

In addition, it was found that the intensity of brownish color was higher for the

membranes with longer immersion time which inarguably denotes the presence of higher

concentration of silver ions at the prolonged immersion time scales.

EDX analysis was conducted to further elucidate the state and adsorption of silver

ions into the nanocomposite membrane. Figure 3.4. shows a comparison in profiles of

elementary analysis for Mg and Ag elements which was performed at the cross-section of

127
membranes with a 20wt% MgO loading after silver treatment for different periods of

time. The result of silver tracing inside the silver treated neat Matrimid membrane is

depicted in this figure. Application of elemental line mapping at the cross-section of

membranes suggests that silver cations have successfully penetrated into the whole

structure of nanocomposite membranes. One can see that the intensity of the peaks was

increased for the samples with longer retention time; indicating an increase in population

of silver cations as a function of time. However, it is clear in Figure 3.4. that much less

silver ions have penetrated into the neat Matrimid membrane. This may suggest that there

is almost no sensible interaction between the elements in polymer chains and silver ions.

In other words, the polymer itself may not exert any attractive force to draw the silver

ions into the membrane. These results lead us to conclude that the main driving force

causing the migration of silver ions from the solution into the nanocomposite membranes

have been induced by MgO nanoparticles. This can be confirmed by careful scrutiny of

EDX profiles where the peaks of Mg and Ag elements have been appeared in

corresponding locations and follow the same trend. As a result, a close consistency in the

locus and concentration of these two elements can be drawn. This indicates that the major

adsorption of silver elements takes place onto the sites provided by MgO nanoparticles.

The surface properties of neat Matrimid and nanocomposite membranes were

studied using XPS. Data in Table 3.4. present the binding energy values of C, N, and O

elements for the neat Matrimid membrane before and after silver treatment. As it was

expected based on the previous results, the amount of silver ions attached to the

membrane was not high enough to be detected by XPS. In addition, no significant

128
Figure. 3.4. The elementary profiles of EDX analysis for the pure Matrimid and nanocomposite membranes (20wt% MgO

loading) after silver treatment for different periods of time (Upper row: Ag element, Lower row: MgO element).

Matrimid-0%-1 day Matrimid-20%-2 days Matrimid-20%-5 days Matrimid-20%-10 days

15- 90-90
30-30
20- Ag Ag
10- 60-60
20-20

10-
5- 30-30
10-10

0- 0- 0-0
0-0

150-
200- 200-

100-

100- 100-
50-

0- 0- 0-

129
Table 3.4. XPS element binding energies of the neat Matrimid membrane before and after

silver treatment for 2 days.

Element Binding energy (eV) Shift (eV)


Before treatment After treatment
C 1s 282.9 283.2 0.3

O 1s 529.8 530.0 0.2

N 1s 398.1 398.7 0.6

changes are found in the position of binding energy profile of polymer elements after

silver treatment. This might be the evidence to rules out the presence of any interaction

between polymer and silver ions. Meanwhile, XPS Ag 3d spectra of nanocomposite

membranes containing 20wt% MgO after silver treatment are shown in Figure 3.5. A

step-wise increment in intensity and area under the index peaks of Ag element can be

observed as a function of immersion time. In addition, the analytical calculations on the

XPS spectra revealed the fact that mass percentage of silver ions underwent a change

from 21.28% up to 36.96% with an increase in immersion time from 2 to 10 days.

Therefore, XPS analysis confirms that the adsorption of silver ions into the

nanocomposite membranes is mainly due to the presence of MgO nanoparticles, which is

in good agreement with the results obtained from EDX analysis.

In order to understand the mechanism(s) involved in the migration of silver ions

into the membrane, probably it is needed to focus on the possible interactions between

MgO and Ag and their interfacial organization. Interestingly, due to peculiar

characteristics of the Ag/MgO hybrid system and its technological importance in

130
composite materials and microelectronics industry, the structure and morphology of

Ag/MgO interface has been the subject of extensive studies [40]. One of the reasons is

that the surface properties of MgO is basically considered as one of the prototype systems

for studying adsorption of ionic surfaces [41]. The physical and chemical properties of

magnesium oxide are primarily governed by the source of the precursor, that is, derived

from magnesite or precipitated from brine or seawater.

×102

Matrimid+MgO (20%)+2 days silver treatment


Matrimid+MgO (20%)+ 5 days silver treatment
Matrimid+MgO (20%)+10 days silver treatment
120
Intensity (CPS)

80

40

380 376 372 368 364 360


Binding Energy (eV)

Figure 3.5. XPS Ag 3d spectra of the nanocomposite membranes with 20wt% MgO

nanoparticles after silver treatment for different immersion times.

Basically, Ag is a noble metal, and hence cannot be involved in any chemical

reaction with other materials. Therefore, the possibility of interaction in the form of

chemical adsorption between Ag ions and MgO is very low. High surface activity and

131
strong basicity characteristics of MgO surface may suggest the physisorption mechanism

for this process of Ag adsorption onto the MgO which is confirmed by other researchers

[42]. Interestingly, this mechanism is in agreement with the findings in other works. For

example, interfacial study on the Ag/MgO using cluster models has determined the

presence of electrostatic bonding rather than covalent at Ag/MgO interface [43].

Considering the Face-Centered Cubic structure of MgO crystal, three possible

sites have been proposed for the adsorption of Ag ions onto the MgO viz. above O ions,

above Mg ions, or in between the octahedral sites [44]. The majority of interfacial studies

on Ag/MgO systems have confirmed that the main attractive forces from MgO are

offered by oxygen ions rather than magnesium ions [45]. In other words, the attractive

energy between distinguished ions causes the migration of Ag cations into the

membranes and finally they are deposited onto the anionic oxygen sites. In this respect,

Figure 3.6. might be an appropriate representative model for molecular structure of

adsorption of Ag ions onto the surface of MgO nanoparticles in which preferential

adsorption sites are assigned to the oxygen atoms of MgO. This model is quite similar to

the one proposed by Qin et al.[40].

In addition, Flank et al. demonstrated by their research study that the electrostatic

attractive force between MgO and Ag is so strong that may support the Ag ion over the

range of monolayer Ag deposits, which can lead to a formation of three-dimensional Ag

clusters over the MgO surface [46]. This fact is well considered in our proposed model

132
(Figure 3.6.) and dotted lines represent the direct interaction between oxygen and silver

atoms, while dashed lines indicate the over-layer interactions.

Legend
Ag
Mg
O

Figure 3.6. The schematic representation and molecular model for the adsorption of Ag

ions onto the oxygen sites at the surface of MgO. (Type of Interaction: Direct: dotted

lines, Secondary: dashed lines)

In literature, the calculated distance for Ag-O interaction is reported to be from 2.28 to

2.70Å which is determined through ab initio method [40]. It could be seen from our

characterization results that the amount of silver ions inside the membranes shows a

gradual increase even after 5-day immersion. This may suggest that the process of Ag+

adsorption onto the MgO surface is a rate dependant process in which adsorption sites are

occupied progressively with the retention time.

133
3.3.4 Effect of silver treatment on the gas separation performance of nanocomposite

membranes

The gas permeation test was carried out on the nanocomposite membranes with

20wt% MgO content to investigate the effect of silver treatment on the gas separation

performance. The pure gas permeability and ideal selectivity of membranes treated with

silver ions are presented in Table 3.5. Results show that silver treatment for 2 days causes

a significant decrease in permeability of all gases. It is interesting to note that the gas

permeability results are in correlation with the size of gas molecules. In other words, the

permeability increases with decreasing gas kinetic diameters in the sequence of CH 4 , N 2 ,

O 2 , CO 2 , H 2 and He. Moreover, a further decrease in permeability is observed by

prolonging the silver treatment. Thus, the minimum permeability is obtained for the

nanocomposite membranes with 10 days silver treatment time. However, surprisingly, a

reverse trend is observed for the light gases (i.e., He and H 2 ) and the immersion time has

an increasing effect on permeability of the membranes toward these two gases. The direct

consequence of this strange behavior is the achievement of significantly enhanced

selectivity of light gases over nitrogen. For example, H 2 /N 2 selectivity has reached to

146, which is 70% higher than that of pure polymer membrane. Likewise, He/N 2

selectivity of the nanocomposite membrane with 10-day treatment has increased up to

twofold compared to the untreated sample membrane. It also can be found that, CO 2 /CH 4

selectivity (which is of the paramount industrial importance) reaches up to 42.3 after 10-

day silver treatment which is superior to both neat Matrimid and Ag+-untreated

nanocomposite membranes. The following hypotheses can be drawn in order to explain

the enhanced gas separation performance of membranes after silver treatment.

134
Table 3.5. The effect of silver treatment on the permeability and selectivity of Matrimid-MgO nanocomposite membranes.

Polymer matrix- Permeability (Barrer) Ideal Selectivity


MgO loading (%)-
Retention time (day)
He H2 O2 N2 CH4 CO2 H2/N2 He/N2 CO2/ CH4 O2/N2 H2/ CO2
Matrimid-20%- 0 day 31.0 32.2 2.27 0.360 0.278 8.27 89.5 86.0 29.8 6.31 3.89
Matrimid-20%- 2 days 20.7 19.8 1.28 0.183 0.130 5.13 108.2 112.8 39.4 7.00 3.86
Matrimid-20%- 5 days 23.0 22.5 1.26 0.173 0.121 5.05 129.9 132.8 41.7 7.28 4.45
Matrimid-20%- 10 days 25.9 22.7 1.17 0.155 0.102 4.31 146.5 166.9 42.3 7.54 5.26
1 Barrer = 1*10-10 cm3(STP) cm/cm2 s cm Hg

Testing Temperature 35oC

135
As explained in previous sections, the average pore size of the MgO nanoparticles is far

larger than the size range of gas molecules and thus results in the low selectivity

performance for the gas pairs. Therefore, this phenomenon exterminates the notion of

exploiting advantageous affinity of MgO toward CO 2 for gas separation. Revitalizing the

advantageous properties of MgO for gas separation would not be possible unless some

strategies be adopted to tailor the porous structure of MgO nanoparticles and diminish the

pore sizes to the order of gas molecules. It seems that the surface coating of MgO with

multilayer silver ions provides this opportunity to perform this modification. Although,

the addition of silver ions may activate the facilitate transport role of MgO; however the

narrowed pathways for the transport of gas molecules caused by Ag layers also brings

about the size exclusion mechanism in the membranes. Therefore, one can notice from

the permeability results that positive role of facilitated transport in permeability is highly

influenced by the negative role of size exclusion mechanism. The constant decrease in

membrane permeability and increase in He/N 2 , H 2 /N 2 , O 2 /N 2 and CO 2 /CH 4 selectivity

by prolonging the treatment step clearly support the fact that the sizes of available

pathways for gas transport may be reduced to the level of gas molecule dimensions, thus

distinguishing the size difference of these gas pairs. Furthermore, presence of MgO

facilitates the transport of CO 2 while size exclusion provides more resistance for the

transport of CH 4 due to its larger kinetic diameter compared to CO 2 . As a result, the

combination of these two mechanisms results in advantageous properties of silver treated

nanocomposite membranes for CO 2 /CH 4 separation.

136
It is worthwhile to note that the presence of Ag ions inside the membranes may

also affect the transport properties of gas molecules within the membrane. However,

despite the abundance of studies on the interaction between olefins with silver ions,

interestingly, little work has been done about the interaction of silver ions with CO 2 and

other gases used in this study. In agreement with other reports [47], it can be concluded

that the gas transport properties of silver doped nanocomposite membranes is governed

by the combination of solution-diffusion (in the polymeric medium), size exclusion (in

the silver coated porous structure of MgO), and facilitated transport mechanisms (by

interaction with MgO and/or Ag). However, the latter is only effective for the transport of

CO 2 .

The sorption isotherms of CO 2 gas in the nanocomposite membranes before and

after Ag treatment are shown in Figure 3.7. It can be seen that sorption behaviors abide

the dual-mode sorption model and the sorption capacity of nanocomposite membrane has

been decreased after doping with silver ions. It is known that MgO has the capability to

presents both reversible and irreversible affinity toward CO 2 through interaction with

oxygen sites of MgO. The irreversible adsorption of CO 2 onto the MgO sites can reduce

the permeability of CO 2 . It was shown that upon doping, Ag ions interact strongly with

oxygen sites. Therefore, it seems there remain far less available sites for CO 2 molecules

to be adsorbed onto the MgO surface since the surface sites of MgO is already occupied

with Ag ions. As a result, it is reasonable to observe less sorption values for CO 2 in

Mat+MgO+Ag membranes compared to the sample without Ag and the results already

137
corroborate this fact by showing the decline in the sorption values of

Matrimid+MgO+Ag.

45

C [cm3(STP)/cm3(membrane)] 40

35

30

25

20

15

10
Mat+MgO(20%)
5 Mat+MgO(20%)+Ag(10 days)

0
0.0 5.0 10.0 15.0
P [atm]

Figure 3.7. CO 2 sorption isotherms for the nanocomposite membranes before and after

silver doping.

The transport of CO 2 may be affected by MgO and/or Ag. The enhanced transport

of CO 2 due to the presence of Ag proposed in this study is in accordance with the recent

study [47] in our group indicating the facilitated transport of CO 2 molecules. Considering

the case of MgO, it was indicated that MgO has the ability to interact electrostatically

over multilayers, so there might be the possibility that MgO nanoparticles interact with

CO 2 molecules passing over the Ag layer and enhance the transport of CO 2 molecules.

Therefore, one may also consider that the presence of Ag coated MgO nanoparticles

throughout the membrane may provide a driving force for the transport of CO 2 gas

138
molecules. Although it is difficult to quantitatively provide a measure for the phenomena,

but we deem it is better not to totally rule out the possibility of presence of facilitated

transport caused by MgO and/or Ag along with other obvious separation mechanisms

introduced.

Structural analysis of the silver-treated membranes by XRD validated our

hypothesis about the microstructural changes and role of size exclusion mechanism after

silver treatment. Figure 3.8. shows the XRD spectra of membranes before and after silver

treatment. It can be seen that almost no peak shift was occurred in the spectra of the

3000

2.36 Å Pure Matrimid


Matrimid+MgO(20%)
2500
Matrimid+MgO(20%)+2 days silver treatment
Matrimid+MgO(20%)+5 days silver treatment
Matrimid+MgO(20%)+10 days silver treatment

2000
Intensity

1500

3.78 Å 2.04 Å

1.44 Å
1.23 Å
1000
1.179 Å

500

0
5 10 15 20 25 30 35 40 45 50 55 60 65 70 75 80 85
2 Theta (degree)

Figure 3.8. The XRD spectra of the membranes before and after silver treatment with

different profiles.

139
polymer after silver treatment. This implies that the d-spacing of the polymer matrix has

almost remained unchanged. However, analysis of the spectra shows that some new

peaks have been appeared by incorporation of nanoparticles which can be ascribed to the

crystalline structure of MgO particles. These peaks are distributed in the region between

1.179 to 2.36 Å. It is clearly depicted that the silver treatment has further altered the

morphology of the membranes. In fact, after a 2-day silver treatment, the intensity of

these peaks tended to increase and the maximum values were obtained for the membranes

with 10-day treatment period. Since the location of the peaks remains unchanged, the

intensity increase observed for the peaks may indicate the growth of the crystalline

structures at the vicinity of MgO nanoparticles. It also seems that attractive force between

silver cations and oxygen anions promote the integrity of the whole structure around the

nanoparticles. These results further corroborate our hypothesis on the development of

finer pathway for the gas transport through the membranes which brings about the

stronger size exclusion effect to the system.

3.4 CONCLUSIONS

Flat nanocomposite membranes were fabricated by the introduction of magnesium

oxide nanoparticles into the Matrimid polymer matrix. The increment in the gas

permeability as a function of particle content and the drop in gas selectivity reflected the

fact that the average pore size of MgO (30 Å) is much larger than the kinetic diameters of

gas molecules. The heat treatment at high temperatures followed by quenching resulted in

the highest gas permeability while retaining selectivity. This was ascribed to the increase

140
in fraction of free volumes in the polymeric medium of membrane structure. Silver

treatment was carried out in order to tailor the performance of membranes. XPS and EDX

analyses revealed that in contrast to the polymer, it is MgO nanoparticles that attract the

silver ions, thus resulting in the penetration of Ag cations into nanocomposite membranes

followed by adsorption onto the surface of MgO nanoparticles. The amount of adsorption

was monotonically increased with the retention time. It is found that silver treated

nanocomposite membranes exhibit better gas separation performance for all gas pairs and

H 2 /N 2 and CO 2 /CH 4 selectivity of nanocomposite membranes after 10-day silver

treatment were about 146.5 and 42.3, respectively. The role of silver ions was recognized

to be covering the porous structure and control of the pore size of MgO nanoparticles. As

a result, in addition to the solution-diffusion mechanism, the size exclusion mechanism at

MgO-Ag+ structures as well as the presence of facilitated transport due to special

interaction between CO 2 and MgO were suggested by experimental analyses to be the

governing mechanisms in the superior separation performance of the newly developed

Matrimid-MgO-Ag+ nanocomposite membranes.

3.5 REFERENCES

[1] L.M. Robeson, Correlation of separation factor versus permeability for polymeric

membranes, J. Membr. Sci. 62 (1991) 165.

[2] G.W. Peng, F. Qiu, V.V. Ginzburg, D. Jasnow, A.C. Balazs, Forming supramolecular

networks from nanoscale rods in binary, phase-separating mixtures, Science 288

(2000) 1802

141
[3] T. J. Pinnavaia, Intercalated Clay Catalysts, Science 220 (1983) 365

[4] Q. Hu, E. Marand, S. Dhingra, D. Fritsch, J. Wen, G. Wilkes, Poly(amide-

imide)/TiO2 nano-composite gas separation membranes: Fabrication and

characterization, J. Membr. Sci. 135 (1997) 65.

[5] H.J.C. Te Hennepe, Zeolite Filled Polymeric Membranes: A New Concept in

Separation Science, Thesis, University of Twente, 1988.

[6] M. Jia, K.V. Peinemann, R.D. Behling, Molecular sieving effect of zeolite filled

silicone rubber membranes, J. Membr. Sci. 57 (1991) 289.

[7] J. M. Duval, B. Folkers, M. H. V. Mulder, G. Desgrandchamps, C.A. Smolders,

Adsorbent-filled membranes for gas separation. Part 1. Improvement of the gas

separation properties of polymeric membranes by incorporation of microporous

adsorbents, J. Membr. Sci. 80 (1993) 189.

[8] D.R. Paul, D.R. Kemp, The diffusion time lag in polymer membranes containing

adsorptive fillers, J. Polym. Sci. Symp. 41 (1973) 79.

[9] R. Mahajan, W.J. Koros, Factors controlling successful formation of mixed-matrix

gas separation materials, Ind. Eng. Chem. Res. 39 (2000) 2692.

[10] L. Y. Jiang, T. S. Chung¸ S. Kulprathipanja, Fabrication of mixed matrix hollow

fibers with intimate polymer-zeolite interface for gas separation, AIChE J. 52

(2006) 2898.

[11] R. Mahajan, R. Burns, M. Schaeffer, W.J. Koros, Challenges in forming successful

mixed matrix membranes with rigid polymeric materials, J. App. Polym. Sci. 86

(2002) 881.

142
[12] L. Y. Jiang, T.S. Chung¸ S. Kulprathipanja, An investigation to revitalize the

separation performance of hollow fibers with a thin mixed matrix composite skin

for gas separation, J. Membr. Sci. 276 (2006) 113.

[13] Ş. B. Tantekin-Ersolmaz, L. Şenorkyan, N. Kalaonra, M. Tatlıer, A. Erdem-

Şenatalar, n-Pentane/i-pentane separation by using zeolite-PDMS mixed matrix

membranes, J. Membr. Sci. 189 (2001) 59.

[14] T. M. Gür, Permselectivity of zeolite filled polysulfone gas separation membranes, J.

Membr. Sci. 93 (1994) 283.

[15] Y. Li, T. S. Chung, C. Cao, S. Kulprathipanja, The effects of polymer chain

rigidification, zeolite pore size and pore blockage on polyethersulfone (PES)-zeolite

A mixed matrix membranes, J. Membr. Sci. 260 (2005) 45.

[16] Y. Li, H. M. H. Guan, T. S. Chung, S. Kulprathipanja, Effects of novel silane

modification of zeolite surface on polymer chain rigidification and partial pore

blockage in polyethersulfone (PES)–zeolite A mixed matrix membranes, J. Membr.

Sci. 275 (2006) 17.

[17] H. Suda, K. Haraya, Gas permeation throughmicropores of carbon molecular sieve

membranes derived from Kapton polyimide, J. Phys. Chem. B 101 (1997) 3988.

[18] J. Petersen, M. Matsuda, K. Haraya, Capillary carbon molecular sieve membranes

derived from Kapton for high temperature gas separation, J. Membr. Sci.131 (1997)

85.

[19] P. S. Tin, Y.C. Xiao, T.S. Chung, Polyimide-carbonized membranes for gas

separation: structural, composition and morphological control of precursors, Sep. &

Pur. Rev., 35 (2006) 285.

143
[20] P. S. Tin, T. S. Chung, A. J. Hill, Advanced fabrication of carbon molecular sieve

membranes by nonsolvent pretreatment of precursor polymers, Ind. Eng. Chem.

Res. 43 (2004) 6476.

[21] Y. K. Kim, H. B. Park, Y. M. Lee, Gas separation properties of carbon molecular

sieve membranes derived from polyimide/polyvinylpyrrolidone blends: effect of the

molecular weight of polyvinylpyrrolidone, J. Membr. Sci. 251 (2005) 159.

[22] Y. K. Kim, H. B. Park, Y. M. Lee, Preparation and characterization of carbon

molecular sieve membranes derived from BTDA–ODA polyimide and their gas

separation properties, J. Membr. Sci. 255 (2005) 265.

[23] D. Q. Vu, W. J. Koros, S. J. Miller, Mixed matrix membranes using carbon

molecular sieves I. Preparation and experimental results, J. Membr. Sci. 211 (2003)

311.

[24] M. Moaddeb, W. J. Koros, Gas transport properties of thin polymeric membranes in

the presence of silicon dioxide particles, J. Membr. Sci. 125 (1997) 143.

[25] H. Lin, S. Matteucci, B. D. Freeman, S. Kalakkunnath, D. S. Kalika, Novel

membrane materials for CO 2 removal from mixtures with H 2 , Preprints of

Symposia-American Chemical Society, Division of Fuel Chemistry 50 (2005) 617.

[26] S. Hess, C. S.-Bickel, R.N. Lichtenthaler, Propene/propane separation with

copolyimide membranes containing silver ions, J. Membr. Sci. 275 (2006) 52.

[27] E.F. Steigehnenn and R.D. Hughes, Process for separation of unsaturated

hydrocarbons, US Patent 3,758,603 (1973).

[28] E.F. Steigehuann and R.D. Hughes, Semi-permeable membrane compositions based

on blends of polyamides and polyvinyl alcohols, US Patent 3,980,605 (1976).

144
[29] S. A. Chai, R. R. Beitle, and M. R. Coleman, Facilitated Transport Metal Affinity

Membranes, Int. J. Biochromatography, 2 (1996) 125.

[30] J. D. Way, R. D. Noble, D. L. Reed, G. M. Ginley, L. A. Jan, Facilitated transport of

CO 2 in ion exchange membranes, AIChE J. 33 (1987) 480.

[31] S. W. Kang, J. H. Kim, K. Char, J. Won, Y. S. Kang, Nanocomposite silver polymer

electrolytes as facilitated olefin transport membranes, J. Membr. Sci. 285 (2006)

102.

[32] J. H. Kim, D. H. Lee, J. Won, H. Jinnai, Y. S. Kang, The structural transitions of π-

complexes of poly(styrene-b-butadiene-b-styrene) block copolymers with silver

salts and their relation to facilitated olefin transport, J. Membr. Sci. 281 (2006) 369.

[33] G. Pacchioni, Physisorbed and chemisorbed CO2 at surface and step sites of the

MgO(100) surface, Surf. Sci., 281 (1993) 207.

[34] K. J. Klabunde, A. Khaleel, D. Park, Overlayer of iron oxide on nanoscale

magnesium oxide crystallites. Chlorocarbon destruction by catalytic solid-state

ion/ion exchange at 400 C, High Temp. Mater. Sci. 33 (1995) 99.

[35] T.H. Kim, W.J. Koros, G.R. Husk and K.C. O'Brian, Relationship between gas

separation properties and chemical structure in a series of aromatic polyimides, J.

Membr. Sci., 37 (1988) 45.

[36] L. Y. Jiang, T.S. Chung, D.F. Li, C. Cao, S. Kulprathipanja, Fabrication of

Matrimid/ polyethersulfone dual-layer hollow fiber membranes for gas separation,

J. Membr. Sci., 240 (2004) 91.

145
[37] P.S. Tin, T.S. Chung, Y. Liu, R. Wang, S.L. Liu, K.P. Pramoda, Effects of cross-

linking modification on gas separation performance of Matrimid membranes, J.

Membr. Sci., 225 (2003) 77.

[38] W.H. Lin, R.H. Vora, T.S. Chung, Gas transport properties of 6FDA-durene/1,4-

phenylenediamine (pPDA) copolyimides, J. Polym. Sci. Part B: Polym. Phys., 38

(2000) 2703.

[39] Y. Tsujita, Gas sorption and permeation of glassy polymers with microvoids, Prog.

Polym. Sci. 28 (2003) 1377.

[40] C. Qin, L. S. Sremaniak, J. L. Whitten, CO adsorption on Ag(100) and

Ag/MgO(100), J. Phys. Chem. B, 110 (2006) 11272.

[41] M. Sterrer, T. Risse, H.-J. Freund, CO adsorption on the surface of MgO(001) thin

films, Appl. Catalysis A: General, 307 (2006) 58.

[42] H. Guan, P. Wang, H. Wang, B. Zhao, Y. Zhu, Y. Xie, Preparation of nanometer

magnesia with high surface area and study on the influencing factors of the

preparation process, Acta physico-chemica sinica, 22 (2006) 804.

[43] H. X. Liu, H. L. Zhang, H. L. Ren, S. X. Ouyang, R. Z. Yuan, The influence of the

cluster models on the study of electronic structure of MgO/Ag interface, Ceram. Int.

22 (1996) 79.

[44] M. Shibata, Porous silica modified by phenyltetraethoxysilane, Maku 19 (1994) 332.

[45] O. Robach, G. Renaud, A. Barbier, Structure and morphology of the Ag/MgO(001)

interface during in situ growth at room temperature, Phys. Rev. B, 60 (1999) 5858.

[46] A. M. Flank, R. Delaunay, P. Lagarde, M. Pompa, J. Jupille, Epitaxial silver layer at

the MgO(100) surface, Phys. Rev. B, 53 (1996) 1737.

146
[47] Y. Li, T.S. Chung, S. Kulprathipanja, Novel Ag+-zeolite/polymer mixed matrix

membranes with a high CO2/CH4 selectivity, AIChE J., 53 (2007) 610.

147
CHAPTER FOUR

HYDROGEN SEPARATION AND PURIFICATION IN MEMBRANES OF

MISCIBLE POLYMER BLENDS WITH INTERPENETRATION NETWORKS

4.1 INTRODUCTION

Hydrogen is anticipated to play an inevitable role as one of the promising sources

of energy in the future. According to the latest statistics, the U.S. annual production of

hydrogen is appraised to about 9 million tons [1]. The major proportion of produced

hydrogen is used to supply the refinery and fertilizer industries. The constant increasing

demand for highly purified hydrogen has resulted in attracting more attentions for

extensive research studies on advancement of the technologies for generation of

hydrogen with higher efficiency and lower production cost.

At present, steam reforming of hydrocarbons is predominantly the established

technology for generation of hydrogen and it supplies about half of the world’s demand

for this viable gas. In this process, a stream of natural gas mainly composed of methane

undergoes a water gas shift reaction. The yield product has to be further processed in a

separation process in order to remove unwanted carbon dioxide and other by-products

from the mixtures with hydrogen. Currently, commercially available separation processes

148
for this purpose are pressure swing adsorption, amine absorption and absorption using

aqueous solution of potassium carbonate which are highly energy intensive [2-4].

Membranes have received global attention as a promising technology for gas

separation and purification. This mainly stems from the small footprint, reliability,

considerable high energy efficiency and low capital cost offered in membranes compared

to other conventional separation processes [5,6]. It is also worth noting that hydrogen

recovery has been among the first commercial applications of membranes in the field of

gas separation [7]. The major application of membranes for hydrogen separation includes

hydrogen purifications, fuel cell technology and membrane reactor processes [8].

Recovery of hydrogen from hydrocarbon streams, recovery of hydrogen from ammonia

streams, adjusting the composition of hydrogen in synthetic gases are among the major

industrial and petrochemical applications of hydrogen purification [9]. Interestingly,

membrane separation systems have good potentials for integration with hydrogen

generation equipments that permits the processes to be entirely carried out in a single

processing unit.

In recent years, research directions have been on both inorganic and organic

materials to develop membranes with desirable hydrogen separation properties. There are

different approaches for the use of inorganic membranes in hydrogen separation. One is

based on a dense metal layer deposited on a ceramic support. In this category, palladium-

based membranes are more attractive and generally exhibit high hydrogen permeability

and selectivity [10]. Zhang et al. [11] studied the hydrogen permeation behavior of dense

149
palladium membranes prepared by microfabrication technique. According to their study,

this technique offers high durability to membrane for operation in elevated temperatures.

The application of other metals also has also been examined. For example, Ernst et al. [8]

successfully fabricated a nickel/ceramic composite membrane which possesses high

permeability and selectivity to hydrogen. However, inorganic membranes are fraught

with complexities in fabrication, difficulties in prediction of membrane permeation

properties and generally require high operating costs [12].

On the other hand, the advantageous properties of polymeric membranes like ease

of processing and appropriate robustness may dominate the application of organic

membranes over inorganic counterparts. However, the main drawback of polymeric

membranes for hydrogen separation compared to inorganic is the operating temperature

range that can be used. The coincidence of unfavorable diffusivity selectivity and

solubility selectivity for selected gas pairs like H 2 and CO 2 has also restrained the

application of polymeric membranes for separation of such molecules. This mainly stems

from the negligible difference in kinetic diameters of gas molecules. Swelling of the

interstitial spaces between polymer chains due to penetrant-induced plasticization may

also play a part which restrains the performance of separation based on size exclusion

mechanism [13]. It is well shown that CO 2 induced plasticization severely deteriorates

the separation properties of membranes prepared from glassy polymers including

polyimides [14]. The presence of challenges has resulted in limited number of reports on

the development of polymeric membranes for hydrogen separation. In a recent research

study in our group [15] polymeric membranes were successfully fabricated with high

150
selectivity for H 2 /CO 2 . It was observed that chemical surface modification of 6FDA-

durene offers membranes with gas selectivity of over 100 for H 2 /CO 2 in pure gas tests

and about 42 in mixed gas tests.

The objective of this study is to fabricate high performance polymeric membranes

suitable for separation and purification of hydrogen. Dense flat membranes were

fabricated using the solution blending technique using Matrimid and PBI. Matrimid is a

polyimide with high gas permeability and selectivity, high thermal stability and solvent

resistance [16-17]. Similarly, PBI is a high performance polymer possessing high glass

transition temperature, outstanding thermal stability and chemical resistance that makes it

suitable candidate for advanced technologies like fuel cells [18-20]. The gas permeation

characteristics of Matrimid /polybenzimidazole (PBI) homogenous blend membranes are

evaluated. The effects of composition on miscibility, microstructure and gas separation

performance of membranes are also investigated. Gas separation performance of

membranes was further ameliorated through chemical cross-linking modification of blend

constituents using distinct types of cross-linking agents. To our best knowledge, this is

the first study focusing the exploitation of unique properties of blending high

performance polymers of Matrimid and PBI for the development of specialty membranes

for hydrogen separation and purification.

4.2 EXPERIMENTAL

4.2.1 Materials

151
The polyimide (Matrimid5218) (3,3’,4,4’-benzophenone tetracarboxylic

dianhydride and diamino-phenylindane) was purchased from Vantico Inc., (Luxemburg)

and poly [2,2’-(1,3-phenylene)-5,5’-bibenzimidazole] (PBI) was obtained from Aldrich

Chemical Company Inc. (Milwaukee, USA). Both polymer powders were dried for

overnight at 120oC under vacuum prior to use. The chemical structure of polymers is

shown in Figure 4.1. N-methyl-2-pyrolidone (NMP) and methanol were supplied by

Merck and used as solvent for polymers. Cross-linking agents used in this study were p-

xylenediamine and p-xylenedichloride which were supplied by Merck, (Germany) and

TCI Inc., (Japan), respectively. Ethanol (analytical grade) was supplied by Merck and

was used for density measurements.

Matrimid

PBI

Donor: C=O (Matrimid)


Acceptor: N-H (BPI)

Figure 4.1. Schematic representation of the hydrogen bonding interaction between

functional groups of Matrimid and PBI.

152
4.2.2 Preparation of dense flat membranes

Polymer solutions (2 wt.% polymer/ 98 wt.% NMP) with various compositions of

25/75, 50/50, 75/25 wt% were prepared from Matrimid and PBI. Firstly, PBI with

stipulated quantity was dissolved in NMP at 150°C using a magnetic stirrer and it took

about 2 days to obtain a complete dissolution. The polymer solution was filtered using a 2

μm filter to eliminate any undissolved polymer and dusts. Subsequently, Matrimid

powders were added and stirring was continued for few days to allow the complete

mixing of blend constituents. After degassing, polymer solutions were poured onto a

silicon wafer surrounded by a metal ring. The wafers were placed in a vacuum oven and

the temperature of oven was set to 100 °C to allow slow evaporation of NMP. This

process was continued for about five days. While still under vacuum, the temperature was

gradually increased at the rate of 20 °C/0.5 hr up to 250 °C and membranes were kept at

the final temperature for overnight to assure the residual solvent removal. The as-cast

membranes were collected after natural cooling down.

4.2.3 Chemical modification of membranes

Round-cut pieces of membranes underwent cross-linking using a simple dipping

method. A 10 wt.% solution of p-xylenediamine in methanol was prepared which was

aimed for cross-linking of the polyimide. On the other hand, a solution composed of 2

wt.% p-xylenedichloride in methanol was used to cross-link PBI components.

Membranes were immersed in the solutions for stipulated periods of 5 and 10 days. After

taking out, membranes were immersed in fresh methanol to wash out unreacted

molecules followed by drying in vacuum oven at 120 °C for overnight.

153
4.2.4 Characterization techniques

The miscibility of membrane constituents was examined using differential

scanning calorimetry (DSC). Glass transition temperature (T g ) of polymer films was

evaluated on a Perkin-Elmer, DSC-Pyris 1 calorimeter. Approximately 5-10 mg of

membrane was put in aluminum pans. Two consecutive scans from 50 to 450 °C at the

heating rate 10 °C/min were performed for each sample. The T g of samples was

determined as mid-point transition temperature of the second scan.

A standard displacement technique was employed to measure the density of blend

films [21]. Round pieces of films were punched and weighed in air ( wair ) and then in

ethanol ( wEtOH ), respectively. The density of polymer film was calculated using the

following equation:

wair
ρ polymer = ρ EtOH (4.1)
wair − wEtOH

where ρ EtOH is the density of ethanol. Five pieces of films were prepared for each

sample.

The structural properties and inter-chain spacing of membranes were analyzed

qualitatively on an X’Pert PRO-MRD wide-angle X-ray diffractometer (XRD) from

PANalytical. The measurements were carried out by using Cu Kα radiation with a

wavelength of λ = 1.54 Å at room temperature. The scan range was from 4 to 70


˚ with a

step increment of 0.04̊/s. Average d-spacing was determined based on the Bragg’s law,

viz.

154
nλ = 2d sin θ (4.2)

where n is an integral number (1, 2, 3, ...), λ denotes the X-ray wavelength, d stands for

the intersegmental spacing between two polymer chains and θ indicates the diffraction

angle.

Molecular simulation was performed using the Amorphous Cell procedure

developed by Material Studio software packages, Accelrys Inc. The most stable

energy-minimized configurations of polymer chains were simulated to qualitatively

examine the possible chain morphology for crosslinking.

4.2.5 Gas permeability measurements

The gas permeation properties of membranes were determined by a variable-

pressure constant-volume method. Detailed experimental design and procedures can be

found elsewhere [22]. The rate of pressure increase (dp/dt) at steady state was used for

the calculation of gas permeability with the following relationship:

273.15 × 1010 Vl  dp 
P=   (4.3)
760 AT (( P0 × 76) / 14.7)  dt 

where P is the gas permeability of a membrane in Barrer (1Barrer = 1×10-10 cm3 (STP)-

cm/cm2.sec.cmHg), V is the volume of the down-stream chamber (cm3), A refers to

effective area of the membrane (cm2), l is the membrane thickness (cm), T is the

operating temperature (K) and the feed gas pressure in the up-stream is given by P 0 in

psia.

The pure gas permeability was obtained in a sequence of H 2 , N 2 , CH 4 and CO 2 at

35 °C for each piece of membrane. The H 2 permeability test was conducted at 3.5 atm,

155
while the testing pressure for other gases was 10 atm. The ideal selectivity of a membrane

for gas A to gas B was evaluated as follows:

PA
α A/ B = (4.4)
PB

The apparent diffusion coefficient (D app ) is obtained by using time-lag (θ) method as

follows:

L2
Dapp = (4.5)

where in this expression, θ is the diffusion time-lag. The apparent solubility coefficient

(S app ) was subsequently calculated through the following expression:

P
S app = (4.6)
Dapp

4.3 RESULTS AND DISCUSSION

4.3.1 The miscibility and other characteristics of blends

Extensive studies have given countenance to the marked role of constituents’

compatibility on the phase morphology and consequently gas separation performance of

blend membranes [23]. Hence, it is well accepted that miscibility analysis should be set

as a prerequisite in the design of superior gas separation membranes. The miscibility of

polyimides and PBI has been the subject of few studies [18,24-26]. All these reports

unanimously corroborate the miscibility of these polymer pairs. However, referring to

those reports, either the analyses have been carried out in a limited range or the grade of

materials is found to be different from what was used in this study. Therefore, the

156
miscibility of as-cast blend membranes was examined over the various ranges of

compositions. Obtaining clear and transparent films and the presence of a single T g in

calorimetry thermograms have been verified to be appropriate indicatives of miscibility

in the molecular level. All the films prepared from the blends of Matrimid and PBI were

transparent and homogenous and no symptom of phase separation was detected by visual

observation. The DSC thermograms of the blend membranes with various compositions

are depicted in Fig. 2. These results also confirm the miscibility of blends by showing the

presence of composition dependent single T g that all fall in the midst of each individual

components’ T g value. In addition, a positive shift in T g was observed with an increase

in PBI composition which was in fact due to the higher T g of PBI.


Heat Flow (mW)

b
c
d
e

300 330 360 390 420 450


Temperature (C)

Figure 4.2. DSC thermograms of the polymer blend membranes and individual

constituents. (a) 0, (b) 25, (c) 50, (d) 75, (e) 100 wt% PBI composition.

The glass transition temperature of polymer blends was calculated theoretically by using

the Fox equation which is expressed as follows:


157
1 W1 W2
= + (4.7)
Tg Tg Tg
1 2

In this Eq., T g1 and T g2 are glass transition temperatures (K) of individual polymers while

W1 and W 2 indicate mass fractions of each component in the blend.

The T g -composition curve of blend system is shown in Figure 4.3. It can be seen

that the connection of experimental data points forms a sigmoidal curve. The comparison

of experimental and theoretical results indicate that the Fox equation can almost

successfully predict the T g values for the blends rich in Matrimid. However, the trend is

followed by a negative deviation in prediction of T g as the composition of PBI is further

increased. Interestingly, the deviation turns to the positive prediction once the

composition of PBI in the blend system exceeds 65 wt.%. It should be noted that

deviation from the theoretical results is a common phenomenon which is seen in behavior

of majority of polymer blends and is generally a reflection of the specific interactions

between the components. The nature of miscibility of Matrimid/PBI is believed to be the

intense hydrogen bonding between N-H group of PBI and C=O group of Martimid which

provides a specific interaction for intermolecular compatibility [18, 24-25]. This strong

interaction has been confirmed by various techniques including FTIR where Karasz et al.

[27] observed shifts in N-H bond of PBI and carbonyl group of Matrimid. The schematic

representation of this interaction is shown using dashed lines in Figure 4.1.

4.3.2 Gas separation characteristics of blend membranes

Basically, potential use of polymeric blends for the transport and separation of

gaseous systems entails a dexterous control over the phase morphology and interfacial

158
interaction of blend components [28-29]. The confirmed miscibility may satisfy this

prerequisite criterion for the use of Matrimid/PBI blends for gas separation. Table 4.1.

tabulates the results of gas permeability measurements and ideal selectivity of blend

membranes. The results denote a general decline in gas permeability with an increase in

PBI concentration.

440

420 Experimental
Theoretical-Fox Eq.
400

380
Tg

360

340

320

300
0 20 40 60 80 100
Matrim id Com position (%)

Figure 4.3. Composition dependency of T g and comparison with theoretical values

predicted by Fox equation.

It can be found that the decrease in gas permeability of membranes commensurate

with kinetic diameter of gas molecules. In other words, gases with larger kinetic diameter

experience more decrease in permeability with the increase of PBI content. The

instrumental consequence of this behavior was the melioration in selectivity of selected

gas pairs. For example, data in Table 1 show that the selectivity of H 2 /CO 2 and H 2 /N 2

159
increases by about 1.5 fold and reaches to about 9.43 and 260.47, respectively.

Interestingly, quite similar improving trend was observed in the selectivity of CO 2 /CH 4

which there is a minor difference in size of these molecules and ideal selectivity of this

gas pair was reached to about 60. It is worthwhile to mention that the separation of this

gas pair is of paramount importance in processing of natural gas and considering the large

amount of production, any small improvement in efficiency of the process could result in

substantial cost reduction in processing of natural gas.

Table 4.1. The permeability and ideal gas selectivity of polymer blend membranes with

various compositions.

Membrane Permeability (Barrer) Ideal Selectivity


constituents
H2 N2 CH4 CO2 H2/N2 H2/ CO2 CO2/ CH4
Matrimid®5218 27.16 0.280 0.210 7.00 97.00 3.88 33.33
Matrimid/PBI(75/25%) 19.72 0.163 0.130 4.19 120.98 4.07 32.23
Matrimid/PBI(50/50%) 13.06 0.072 0.045 2.16 181.38 6.05 48.00
Matrimid/PBI(25/75%) 5.47 0.021 0.0097 0.58 260.47 9.43 59.79
PBIa 0.6 0.0048 0.0018 0.16 125 3.75 88.88

1 Barrer = 1*10-10 cm3(STP) cm/cm2 s cm Hg


Testing Temperature: 35ºC
a From reference [34]

The gas selectivity of a membrane for a gas pair is the ratio of their individual

permeability coefficients, while the gas permeability (P) is a product of solubility (S) and

diffusivity (D) coefficients which are highly dependent on the nature of penetrant and the

membrane material and morphology. These coefficients jointly control the overall

permeability of a membrane. Table 4.2 provides the calculated values of the diffusivity

160
and solubility coefficients of the membranes developed in this study. It seems that the

interrelation between the size of gas molecules and changes in permeability as well as the

presence of interaction between the components fortify the idea that the possible reason

for changes in transport properties of membranes might have been the effect of PBI on

membranes morphology in a molecular level. Based on the solution-diffusion

mechanism, the present experimental results imply that the incorporation of PBI results in

the enhanced stiffness and decreased spacing between polymer chains with little changes

in solubility selectivity. In addition, it is well documented that free volume is a

determining factor in gas transport properties of a membrane [30-31].

Table 4.2. Gas diffusion coefficients and solubility coefficients of polyimide, blend

membranes and modified blend membranes.

D (×10-8 cm2/s) S (×10-2 cm3(STP)/cm3 cmHg)


Membrane constituents
H2 N2 CH4 CO2 H2 N2 CH4 CO2
Matrimid®5218 174.0 0.413 0.093 0.897 0.16 0.68 2.26 7.80
Matrimid/PBI (75/25%) 136.0 0.312 0.061 0.572 0.15 0.52 2.12 7.32
Matrimid/PBI (50/50%) 94.0 0.144 0.022 0.311 0.14 0.50 2.03 6.95
Matrimid/PBI (25/75%) 44.5 0.044 0.005 0.088 0.12 0.48 1.85 6.62
Matrimid/PBI (25/75%)-5 days PXDC 43.8 0.040 0.009 0.070 0.12 0.47 1.87 6.50
Matrimid/PBI (25/75%)-10 days PXDC 33.9 0.030 0.008 0.048 0.12 0.46 1.86 6.41
Matrimid/PBI (25/75%)-5 days PXDA 34.4 0.032 0.003 0.034 0.12 0.47 1.82 6.13
Matrimid/PBI (25/75%)-10 days PXDA 32.1 0.029 0.002 0.023 0.11 0.46 1.81 6.22

Therefore, the free volume analysis was performed in order to investigate the effect of

incorporation of PBI on microstructure and how the strong interaction between blend

components can alter the transport properties of membranes.

161
The fractional free volume (FFV) calculations were based on following equations:

Vsp − V0
FFV = (4.8)
Vsp

V0 = 1.3VW (4.9)

Vsp = 1/ ρ (4.10)

where Vsp is the specific volume of polymer film (cm3/g) and V0 is the volume occupied

by the chains (cm3/g). The values of Van der Waals volumes (V W ) for individual

polymers were obtained by applying group contribution method [32]. According to Bondi

[33], Eq. (9) provides a good approximation for the evaluation of V0 based on the Van

der Waals volume. Moreover, ρ represents the density (g/cm3) of polymer films which

was measured experimentally. The corresponding values of V W and M for blends were

calculated using the mixing rules. The results of measurements and calculations are

shown in Table 4.3. Data in this table indicate that Matrimid and PBI possess the

maximum (0.268) and the minimum (0.116) fractional free volumes, respectively. These

values are in good agreement with other reports [14, 34]. The difference in FFV of these

materials possibly stems from the difference in chemical structure [30]. It can be seen

from Figure 4.1. that Matrimid and PBI are quite similar in structure of backbone and

repeating unit. However, it is expected that the presence of few side oxygen groups as

well as large methyl groups in the structure of Matrimid does not allow the polymer

chains to come closer. Furthermore, molecular simulation is found to be a very useful

tool in analysis of the properties of membranes [35].

Application of molecular simulation study on the configuration and

conformations of polymer chains corroborated our hypothesis. As it is presented in Fig. 4,


162
clearly a better chain packing can be observed between the PBI macromolecules.

However, in contrast to PBI, there exists a relatively larger free volume formed in the

structure of Matrimid chains which mainly occurs in the junction of two repeating units

where the bulky methyl groups provide spatial hindrance against well chain packing.

Table 4.3. The effect of blend composition on the physical properties and fraction of free

volume of membranes.

Membrane constituents M (g/mol) ρ (g/cm3) Vw (cm3/mol) V0 (cm3/g) Vsp (cm3/g) FFV

Matrimid®5218 568.6 1.172 273.1 0.624 0.853 0.268

Matrimid/PBI (75/25%) 469.5 1.192 229.9 0.637 0.839 0.241

Matrimid/PBI (50/50%) 399.6 1.244 199.6 0.649 0.804 0.193

Matrimid/PBI (25/75%) 347.9 1.274 177.1 0.662 0.785 0.157

PBI 308 1.311 159.8 0.674 0.763 0.116

It should be noted that intensity of this hindrance is large enough that results in

the chain conformational rotation at the junction of repeating units in the structure of

Matrimid. Therefore, in consistence with data in Table 4.1., the FFV of membranes films

undergoes a stepwise decline by an increase in PBI composition. It should be noted that

in addition to this phenomenon, the formation of strong hydrogen bonding between blend

components which was not present in any of individual polymers may also contribute to

the diminishment of FFV by keeping the distinct polymer chains in the minimum

interstitial distances. This hydrogen bonding also restrains the polymer chain from free

rotation and mobility. The augmentation of glass transition temperature by increasing in

PBI composition is a sign of enhancing chain stiffness. The increase in density as a result

163
of increase in the PBI content in the blend membrane can be considered as another

evidence for amplified polymer chain packing throughout the microstructure of resultant

membranes.

(a)

(b)

Formation of free volume at the


junction of two repeating units

Figure 4.4. Analysis and comparison of the status of free volume in the structure of (a)

PBI and (b) Matrimid membranes using molecular simulation.

The XRD analysis was performed in order to obtain further information about the

effects of blending on microstructure and d-spacing of membranes. As shown in Figure

4.5, the indicative peak of Matrimid is a merger of two amorphous peaks with d-spacings

of 3.86 and 5.53 Å. An increase in PBI concentration gradually vanishes the peak of

Matirmid at 5.53 Å and finally the sample containing 75 wt.% PBI possesses almost a

mere peak at 3.86 Å. The shift in d-spacing indicates a shift of interstitial chain-chain

distance and free volume distribution toward a tighter and narrower structure. The results
164
of microstructural analysis of polymer blends indicate that the decrease in permeability

and enhancement in gas separation performance can be attributed to the contribution of

reduction in fractional free volumes and d-spacings, restrictions in mobility of polymer

chains and increase in chain packing density.

2000

1800
Matrimid/PBI (75/25)
1600 3.86 Å Matrimid/PBI (50/50)
Matrimid/PBI (25/75)
1400
Pure Matrimid
1200 5.53 Å
Intensity

1000

800

600

400

200

0
4 14 24 34 44 54 64

2 Theta (degree)

Figure 4.5. The XRD spectra of blends depicting the effect of composition on

microstructure of membranes.

These changes highly affect the diffusivity selectivity and stabilize the membrane

structure against CO 2 -induced swelling phenomena and consequently result in significant

increment in selectivity for H 2 /CO 2 , H 2 /N 2 , and CO 2 /CH 4 gas separation. In fact, one of

the main advantages of developed blend membranes in this study is exploiting benign

properties of PBI for gas separation that otherwise could not be fabricated as a free-

standing membrane due to its poor processibility and high degree of brittleness.

165
4.3.3 Effect of chemical modification on performance of membranes

Abundance of reports [16, 36-38] on the successful improvement of gas

separation performance of chemically modified polymeric membranes initiated the idea

of applying this technique in this study. The presence of five carboxyl group in repeating

unit of Matrimid and two N-H groups in repeating unit of PBI provides a great potential

that these polymers easily be involved in any modification reactions. Thus, Matimid/PBI

blend membranes with composition of (25/75 wt.%) which exhibited the superior

performance to other compositions were chemically modified using two different cross-

linking agents. The proposed mechanisms for cross-linking process are shown in Figures

4.6. and 4.7. Moreover, the most stable energy-minimized configurations for cross-linked

PBI and Matrimid are also illustrated in these figures. According to these schematics, in

the course of chemical modification, p-xylenedichloride molecules attack to the N-H

groups of PBI and p-xylenediamine molecules react with amide groups of Matrimid.

Obviously, it seems that the chemical structure of Matrimid contains a polar bond (ether

groups) which may result in lesser linearity of the polymer backbone as well as a lesser

degree of molecular packing compared to the PBI structure. The chemical modification

of polyimides [36-37] and specifically Matrimid [16] using p-xylenediamine have been

extensively investigated and verified. Similarly, the successfulness of chemical

modification of PBI with p-xylenedichloride through this mechanism was previously

confirmed in our group where Wang et al. [39], employed FTIR and XPS techniques and

confirmed the cross-linking of PBI membranes post-treated for up to 24 hrs. Therefore, it

is anticipated that the reaction starts with modification of functional groups at the

outermost surface of membranes which are easily accessible to the cross-linking agents.

166
(a)

(b) (c)
(c)

Figure 4.6. (a) Proposed mechanism for the chemical cross-linking modification of

Matrimid component of blend using p-xylene diamine. (b) Chemical structure of p-xylene

diamine, (c) possible chain morphology and configuration of p-xylene diamine cross-

linked Matrimid. (cross-linking agents are specified by circles).

In the meantime, swelling of membranes structure by the carrier (i.e., methanol) provides

the opportunity for the transport of the cross-linking agents to penetrate inside the

membranes and access the functional groups of polymers within the membrane structure.

The process of cross-linking at the internal structure of membrane is expected to decline

with the progress of superficial modification which seems to impede further inward

penetration of agents.

167
(a)

(b) (c)

Figure 4.7. (a) Proposed mechanism for the chemical cross-linking modification of PBI

component of blend using p-xylene dichloride. (b) Chemical structure of p-xylene

dichloride, (c) possible chain morphology and configuraion of p-xylene dichloride cross-

linked PBI. (cross-linking agents are specified by circles).

Gas permeability and ideal gas selectivity of modified membranes with p-

xylenedichloride and p-xylenediamine are presented in Tables 4.4. and 4.5., respectively.

It can be seen from the results that chemical modification of membranes clearly affects

the gas permeability of membranes and a gradual decline in gas permeability was resulted

by prolonging the immersion time. Interestingly, the extent of depletion in permeability

follows the similar trend of dependence to the size of gas molecules as it was observed in

permeability of blend films. Consequently, this characteristic phenomenon results in

variation in selectivity performance of membranes.

168
Table 4.4. The effect of chemical modification using p-xylenedichloride on gas

permeability and selectivity of blend membranes.

Membrane Permeability (Barrer) Ideal Selectivity


constituents

H2 N2 CH4 CO2 H2/N2 H2/ CO2 CO2/ CH4


Matrimid®5218 27.16 0.280 0.210 7.00 97.00 3.88 33.33
Matrimid+PBI(25/75) 5.47 0.021 0.0097 0.580 260.47 9.43 59.79

Matrimid+PBI(25/75)
5.34 0.019 0.0175 0.453 281.05 11.79 31.03
duration= 5 days

Matrimid+PBI(25/75)
4.04 0.014 0.0158 0.306 288.57 13.02 19.36
duration= 10 days

1 Barrer = 1*10-10 cm3(STP) cm/cm2 s cm Hg

Testing Temperature: 35ºC

Table 4.5. The effect of chemical modification using p-xylenediamine on gas

permeability and selectivity of blend membranes.

Membrane Permeability (Barrer) Ideal Selectivity


constituents
H2 N2 CH4 CO2 H2/N2 H2/ CO2 CO2/ CH4
Matrimid®5218 27.16 0.280 0.210 7.00 97.00 3.88 33.33
Matrimid+PBI(25/75) 5.47 0.021 0.0097 0.580 260.47 9.43 59.79

Matrimid+PBI(25/75)
4.09 0.0152 0.0046 0.209 269.1 19.56 45.43
duration= 5 days

Matrimid+PBI(25/75)
3.60 0.0132 0.0031 0.138 271.2 26.09 44.51
duration= 10 days

1 Barrer = 1*10-10 cm3(STP) cm/cm2 s cm Hg

Testing Temperature: 35ºC

According to Table 4.4., slight improvements were achieved in selectivity of

H 2 /N 2 and H 2 /CO 2 after cross-linking of PBI, but the selectivity of CO 2 /CH 4 underwent

a slight decrease. On the other hand, as shown in Table 4.5., a similar declining trend is

169
observed for the membranes upon cross-linking of Matrimid phase. Considering the

structural models provided in Figures 4.6. and 4.7. and the results in Table 4.2., the

opposite trends in the selectivity of gas pairs upon cross-linking can be attributed to the

combinatory effects the configuration changes in the polymeric structure of membranes

as well as the inherent molecular size difference of gas molecules.

A comparison of the results in Tables 4.4. and 4.5. for the corresponding

chemically modified membranes reveals the fact that treatment with p-xylenediamine has

spelled greater effect on the transport properties of membranes noticed from the level of

changes in permeability values. As a result, the H 2 /CO 2 selectivity of the resultant

membrane has reached up to about 26 after 10 days modification with p-xylenediamine

which is considerably higher than that of other samples and corresponding individual

components.

The effect of chemical cross-linking on gas permeability decay of membranes is

probably due to the tighter interstitial spaces among the chains and restriction in vibration

and mobility of polymer chains and side groups attached to the backbone. As a result, the

diffusion of penetrant gas molecules through the membrane is hindered and gas

permeability decreases. Therefore, in agreement with a similar report [37], the enhanced

selectivity performance of membranes is mainly achieved by the diffusivity selectivity of

membranes. This was confirmed by examination of the d-spacing between the polymer

chains through XRD analysis. According to Figure 4.8., upon cross linking, the

interstitial space between the polymer chains shifted from the original value of 3.86 Ǻ

170
2000

1800 Matrimid/PBI (25/75)


3.86 Å
3.36 Å X-link-PXDC-5 days
1600 X-link-PXDC-10 days
X-link-PXDA-5 days
1400
X-link-PXDA-10 days

1200
Intensity

1000

800

600

400

200

0
4 14 24 34 44 54 64

2 Theta (degree)

Figure 4.8. The XRD spectra of blend membranes cross-linked with p-xylene dichloride

(PXDC) and p-xylenediamine (PXDA) for different periods of time.

down to minimum 3.36Ǻ after cross -linking with p-xylenediamine for 10 days. The

order of shift is proportional to the period of modification and it is more pronounced for

samples modified with p-xylenediamine. Furthermore, the molecular modeling of the

systems show that cross-linking of Matrimid results in relatively more compact structure

compared to that of PBI. It seems that bond breaking reaction at the imidazole sites of

Matrimid during the cross-linking provides this flexibility for polymer chains to find

better configuration. However, this phenomenon is not observed for cross-linking of PBI

due to its relatively rigid polymer structure. As a result, the cross-linking agent seems to

act as a bridge between the lateral polymer chains.

171
Furthermore, it is anticipated that cross-linking of polymer chains offers a special

integrity and structural stability which protects the membrane from CO 2 induced swelling

and plasticization. One possibility is that more intense influence of cross-linking with p-

xylenediamine compared to p-xylenedichloride might be due to difference in

concentration of the cross-linking agents in the solution. This also could be ascribed to

the difference in activity of functional groups of constituents towards corresponding

cross-linking agents. In such a case, results may indicate that the cross-linking efficiency

of Matrimid is considerably higher than that of PBI.

It seems also worthwhile to consider this issue with respect to the size of the

crosslinking agents. Structural analysis of p-xylenedichloride and p-xylenediamine by

molecular simulation denotes the fact that while these chemical agents have similar

molecular width (around 3Ǻ measured by Material Studio software), there is only a minor

molecular length different between these two cross-linking agents, (7.9 and 8.0Ǻ, f or p-

xylenedichloride and p-xylenediamine respectively). In Fig.9, the structure and

configuration of the cross-linking agents are provided at the minimum energy. According

to this figure, it can be seen that except the benzene group that both cross-linking agents

have in common, the rest of the molecules have not much difference in the spatial

configuration. As a result, the difference in the size of these molecules is not large

enough to be considered into the account to justify about the penetration of these

molecules into the membrane structure.

172
(a) H H
C C H
N H N
H C C
H
H C C H

H H

(b) H H
C C H
H
Cl C C C C Cl
H C H
C
H H

Figure 4.9. Molecular simulation analysis of chemical structure of cross-linking agents

(a) p-xylenediamine (b) p-xylenedichloride

4.4 CONCLUSIONS

Blend membranes prepared from Matrimid and PBI are completely miscible in

the whole range of compositions in molecular level. This property is given to the blends

by strong hydrogen bonding interaction among the functional groups of components.

Therefore, incorporation of PBI resulted in improvements in gas separation performance

of Matrimid which was caused by increase in chain packing density and hindrance in

segmental mobility of polymer chains and its consequence effects on diffusivity

selectivity. This was confirmed by analysis of microstructure of membranes. Chemical

modification was found to be an instrumental means for tailoring performance of

membranes for separation of selected gas pairs. However, distinguished behaviors were

173
observed from the membrane toward cross-linking agents. The effect of p-xylenediamine

in progressive performance of membranes was more pronounced compared to p-

xylenedichloride. The Matrimid/PBI (25/75 wt.%) blend membrane cross-linked with p-

xylenediamine exhibits the best performance with H 2 /CO 2 selectivity of about 26 which

was due to high chain packing density and restrictions in segmental mobility of polymer

chains.

4.5 REFERENCES

[1] Today’s Hydrogen Production Industry, Website of U.S. Department of Energy,

http://www.fossil.energy.gov/programs/fuels/hydrogen/currenttechnology.html

[2] A.L. Kohl, R. Nielson, Gas purification, (5th ed.), Gulf Publishing, Houston, TX,

1997.

[3] S. Adhikari, S. Fernando, Hydrogen membrane separation techniques, Ind. Eng.

Chem. Res. 45 (2006) 875.

[4] S. Peramanu, B.G. Cox, B.B. Pruden, Economics of hydrogen recovery processes for

the purification of hydroprocessor purge and off-gases, Int. J. Hydrogen Energy 24

(1999) 405.

[5] D.R. Paul, Y.P. Yampol’skii (Eds.), Polymeric gas separation membranes, CRC

Press, Boca Raton, 1994.

[6] S.A. Stern, Polymers for gas separation: the next decade, J. Membr. Sci. 94 (1994) 1.

[7] R.G. Gardner, R.A. Crane, J.F. Hannan, Hollow fiber permeator for separating gases,

Chem. Eng. Prog. 73 (1977) 76.

174
[8] B. Ernst, S. Haag, M. Burgard, Permeselectivity of a nickle/ceramic composite

membranes at elevated temperatures: A new prospect in hydrogen separation?, J.

Membr. Sci. 288 (2007) 208.

[9] R.R. Zolandz, G.K. Fleming, gas permeation applications, in: W.S.W. Ho, K.K.

Sirkar, (Eds.), Membrane handbook, Chapman & Hall, New York, 1992, 78.

[10] A.W. Li, W. Liang, R. Hughes, Characterization and permeation properties of

palladium/stainless steel composite membranes, J. Membr. Sci. 149 (1998) 259.

[11] Y. Zhang, J. Gwak, Y. Murakoshi, T. Ikehara, R. Maeda and C. Nishimura,

Hydrogen permeation characteristics of thin Pd membrane prepared by

microfabrication technology, J. Membr. Sci. 277 (2006) 203.

[12] S. Tosti, L. Bettinali, V. Violante, Rolled thin Pd-Ag membranes for hydrogen

separation and purification, Int. J. Hydrogen Energy 25 (2000) 319.

[13] J.S. Chiou, D.R. Paul, Effects of CO 2 exposure on gas transport properties of glassy

polymers, J. Membr. Sci. 32 (1987) 195.

[14] A. Bos, I.G.M. Pünt, M. Wessling, H. Strathmann, CO 2 -induced plasticization

phenomena in glassy polymers, J. Membr. Sci. 155 (1999) 67.

[15] T.S. Chung, L. Shao, P.S. Tin, Surface modification of polyimide membranes by

diamine for H 2 and CO 2 separation, Macromol. Rapid Commun. 27 (2006) 998.

[16] P.S. Tin, T.S. Chung, Y. Liu, R .Wang, S.L. Liu, K.P. Pramoda, Effect of cross-

linking modification on gas separation performance of Matrimid membranes, J.

Membr. Sci. 225 (2003) 77.

175
[17] L.Y. Jiang, T.S. Chung, D.F. Li, C. Cao, S. Kulprathipanja, Fabrication of

Matrimid/Polyethersulfone dual-layer hollow fiber membranes for gas separation ,

J. Membr. Sci. 240 (2004) 91.

[18] T.S. Chung, W.F. Guo, Y. Liu, Enhanced Matrimid membranes for pervaporation by

homogenous blends with polybenzimidazole (PBI), J. Membr. Sci. 271 (2006) 221.

[19] R. He, Q. F. Li, A. Bach, J.O. Jensen, N.J. Bjerrum, Physiochemical properties of

phosphoric acid doped polybenzimidazole membranes for fuel cells, J. Membr. Sci.

277 (2006) 38.

[20] Khare, V.; Greenberg, A. R.; Balakrishnan, A.; Diani J.; Brahmandam S., Gas

transport and mechanical behavior of polymeric dense films at elevated

temperatures. 230th ACS National Meeting, Washington, DC, United States, Aug.

28-Sept. 1, 2005.

[21] Standard test methods for specific gravity (relative density) and density of plastics

by displacements, in: ASTM Annual Book of Standards, 1993, D-792.

[22] W.H. Lin, R.H. Vora, T.S. Chung, Gas transport properties of 6FDA-durene/1,4-

phenylenediamine (pPDA) copolyimides, J. Polym. Sci. Part B: Polym. Phys., 38

(2000) 2703.

[23] M.H. Kim, J.H. Kim, C.K. Kim, Y.S. Kang, H.C. Park, J.O. Won, Control of phase

separation behavior of PC/PMMA blends and their application to the gas separation

membranes, J. Polym. Sci.: Polym. Phys., 37 (1999) 2950.

[24] G. Guerra, S. Choe, D.J. Williams, F.E. Karasz, W.J. MacKnight, Fourier transform

infrared spectroscopy of some miscible polybenzimidazole/polyimide blends,

Macromol. 21 (1988) 231.

176
[25] E. Földes, E. Fekete, F.E. Karasz, B. Pukánszky, Interaction, miscibility and phase

inversion in PBI/PI blends, Polymer 41 (2000) 975.

[26] Y. Wang, S.H. Goh, T.S. Chung, Miscibility study of Torlon polyamide-imide with

Matrimid 5218 polyimide and polybenzimidazole, Polymer 48 (2007) 2901.

[27] P. Musto, F.E. Karasz, W.J. MacKnight, Hydrogen bonding in

polybenzimidazole/polyimide blends: a spectroscopic study, Macromol. 24 (1991)

4762.

[28] P.S.O. Patrício, J.A. de Sales, G.G. Silva, D. Windmöller, J.C. Machado, Effect of

blend composition on microstructure, morphology, and gas permeability in

PU/PMMA blends, J. Membr. Sci. 271 (2006) 177.

[29] S.H. Kim, D. Kim, D.S. Lee, Gas permeation behavior of PS/PPO blends, J. Membr.

Sci. 127 (1997) 9.

[30] W.M. Lee, Selection of barrier materials from molecular structure, Polym. Eng. Sci.

20 (1980) 65.

[31] C.C. Hu, K. R. Lee, R.C. Ruaan, Y.C. Jean, J.Y. Lai, Gas permeation properties in

cyclic olefin copolymer membrane studied by positron annihilation, sorption, and

gas permeation, J. Membr. Sci. 274 (2006) 192.

[32] D.W. Van Krevelen, Properties of Polymers-Their Correlation with Chemical

Structure; Their Numerical Estimation and Prediction from Additive Group

Contributions, Elsevier, Amsterdam, 1990.

[33] A. Bondi, Physical Properties of Molecular Crystals, Liquids, and Glasses, Wiley,

New York, 1968.

177
[34] S.C. Kumbharkar, P.B. Karadkar, U.K. Kharul, Enhancement of gas permeation

properties of polybenzimidazoles by systematic structure architecture, J. Membr.

Sci. 268 (2006) 161.

[35] K.L. Tung, K.T. Lu, R.C. Ruaan, J.Y. Lai, MD and MC simulation analyses on the

effect of solvent types on accessible free volume and gas sorption in PMMA

membranes, Desalination 192 (2006) 391.

[36] C. Cao, T.S. Chung, Y. Liu, R. Wang, K.P. Pramoda, Chemical cross-linking

modification of 6FDA-2,6-DAT hollow fiber membranes for natural gas separation,

J. Membr. Sci. 216 (2003) 257

[37] Y. Liu, R. Wang, T.S. Chung, Chemical cross-linking modification of polyimide

membranes for gas separation, J. Membr. Sci. 189 (2000) 231.

[38] S.S. Kelley, S. Shojaie, R. Peterson, A.R. Greenberg, W.B. Krantz, J. Filley, Effects

of crosslinking on the properties of cellulose ester membranes,. 213th ACS National

Meeting, San Francisco, (1997) April 13-17.

[39] K.Y. Wang, Y. Xiao, T.S. Chung, Chemically modified polybenzimidazole

nanofiltration membrane for the separation of electrolytes and cephalexin, Chem.

Eng. Sci. 61 (2006) 5507.

178
CHAPTER FIVE

CARBON MEMBRANES FROM BLENDS OF PBI AND POLYIMIDES FOR

N 2 /CH 4 AND CO 2 /CH 4 SEPARATION AND HYDROGEN PURIFICATION

5.1 INTRODUCTION

The inevitability and prominent importance of separation technology in various

sectors of energy and chemical industries is well recognized. Particularly, a significant

growth can be realized in development of technologies related to the separation of

gaseous species. This is, perhaps, driven largely by the great demand for purified natural

gas and other gases for industrial and pharmaceutical applications. For instance, by

considering natural gas sector alone, it is estimated that about 41% of the proven gas

reservoirs in the United States are sub-quality thus require upgrading by removal of

excessive CO 2 , N 2 and H 2 S and other impurities in order to meet the requirements for

pipeline transmission or wellhead processing [1].

Despite the presence of numerous established technologies (e.g., liquid

absorption, pressure swing adsorption, etc.) amenable to gas separations, it can be

realized that in recent years, extensive research studies has been devoted to devising new

techniques with potentials to bring cost-effectives and less complexities in process

179
operation and control. The progressive emergence of membranes that enjoy distinctive

inherent features such as high energy efficiency, ease of scale-up and small footprint has

drawn remarkable perspectives for this technology to become a viable choice for wide

variety of gas separation applications. According to the cited records, gas separation

membrane technology has turned to a $150 million/year business (in 2002) and yet

significant growth is anticipated for it in the near future [2]. Therefore, further

achievements toward maturization of membrane technology require extensive scientific

research and R&D breakthroughs to overcome the challenges.

Besides commercially available polymeric membranes which are gaining more

popularity, carbon membranes constitute a novel and attractive class of membranes with

distinctive features and outstanding gas separation performance. The very good chemical

and thermal stability and capability to surpass the permeability-selectivity trade-off are

considered among the prominent characteristics of carbon membranes [3,4]. From the

technical point of view, carbon membranes perhaps can be considered as true elements in

manifestation and realization of the role of morphology in separation performance. This

is based on the fact that the porous structure allows the high permeability (high

productivity) while the molecular sieving network provides the efficient size and shape

discrimination of molecules (high selectivity) [5]. Carbon membranes are basically

formed through the pyrolysis of polymeric precursors. Extensive studies have revealed

that chemical structure of the polymer, microstructure of the precursor and pyrolysis

process parameters are among the key factors determining transport properties of

resultant carbon membranes [5-10]. In other words, all these parameters play a part in

180
achieving the ultimate goal of obtaining a masterfully structured membrane comprised of

uniquely distributed pores with critical dimensions and desired surface properties.

High attentions have also been paid to the prominent effect of the precursor type;

hence various polymeric materials have been investigated for fabrication of high

performance carbon membranes for gas separation [11-16]. Basically, material selection

is inevitably one of the fundamental criteria for preparation of almost any kind of

membranes and particularly for gas separation. Therefore, apart from the fabrication

technology, it is anticipated the candidate material to possess the prerequisite properties

required for performing the designated functionality. Although, the same concept applies

to the carbon membranes but, in most cases, the candidate materials should meet even

more stringent requirements. For instance, good chemical and thermal stability as well as

amenability to retain the macromolecular structures and network integrity during

pyrolysis process are regarded as additional key prerequisites for achieving the desired

properties in carbon membranes. Of the recent developments, Zhang et al [11] studied the

properties and gas separation performance of a novel poly(phthalazinone ether sulfone

ketone)(PPESK)-based carbon membranes for various gas pairs. Based on their

observations, the resultant carbon membranes could exhibit good separation

performances comparable to polyimide-based carbon membranes. In another work, David

et al. [12] selected polyacrylonitrile (PAN) as the precursor to study the influence of

thermo-stabilization process and soak time leading to fabrication of carbon membranes

with enhanced performance for O 2 /N 2 separation.

181
Abundance of publications, especially in recent years, unanimously corroborate

the prominent position of polyimides as adequate precursors for preparation of carbon

membranes [6,9,17-31]. This is believed to be owing to the distinctive properties of

polyimides like rigidity, high melting point and T g while good chemical and thermal

stability. Thus various classes of polyimides including those containing

hexafluoroisopropylidene (6FDA) groups [20-21], P84 [10, 25, 40], pyromellitic

dianhydride (PMDA, Kapton) [9, 23, 27, 32-34], 1,4,5,8-Naphthalene tetracarboxylic

dianhydride (NTDA)- [35], benzophenone tetracarboxylic dianhydride (BTDA)-

[18,29,36-38] and 2,4,6,-trimethyl-1,3-phenylene diamine, 3,3’,4,4’-

biphenyltetracarboxylic dianhydride (BPDA)-[19,22] based polyimides have undergone

investigations. A comprehensive reference on the carbon membranes derived from the

family of polyimides is tabulated elsewhere [39]. Among all, Matrimid is widely

recognized as it is largely available as a commercial product [18, 29, 37-38, 40]. This is

more highlighted noting the fact that Matrimid have been used as a model polymer for

studying the porosity and its related effects on gas separation properties of carbon

membranes [18]. Jiang et al. employed Matrimid for fabrication of hollow carbon fiber

membranes [40]. Based on this study, although the results of single layer hollow fibers

were not attractive, but the inclusion of Matirmid as an inner layer followed by an outer

PSF-beta mixed matrix layer resulted in significantly enhanced gas separation

performance. In another attempt aiming improvements in gas transport and separation

performance of carbon membranes, Tin et al. [38] applied the chemical modification and

solvent treatment techniques on precursors prepared from Matrimid. The results revealed

182
the effects of swelling caused by solvent and presence of an optimum cross-linking

density.

Vast experiences have revealed the interesting advantages offered by the

involvement of blending technology into the area of polymeric gas separation

membranes. However, there can be found that only few and limited number of studies

have further extended the idea of blending for development of carbon membranes [41-

47]. It should be noted that majority of these reports are published in recent years which

may reflect a positive trend toward this approach. The main objective of this research

study is to explore the advantageous of employing blending technique in enhancing the

gas separation performance of carbon membranes. It is anticipated that blending can be

used as a simple while effective and efficient tool in tailoring the properties of

membranes. The foremost attention was paid in selection of materials amenable to form a

homogeneous matrix and fulfill the expectations. As a result, four high performance

polymers were selected to perform a comprehensive investigation. The aim was further

enriched by systematic investigation and involvement of other parameters including

pyrolysis temperature, blend composition and chemical modification. Membranes were

eventually analyzed with respect to their gas separation performance. Results revealed

that interesting gas permeability and permselectivity could be achieved in carbon

membranes engineered by blending and selection of a proper set of parameters. It is

believed that the knowledge obtained in this study not only provides a useful guideline

for further development of membranes with unique structures but also introduces a new

class of high performance membranes with attractive potentials for various gas

183
separations including nitrogen removal from natural gas streams which has been

suffering from technological limitations among membrane applications.

5.2 EXPERIMENTAL

5.2.1 Materials

Four types of materials were selected and examined in this study.

Polybenzimidazole (PBI, poly [2,2’-(1,3-phenylene)-5,5’-bibenzimidazole], a high

performance polymer with a high glass transition temperature, outstanding thermal

stability and chemical resistance was used as the main polymer. PBI was supplied by

Aldrich Chemical Company Inc. (Milwaukee, USA). Three other commercially available

polymers were selected as partners for preparation of blends with PBI. Matrimid® 5218

(poly [3,3'4,4'-benzophenone tetracarboxylic dianhydride and 5(6)-amino-1-(4'-

aminophenyl-1,3-trimethylindane)], BTDA-DAPI) was purchased from Vantico

(Luxemburg). Torlon® 4000T, a polyamide-imide, was supplied by Solvay Advanced

Polymers and P84 (copolyimide of 3,3’4,4’-benzophenone tetracarboxylic dianhydride

and 80% methylphenylenediamine + 20% methylenediamine) was provided by Lenzing.

The chemical structure of these polymers is shown in Figure 5.1. P-xylene diamine

(PXDA) was supplied by Merck (Germany) and used as the cross-linking agent. N-

Methyl-2-pyrolidone (NMP) and methanol were also procured from Merck and used as

solvents for polymers and the cross-linking agent, respectively.

184
(a)

(b)

(c)

(d)

Figure 5.1. Chemical structure of (a) Polybenzimidazole (PBI) (b) Matrimid® 5218 (c)

Torlon® 4000T and (d) P84.

5.2.2 Preparation of polymer precursors

Polymeric precursors were prepared as dense flat films in three compositions (i.e.,

25/75, 50/50 and 75/25 wt. %). Additional membranes were prepared purely from

individual polymers of Matrimid, Torlon and P84 as standard samples for comparison.

The total solid concentration in each solution was controlled at 2 wt. % in NMP. The

185
detailed preparation procedures including polymer dissolution, casting steps, and solvent

evaporation profile are reported elsewhere [48].

5.2.3 Chemical modification of membranes

Chemical modification of membranes was carried out through application of

cross-linking agent. For this purpose, blend precursors made of Matrimid and PBI were

immersed in solutions comprised of 10 wt.% p-xylene diamine in methanol for two

distinct periods of 5 and 10 days. At the end of the process, each film was washed with

fresh methanol followed by drying in vacuum oven at 120°C for overnight.

5.2.4 Preparation of carbon molecular sieve membranes

The pyrolysis process was performed in a vacuum furnace manufactured by

Centurion Neytech Qex. Polymeric precursors were carbonized under vacuum (2 mm-

Hg) while sandwiched in between two metallic wire meshes. Three different protocols

schematically illustrated in Figure 5.2. were applied. According to the information

provided in this figure, three temperatures of 600°C, 700°C and 800°C were set as final

pyrolysis temperatures attributed as A, B and C protocols, respectively. The heating rates

were controlled following a ramp function, starting with 15°C/min, with consecutive

lower rates of 4°C/min, 2.5°C/min and finally 0.2°C/min. The dwelling time at final

temperature was 2 hrs in all protocols. At the end of the process, the membranes were

cooled down steadily to room temperature.

186
Dwell for 2 hrs

800°C Protocol C
750°C
550°C

Dwell for 2 hrs


Temperature (oC)

250°C
700°C Protocol B
600°C
500°C

Dwell for 2 hrs


250°C
600°C Protocol A
550°C
400°C

250°C

15 oC/min 4 oC/min 2.5 oC/min 0.2 oC/min 0 oC/min

Pyrolysis progress and temperature increment rate

Figure 5.2. Representation of carbonization protocols used in this study with three

distinct final temperatures; protocol A: 600oC; protocol B: 700oC; protocol C: 800oC.

5.2.5 Characterization techniques

Several techniques were used in this study for the characterization of precursors

and their derivative carbon membranes. Differential scanning calorimetry (DSC) was

employed for the evaluation of glass transition temperature (T g ) of individual polymers

and blended precursors. Two consecutive scans from 50 to 450°C at the heating rate

10°C/min (under nitrogen purging) were performed for each sample on a Mettler Toledo

calorimeter (model DSC822). The mid-point of transition temperatures in the second scan

were reported as T g and for evaluation of the miscibility of fabricated precursors.

187
The trend of changes in weight loss during carbonization was simulated using the

Thermogravimetric Analysis (TGA) technique. The analyses were carried out on a

Shimadzu TG-DTA (DTG-60AH) using N 2 as purging gas at the flow rate of 50 ml.min-
1
. The applied temperature range was from 50 up to 900°C.

The microstructural properties of precursors and carbon membranes were

analyzed on X’Pert PRO-MRD wide-angle X-ray diffractometer (XRD) from

PANalytical. The measurements were carried out using Cu Kα radiation with a

wavelength of λ = 1.54 Å at room temperature. The scan range was from 4 to 70


˚ with a

step increment of 0.02˚ s -1. The average d-spacing was determined based on the Bragg’s

law according to the following formula:

nλ = 2d sin θ (5.1)

where n is an integral number (1, 2, 3, ...), λ denotes the X-ray wavelength, d stands for

the inter-segmental spacing between the polymer chains and θ indicates the diffraction

angle. The d-spacing values are interpreted as the average chain spacing.

The elemental analysis was conducted using X-ray photoelectron spectroscopy

(XPS) for scanning carbon, nitrogen and oxygen atoms. The measurements were

performed on a Kratos AXIS-HSi spectrometer with a built-in charge neutralization

system and dual anode Al/Mg X-ray source operating at 300W (15 kV, 20 mA).

5.2.6 Gas permeability and sorption measurements

188
The gas permeation properties of precursors and carbon membranes were

determined by a constant-volume method according to the procedure described in detail

elsewhere [49]. The pure gas permeability for carbon membranes was measured in a

sequence of He, H 2 , N 2 , CH 4 , CO 2 and O 2 at 35 °C. The gas permeability was

determined from the rate of pressure increase (dp/dt) at steady state using the following

relationship:

273.15 × 1010 Vl  dp 
P=   (5.2)
760 AT (( P0 × 76) / 14.7)  dt 

where P is the gas permeability of a membrane in Barrer (1Barrer = 1×10-10 cm3 (STP)-

cm/cm2.sec.cmHg), V is the volume of the down-stream chamber (cm3), A refers to

effective area of the membrane (cm2), l is the membrane thickness (cm), T is the

operating temperature (K) and the feed gas pressure in the up-stream is given by P 0 in

psia. The H 2 permeability test was conducted at 3.5 atm while the testing pressure for

other gases was 10 atm. The ideal selectivity for gas A to gas B was defined as follows:

PA
α A/ B = (5.3)
PB

The sorption properties of carbon membranes were evaluated for N 2 and CH 4

gases using a Cahn D200 microbalance sorption cell. Approximately 60 mg of membrane

was used during each test. The testing temperature was 35°C and sorption equilibriums

data were collected in the pressure range from 0 to 15 atm. The amount of dissolved gas

was calculated after accounting the buoyancy correction factor. The solubility

coefficients were then calculated using the following equation:

189
C
S= (5.4)
p

in which C represents the amount of sorption (cm3 (STP) gas)/(cm3 membrane) at the

corresponding pressure p (10 atm). The diffusion coefficients were calculated using

D=P/S, where P and S are the permeability and solubility coefficient, respectively.

5.3 RESULTS AND DISCUSSION

5.3.1. The physical and transport properties of polymers and blend precursors

PBI is a high performance polymer which its properties well meet the

requirements for preparation of carbon membranes. Nonetheless, seemingly the major

impediment which has hampered the wide application of PBI is the limited solubility in

common solvents. In addition, it has been observed that extraordinary brittleness often

encumbers the formation of free-standing films from PBI solutions [50]. The latter is

ascribed mainly to the presence of excessive aromatic rings in the repeating units of the

polymer as shown in Fig. 1. The good compatibility of PBI with other polymers

(especially polyimides) [51-53] originated the idea of employing blending technique to

exploit the advanced properties of PBI in preparation and development of desirable

carbon molecular sieve membranes. In this regard, the interesting capabilities of PBI to

withstand the elevated temperatures and also its stability and endurance toward adverse

environments are expected to be advantageous [54-55]. On the other hand, Matrimid,

Torlon and P84, with well examined and proven excellent properties for membrane

applications were selected as counterpart for PBI. The chemical structure of these

polymers is shown in Figure 5.1.

190
Conceptually, in contrast to the single phased systems, development of carbon

molecular sieve membranes from polymer blends requires a delicate approach and in

most of the cases degree of miscibility is regarded as the key governing factor [56]. In

addition, in such blend systems special attention has to be paid to the phase morphology

which is highly vulnerable to endothermic conditions due to the presence of different

coefficients of thermal expansions especially at temperatures near and above the T g . In

this study, transparency of the films and glass transition temperature (T g ) were used as

criteria to examine the miscibility of polymeric precursors. Table 5.1. provides the results

of Tg measurements for individual polymers as well as blend precursors. The miscibility

on the molecular level could be confirmed as all blend precursors exhibited a single

composition-dependent glass transition temperature falling in the range between the

values obtained for blend constituents. Moreover, the films were completely transparent

and no sign of phase separation could be observed throughout the entire films; further

Table 5.1. The glass transition temperature of polymers and blend precursors measured

by differential scanning calorimetry (DSC).

Polymer Tg (ºC)
PBI 417.5
Matrimid 323.1
Torlon 274.7
P84 315.0
Blends and composition Tg (ºC)
PBI/Matrimid (50/50 wt.%) 375.2
PBI/Torlon (50/50 wt.%) 355.2
PBI/P84 (50/50 wt.%) 390.1

191
corroborating the results of calorimetric experiments. These findings are in good

agreements with earlier research works [51-53]. Fourier Transform Infra-Red (FTIR)

analysis performed as supplementary experiments could also reveal the miscibility of

blend and existence of interaction in the molecular level. Details of FTIR analysis along

with figures can be found elsewhere thus are not repeated here [53,57].

In principle, the basic idea behind the design of polymer blends is to extract

simultaneous benefits from the complementary advantageous properties of each

constituent. This was targeted to employ this idea for advanced gas separation

applications. For this purpose, it is essential that each polymer be sufficiently efficient.

Gas transport properties and separation performance of PBI, Matrimid, Torlon and P84

are provided in Table 2 with the aid of reference [58]. The overall evaluation of data

indicate the lowest gas permeability belong to PBI; most probably due to its peculiar high

chain packing density and rigid structure. The permeability values for the rest of

polymers are relatively higher, following a general trend of P Matrimid > P P84 > P Torlon .

From the separation performance of point of view, the reported PBI data essentially

exhibits the highest selectivity for N 2 /CH 4 and CO 2 /CH 4 . The O 2 /N 2 selectivities

obtained for P84 and Matrimid membrane films were almost comparable and in the range

of 7~8. However, the reported O 2 /N 2 selectivity for PBI is unexpectedly low compared to

the previous proprietary information [59]. On the other hand, P84 exhibits the highest

H 2 /CO 2 selectivity. Therefore, the results indicate that all these polymers essentially

possess reasonably high separation performance; implying tactful selection of materials

to meet the specified requirements.

192
Table 5.2. Gas permeability and ideal selectivity of individual polymers used in this study. a,b

Membrane material Permeability (Barrer) Ideal Selectivity

He H2 O2 N2 CH4 CO2 H2/N2 N2/CH4 CO2/ CH4 O2/N2 H2/ CO2


PBIc 1.05 0.60 0.015 0.0048 0.0018 0.16 125.1 2.66 88.88 3.125 3.75
Matrimid 26.06 27.16 2.00 0.280 0.210 7.00 97.0 1.33 33.33 7.14 3.88
Torlon 5.53 4.44 0.212 0.037 0.030 0.83 120.0 1.23 27.80 5.73 5.32
P84 11.29 9.09 0.402 0.050 0.028 1.37 181.8 1.78 48.93 8.04 6.63
a1 Barrer = 1*10-10 cm3(STP) cm/cm2 s cm Hg
b Testingtemperature: 35oC
c Data obtained from reference [58] with testing temperature : 35oC

193
5.3.2. The effect of pyrolysis protocol on gas separation performance of carbon

membranes

The pyrolysis and its associated parameters as the key processing step for

fabrication of carbon membranes have been the subject of numerous studies. Through

which remarkable effects of heating protocol, final carbonization temperature and its

duration on the morphology of resultant carbon membranes have been appreciated [7, 19,

36]. Essentially it has been found that a typical final temperature for pyrolysis may fall

within the range of 500°C to 1000°C, depending on the type of precursors. The findings

also unanimously suggest that higher temperature and vacuum conditions may yield

carbon membranes with enhanced selectivity but typically at the expense of permeability

for gas separation applications [4,9,41,60]. However, one should also consider the

presence of an optimum temperature at which a molecular sieving microstructure with

balanced separation performance could be produced [19]. In this respect, the optimum

temperature is speculated to be largely dependent on the properties of the precursor and

pyrolysis condition. Thus, in the light of its importance, attention was paid to the

identification of optimum carbonization temperature. In order to avoid unnecessary

complexities, other parameters involved in the process like heating rate, residence time

and atmosphere were adjusted based on our prior experiences and remained identical for

the protocols. Blend precursors from PBI/Matrimid (50/50 wt. %) were selected as the

model samples. The effect of carbonization temperature on gas permeation properties of

resultant carbon molecular sieve membranes are presented in Table 3. From the results, it

could be found that a higher pyrolysis temperature produced carbon membranes with a

lower permeability; the lowest belonging to the membrane carbonized at 800ºC.

194
Table 5.3. The effect of pyrolysis temperature on gas permeability and selectivity of carbon membranes made of PBI/Matrimid

blend precursors. a,b

Precursor components Permeability (Barrer) Ideal Selectivity


& pyrolysis temperature

He H2 O2 N2 CH4 CO2 H2/N2 N2/CH4 CO2/ CH4 O2/N2 H2/ CO2


PBI/Matrimid
459.5 1337 113.2 15.71 5.84 305.5 85.10 2.69 52.31 7.2 4.37
(50/50 wt.%) @ 600oC
PBI/Matrimid
296.6 940.3 54.2 6.3 1.4 169.6 149.25 4.5 121.14 8.6 5.54
(50/50 wt.%) @ 700oC
PBI/Matrimid
112.12 324.0 10.96 1.26 0.278 36.6 257.1 4.53 131.65 8.7 8.85
(50/50 wt.%) @ 800oC
a 1 Barrer = 1*10-10 cm3(STP) cm/cm2 s cm Hg
b Testing temperature: 35oC

195
On the other hand, the increment in pyrolysis temperature was accompanied by

the enhancements in ideal gas pair selectivity. The effect was particularly more tangible

for particular gas pairs like H 2 /N 2 , CO 2 /CH 4 and H 2 /CO 2 which are of relatively larger

difference in kinetic diameter and the values reached up to about 257, 131 and 8.85,

respectively. This may imply the dominant contribution of size discriminative molecular

sieving rather than other mechanisms usually involved in transport of gas molecules in

carbon membranes. This can be interpreted as reduction in pore size, formation of tighter

and more compact structure, limiting in population of larger pores or sharper pore size

distribution upon rising the pyrolysis temperatures. There also have been some thoughts

that pyrolysis temperature may affect the transport properties of membranes through

changes in structural morphology, pyrolysis kinetics of polymer and pyrolysis kinetics of

degraded byproducts [7]. However, experiences have shown that the determination of the

exact contribution of each component is challenging or requires a tedious task.

Elemental analyses of precursors and carbon membranes prepared at various

temperatures could provide further insights over the issue. The results for PBI/Matrimid

blend systems, as shown in Figure 5.3., suggest that the precursor itself contains the

lowest carbon content (83%) but the highest oxygen content (12%). However, the carbon

content starts to increase upon pyrolysis and application of higher temperatures results in

membranes with higher carbon content. As a result, the carbon content of membrane

pyrolyzed at 800ºC could be reached up to about 94%. The observed trend which is

similar to other reports [19] can confirm the progressive evolution of a more graphite-like

structure at relatively higher temperatures. It is also worthwhile to note that a constant

196
depletion occurs in concentration of nitrogen elements upon the increasing the pyrolysis

temperature. The variation in oxygen content was impalpable, similarly observed by

others, probably due to the surface sorption phenomena [20]. The foregoing analysis may

also provide useful information on the elemental evolutions in carbonization process. It

could be seen that oxygen release is stabilized at temperatures about 600°C while the

major progress during the rest of process is on the evolution of other atoms like nitrogen.

A similar trend has been reported by others [61].

40 96

94
35
92
Elemental content (O & N)(%)

Elemental content (C)(%)


30
90
25
88

20
Oxygen
86
Nitrogen
Carbon 84
15
82
10
80
5
78

0 76
precursor 600 700 800

Carbonization Temperature (oC)

Figure 5.3. The elemental analyses of the polymeric precursor and derived carbon

membranes from PBI/Matrimid (50/50 wt. %) at various temperatures. (♦ : Carbon, ■ :

Oxygen, ▲ : Nitrogen)

197
Further structural characterization on carbon membranes was performed by

obtaining XRD spectra as a function of pyrolysis temperature. As shown in Figure 5.4.,

the XRD pattern for membrane carbonized at 600ºC is a merger of two amorphous peaks

appeared at 22.7° and 25.4°. However the increment of the final pyrolysis temperature

turns the peaks toward a single but intensified peak which is shifted slightly toward the

smaller pore sizes; indicating a decrement in average d-spacing. According to the fact

that the d-spacing for carbon membranes is a measure of interlayer distances, one can

infer the formation of more compact structures with higher packing density. In addition,

intensification of another peak located at 2θ=45.1° may be corresponded to the evolution

of more ordered microstructure with efficiently packed graphite layers upon increment of

0.9

0.8 (a) PBI/Matrimid (50/50 wt.%)@800C


(b) PBI/Matrimid (50/50 wt.%)@700C
0.7 (c) PBI/Matrimid (50/50 wt.%)@600C

0.6
Intensity

0.5 2.00 Å

0.4
(a)
0.3
(b)
0.2 3.47 Å

(c)
3.80 Å
0.1

0
10 20 30 40 50 60 70
2 Theta (degree)

Figure 5.4. The effect of variations in pyrolysis temperature on microstructure of

PBI/Matrimid (50/50 wt.%) blend carbon membranes evaluated by X-ray diffraction

analysis.

198
final pyrolysis temperatures [36]. This is of great significance as basically blend systems

are inclined to phase separation phenomena at temperatures close to glass transition

temperatures. This propensity could be more pronounced as the membranes undergo

carbonization which is well above the T g . This can be an indication of the robustness of

blended carbon membranes developed during the entire process. Therefore, the results of

the experiments could identify 800°C as the best pyrolysis temperature which was used

for preparation of carbon membranes throughout the rest of this study.

5.3.3. The effect of temperature protocol on evolution of carbon molecular sieve

membranes

It is well-known that the pyrolysis process is consisting of changes in membrane

morphology in conjunction with the evolution of some elemental components. Some

aspects of this issue were discussed in the previous section. Therefore, it should be

expected that tracking the composition of membrane elements in the course of the

process to provide useful information and insights about the progress of carbonization at

any corresponding stage. In this respect, thermogravimetric analysis (TGA) is the most

widely used and recommended technique. In fact, TGA has been identified as a suitable

means amenable to simulation of the carbonization process [7,9, 60-61]. According to

various reports, TGA has been effectively employed for determination of the temperature

increment protocol during the evolution of carbon membranes. According to the typical

practice the test is being conducted at a certain temperature rate (e.g., 10°C/min) under

nitrogen atmosphere. However, one should note that pyrolysis process is typically carried

out following different ramp functions and under vacuum condition. Therefore, there

199
exists a subtle discrepancy between two procedures and this originated the idea that

perhaps application of TGA with constant protocol may not be a true indicative of an

actual pyrolysis process. This led us to examine the validity of the approach by applying

the real pyrolysis protocols in TGA and evaluate the outcomes. Figure 5.5. shows the

results for PBI/Matrimid (50/50 wt. %) samples using two protocols; one with constant

ramp (10°C/min) (hereinafter: protocol D) and another one with specified ramp function

(i.e., protocol C, also shown in Figure 5.2.). It is interesting to observe distinctive

behaviors particularly with inception from about 350°C and onwards. While the sample

with protocol D exhibits a typical thermogram profile, a comparison reveals that the

100

90
Weight (%)

80
(a)
(b)

70

60 — (a) Protocol C
— (b) Protocol D (10oC/min)

50
0 100 200 300 400 500 600 700 800
Temperature (oC)

Figure 5.5. The results of the thermogravimetric analysis carried out on blend precursor

PBI/Matrimid (50/50 wt. %) applying different temperature protocols. (a) Protocol C (b)

Protocol D (10oC/min)

200
sample degradation in protocol C starts at relatively a lower temperature followed by a

similar weight loss profile. As a result, this stage of degradation is obviously more

extended for protocol C. The stabilization profile then resembles for both conditions till

to about 750°C where a divergence appears in protocol C. The observed features

essentially reflect the effects caused due to the changes in temperature increment rate of

protocol C. Hence, the results can verily highlight the importance of adopting conditions

consistent with the real case pyrolysis in order to avoid any probable mistake in

interpretation of the pyrolysis process.

5.3.4. The effect of blend component on gas separation performance of precursors and

derived carbon membranes

Enough emphasis was given, in the previous sections, to the conspicuous roles of

material selection in obtaining membranes with desired properties. Attempt in this study

was to exploit the synergistic properties of high performance materials through physical

blending. Table 5.4. provides the measured gas permeation values for precursors

fabricated from blending PBI and three other polymers; namely Matrimid, Torlon and

P84. The results show that permeability of blend precursors essentially fall in the range

between that of corresponding values obtained for individual polymers (Table 5.2.). It is

also evident that incorporation of PBI macromolecules which are consisting of rigid

aromatic chains brings a diminishing effect on permeation properties of membranes. The

influences caused by blending are different for each pair which possibly reflects the

nature and intensity of the interactions involved. The greatest effect could be observed for

PBI/Matrimid blend in which the permeability was dropped in average to about the half

201
Table 5.4. The effect of blend components on gas permeability and ideal selectivity of polymeric precursors. a,b

Membrane material Permeability (Barrer) Ideal Selectivity


& composition
He H2 O2 N2 CH4 CO2 H2/N2 N2/CH4 CO2/ CH4 O2/N2 H2/ CO2
PBI/Matrimid
13.55 13.06 0.58 0.072 0.045 2.16 181.38 1.6 48.00 8.05 6.05
(50/50 wt.%)
PBI/Torlon
4.38 3.75 0.151 0.023 0.021 0.615 163.04 1.1 29.28 6.57 6.10
(50/50 wt.%)
PBI/P84
7.52 6.88 0.389 0.067 0.056 1.60 102.68 1.2 28.61 5.81 4.30
(50/50 wt.%)
a 1 Barrer = 1*10-10 cm3(STP) cm/cm2 s cm Hg
b Testing temperature: 35oC

202
of original value for Matrimid. This is possibly owing to the strong interaction between

the functional groups. In comparison, a less reduction in permeability was observed for

other blends. On the other hand, it seems that no definite trend could be established for

variations in permselectivities.

Taking into the account the eminent role of phase morphology in tailoring the

physical properties and subsequently transport properties of polymeric blend membranes,

the results can be regarded as further indication of successful miscibility in molecular

level. In other words, another implication of the permeation tests is to assure the

successful progress of critical preliminary steps before proceeding to the carbonization

stage. Therefore, the above results could ensure the fulfillment of the requirements. The

gas permeability and selectivity of carbon molecular sieve membranes derived from

blend precursors are presented in Table 5.5. As could be expected from the nature of

carbon membranes, the resultant membranes exhibit much larger permeability compared

to their corresponding precursors. Also, interesting to note is that gas permeability of

carbon membranes follows an opposite trend compared to that observed in the precursors.

For instance, polymeric membrane from PBI/Torlon (50/50 wt.%) offering the lowest gas

permeability among the precursors was found to exhibit that of the largest upon

carbonization. On the other hand, permselectivities in almost all cases were raised to

some extent, depending on the type of gas pair. Table 5.6. provides the gas permeability

and selectivity of carbon membranes fabricated from each individual polymer for

comparison. Essentially, it could be observed that compared to other blended carbon

membranes, the one derived from PBI/Matrimid (50/50 wt.%) possesses the largest

203
Table 5.5. The effect of blend components on gas permeability and ideal selectivity of carbon membranes. a,b,c

Membrane material Permeability (Barrer) Ideal Selectivity


& composition
He H2 O2 N2 CH4 CO2 H2/N2 N2/CH4 CO2/ CH4 O2/N2 H2/ CO2
PBI/Matrimid
112.12 324.0 10.96 1.26 0.278 36.6 257.1 4.53 131.65 8.7 8.85
(50/50 wt.%)
PBI/Torlon
350.02 970.3 56.61 6.24 2.83 279 155.5 2.20 98.58 9.07 3.47
(50/50 wt.%)
PBI/P84
222.25 355.2 13.21 1.83 0.568 60.54 194.1 3.22 106.58 7.21 5.87
(50/50 wt.%)
a 1 Barrer = 1*10-10 cm3(STP) cm/cm2 s cm Hg
b Testing temperature: 35oC
c Pyrolysis temperature: 800ºC

Table 5.6. Gas permeability and ideal selectivity of carbon membranes derived from individual polymers. a,b,c

Membrane material Permeability (Barrer) Ideal Selectivity

He H2 O2 N2 CH4 CO2 H2/N2 N2/CH4 CO2/ CH4 O2/N2 H2/ CO2


Matrimid 398.1 1279 79.37 12.92 4.486 300.9 98.99 2.88 67.08 6.14 4.25
Torlon 136.1 390.1 8.39 3.084 2.351 43.06 126.49 1.31 18.31 2.72 9.06
P84 585.1 1151 37.43 4.90 2.147 144.2 234.89 2.28 67.16 7.63 7.98
a1 Barrer = 1*10-10 cm3(STP) cm/cm2 s cm Hg
b Testingtemperature: 35oC
c Pyrolysis temperature: 800ºC

204
selectivity for majority of gas pairs including H 2 /N 2 , N 2 /CH 4 , CO 2 /CH 4 and H 2 /CO 2 .

However, the O 2 /N 2 selectivity for this membrane is comparable and slightly lower to

that of PBI/Torlon (50/50 wt.%). This trend can be ascribed primarily to the difference in

microstructural properties. According to the XRD patterns shown in Figure 5.6., the

peaks are appeared at almost the same loci (2θ=25.58 & 45.16) indicating apparently no

difference in the average d-spacings (3.47 and 2.00 Å) between the membranes.

However, the higher peak intensities suggest the increase in regularity and packing order

which can effectively explain the observed trend in permeation results. This reveals that

the precursor morphology can play an important contribution in bringing about such final

properties.

0.9
(a) PBI/Matrimid (50/50 wt.%)
0.8
(b) PBI/P84 (50/50 wt.%)
0.7 (c) PBI/Torlon (50/50 wt.%)

0.6
Intensity

0.5
2.00 Å
0.4
(a)
0.3
(b)

0.2 3.47 Å
(c)
0.1

0
10 20 30 40 50 60 70
2 Theta (degree)

Figure 5.6. The effect of blend component on structural characteristics of blend carbon

membranes evaluated by XRD analysis. (Pyrolysis temperature: 800oC)

205
5.3.5. The effect of blend composition on transport properties of carbon membranes

It is widely shown that composition plays an important role in governing the

physical and chemical properties of blend systems. Our prior experiences have also

confirmed that the transport properties could largely be influenced by variation in blend

composition [48]. The subject was further extended to carbon membranes, as part of this

study, to explore whether and how the variation in composition could affect the

membrane microstructure, morphology and subsequently the gas transport properties.

The effect was investigated on blends made of PBI and Matrimid, primarily due to their

superior performance compared to the other systems. Table 5.7. provides the gas

permeability and selectivity evaluated for three different compositions. One can clearly

witness a large difference in permeability values of the samples; essentially the larger

PBI content has brought about a lower permeability. It can be noticed that the gas

permeability has increased up to 10 to 27 folds upon upgrading the Matrimid content

from 25% to 50% and 75%, respectively. The results also show that, compared to other

samples, the one with highest PBI content can be considered as a more viable candidate

for H 2 /CO 2 separation exhibiting the permselectivity of about 9.2. Achieving such good

performance can possibly be due to the contribution of rigid PBI chains with high

packing density that can have large effects on the chain configurations within the

membrane context and also pore formation during the carbonization process. This might

be in conjunction to the fact that compared to Matrimid, less non-carbon elements are

present in the repeating unit of PBI which can assist in retaining the membrane’s original

stability during carbonization process. As a result, formation of a fine porous structure in

carbon membranes containing higher PBI portions are expected to be responsible for

206
Table 5.7. The effect of blend composition on gas permeability and selectivity of resultant carbon membranes. a,b,c

Membrane material Permeability (Barrer) Ideal Selectivity


& composition
He H2 O2 N2 CH4 CO2 H2/N2 N2/CH4 CO2/ CH4 O2/N2 H2/ CO2
PBI/Matrimid
194.10 660.2 30.39 3.78 0.473 96.47 174.6 7.99 203.95 8.04 6.84
(25/75 wt.%)
PBI/Matrimid
112.12 324.0 10.96 1.26 0.278 36.60 257.1 4.53 131.65 8.70 8.85
(50/50 wt.%)
PBI/Matrimid
62.83 148.4 4.22 0.49 0.170 16.13 302.8 2.88 94.88 8.61 9.20
(75/25 wt.%)
a 1 Barrer = 1*10-10 cm3(STP) cm/cm2 s cm Hg
b Testing temperature: 35oC
c Pyrolysis temperature: 800ºC

207
improved discrimination between H 2 (2.89Å) and CO 2 (3.3Å) molecules. The enhanced

separation performance of PBI/Matrimid carbon membranes for H 2 /CO 2 is depicted in

Fig. 5.7. demonstrating that developed membranes are positioned above the trade-off

line.

100

11
10
9
H2/CO2 ideal selectivity

8 7
10 6
5
4
1 PBIa 1 3 2
2 Matrimid
3 PBI/Matrimid (25/75 wt.%)
4 PBI/Matrimid (50/50 wt.%)
1 5 PBI/Matrimid (75/25 wt.%)
6 PBI/Matrimid (25/75 wt.%) carbonized @ 800ºC
7 PBI/Matrimid (50/50 wt.%) carbonized @ 800ºC
8 PBI/Matrimid (75/25 wt.%) carbonized @ 800ºC
9 PBI/Matrimid (25/75 wt.%)+5 days p-xylene diamine X-linking carbonized @ 800ºC
10PBI/Matrimid (50/50 wt.%)+5 days p-xylene diamine X-linking carbonized @ 800ºC
11PBI/Matrimid (75/25 wt.%)+5 days p-xylene diamine X-linking carbonized @ 800ºC
0.1
0.1 1 10 100 1000
H2 permeability (Barrers)

Fig. 5.7. Performance of PBI/Matrimid blend precursors with various compositions and

their corresponding carbon membranes for H 2 /CO 2 separation with respect to trade-off

line. (aData from Ref. [58])

On the other hand, the results obtained by variation in composition indicate that

carbon membranes can also offer high separation performance for gases in the range of

larger molecular size. This would particularly be very attractive for separation of N 2 and

CO 2 from CH 4 . According to the data in Table 5.7, gradual reduction in PBI content of

208
precursors could provide carbon membrane with enhanced N 2 /CH 4 and CO 2 /CH 4

selectivity. Surprisingly, the enhancement in separation performance was very eminent

for N 2 /CH 4 and a selectivity of about 8 was obtained in carbon membrane derived from

PBI/Matrimid (25/75 wt.%). To our best knowledge, this is the highest ever reported

selectivity for this gas pair. Data in Figure 5.8. can clearly demonstrate the high impact of

achieving such high selectivity with enhanced permeability (P N2 =3.78 Barrer).

100
N2/CH4 ideal selectivity

10 6

7
8

1
5
1 PBIa 4
1 3 2
2 Matrimid
3 PBI/Matrimid (25/75 wt.%)
4 PBI/Matrimid (50/50 wt.%)
5 PBI/Matrimid (75/25 wt.%)
6 PBI/Matrimid (25/75 wt.%) carbonized @ 800oC
7 PBI/Matrimid (50/50 wt.%) carbonized @ 800oC
8 PBI/Matrimid (75/25 wt.%) carbonized @ 800oC
0.1
0.001 0.01 0.1 1 10 100 1000
N2 permeability (Barrers)

Figure 5.8. Performance of PBI/Matrimid blend precursors with various compositions

and their corresponding carbon membranes for N 2 /CH 4 separation with respect to trade-

off line. (aData from Ref [58]) (Trade-off line was drawn based on data points retrieved

from Ref. [62])

209
In this respect, study by Cecopieri-Gómeza et al. [62] can also shed lights on the

limitations of polymeric membranes and particularly polyimides for separation of

N 2 /CH 4 and similar gas pairs. Data in Figure 5.9 readily demonstrates the enhanced

CO 2 /CH 4 performance of carbon membranes compared to the trade-off line for this gas

pair.

1000

6
7
CO2/CH4 ideal selectivity

8
100
1

1 PBIa 5
10 2 Matrimid 4
3 PBI/Matrimid (25/75 wt.%)
4 PBI/Matrimid (50/50 wt.%) 3
5 PBI/Matrimid (75/25 wt.%)
6 PBI/Matrimid (25/75 wt.%) carbonized @ 800oC
7 PBI/Matrimid (50/50 wt.%) carbonized @ 800oC
8 PBI/Matrimid (75/25 wt.%) carbonized @ 800oC
1
0.01 0.1 1 10 100 1000

CO2 permeability (Barrers)

Figure 5.9. Performance of PBI/Matrimid blend precursors with various compositions

and their corresponding carbon membranes for CO 2 /CH 4 separation with respect to trade-

off line. (aData from Ref. [58])

The interesting properties of developed carbon membranes for the separation of

N 2 and CO 2 from CH 4 can possibly be attributed to the effective contribution of

diffusivity selectivity. This hypothesis was drawn from the careful scrutiny of the

210
spectacular transport properties of carbon membranes. According to an earlier study by

Singh et al. [63], the enhanced selectivity in carbon membranes is found to be due to the

prominent effect of diffusivity selectivity than the solubility selectivity. It should be

noted that the situation is however, different in polymeric membranes in which the effects

of both solubility and diffusivity selectivities are considerable. In fact, the concept of

diffusivity selectivity in carbon membranes has been extended by applying the transition

state theory and taking into account the physical properties of gas molecules. The results

indicated that among the two components of diffusivity selectivity, the effect entropic

selectivity was more significant compared to that of energetic selectivity. This was

attributed to the special provision of preferential transport in finely tuned pathways

within the graphite micro-crystals of carbon membranes [63].

The results obtained from the sorption tests further verified the significant role of

diffusivity selectivity in enhanced separation performance of derived carbon membranes.

The experiments were carried out on carbon membrane derived from PBI/Matrimid

(25/75 wt.%) using N 2 and CH 4 gases. Figure 5.10. presents the amount of sorption at

equilibrium against the testing pressure. It can be seen that the membrane has essentially

exhibited a greater sorption affinity toward CH 4 compared to N 2 over the studied

pressure range. The difference in sorption directly translates it effect into solubility

coefficients and as a result, membrane exhibited a solubility selectivity equivalent to

α S,N2/S,CH4 = 0.66 (S N2 =4.9 and S CH4 =7.4 (cm3(STP)/(cm3.atm)). Besides the effect of

membrane properties, the favorable solubility of methane is partially due to the larger

condensability of CH 4 molecules. (T b,CH4 =112K; T b,N2 =77K). The above results are in

211
good agreement with our hypothesis on the dominant effect of diffusivity selectivity (i.e.,

α D,N2/D,CH4 = 12.1) in achieving the enhanced separation performance of membranes for

N 2 /CH 4 . Possessing such enhanced diffusivity selectivity may indicate the formation of

finely tuned pores in the microstructure of carbon membranes that can constrict or retard

the transport of large molecules like CH 4 . The molecular passage in such network is

possibly permitted only to those molecules which have given up certain degrees of

translational/rotational /vibrational freedom or are aligned in a favorable direction with

respect to the pore geometry.


C [cm 3(STP)/cm3(membrane)]

CH4
N2

Figure 5.10. Plots of N 2 and CH 4 sorption isotherms for the carbon membrane derived

from the blend precursor composed of PBI/Matrimid (25/75 wt.%). (Pyrolysis

temperature: 800oC)

212
The results of XRD analysis on these carbon membranes (Figure 5.11.) can also

provide further support to our hypothesis. Although almost no tangible shift was occurred

in the position of indicative peaks, but evolution of more ordered structural morphologies

could be witnessed by the increments in intensity of the peaks upon increase in Matrimid

content. Therefore, developed carbon molecular sieve membranes from blends of PBI

and Matrimid based on the prescribed procedures can be considered as potential

breakthroughs for practical implementation of membranes for nitrogen rejection from

natural gas. This is particularly more interesting for real case application as favorable

permeation of nitrogen than methane can bring substantial savings in costly and energy

intensive recompression of natural gas.

0.9 (a) PBI/Matrimid (25/75 wt.%)


(b) PBI/Matrimid (50/50 wt.%)
0.8
(c) PBI/Matrimid (75/25 wt.%)
0.7

0.6
Intensity

0.5

0.4
(a)
0.3 (b)

(c)
0.2

0.1

0
10 20 30 40 50 60 70
2 Theta (degree)

Figure 5.11. The effect of blend composition on structural characteristics of

PBI/Matrimid blend carbon membranes evaluated by XRD analysis. (Pyrolysis

temperature: 800oC)

213
Figures 5.12 (a), (b) and (c) demonstrate the performance standing of the carbon

membranes for O 2 /N 2 , H 2 /N 2 and H 2 /CH 4 separations.

100
(a)
O2/N2 ideal selectivity

10 8 7 6

5 4
3 2

1 PBIa
1 2 Matrimid
3 PBI/Matrimid (25/75 wt.%)
4 PBI/Matrimid (50/50 wt.%)
5 PBI/Matrimid (75/25 wt.%)
6 PBI/Matrimid (25/75 wt.%) carbonized @ 800oC
7 PBI/Matrimid (50/50 wt.%) carbonized @ 800oC
8 PBI/Matrimid (75/25 wt.%) carbonized @ 800oC
0.1
0.001 0.01 0.1 1 10 100 1000
O2 permeability (Barrers)

1000
(b)
11
10
9
H2/N2 ideal selectivity

8
5 7

4 6

100 1 PBIa 1 3
2 Matrimid 2
3 PBI/Matrimid (25/75 wt.%)
4 PBI/Matrimid (50/50 wt.%)
5 PBI/Matrimid (75/25 wt.%)
6 PBI/Matrimid (25/75 wt.%) carbonized @ 800ºC
7 PBI/Matrimid (50/50 wt.%) carbonized @ 800ºC
8 PBI/Matrimid (75/25 wt.%) carbonized @ 800ºC
9 PBI/Matrimid (25/75 wt.%)+5 days p-xylene diamine X-linking carbonized @ 800ºC
10PBI/Matrimid (50/50 wt.%)+5 days p-xylene diamine X-linking carbonized @ 800ºC
11PBI/Matrimid (75/25 wt.%)+5 days p-xylene diamine X-linking carbonized @ 800ºC
10
0.1 1 10 100 1000
H2 permeability (Barrers)

214
10000
(c)

H2/CH4 ideal selectivity 9


11 10
1000 7 6

5
1 4 8
3
100 1 PBIa 2
2 Matrimid
3 PBI/Matrimid (25/75 wt.%)
4 PBI/Matrimid (50/50 wt.%)
5 PBI/Matrimid (75/25 wt.%)
6 PBI/Matrimid (25/75 wt.%) carbonized @ 800ºC
10
7 PBI/Matrimid (50/50 wt.%) carbonized @ 800ºC
8 PBI/Matrimid (75/25 wt.%) carbonized @ 800ºC
9 PBI/Matrimid (25/75 wt.%)+5 days p-xylene diamine X-linking carbonized @ 800ºC
10PBI/Matrimid (50/50 wt.%)+5 days p-xylene diamine X-linking carbonized @ 800ºC
11PBI/Matrimid (75/25 wt.%)+5 days p-xylene diamine X-linking carbonized @ 800ºC
1
0.01 0.1 1 10 100 1000
H2 permeability (Barrers)

Figure 5.12. Performance of PBI/Matrimid blend precursors with various compositions

and their corresponding carbon membranes with respect to trade-off line for various gas

pairs. (a) O 2 /N 2 (b) H 2 /N 2 (c) H 2 /CH 4 . (aData from Ref. [58])

5.3.6. The effect of chemical modification on properties and gas separation

performance of carbon membranes

Various types of cross-linking methods have attracted attentions as viable means

for tailoring the properties of membranes, and particularly to modify the substructure in

micro- and nano-domain levels. This subject was highlighted in an earlier report

describing the various aspects and consequences of applying such chemical modification

on polymeric materials. For instance, an altered morphology and improved gas separation

performance could be achieved in carbon membranes derived from treated precursor [64].

215
However, it can be realized that this attempt has only been practiced on membranes

composed of a single polymeric component. On the other hand, the experiences obtained

from employing similar approach on polymeric blend membrane has revealed that

chemically modified blend membranes could potentially yield an improved separation

performance [48]. Interestingly, it was revealed in the course of that study that each

component in the blend may respond differently to the modifying agents. Specifically, in

membranes prepared from Matrimid/PBI, more favorable results could be achieved by

the modification of Matrimid phase compared to that of PBI.

An idea was originated to further extend the concept and explore the applicability

of the approach to identify the effects of chemical modification prior to carbonization and

its subsequent influences on the properties and performance of resultant carbon

membranes derived from polymeric blend precursors. P-xylene diamine was selected as

the chemical agent for cross-linking due to its high degree of solubility in the medium

(i.e., methanol, a non-solvent for polymers) and high reaction efficiency. The carbon

membranes derived from the cross-linked precursors were tested and the measured gas

permeability and selectivity values are provided in Table 5.8. Comparison of the results

with corresponding unmodified counterparts (Table 5.7.) indicates a significant decline in

permeability values upon modification. Additionally, the selectivity values have also

been affected and two distinct trends could be recognized. Firstly, the permselectivity for

the gas pairs with negligible difference in molecular dimensions (e.g. N 2 /CH 4 , CO 2 /CH 4

and O 2 /N 2 ) are essentially declined, seemingly as a result of tighter structures provided

by the linkages among macromolecules. On the other hand, favorable enhancements were

216
Table 5.8. The effect of application of a 5-day cross-linking modification by p-xylene diamine (PXDA) on the gas transport

properties of blend carbon membranes with various compositions. a,b,c

Membrane material Permeability (Barrer) Ideal Selectivity


& composition
He H2 O2 N2 CH4 CO2 H2/N2 N2/CH4 CO2/ CH4 O2/N2 H2/ CO2
PBI/Matrimid
71.16 182.7 2.615 0.392 0.195 11.00 466.1 2.01 56.41 6.67 16.61
(25/75 wt.%)
PBI/Matrimid
34.77 91.0 0.826 0.172 0.104 3.71 529.1 1.65 30.50 4.80 24.52
(50/50 wt.%)
PBI/Matrimid
23.91 63.2 0.357 0.109 0.081 1.89 579.8 1.34 13.23 3.28 33.44
(75/25 wt.%)
a 1 Barrer = 1*10-10 cm3(STP) cm/cm2 s cm Hg
b Testing temperature: 35oC
c Pyrolysis temperature: 800ºC

217
achieved in selectivity for gas pairs possessing larger differences in dimensions. For

instance, depending on the blend composition, the gas pair selectivity for both H 2 /N 2 and

H 2 /CO 2 has boosted up to the values in the ranges of 466.1~579.1 and 16.61~33.44,

respectively.

These results can be explained by taking into account the specific role and

contributions of the cross-linking agent in the membranes. Based on the simulation

studies [48], it has been verified that cross-linking molecules not only perform as spaces

filling materials within the membrane context but also their presence can bring about

macromolecular reconfiguration and thus influence the membrane morphology through

establishment of covalent bonding between the adjacent macromolecules. This is

expected to improve the structural stability of the membranes to withstand high

temperatures associated with pyrolysis process. This can take effect through among other

contributing mechanisms such as prevention from the sudden structural collapse, tuning

the interstitial spaces between molecules and preserving the structural integrity after

pyrolysis.

By comparing the membranes derived from after cross-linking modification, one

could perhaps anticipate a better separation performance in those containing larger

amount of Matrimid. This is due to the possible higher degree of modification that can

occur during the treatment of precursors. In this regard, the results in Table 8 may look

counterintuitive as upon reduction in Matrimid content, a stepwise improvement in

selectivity could be achieved for gas pairs like H 2 /CO 2 and H 2 /N 2 .

218
The occurrence of such impressive separation performance in carbon membranes

from BPI/Matimrid (75/25 wt.%) compared to the counterparts can be realized by

simultaneous considerations of the cross-linking phenomena and variations in blend

composition. Basically, in one hand, by applying cross-linking modification, all three

membranes are expected to exhibit improved separation performance; the higher

Matrimid content thus should result in the larger improvement. This could be confirmed

by observation of the XRD patterns (Figure 5.13.) which revealed relatively intensified

indicative peaks for these membranes compared to the unmodified ones shown in Figure

5.11.

1.4

(a) X-linked PBI/Matrimid (25/75 wt.%)


1.2
(b) X-linked PBI/Matrimid (50/50 wt.%)
(c) X-linked PBI/Matrimid (75/25 wt.%)
1

0.8
Intensity

0.6

(a)
0.4 (b)
(c)

0.2

0
10 20 30 40 50 60 70
2 Theta (degree)

Figure 5.13. The effect of blend composition on structural characteristics of

PBI/Matrimid blend carbon membranes derived from cross-linked precursors. (Pyrolysis

temperature: 800oC)

219
However, on the other hand, the higher Matrimid content translates to the less PBI

content in the blend which according to the discussions in the previous section (Table

5.7.) should result in diminishing effects in separation performance. In other words,

introduction of large Matrimid for achieving high cross-linking efficiency comes at the

expense of benefits offered by inherent properties of PBI phase. Thus, these two factors

compete against each other and alter the membrane performance in opposite directions.

In overall, the results in Table 5.8., imply the more effective and dominant influence of

blend component over chemical modification in controlling the transport and separation

properties of carbon membranes derived from modified precursors. This is in agreement

with the fact that modification through chemical cross-linking is perceived to influence

the limited outermost surfaces of the membranes, whereas the phase composition has a

global effect over the entire membrane structure. As a result, the high separation

performance in BPI/Matimrid (75/25 wt.%) can be ascribed to the enhanced

morphological properties by partial contributions of PBI chains and chemically integrated

Matrimid macromolecules. As illustrated in Figure 5.7. and Figures 5.12. (b) & (C),

carbon membranes derived from chemically-modified precursors stand well above the

trade-off line specifically for H 2 /CO 2 , H 2 /N 2 and H 2 /CH 4 separation.

5.4 CONCLUSIONS

Advanced carbon molecular sieving membranes were fabricated from the blend of

PBI and various polyimides for gas separation applications. Polymeric films fabricated

from blends of Matirmid, P84 and Torlon with PBI were found exhibiting desirable gas

220
transport properties in conjunction with homogenous structure which were believed to fit

the prerequisites for further development and derivation of carbon membranes.

Examination of three carbonization protocols with distinctive final temperatures revealed

achieving a higher efficiency in carbon membranes obtained at elevated temperatures.

This was mainly ascribed to the extent of progress in processing at higher temperatures

which could bring about microstructures with higher carbon concentrations and more

effective pore sizes with desirable distributions. In addition, according to the analyses

carried out in this study, it was found that proper parameter settings might be essential in

order to achieve more realistic insights about the governing pyrolysis process through the

application of TGA.

Derived carbon membranes were found more permeable to gases compared to

their respective precursors. More interestingly, precursors with tighter polymeric

microstructures where found to offer carbon membranes with higher permeability.

Carbon membranes derived from PBI/Matirmid (50/50 wt. %) exhibited more attractive

performance than the other blends. Further investigations on this promising blend system

revealed higher Matirmid content could result in improved permeability while enhanced

selectivity particularly for separation of N 2 and CO 2 from CH 4 . In addition, further

chemical modification of PBI/Matrimid precursors prior to carbonization produced

membranes with enhanced performance for separation of H 2 molecules from N 2 and

CO 2 . Carbon molecular sieve membranes developed in this study surpass various trade-

off lines and are of great potentials for industrial applications.

221
5.5 REFERENCES

[1] H.S. Meyer, Volume and distribution of subquality natural gas in the United States,

GasTIPS, Gas Research Institute, Winter (2000) p.10.

[2] R.W. Baker, Future directions of membrane gas separation technology, Ind. Eng.

Chem. Res., 41 (2002) 1393.

[3] A.F. Ismail, L.I.B. David, A review on the latest development of carbon membranes

for gas separation, J. Membr. Sci. 193 (2001) 1.

[4] P.S. Tin, Y. Xiao, T.S. Chung, Polyimide-carbonized membranes for gas separation:

structural, composition, and morphological control of precursors, Sep. Purif. Rev.,

35 (2006) 285.

[5] J.E. Koresh, Soffer A. Study of molecular sieve carbons, Part 1: pore structure,

gradual pore opening and mechanism of molecular sieving, J. Chem. Soc. Faraday

I, 76 (1980) 2457.

[6] H.B. Park, Y.K. Kim, J.M. Lee, S.Y. Lee, Y.M. Lee, Relationship between chemical

structure of aromatic polyimides and gas permeation properties of their carbon

molecular sieve membranes. J. Membr. Sci. 229 (2004) 117.

[7] V.C. Geiszler, W.J. Koros, Effects of polyimide pyrolysis conditions on carbon

molecular sieve membrane properties, Ind. Eng. Chem. Res. 35 (1996) 2999.

[8] J.E. Koresh, A. Soffer, The carbon molecular sieve membranes, general properties

and the permeability of CH 4 /H 2 mixture, Sep. Sci. Technol. 22 (1987) 973.

[9] H. Suda, K. Haraya, Gas permeation through micropores of carbon molecular sieve

membranes derived from Kapton polyimide, J. Phys. Chem.: B, 101 (1997) 3988.

222
[10] P.S. Tin, T.S. Chung, A.J. Hill, Advanced fabrication of carbon molecular sieve

membranes by non-solvent pretreatment of precursor polymers, Ind. Eng. Chem.

Res., 43 (2004) 6476.

[11] B. Zhang, T. Wang, S. Zhang, J. Qiu, X. Jian, Preparation and characterization of

carbon membranes made from poly(phthalazinone ether sulfone ketone), Carbon,

44 (2006) 2764.

[12] L.I.B. David, A.F. Ismail, Influence of thermastabilization process and soak time

during pyrolysis process on the polyacrylonitrile carbon membranes for O 2 /N 2

separation, J. Membr. Sci., 213 (2003) 285.

[13] T.A. Centeno, A.B. Fuertes, Carbon molecular sieve gas separation membranes

based on poly(vinylidene chloride-co-vinyl chloride), Carbon, 38 (2000) 1067.

[14] F. K. Katsaros, T.A. Steriotis, A.K. Stubos, A. Mitropoulos, N.K. Kanellopoulos, S.

Tennison, High pressure gas permeability of microporous carbon membranes

Micro. Mat., 8 (1997) 171.

[15] H. Wang, L. Zhang, G.R. Gavalas, Preparation of supported carbon membranes from

furfuryl alcohol by vapor deposition polymerization, J. Membr. Sci., 177 (2000) 25.

[16] Y.D. Cheng, R.T. Yang, Preparation of carbon molecular sieve membrane and

diffusion of binary mixtures in the membrane, Ind. Eng. Chem. Res., 33 (1994)

3146.

[17] V.C. Geiszler, Polyimide precursors for carbon membranes, University of Texas,

PhD thesis, 1997.

[18] K.M. Steel, W.J. Koros, Investigation of porosity of carbon materials and related

effects on gas separation properties, Carbon, 41 (2003) 253.

223
[19] J. Hayashi, M. Yamamoto, K. Kusakabe, S. Morooka, Effect of oxidation on gas

permeation of carbon molecular sieving membranes based on BPDA-pp’ODA

polyimide, Ind. Eng. Chem. Res., 36 (1997) 2134.

[20] C.W. Jones, W.J. Koros, Carbon molecular sieve gas separation membranes, I.

Preparation and characterization based on polyimide precursors, Carbon, 32 (1994)

1419.

[21] A. Singh-Ghosal, W.J. Koros, Air separation properties of flat sheet homogeneous

pyrolytic carbon membranes, J. Membr. Sci., 174 (2000) 177.

[22] J. Hayashi, H. Mizuta, M. Yamamoto, K. Kusakabe, S. Morooka, Pore size control

of carbonized BPDA-pp’ODA polyimide membrane by chemical vapor deposition

of carbon. J. Membr. Sci., 124 (1996) 243.

[23] H. Hatori, Y. Yamada, M. Shiraishi, H. Nakata, S. Yoshitomi, Carbon molecular

sieve films from polyimide, Carbon, 30 (1992) 305.

[24] T. Takeichi, Y. Eguchi, Y. Kaburagi, Y. Hishiyama, M. Inagaki, Carbonization and

graphitization of BPDA/PDA polyimide films: effect of structure of polyimide

precursor, carbon, 37 (1999) 569.

[25] P.S. Tin, T.S. Chung, Y. Liu, R. Wang, Separation of CO 2 /CH 4 through carbon

molecular sieve membranes derived from P84 polyimide, Carbon, 42 (2004) 3123.

[26] K. Okamoto, S. Kawamura, M. Yoshino, H. Kita, Y. Hirayama, N. Tanihara, Y.

Kusuki, Olefin/paraffin separation through carbonized membranes derived from an

asymmetric polyimide hollow fiber membrane. Ind. Eng. Chem. Res., 38 (1999)

4424.

224
[27] K. Wang, H. Suda, K. Haraya, The characterization of CO 2 permeation in a CMSM

derived from polyimide, Sep. Purif. Technol., 31 (2003) 61.

[28] M.G. Sedigh, L. Xu, T.T. Tsotsis, M. Sahimi, Transport and morphological

characteristics of polyetherimide-based carbon molecular sieve membranes, Ind.

Eng. Chem. Res., 38 (1999) 3367.

[29] A.B. Fuertes, D.M. Nevskaia, T.A. Centeno, Carbon composite membranes from

Matrimid and Kapton polyimides for gas separation. Micro. Meso. Mater., 33

(1999) 115.

[30] Y. Xiao, Y. Dai, T.S. Chung, M.D. Guiver, Effects of brominating Matrimid

polyimide on the physical and gas transport properties of derived carbon

membranes, Macromol., 38 (2005) 10042.

[31] J.N. Barsema, N.F.A. van der Vegt, G.H. Koops, M. Wessling , Carbon molecular

sieve membranes prepared from porous fiber precursor, J. Membr. Sci., 205 (2002)

239.

[32] J. Petersen, M. Matsuda, K. Haraya, Capillary carbon molecular sieve membranes

derived from Kapton for high temperature gas separation, J. Membr. Sci., 131

(1997) 85.

[33] H. Suda, K. Haraya, Molecular sieving effect of carbonized kapton polyimide

membrane J. Chem. Soc., Chem. Commun., 11 (1995) 1179.

[34] A.C. Lua, J. Su, Effects of carbonisation on pore evolution and gas permeation

properties of carbon membranes from Kapton polyimide, Carbon, 44 (2006) 2964.

225
[35] M.N. Islam, W. Zhou, T. Honda, K. Tanaka, H. Kita, K. Okamoto, Preparation and

gas separation performance of flexible pyrolytic membranes by low-temperature

pyrolysis of sulfonated polyimides, J. Membr. Sci., 261 (2005) 17.

[36] Y.K. Kim, H.B. Park, Y.M. Lee, Preparation and characterization of carbon

molecular sieve membranes derived from BTDA–ODA polyimide and their gas

separation properties, J. Membr. Sci., 255 (2005) 265.

[37] J.N. Barsema, S.D. Klijnstra, J.H. Balster, N.F.A. van der Vegt, G.H. Koops, M.

Wessling, Intermediate polymer to carbon gas separation membranes based on

Matrimid PI. J. Membr. Sci., 238 (2004) 93.

[38] P.S. Tin, T.S. Chung, S. Kawi, M.D. Guiver, Novel approaches to fabricate carbon

molecular sieve membranes based on chemical modified and solvent treated

polyimides, Micro. Meso. Mater., 73 (2004) 151.

[39] S.M. Saufi, A.F. Ismail, Fabrication of carbon membranes for gas separation-a

review, Carbon, 42 (2004) 241.

[40] L.Y. Jiang, T.S. Chung, R. Rajagopalan, Dual-layer hollow carbon fiber membranes

for gas separation consisting of carbon and mixed matrix layers, Carbon 45 (2006)

166.

[41] Y.K. Kim, H.B. Park, Y.M. Lee, Gas separation properties of carbon molecular sieve

membranes derived from polyimide/polyvinylpyrrolidone blends: effect of the

molecular weight of polyvinylpyrrolidone, J. Membr. Sci., 251 (2005) 159.

[42] H.J. Lee, H. Suda, K. Haraya, S.H. Moon, Gas permeation properties of carbon

molecular sieving membranes derived from the polymer blend of polyphenylene

oxide (PPO)/polyvinylpyrrolidone (PVP), J. Membr. Sci., 296 (2007) 139.

226
[43] X. Zhang, H. Hua, Y. Zhua, S. Zhua, Carbon molecular sieve membranes derived

from phenol formaldehyde novolac resin blended with poly(ethylene glycol) J.

Membr. Sci., 289 (2007) 86.

[44] J.N. Barsema, N.F.A. van der Vegt, G.H. Koops, M. Wessling, Ag functionalized

carbon molecular-sieve membranes based on polyelectrolyte/polyimide blend

precursors, Adv. Funct. Mater., 15 (2005) 69.

[45] H. Hatori,1 T. Kobayashi, Y. Hanzawa, Y. Yamada, Y. Iimura, T. Kimura, M.

Shiraishi, Mesoporous carbon membranes from polyimide blended with

poly(ethylene glycol), J. Appl. Polym. Sci., 79 (2001) 836.

[46] Y.K. Kim, H.B. Park, Y.M. Lee, Carbon molecular sieve membranes derived from

thermally labile polymer containing blend polymers and their gas separation

properties, J. Membr. Sci., 243 (2004) 9.

[47] P.S. Rao, M.Y. Wey, H.H. Tseng, I.A. Kumar, T.H. Weng, A comparison of

carbon/nanotube molecular sieve membranes with polymer blend carbon molecular

sieve membranes for the gas permeation application, Micro. Meso. Mater., (2008)

In press.

[48] S.S. Hosseini, M.M. Teoh, T.S. Chung, Hydrogen separation and purification in

membranes of miscible polymer blends with interpenetration networks, Polymer, 49

(2008) 1594.

[49] W.H. Lin, R.H. Vora, T.S. Chung, Gas transport properties of 6FDA-durene/1,4-

phenylenediamine (pPDA) copolyimides. J. Polym. Sci.: Polym. Phys., 38 (2000)

2703.

227
[50] D.R. Pesiri, B. Jorgensen, R.C. Dye, Thermal optimization of polybenzimidazole

meniscus membranes for the separation of hydrogen, methane, and carbon dioxide,

J. Membr. Sci. 218 (2003) 11.

[51] E. Földes, E. Fekete, F.E. Karasz, B. Pukánszky, Interaction, miscibility and phase

inversion in PBI/PI blends, Polymer, 41 (2000) 975.

[52] G. Guerra, S. Choe, D.J. Williams, F.E. Karasz, W.J. MacKnight, Fourier transform

infrared spectroscopy of some miscible polybenzimidazole/polyimide blends,

Macromol., 21 (1988) 231.

[53] Y. Wang, S.H. Goh, T.S. Chung, Miscibility study of Torlon polyamide-imide with

Matrimid 5218 polyimide and polybenzimidazole, Polymer, 48 (2007) 2901.

[54] V. Khare, A.R. Greenberg, A. Balakrishnan, J. Diani, S. Brahmandam, Gas transport

and mechanical behavior of polymeric dense films at elevated temperatures. 230th

ACS National Meeting, Washington, DC, United States, Aug. 28-Sept. 1, 2005.

[55] T.S. Chung, A critical review of polybenzimidazoles: historical development and

future R&D, J. Macromol. Sci. Polym. Rev. C37 (1997) 277.

[56] M.H. Kim, J.H. Kim, C.K. Kim, Y.S. Kang, H.C. Park, J.O. Won, Control of phase

separation behavior of PC/PMMA blends and their application to the gas separation

membranes, J. Polym. Sci.: Polym. Phys., 37 (1999) 2950.

[57] T.S. Chung, W.F. Guo, Y. Liu, Enhanced Matrimid membranes for pervaporation by

homogenous blends with polybenzimidazole (PBI), J. Membr. Sci. 271 (2006) 221.

[58] S.C. Kumbharkar, P.B. Karadkar, U.K. Kharul, Enhancement of gas permeation

properties of polybenzimidazoles by systematic structure architecture, J. Membr.

Sci. 286 (2006) 161.

228
[59] Hoechst Celanese proprietary information.

[60] L. Shao, T.S. Chung, K.P. Pramoda, The evolution of physicochemical and transport

properties of 6FDA-durene toward carbon membranes; from polymer, intermediate

to carbon, Micro. Meso. Mater., 84 (2005) 59.

[61] E. Barbosa-Coutinho, V.M.M. Salim, C.P. Borges, Preparation of carbon hollow

fiber membranes by pyrolysis of polyetherimide, Carbon, 41 (2003) 1707.

[62] M.L. Cecopieri-Gómeza, J. Palacios-Alquisira, J.M. Domínguez, On the limits of

gas separation in CO 2 /CH 4 , N 2 /CH 4 and CO 2 /N 2 binary mixtures using polyimide

membranes, J. Membr. Sci., 293 (2007) 53.

[63] A. Singh, W.J. Koros, Significance of entropic selectivity for advanced gas

separation membranes, Ind. Eng. Chem. Res., 35 (1996) 1231.

[64] P.S. Tin, T.S. Chung, Novel approach to fabricate carbon molecular sieve

membranes based on consideration of interpenetrating network, Macromol. Rapid

Commun., 25 (2004) 1247.

229
CHAPTER SIX

GAS SEPARATION MEMBRANES DEVELOPED THROUGH INTEGRATION

OF POLYMER BLENDING AND DUAL-LAYER HOLLOW FIBER SPINNING

PROCESS FOR HYDROGEN AND NATURAL GAS ENRICHMENTS

6.1 INTRODUCTION

The higher efficiency and beneficial advantages offered by gas separation

membranes especially in terms of capital investments, operational costs and also energy

requirements have led to the attraction of many attentions toward this process. The

accelerated popularity of using membranes for various applications can largely be

attributed to the growing concerns over global energy resources and the green house

effects caused by CO 2 . This has, however, been in conjunction with the progressing

improvements in the science and technology of membranes and membrane processes.

Some of the most established membranebased gas separation processes include:

separation of hydrogen from its mixtures with nitrogen or hydrocarbons, pre-combustion

and post-combustion capture of CO 2 , nitrogen/oxygen separationand purification from

air, removal of acid gases and water from natural gas streams, organic vapor removal

from air or nitrogen streams, etc.

A chronological review reveals that the current state of gas separation membranes

230
is largely beholden to the notable breakthroughs in both material science and engineering

and also the membrane fabrication technology. Majority of the gas separation

membranes, in early days, were largely fabricated in the form of asymmetric composite

flat or hollow fiber membranes with a selective layer made from a neat material. As a

result, the application of molecular design using blend materials with synergistic

properties was not often reported.

The trend of fabrication of gas separation membranes was revolutionized by the

introduction of phase inversion process in 1960s [1]. The significance of this

technological breakthrough was more realized upon its successful extension to hollow

fibers as the more preferred geometry offering a large surface area per unit volume of

membrane material. The asymmetric structure of the hollow fibers not only could

significantly reduce the resistance against the transport of the species but also could offer

the membrane a good mechanical support.

Further advancement to the membrane fabrication technology was accomplished

by the advent of the co-extrusion technique that could be used for fabrication of dual-

layer hollow fiber membranes. This technological breakthrough was a satisfactory

response, at least at that time, to the high demands for suitable means to exploit the

potential advantages of the newly developed high performance (yet expensive) materials

for membrane separation applications. The new features offered by the dual-layer hollow

fiber membranes, in comparison to the traditional single-layer hollow fibers, created great

potentials for remarkable savings in membrane costs through substituting the functional

231
material with inexpensive alternatives in the support layer. In addition, the dual-layer

architecture provided a greater flexibility for tailoring the membrane morphology and

spinning parameters.

Nonetheless, the fabrication of the dual-layer hollow fibermembranes with

desirable characteristics is still not a trivial task and requires careful considerations from

the physicochemical properties of materials throughout the entire chain of dope

formulation, spinning and phase inversion process. However, the complexity often arises

in the simultaneous co-precipitation of two distinct materials with different formulations.

An ideal dual-layer hollow fiber membrane consisted of an ultrathin dense functional

layer supported by a porous substructure to enable the membrane withstanding high feed

gas pressures. Equally important to the above requirements is that the structure of the

dual-layer hollow fiber membrane must be free from any delamination at the interfacial

region where the two layers meet. It is shown that the delamination problems can often be

minimized or overcome through promoting the integrity of two layers by choosing

compatible materials or using spinnerets with modified designs [2,3].

As highlighted, one of the key benefits of the dual-layer hollow fiber technology

lies in the great opportunities provided for the exploitation of the wide range of high

performance materials for membrane-based separation applications. Liu et al. [4]

developed polyimide/polyethersulfone dual-layer hollow fiber membranes with the

enhanced performance for separation of CO 2 /N 2 and CO 2 /CH 4 . They also found that

chemical modification of the outer layer offered membranes with enhanced stability

232
against CO 2 -induced plasticization. As a result, the desirable performances obtained in

their hollow fibers promoted the viability of using expensive fluoropolyimide for gas

separation applications. In another work by Li et al. [5], dual-layer hollow fibers

membranes were developed by using zeolite-functionalized-PES as the outer layer

supported by the P84 co-polyimide. The fabricated hollow fibers exhibited enhanced

O 2 /N 2 and CO 2 /CH 4 selectivity in the order of around 10–20% compared to the PES

dense films. Jiang et al. [6] investigated the effect of various spinning parameters on

fabrication of defect-free dual-layer hollow fiber membranes using

Matrimid/polyethersulfone (PES) and fiber membranes and evaluated their performance

for O 2 /N 2 and CO 2 /CH 4 separations. They claimed the occurrence of a faster inter-layer

diffusion when the fibers were spun at elevated spinneret temperatures. Another similar

study was carried out by Ding et al. [7] who fabricated hollow fiber membranes using

Matrimid as the outer layer while polysulfone as the inner layer material for O 2 /N 2

separation.

Interestingly, almost all the prior studies on the development of the hollow fiber

membrane were aimed to the separation of H 2 /N 2 , O 2 /N 2 , CO 2 /N 2 , CO 2 /CH 4 but

H 2 /CO 2 . This can possibly be due to the fact that most polymeric membranes show a

relatively poor selectivity for separation of H 2 /CO 2 . The problem arises from the fact that

despite a noticeable difference in the molecular size (H 2 : 2.89Å vs. CO 2 : 3.30 Å) of the

two gases, the solubility of CO 2 in polymers is much higher than that of H 2 .

Consequently, a relatively low permselectivity is achieved owing to the presence of these

competing effects. In addition, CO 2 is a highly condensable gas and its sorption can

233
plasticize the polymeric membrane and increase the interstitial chain spaces. The CO 2 -

induced plasticization typically results in reduced efficiency through deterioration of the

size-exclusion mechanism. As a result, novel membrane materials and fabrication

techniques must be advanced to overcome the obstacles associated with the separation of

H 2 /CO 2 using polymeric membranes. We selected polymer blend technology as a

promising strategy to address these issues. So far, no study could be found on the use of

polymeric blends as the functional material for preparation of dual-layer hollow fiber

membranes consequently no one has reported the use of this type of dual-layer

membranes for H 2 /CO 2 separation. Only limited studies could be found that have used

polymeric blend systems in the form of single-layer hollow fiber membranes [8–10], but

none of them dealt with H 2 /CO 2 separation. This is quite surprising considering the

abundance of numerous studies that have extensively demonstrated the promising

features of polymer blends for a variety of membrane-based gas/liquid separation

applications [11–16]. In addition to material limitation, the other possible reasons for

underutilization of polymerblendsmayinclude the increase in complexity of the dual-layer

hollow fiber spinning process and the difficulty in prediction and control of various

parameters.

To our best knowledge, this is the first study on the application of miscible

polymer blends for fabrication of dual-layer hollow fiber membranes. The main objective

is to develop a novel approach that encompasses the synergistic advantages of the

polymer blending technology and the membrane fabrication process for high performance

gas separation applications. Indeed, this work can be regarded as an extension to our

234
earlier studies [11,12] that reported the development of high performance dense flat

membranes with tunable properties from selected polymer blends. Attempts were made in

this study to unravel the underlying mechanisms and engineering to scale up the flat

dense membranes to dual-layer hollow fiber membranes with desirable features. In this

respect, the effect of various important parameters such as dope flow rate, air-gap

distance and take-up speed on the morphological and gas transport properties of hollow

fiber membranes were investigated. The resultant hollow fiber membranes were analyzed

based on their separation performance for industrially important gas pairs. It is believed

that the knowledge and experience gained throughout this exemplary study can create

greater opportunities for further advancements in the field of membranes and membrane

materials for various gas separation applications.

6.2 EXPERIMENTAL

6.2.1. Materials

A polymer blend composed of Matrimid and poly(benzimidazole) (PBI) was

prepared and used as the outer-layer material for the fabrication of dual-layer hollow

fiber membranes. Matrimid is a commercially available polyimide with attractive

properties and gas separation performance [17-19]. Matrimid® 5218 (poly [3,3’4,4’-

benzophenone tetracarboxylic dianhydride and 5(6)-amino-1-(4’-aminophenyl-1,3-

trimethylindane)], BTDA-DAPI) was purchased from Vantico (Luxemburg) in the form

of powders. PBI (poly [2,2’-(1,3-phenylene)-5,5’-bibenzimidazole]) is also a high

performance polymer with a relatively high glass transition temperature (~418ºC) and

235
outstanding thermal and chemical stabilities. The PBI used in this study was supplied by

Hoechst Celanese Corporation, (NJ,USA) in the solution state with the following

formulation: 25.6 wt.% PBI, 72.4 wt.% DMAc and 2.0 wt.% LiCl.

Polysulfone (PSf) was selected as the inner supporting material. PSf is an

amorphous glassy polymer possessing a reasonably high T g and good stability against

environmental oxidation [20]. Polysulfone (UDEL) was provided by Solvay Advanced

Polymers, Singapore. The chemical structure of PBI, Matrimid and polysulfone is shown

in Figure 6.1.

(a)

(b)

(c)

Figure 6.1. The chemical structure of (a) Polybenzimidazole (PBI) (b) Matrimid® 5218

and (c) UDEL Polysulfone (PSf).

236
Besides polymers, various chemical reagents were also used. N-Methyl-2-

pyrolidone (NMP) and N,N-dimethylacetimide (DMAc) were procured from Merck and

Tedia Chemical Inc., respectively and used for the polymer dissolution. Methanol and

hexane (obtained from Merck) were used during the solvent exchange. p-

Xylylenediamine (PXDA) from Sigma-Aldrich (Germany) was used for chemical cross-

linking modification of the hollow fibers. Methanol was used as the medium for this

process. All chemical reagents were used as received.

6.2.2. Dope formulation and preparation

The outer-layer dope was prepared from a blend comprised of PBI and Matrimid

in 1:1 (weight ratio) composition. According to the literature [21,22], one potentialwayto

eliminate or at least minimize the defects in the dense-selective layer is to properly adjust

the dope formulation. Figure 6.2 demonstrates the trend of changes in viscosity of dope

solutions as a function of polymer concentration. Data were obtained using an ARES

rheometer at the shear rate of 10 s−1. This figure also includes the critical polymer

concentrations (c.p.c.) identified for the individual polymer dopes. Based on these results,

the concentration of 22 wt.% was selected for the preparation of outer-layer dope to

ensure achieving the desired morphology in this functional layer. On the other hand, the

concentration of inner dope was adjusted to 25 wt.% that is sufficiently lower than its

corresponding c.p.c. value (i.e., 29 wt.%). The objective was to minimize the effect of

substructure resistance on the membranes’ performance.

237
Both dope solutions were prepared by adding stipulated amounts of polymers into

the respective solvent followed by rigorous stirring until complete dissolution. Both dope

solutions were degassed in the syringe pumps for at least 24 h prior to spinning.

Figure 6.2. The trend of changes in dope viscosity as a function of polymer concentration

in the respective solvent. (♦ PBI/DMAc, ■ Matrimid/DMAc, ▲ PSf/NMP, ●

PBI:Matrimid (1:1)/DMAc).

6.2.3. Dual-layer hollow fiber spinning process

Dual-layer hollow fiber membranes were fabricated by adopting the dry-jet wet

spinning process. Figure 6.3 depicts the schematic diagrams for the spinning set-up and

the triple-orifice spinneret used in this study. As shown, three high precisions syringe

pumps (ISCO, USA) were employed to co-extrude dope solutions and the bore fluid

through the spinneret at specified flow rates. The solutions were filtered before flowing

238
into the spinneret channels. Table 6.1 provides a summary of the spinning parameters and

conditions. The bore fluid was a mixture of NMP/water (95/5 wt.%) whereas the external

coagulant was merely tap water.

Inner Bore fluid


channel

a b a
Outer
channel

c
2.4 mm
1.4 mm
c c
d

Figure 6.3. Schematics of dual-layer hollow fiber spinning set-up and triple-orifice

spinneret. (a) dope fluid tank and pump; (b) bore fluid tank and pump; (c) filter; (d)

spinneret; (e) coagulation bath and (f) take-up drum.

Table 6.1. The spinning parameters and conditions used in preparation of dual-layer

hollow fiber membranes.

239
Sample ID Outer dope flow Inner dope flow Bore fluid flow Air-gap Take-up speed Fiber collection state
rate (cm3.min-1) rate (cm3.min-1) rate (cm3.min-1) (cm) (cm.min-1)
A 0.20 2.00 1.00 0.0 327 Free fall
B 0.40 2.00 1.00 0.0 327 Free fall
C 0.20 2.00 1.00 1.0 371 Free fall
D 0.20 2.00 1.00 1.0 681 Elongational draw
X 0.20 2.00 1.00 2.0 449 Free fall
Y 0.20 2.00 1.00 3.0 560 Free fall
Outer dope: PBI/Matrimid (1:1), 22 (wt.%) in DMAc;
Inner dope: Polysulfone, 25 (wt.%) in NMP;
Spinning temperature: 25oC

Nascent hollow fibers were collected around a rotating drum after passing through the

coagulation bath. This drum was equipped with a gearing system enabling spinning at

various take-up speeds. As-spun hollow fiber membranes were cut into small pieces and

placed in clean water for about two days. The solvent exchange was carried out by

immersing hollow fibers in methanol for three times, each lasting for about 30 min. This

procedure was repeated using hexane. The fibers were finally dried in ambient

environment.

6.2.4. Silicone rubber coating and chemical modification

Silicone rubber coating was applied, using a solution composed of silicone rubber

(Sylgard-184) in hexane (2 wt.%), to seal the defects and imperfections of the

membranes. Hollow fibers, in the form of modules, were immersed in the solution for 2

min. For chemical modification, hollow fibers were immersed in the solutions of p-

xylylenediamine in methanol (10 wt.%) for stipulated periods of time. The fibers were

then removed and rinsed in fresh methanol in order to strip the unreacted molecules away

from the membrane’s surface.

6.2.5. Membrane characterization and gas permeation tests

240
The morphology of dual-layer hollow fiber membranes was examined using a

JEOL JSM-5600LV scanning electron microscope (SEM) and a JEOL JSM-6700F field

emission SEM. Samples were fractured in liquid nitrogen and then coated with platinum

before observation. The gas permeation properties of the hollow fiber membranes were

analyzed using the variable-pressure constantvolumemethod. This method is described in

details elsewhere [23]. The feed gas was supplied to the shell side of the membrane

module and the permeate side was connected to a vacuumed chamber. All pure gas

permeation tests were carried out at 35oC with the testing pressure of 10 atm for CH 4 and

CO 2 , and 3.5 atm for H 2 . However, during the CO 2 plasticization experiments, the

testing pressure was raised up to about 15 atm. The permeance, P/L, was determined

using the following equation:

𝑃𝑃 𝑄𝑄 𝑄𝑄
= = (6.1)
𝐿𝐿 𝐴𝐴∆𝑃𝑃 𝑛𝑛𝑛𝑛𝑛𝑛𝑛𝑛 ∆𝑝𝑝

where P is the permeability of separating layer (Barrer), L is the thickness of the apparent

dense-selective layer (cm) and Q is the pure gas flux (cm3.s-1). On the right hand side, n

specifies the number of fibers in the testing module, D is the outer diameter of the fiber

(cm), l is the effective length of the fibers (cm), and Δp is the gas pressure difference

across the membrane (cm Hg). The permeance is reported in GPU (1 GPU = 1×106 cm3

(STP)/cm2.s.cm Hg).

The ideal separation factor (α A/B ) was calculated using the following equation:

𝑃𝑃 𝑃𝑃
𝛼𝛼𝐴𝐴⁄𝐵𝐵 = ( )𝐴𝐴 /( )𝐵𝐵 (6.2)
𝐿𝐿 𝐿𝐿

6.3 RESULTS AND DISCUSSION

241
6.3.1. Morphological characteristics of dual-layer hollow fiber membranes

Figure 6.4 illustrates the typical morphology of as-spun dual-layer hollow fiber

membranes (sample B). The analysis of SEM images revealed the presence of no major

differences in overall morphology of fabricated hollow fibers, from sample A to sample

Y. However, two distinct morphologies could be seen for the inner and outer layers.

According to Figures 6.4 (d) and (e), the outer layer of the hollow fibers was in the form

of an asymmetric structure; constructed of spongy-like cells surrounded by a thin dense-

selective layer at the outermost edge. No macrovoid could be found in the crosssection

morphology of the outer layer. On the other hand, the inner supporting layer was

considerably larger in thickness with a crosssection occupied by open cell pores and

disrupted by finger-like macrovoids.

d c

X6,000 2 μm X2,500 10 μm

e a b

X10,000 1 μm X500 50 μm

f g

X 25,000 1 μm X 1000 10 μm

242
Figure 6.4. SEM images illustrating various morphological aspects of a typical dual-layer

hollow fiber membrane (sample B): (a) membrane cross-section; (b) membrane sector; (c

and d) inner and outer layers; (e) interface region; (f) outer surface of the outer layer and

(g) inner surface of the inner layer.

The morphological difference between the inner and outer layers can be explained

by taking into account the chemistry of dope components and the coagulation agent.

Table 6.2 provides the solubility parameter values for the polymers, solvents and the

coagulant. According to this table, one could find a negligible difference in the solubility

parameters of NMP and DMAc. On this basis, it was expected for these two solvents to

exhibit a similar interaction with water, especially in terms of diffusion coefficient during

the phase inversion process. On one hand, the data indicate the presence of a relatively

small difference in the solubility parameters of water and the polymer blend compared to

that of water and polysulfone. This results in the slower precipitation rate for the outer

layer. In addition, the characteristics of strong hydrogen bonding provided by PBI and

high viscosity of the outer-layer dope may also inhibit macrovoid formation at the outer

layer [20,27,28]. On the other hand, the greater solubility parameter difference between

water and polysulfone, the low viscosity of the inner dope, and hydrophobicity of

polysulfone were hold responsible for the increased tendency of macrovoid formation in

the structure of inner layer through unbalanced localized stresses, rapid rate of solvent

exchange and the accelerated precipitation rate [20,28–32].

243
Table 6.2. Solubility parameters of the solvents, the coagulation agents and the polymers,

coagulation agents and polymers.*

Component Solubility parameter,


δt (J.cm-3)1/2

NMP 22.9
DMAc 22.7
Water 47.8
PBI 23.3
Matrimid 22.9
Polysulfone 18.0
*Data obtained and calculated using Refs. [24-26].

It could be found from Figure 6.4(e) that as-spun hollow fibers were free from

any delamination at the interface between the inner and outer layers. The delamination-

free structure is usually regarded as one of the key pre-requisites in the fabrication of

dual-layer hollow fiber membranes to ensure consistency in performance of gas

separation membranes throughout the service life. This is due to the fact that delaminated

membranes are highly susceptible to failure and structural collapse at high feed pressures

[2,3,33]. As a result, this subject has attracted special attentions among membrane

specialists. The delamination-free structure of the newly developed dual-layer hollow

fibers can be attributed to (1) good miscibility in molecular level among PBI and

Matrimid polymer chains provided by the strong hydrogen bonding [34–36] and (2) inter-

diffusion of solvents and polymeric chains at the interface driven by the chemical

potential differences and close solubility parameters between two dopes. These unique

combinations can easily eliminate the necessity for using sophisticated spinneret designs

in fabrication of delamination-free hollow fiber membranes [33].

244
According to Figure 6.4(f), the outer skin of the outer layer was essentially dense

with some minor defects scattered over the areas of observation. The dense-selective

layer in asymmetric membranes can often be defective mainly due to the rapid

coalescence of polymermolecules and irregular packing of kinked polymer chains [37].

Number of approaches such as proper dope formulation with favorable rheological

properties [28,37,38], using dual coagulant bath [39], moderate elongation take-up rate

[40], modification of the spinneret design, heat treatment, silicone rubber coating, etc.

have been proposed in order to produce defect-free hollow fibers for gas separation. To

simplify the dope composition and spinning process, silicon rubber coating was selected

to seal the membrane defects before conducting the gas permeation tests in this study.

6.3.2. The effect ofspinning type and air-gap distance on the morphology and gas

transport properties of dual-layer hollow fiber membranes

Figure 6.5 shows the cross-section morphology of dual-layer hollow fiber

membranes spun at various air-gap distances. Higher magnification pictures are provided

below each image. A consistent similarity could be found in the overall morphology of

these hollow fibers.

Hollow fiber membranes were examined for their gas transport properties and

separation performance. Table 6.3 provides the results of gas permeation tests for the

fibers spun at different air-gaps. Data in this table are categorized into two groups of

before and after the application of silicone rubber coating. A relatively poor separation

245
performance was found for the pristine membranes indicating the possible presence of

defective sites in the membranes. Therefore, silicone rubber coating was applied in order

to exclude the effects of defects and imperfections in the analysis of the performance of

the membranes. A significant decline was found in the gas permeance of silicon rubber

coated hollow fibers upon switching from wet to dry-jet spinning. This could possibly be

attributed to the effects of elongational stretch and die swell (will be discussed later).

However, it was noticed that the provision of air-gap resulted in the formation of

amembrane with considerably higher separation performance for H 2 /CO 2 with selectivity

of 8.96 at an air-gap of 1 cm. The selectivity was further improved to about 11.11 for the

246
Figure 6.5. The cross-section morphology of dual-layer hollow fiber membranes spun at different air gap distances.

Sample A (Air gap: 0 cm) Sample C (Air gap: 1 cm) Sample X (Air gap: 2 cm) Sample Y (Air gap: 3 cm)

X80 200 μm

X500 50 μm

247
membrane prepared in 3-cm air-gap. However, spinning with air-gap had a negative

effect on the CO 2 /CH 4 selectivity of membranes.

Table 6.3. The effect of air gap distance on the gas permeance and separation

performance of dual-layer hollow fiber membranes.

Sample ID Air-gap Permeance (GPU) Selectivity (α)


(cm)
H2 CH4 CO2 H2/CO2 CO2/CH4
(±max4.5%) (±max2%) (±max3.2%) (±max3.1%) (±max3.5%)
Before silicone rubber coating
A 0.0 43.22 1.458 7.34 5.89 5.04
C 1.0 30.29 3.542 4.90 6.18 1.38
X 2.0 36.51 2.127 5.48 6.66 2.58
Y 3.0 38.67 1.850 5.65 6.85 3.05

After silicone rubber coating


A 0.0 31.55 0.223 4.37 7.22 19.60
C 1.0 17.84 0.199 1.99 8.96 10.01
X 2.0 26.45 0.273 2.49 10.62 9.12
Y 3.0 29.26 0.328 2.63 11.11 8.03
Spinning Conditions:
Outer dope: PBI/Matrimid (1:1), 22 (wt.%) in DMAc;
Inner dope: Polysulfone 25 (wt.%) in NMP;
Outer dope flow rate: 0.20 cm3.min-1;
Inner dope flow rate: 2.00 cm3.min-1;
Bore fluid flow rate: 1.00 cm3.min-1;

The effect of spinning type (i.e., wet-spun vs. dry-jet) on the microstructural and

transport properties of hollow fiber membranes can be explained by careful examination

of key spinning parameters and their role in phase inversion process. It is known that

flexible polymer coils are fully extended while present in a good solvent but they tend to

contract upon the introduction of a non-solvent [41]. The contraction rate is largely

dependent on the solubility parameter of the system components as well as the amount of

non-solvent. For the fibers spun at zero air-gap distance, the fast and vigorous diffusion

of the water molecules (strong coagulant) results in the sudden contraction of polymer

chains with minimal chance for any orientation or configurations. As a result, the

248
population of the fine pores and free volumes in the membrane’s microstructure are

expected to grow due to the random entanglement of polymer chains [41–44]. This

especially occurs during the solvent exchange process during which the solvent and non-

solvent molecules entrapped among the frozen polymer coils are extracted.

On the other hand, the phase inversion process for the fibers fabricated through

dry-jet wet spinning follows a slightly different pathway [44,45]. In this case, the

properties of hollow fibers are mostly governed by the primary phase inversion in the air-

gap region and also polymer chain orientation. It is documented that moisture-induced

phase separation results in the formation of a relatively dense layer at the outermost skin

of the hollow fibers [44]. The polymer chains in this dense layer are highly oriented due

to the gravity forces induced by the weight of the nascent fiber traveling in the air-gap.

The presence of such a highly oriented and dense skin layer can be regarded as the main

cause for the reduced gas permeance but enhanced gas pair selectivity of the membrane.

In addition, further evolution of the membrane microstructure and the coarsening process

at regions beneath the skin layer is also expected to be affected due to the substantial

reduction in the rate of mass transfer between the solvent and non-solvent (in the

coagulation bath) caused by the skin layer.

It is also interesting to note that a further increase in air-gap distance from 1 cm to

3cm brings about improvement in gas permeance, more noticeably toward H 2 among all.

In terms of separation performance, a larger air-gap further improves the H 2 /CO 2

selectivity of membranes, but at the same time decreases the CO 2 /CH 4 selectivity. One

249
possible reason for the increase in gas permeance is the relatively enlarged interstitial

space of around H 2 size between the polymer chains induced by the weight of the nascent

fiber travelling in the air-gap region. However, seemingly the effect of air-gap distance is

quite limited in provision of more spacious gap between the polymer chains that can

favor passage of all gas molecules. The pronounced improvement in hydrogen permeance

is owing to its considerably smaller kinetic diameter among the others. Consequently, a

higher increase in H 2 permeance compared to that of CO 2 results in an overall increased

H 2 /CO 2 selectivity as a function of air-gap distance.On the other hand, a less enhanced

permeance of CO 2 compared to that of CH 4 results in a less effective discrimination

between CO 2 and CH 4 molecules.

Another possible reason for the effect of air-gap distance can be the increased

intensity of orientation and degree of chain packing, simply due to the larger gravitational

forces induced to the nascent fiber. The occurrence of such phenomena at high air-gaps

provides more resistance to the diffusion of water vapor during the vapor-induced phase

inversion, thus retarding the progress of the precipitation front into the underlying

structure. As a result, fibers spun at the higher air-gaps may possess a thinner dense skin

layer and subsequently promoted gas permeance.

It is noteworthy that the hollow fiber membranes spun at the air-gap distance of 3

cm exhibited the highest separation performance for H 2 /CO 2 (α= 11.11) among all,

exceeding the intrinsic value reported for the dense flat membrane made from the same

outer-layer material (see Table 6.4) [11]. This can possibly be attributed to the combined

250
effects of shear and elongational forces in changing the interaction and micro-

configuration of PBI and Matrimid polymer chains to the state that is much different from

the one dominant in dense flat membranes. In addition, we believe that the presence of

PBI in the polymer blend plays an inevitable role in promoting the overall diffusivity

selectivity of the membranes through (1) the modification of free volumes distribution,

(2) formation of a higher chain packing density, and (3) causing hindrance in segmental

mobility of polymer chains. Occurrence of these effects can be associated to the special

molecular structure of PBI, presence of aromatic rings in the backbone (assistance in

formation of charge transfer complex) and also functional groups capable of forming

strong hydrogen bonding with Matrimid chains. It should be mentioned that despite the

good separation performance obtained in pure gas experiments of membranes, the mixed

gas selectivity (feed gas: H 2 –CO 2 (1:1)) could reach to only about 3.17 possibly owing to

the competitive sorption dominated by CO 2 molecules.

Table 6.4. The intrinsic gas transport properties of the dense flat membrane made from

the polymer blend and respective individual polymers.a

Membrane constituents Permeability (Barrer) Ideal Selectivity

H2 CH4 CO2 H2/ CO2 CO2/ CH4


Matrimid 27.16 0.210 7.00 3.88 33.33
PBI/Matrimid (1:1) 13.06 0.045 2.16 6.05 48.00
PBI 0.6 0.0018 0.16 3.75 88.88
1 Barrer = 1*10-10 cm3(STP) cm/cm2 s cm Hg
Testing Temperature: 35ºC
a Data reproduced from Ref. [11]

251
6.3.3. The effect of elongational drawing on the transport properties of dual-layer

hollow fiber membranes

Table 6.5 provides the results of gas permeance and separation factor for

membranes spun at the free-fall state and those spun with elongational drawing, before

and after the application of silicone rubber coating. The results indicate that hollow fiber

spun with elongational drawing exhibits an improved permeance for both hydrogen and

carbon dioxide, but a reduced permeance for CH 4 . This trend of changes consequently

brings about improvements in separation performance of membranes for CO 2 /CH 4 with

enhanced selectivity of 41.81, but reduces the separation performance for H 2 /CO 2 from

8.96 down to 6.79. The membrane was also tested using a mixed gas ofCO 2 –CH 4 (1:1) as

the feed (at 10 atm and ambient temperature) and CO 2 /CH 4 selectivities as high as 26.63

could be achieved.

Table 6.5. The effect of application of elongational drawing on the gas permeance and

separation performance of dual-layer hollow fiber membranes.

Sample ID Fiber collection state Permeance (GPU) Selectivity, α

H2 CH4 CO2 H2/CO2 CO2/CH4


(±max3.8%) (±max1.5%) (±max3.5%) (±max3.5%) (±max2.9%)
Before silicone rubber coating
C Free fall 30.29 3.542 4.90 6.18 1.38
D Elongational draw 39.00 0.527 5.76 6.77 10.93

After silicone rubber coating


C Free fall 17.84 0.199 1.99 8.96 10.01
D Elongational draw 32.66 0.115 4.81 6.79 41.81
Spinning Conditions:
Outer dope: PBI/Matrimid (1:1), 22 (wt.%) in DMAc;
Inner dope: Polysulfone 25 (wt.%) in NMP;
Outer dope flow rate: 0.20 cm3.min-1
Inner dope flow rate: 2.00 cm3.min-1;
Bore fluid flow rate: 1.00 cm3.min-1;
Air gap: 1.0 cm;

252
A comparison between the gas permeance values in Tables 6.3 and 6.5 revealed

that the application of elongational drawing is more effective than raising the air-gap

distance in altering the microstructure of the membranes, degree of chain packing and

molecular orientation in the dense skin layer. Therefore, explanations provided in the

previous section can be extended to analyze the effect of elongational drawing on the

performance of the membranes. On this basis, we would like to hypothesize that in

addition to the orientation induced to the polymer chains, the application of elongational

drawing could effectively change the microstructure of the membranes, especially at the

skin layer, to be constructed of tiny pores/free volumes elongated in the direction parallel

to the induced forces. These elongated pores/free volumes are expected to provide a

better pathway for the passage of small and linear molecules (e.g., H 2 and CO 2 ), but to

hinder the passage of molecules possessing polyhedral structures (e.g., CH 4 ).

The above phenomena can clearly demonstrate that spinning at high air-gaps or

with elongational drawing can effectively be employed to fabricate hollow fiber

membranes for specific separation applications. In other words, membrane spun at a high

air-gap distance ismorefavorable for H 2 /CO 2 separation (α H2/CO2 = 11.11), whereas the

one spun with elongational drawing is more suitable for CO 2 /CH 4 separation (α CO2/CH4 =

41.81). It also implies that despite the similarity in the nature of two mechanisms, the

evolution of membrane microstructure during the phase inversion process may follow an

entirely different pathway that is largely governed by the characteristics of the membrane

materials (e.g., rheology, hydro-phobicity/philicity, and coil size), state of

macromolecular orientation, travel time in air-gap region and other important parameters.

253
6.3.4. The effect of outer-layer dope flow rate on the transport properties of dual-layer

hollow fiber membranes

It is very obvious that using a low dope flow rate in fabrication of dual-layer

hollow fiber membranes can often be very advantageous especially in terms of the

savings in material costs. However, membranes with an extremely thin functional layer

tend to be more vulnerable to skin defects and other possible imperfections. For instance,

an unbalanced or partial coverage of the outer layer over the supporting layer can easily

destroy the membrane’s functionality. On the other hand, aside from its economical

drawbacks, spinning with excessive outer dope flow rates can result in membranes with

reduced flux and productivity. Therefore, due to its importance, attempts were made in

this study to examine the effect of outer dope flow rate on the gas transport and

separation performance of dual-layer hollow fiber membranes.

Table 6.6 provides the data on the gas permeance and selectivity of hollow fiber

membranes spun at different outer dope flow rates. It was found from the results that an

increase in outer dope flow rate results in considerable reductions in gas permeance of the

membranes. It is also interesting to note amore pronounced drop in membrane’s

permeance toward CH 4 compared to the other gases. For instance, sample B exhibits a

CH 4 permeance of 0.090 GPU which is about 16.2 times lower than the vale for its

counterpart membrane (i.e., sample A). The occurrence of similar phenomena has been

reported in literature in which single-layer hollow fiber membranes spun with higher

dope flow rates exhibited a thicker and denser skin layer with a tighter substructure [46].

254
Table 6.6. The effect of variation in outer dope flow rates on the gas permeance and

separation performance of wet-spun dual-layer hollow fiber membranes.

Sample ID Outer dope flow rate Permeance (GPU) Selectivity, α


(cm3.min-1)
H2 CH4 CO2 H2/CO2 CO2/CH4
(±max3.8%) (±max2.1%) (±max3.1%) (±max3.0%) (±max2.9%)
Before silicone rubber coating
A 0.20 43.22 1.458 7.34 5.89 5.04
B 0.40 22.07 0.090 4.22 5.23 46.84

After silicone rubber coating


A 0.20 31.55 0.223 4.37 7.22 19.60
B 0.40 18.89 0.085 2.97 6.36 34.82
Spinning Conditions:
Outer dope: PBI/Matrimid (1:1), 22 (wt.%) in DMAc;
Inner dope: Polysulfone 25 (wt.%) in NMP;
Inner dope flow rate: 2.00 cm3.min-1;
Bore fluid flow rate: 1.00 cm3.min-1;

The above trends of changes in gas permeance subsequently affect the separation

performance of hollow fiber membranes. In fact, an increase in outer dope flow rate

results in substantial improvements in CO 2 /CH 4 selectivity reaching the values as high as

46.84. However, this is accompanied with a reduction in H 2 /CO 2 separation performance.

The mixed gas selectivity of this membrane for CO 2 /CH 4 is 35.63 when using a mixed

gas of CO 2 –CH 4 (1:1) as the feed at 10 atm and ambient temperature.

The observed trends can possibly be attributed to the larger gas transport

resistance of membranes due to the formation of a thicker dense skin layer. In addition, in

the course of spinning, the polymer dope flows through an annular channel within which

it receives shear stresses through its contacts with the surrounding walls. Due to the fixed

geometry of the channel, a larger dope flow rate translates into a larger axial velocity and

consequently induction of a larger shear stress to polymer chains [47–49]. Therefore, a

dense skin layer is formed possessing a higher degree of molecular orientation and chain

255
packing and evolution of a microstructure that can considerably enhance the distinction

between CO 2 and CH 4 gas molecules. The resultant membrane exhibits promising

performance for natural gas enrichment by having the enhanced CO 2 /CH 4 selectivity as

high as 46.84 without any post treatment.

As anticipated, application of silicone rubber coating on membranes spun with

outer dope flow rate of 0.40 cm3 min-1 resulted in further declines in gas permeance

through sealing the membrane’s defect and also non-selective pores. In addition, silicone

rubber coating was found more effective in improving the H 2 /CO 2 separation of hollow

fiber membranes from about 5.23 up to 6.36.

6.3.5. The analysis of the anti-plasticization behavior of dual-layer hollow fiber

membranes

One of the important factors in evaluation of the long-term performance and

reliability of the gas separation polymeric membranes is their stability against the

plasticization. It is often reported that the plasticization is caused by highly polar gases

such as CO 2 molecules especially when present at relatively high partial pressures. The

incipient point for the plasticization is known as plasticization pressure (P plast ).

Plasticization phenomena are deteriorative to membrane performance through dissolution

of gas molecules into the polymer matrix followed by the disruption of chain packing and

enhancement of their inter-segmental mobility. In this respect, dual-layer hollow fiber

membranes were examined in terms of their anti-plasticization characteristics. This was

achieved through monitoring the trend of changes in the CO 2 permeance of the

256
membranes at various feed gas pressures up to 15 atm. This was the highest pressure that

membranes could withstand. The possible reason for the collapse of membranes at

pressures beyond 15 atm could be the relatively high D/Δh ratio (D: membrane diameter;

Δh: wall thickness) and its detrimental effect on the mechanical stability of the hollow

fiber membrane as also reported elsewhere [50,51].

It can be found from the results (shown in Figure 6.6) that upon increase in feed

gas pressure, the CO 2 permeance of the hollow fibers is gradually decreased. This trend

can be attributed to the saturation of Langmuir sites. However, it also indicates the

prevalence of a good resistance against plasticization in the fabricated membranes.

Figure 6.6. The trend of changes in CO 2 permeance as a function of upstream pressure

for silicone rubber coated dual-layer hollow fiber membranes.

Y
CO2 permeance (GPU)
CO2 permeance (GPU)

Feed gas pressure (atm)

257
According to the literature [52,53], dense flat membranes prepared from Matrimid tend to

plasticize at pressures around 12–15 atm; witnessed by an upward reflection in the

membrane’s permeability. This is in addition to the fact that there exist a number of

studies highlighting the severity of the plasticization in asymmetric hollow fibers

membranes compared to the ones in flat geometry [54–56]. This can mainly be attributed

to the higher vulnerability of the loosely packed structure of the thin dense-selective layer

in hollow fiber membranes toward swelling upon the sorption of CO 2 molecules. In

overall, the anti-plasticization behavior of dual-layer hollow fiber membranes could be

attributed to both (1) an improved microstructural properties evolved in the dense skin

layer and (2) inevitable role of PBI macromolecules in providing an integrated

interpenetration networks through establishment of the strong hydrogen bonding

interactions with Matrimid chains.

6.3.6. The effect of chemical modification on the gas transport properties of dual-layer

hollow fiber membranes

Over the years, various kinds of chemical modification techniques have

extensively been used as effective tools for altering the properties of polymeric materials

andmembranesprepared thereof [12,57–60]. For instance, it was demonstrated in our

earlier studies [11,12] that dense flat membranes prepared from PBI/Matrimid blends

underwent significant improvements in separation performance upon application of

cross-linking modification. In this respect, the same approach was extended to this study

in order to examine its viability and effectiveness for hollow fiber membranes. p-

Xylylene diamine (PXDA) is known to be reactive to the carbonyl functional groups

258
present in the chemical structure of Matrimid and thus was selected as the cross-linking

agent for modification of the membranes. Data in Table 6.7 show that the chemical cross-

linking modification greatly affected the properties of the selected membrane even upon

treatment for a very short period of time. Similar phenomena have been reported in the

case of hollow fiber membranes made from 6FDA-2,6-DAT [57]. The gas permeance of

the hollow fiber was reduced, on average about 84–93%, after cross-linking for only 30 s

and resulted in obtaining membranes with improved performance for H 2 /CO 2 separation.

This may clearly highlight the effective role of cross-linking modification on promoting

the diffusivity selectivity through formation of a tighter chain packing, restriction in

mobility and interstitial spaces among the polymer chains present in the dense functional

layer of the membrane. It was also noted that the CO 2 /CH 4 separation performance of the

membranes was negatively affected. Similarly, this could be attributed to the diminished

size-exclusion capability of the membranes for distinguishing the gas molecules with

close kinetic diameters.

Table 6.7. The effect of chemical cross-linking modification and its period on the gas

permeance and separation factor of dual-layer hollow fiber membranes.

Permeance (GPU) Selectivity, α


Sample ID and
Cross-linking period H2 CH4 CO2 H2/CO2 CO2/CH4
(±max3.7%) (±max2.0%) (±max3.1%) (±max3.0%) (±max2.3%)
Y-0 Sec (pristine sample) 38.67 1.850 5.65 6.85 3.05
Y-30 Sec 6.10 0.187 0.421 14.49 2.25
Y-1.0 min 5.13 0.173 0.368 13.94 2.13
Y-5.0 min 0.562 0.035 0.061 9.21 1.74
Spinning Conditions:
Outer dope: PBI/Matrimid (1:1), 22 (wt.%) in DMAc;
Inner dope: Polysulfone, 25 (wt.%) in NMP;

259
In addition, data in Table 6.7 also suggest that prolonging the cross-linking period

had undesirable effects on both gas permeance and selectivity. This could be attributed to

the intensive effect of cross-linking modification on the membrane microstructure

through excessive reduction in the size of free volumes and interstitial distances between

the polymer chains that could create similar degree of hindrance toward the transport of

gas molecules. Interestingly, a comparison between these data and other reported results

[11] on cross-linking of similar membrane materials could imply the important role of

membrane geometry in effectiveness of the modification process. Perhaps, one may be

able to relate the greater yield of cross-linking in hollow fibers, compared to that in dense

flat membranes, to the higher surface area per unit volume of membranes, easier swelling

of the thinner dense skin layer and its promoting effect on the penetration of the cross-

linking agent into the membrane structure.

6.4 CONCLUSIONS

Following conclusions could be drawn from this study:

1. Novel and high performance dual-layer hollow fiber membranes were

successfully developed by using a polymer blend as the functional material. The

emergence of the two distinct morphologies for inner and outer layers was attributed to

various factors including the characteristics of dope constituents, their composition and

the materials’ chemistry in the course of the phase inversion.

2. The delamination-free structure of the membranes was mainly attributed to the

good miscibility among PBI and Matrimid chains provided by the strong hydrogen

260
bonding, the interpenetration of the solvents and polymeric chains at the interface via

interdiffusion and close solubility parameters between the inner and outer-layer dopes.

3. The separation performance of the hollow fiber membranes was found to be

highly influenced by the type of spinning (i.e., wet-spun/dry-jet) and also the air-gap

distance. The enhanced permeance and selectivity of the membranes, especially in case of

H 2 /CO 2 , at elevated air-gaps could be attributed to the enlarged interstitial spaces

between the polymer chains and the increased intensity in orientation degree of chain

packing induced by the gravitational forces. These were in addition to the possible effect

of shear and elongational forces on the interaction and configuration of the blend chains.

4. The application of the elongational drawing influenced the transport properties

of the dual-layer membranes in almost a similar manner to the increase in the air-gap

distance, but with greater impacts on CO 2 /CH 4 separation performance. This was mainly

attributed to the additional effects of the elongational drawing on the membrane’s outer-

lay microstructure through creating elongated pores in the direction parallel to the

induced forces with resultant hindrance to the passage of methane molecules with

polyhedral structure.

5. The increase in outer dope flow rate could considerably improve the CO 2 /CH 4

separation performance of the membranes, through at the expense of the permeance. The

observed effects were attributed to the formation of a thicker dense skin layer and also

higher degree of molecular orientation and chain packing at the outermost skin induced

by high shear stress.

6. Majority of the hollow fiber membranes exhibited a very good resistance

toward the CO 2 -induced plasticization notably attributed to both the improved

261
microstrucural properties evolved in the dense skin layer and also inevitable role of PBI

macromolecules in providing an integrated interpenetration networks through strong

hydrogen bonding interactions with Matrimid chains.

7. Chemical cross-linking modification was found effective in promoting the

diffusivity selectivity of dual-layer membranes only if applied for a very short period of

time. As a result, H 2 /CO 2 selectivity of dual-layer hollow fiber membranes could be

boosted up to about 14.49.

6.5 REFERENCES

[1] S. Loeb, S. Sourirajan, Sea water demineralization by means of an osmotic

membrane, Adv. Chem. Ser. 38 (1963) 117.

[2] D.F. Li, T.S. Chung, R. Wang, Morphological aspects and structure control of dual-

layer asymmetric hollow fiber membranes formed by a simultaneous co-extrusion

approach, J. Membr. Sci., 243(2004) 155.

[3] N. Widjojo, T.S. Chung, W.B. Krantz, A morphological and structural study of

Ultem/P84 copolyimide dual-layer hollow fiber membranes with delamination-free

morphology, J. Membr. Sci., 294 (2007) 132.

[4] Y. Liu, T.S. Chung, D.F. Li and R. Wang, High-selectivity anti-plasticization cross-

linked polyimide/polyethersulfone dual-layer hollow fiber membranes, Ind. Eng.

Chem. Res., 42 (2003) 1190.

[5] Y. Li, T.S. Chung, Z. Huang, S. Kulprathipanja, Dual-layer polyethersulfone

(PES)/BTDA-TDI/MDI co-polyimide (P84) hollow fiber membranes with a

262
submicron PES–zeolite beta mixed matrix dense-selective layer for gas separation, J.

Membr. Sci., 277 (2006) 28.

[6] L.Y. Jiang, T.S. Chung, D.F. Li, C. Cao, S. Kulprathipanja, Fabrication of

Matrimid/polyethersulfone dual-layer hollow fiber membranes for gas separation, J.

Membr. Sci., 240 (2004) 91.

[7] X. Ding, Y. Cao, H. Zhao, L. Wang, Q. Yuan, Fabrication of high performance

Matrimid/polysulfone dual-layer hollow fiber membranes for O 2 /N 2 separation, J.

Membr. Sci., 323 (2008) 352.

[8] T.S. Chung, Z.L. Xu, Asymmetric hollow fiber membranes prepared from miscible

polybenzimidazole and polyetherimide blends, J. Membr. Sci., 147 (1998) 35.

[9] G.C. Kapantaidakis, G.H. Koops, High flux polyethersulfone–polyimide blend hollow

fiber membranes for gas separation, J. Membr. Sci., 204 (2002) 153.

[10] G.C. Kapantaidakis, G.H. Koops, M. Wessling, Effect of spinning conditions on the

structure and the gas permeation properties of high flux polyethersulfonepolyimide

blend hollow fibers, Desalination, 144 (2002) 121.

[11] S.S. Hosseini, M.M. Teoh, T.S. Chung, Hydrogen separation and purification in

membranes of miscible polymer blends with interpenetration networks, Polymer, 49

(2008) 1594.

[12] S.S. Hosseini, T.S. Chung, Carbon membranes from blends of PBI and polyimides

for N2/CH4 and CO2/CH4 separation and hydrogen purification, J. Membr. Sci.,

328 (2009) 174.

263
[13] Y.K. Kim, H.B. Park, Y.M. Lee, Carbon molecular sieve membranes derived from

thermally labile polymer containing blend polymers and their gas separation

properties, J. Membr. Sci., 243 (2004) 9.

[14] Y.K. Kim, H.B. Park, Y.M. Lee, Gas separation properties of carbon molecular sieve

membranes derived from polyimide/polyvinylpyrrolidone blends: effect of the

molecular weight of polyvinylpyrrolidone, J. Membr. Sci., 251 (2005) 159.

[15] P.S.O. Patrício, J.A. de Sales, G.G. Silva, D. Windmöller, J.C. Machado, Effect of

blend composition on microstructure, morphology, and gas permeability in

PU/PMMA blends, J. Membr. Sci., 271 (2006) 177.

[16] S.H. Kim, D. Kim, D.S. Lee, Gas permeation behavior of PS/PPO blends, J. Membr.

Sci., 127 (1997) 9.

[17] A.B. Fuertes, D.M. Nevskaia, T.A. Centeno, Carbon composite membranes from

Matrimid and Kapton polyimides for gas separation. Micro. Meso. Mater., 33 (1999)

115.

[18] P.S. Tin, T.S. Chung, S. Kawi, M.D. Guiver, Novel approaches to fabricate carbon

molecular sieve membranes based on chemical modified and solvent treated

polyimides, Micro. Meso. Mater., 73 (2004) 151.

[19] L.Y. Jiang, T.S. Chung, R. Rajagopalan, Dual-layer hollow carbon fiber membranes

for gas separation consisting of carbon and mixed matrix layers, Carbon, 45 (2006)

166.

[20] R.E. Kesting, A.K. Fritzsche, Polymeric Gas Separation Membranes, John Wiley &

Sons, New York, 1993.

264
[21] T.S. Chung, S.K. Teoh, X.D. Hu, Formation of ultrathin high-performance

polyethersulfone hollow-fiber membranes, J. Membr. Sci., 133 (1997) 161.

[22] C. Cao, R. Wang T.S. Chung, Y. Liu, Formation of high-performance 6FDA-2,6-

DAT asymmetric composite hollow fiber membranes for CO2/CH4 separation, J.

Membr. Sci., 209 (2002) 309.

[23] Y. Li, C. Cao, T.S. Chung, K.P. Pramoda, Fabrication of dual-layer polyethersulfone

(PES) hollow fiber membranes with an ultrathin dense-selective layer for gas

separation, J. Membr. Sci., 245 (2004) 53.

[24] D.W. Van Krevelen, Properties of Polymers, Elsevier, Amsterdam, 1990.

[25] A.F.M. Barton, Solubility Parameters and Other Cohesion Parameters, CRC Press,

1983.

[26] C.M. Hansen, Hansen Solubility Parameter, A User’s Handbook, CRC Press, 1999.

[27] R. E. Kesting, A. K. Fritzsche, A. Cruse, M. K. Murphy, A. C. Handermann, R. F.

Malon, M. D. Moore, The second generation polysulfone gas-separation membranes.

I: The use of Lewis acid: base complexes as transient templates top increase free

volume, J. Appl. Polym. Sci., 40 (1990) 1557.

[28] T. S. Chung, E. R. Kafchinski, R. Vora, Development of a defect-free 6FDA-durene

asymmetric hollow fiber and its composite hollow fibers, J. Membr. Sci., 88 (1994)

21.

[29] C. A. Smolders, A. J. Reuvers, R. M. Boom, I. M. Wienk, Microstructures in phase-

inversion membranes. Part 1. Formation of macrovoids, J. Membr. Sci., 73 (1992)

259.

265
[30] H. Strathmann, K. Kock, P. Amar, R. W. Waker, The formation mechanism of

asymmetric membranes, Desalination, 16 (1975) 179.

[31] I. Cabasso, E. Klein, J. K. Smith, Polysulfone hollow fibers. II. Morphology, J.

Appl. Polym. Sci. 21 (1977) 165.

[32] K. Y. Wang, T, Matsuura, T. S. Chung, W. F. Guo, The effect of flow angle and

shear rate within the spinneret on the separation performance of poly(ethersulfone)

(PES) ultrafiltration hollow fiber membranes, J. Membr. Sci., 240 (2004) 67.

[33] D.F. Li, T.S. Chung, R. Wang, Y. Liu, Fabrication of fluoropolyimide/polyether-

sulfone (PES) dual-layer asymmetric hollow fiber membranes for gas separation, J.

Membr. Sci. 198 (2002) 211.

[34] G. Guerra, S. Choe, D. J. Williams, F. E. Karasz, W. J. MacKnight, Fourier

transform infrared spectroscopy of some miscible polybenzimidazole/polyimide

blends. Macromolecules 21 (1988) 231.

[35] M. Jaffe, P. Chen, E. W. Choe, T. S. Chung, S. Makhija, High performance polymer

blends, in "Advances in Polymer Science" edited by P. Hergenroth, vol. 117,

Springer Verlag, N.Y., 1994, p.297,.

[36] L. Y. Jiang, H. M. Chen, Y. C. Jean, T, S. Chung, Ultra-thin polymeric

interpenetration network with enhanced separation performance approaching

ceramic membranes for biofuel, AIChE J., 55 (2009) 75.

[37] I. Pinnau, W. J. Koros, Gas-permeation properties of asymmetric polycarbonate,

polyestercarbonate, and fluorinated polyimide membranes prepared by the

generalized dry-wet phase inversion process, J. Appl. Polym. Sci., 46 (1992) 1195.

266
[38] D.T. Clausi, W.J. Koros, Formation of defect-free polyimide hollow fiber

membranes for gas separations, J. Membr. Sci., 167 (2000) 79.

[39] J.A. Van‘t Hof, A.J. Reuvers, R.M. Boom, H.H.M. Rolevink and C.A. Smolders,

Preparation of asymmetric gas separation membranes with high selectivity by a dual-

bath coagulation method, J. Membr. Sci., 70 (1992) 17.

[40] N. Peng, T. S. Chung, The effects of spinneret dimension and hollow fiber

dimension on gas separation performance of ultra-thin defect-free Torlon® hollow

fiber membranes, J. Membr. Sci., 310 (2008) 455.

[41] L. H. Sperling, Introduction to physical polymer science, Wiley, New York, NY,

2001.

[42] G.C. East, J.E. Mclntyre, V. Rogers, S.C. Senn, Production of porous hollow

polysulfone fibers for gas separation, 4th BOC Priestly Conference, Royal Society of

Chemistry, London, 1986.

[43] S.J. Shilton, G. Bell, J. Ferguson, The rheology of fiber spinning and the properties

of hollow fiber membranes for gas separation, Polymer 35 (1994) 5327.

[44] H.A. Tsai, C.Y. Kuo, J.H. Lin, D.M. Wang, A. Deratani, C. Pochat-Bohatier, K.R.

Lee, J.Y. Lai, Morphology control of polysulfone hollow fiber membranes via water

vapor induced phase separation, J. Membr. Sci. 278 (2006) 390.

[45] T.S. Chung, X. Hu, Effect of air-gap distance on the morphology and thermal

properties of polyethersulfone hollow fibers, J. Appl. Polym. Sci., 66 (1997) 1067.

[46] J. Qin, T.S. Chung, Effect of dope flow rate on the morphology, separation

performance, thermal and mechanical properties of ultrafiltration hollow fibre

membranes, J. Membr. Sci., 157 (1999) 35.

267
[47] T.S. Chung, W.H. Lin, R.H. Vora, The effect of shear rates on gas separation

performance of 6FDA-durene polyimide hollow fiber, J. Membr. Sci., 167 (2000)

55.

[48] A.F. Ismail, S.J. Shilton, I.R. Dunkin, S.L. Gallivan, Direct measurement of

rheologically induced molecular orientation in gas separation hollow fiber

membranes and effects on selectivity, J. Membr. Sci., 126 (1997) 133.

[49] T.S. Chung, S.K. Teoh, W.W.Y. Lau, M.P. Srinivasan, Effect of shear stress within

the spinneret on hollow fiber membrane morphology and separation performance,

Ind. Eng. Chem. Res., 37 (1998) 3930, with a subsequent correction at Ind. Eng.

Chem. Res. 37 (1998) 4903.

[50] S.A. McKelvey, D.T. Clausi, W.J. Koros, A guide to establishing hollow fiber

macroscopic properties for membrane applications, J. Membr. Sci., 124 (1997) 223.

[51] O.M. Ekiner, G. Vassilatos, Polyaramide hollow fibers for hydrogen/methane

separation : spinning and properties, J. Membr. Sci., 53 (1990) 259.

[52] P.S. Tin, T.S. Chung, Y. Liu, S.L. Liu, K.P. Pramoda, Effects of cross-linking

modification on gas separation performance of Matrimid membranes, J. Membr.

Sci., 225 (2003) 77.

[53] A. Bos, High pressure CO2/CH4 separation with glassy polymer membranes, Ph.D.

thesis, University of Twente, 1996.

[54] J.Z. Ren, R. Wang, T.S. Chung, D.F. Li, Y. Liu, The effects of chemical

modifications on morphology and performance of 6FDA-ODA/NDA hollow fiber

membranes for CO2/CH4 separation, J. Membr. Sci., 222 (2003) 133.

268
[55] C. Cao, R. Wang, T.-S. Chung, Y. Liu, Formation of high performance 6FDA-2, 6-

DAT asymmetric composite hollow fiber membranes for CO2/CH4 separation, J.

Membr. Sci., 209 (2002) 309.

[56] S.M. Jordan, M.A. Henson, W.J. Koros, The effect of carbon dioxide conditioning

on the permeation behavior of hollow fiber asymmetric membranes, J. Membr. Sci.,

54 (1990) 103.

[57] C. Cao, T.S. Chung, Y. Liu, R. Wang, K.P. Pramoda, Chemical cross-linking

modification of 6FDA-2,6-DAT hollow fiber membranes for natural gas separation,

J. Membr. Sci. 216 (2003) 257.

[58] T.S. Chung, L. Shao, P.S. Tin, Surface modification of polyimide membranes by

diamines for H2 and CO2 separation, Macromol. Rapid Comm., 27 (2006) 998-100.

[59] P.S. Tin, T.S. Chung, Y. Liu, R. Wang, S.L. Liu, K.P. Pramoda, Effect of

crosslinking modification on gas separation performance of Matrimid membranes, J.

Membr. Sci., 225 (2003) 77.

[60] Y. Liu, R. Wang, T.S. Chung, Chemical cross-linking modification of polyimide

membranes for gas separation, J. Membr. Sci., 189 (2000) 231.

269
CHAPTER SEVEN

CONCLUSIONS AND RECOMMENDATIONS

7.1 CONCLUSIONS

In recent years, current available membrane materials have reached Robeson’s

upper bound trade-off limit. Therefore, both the development of membranes materials

with superior separation properties for gas separation and the innovation of membrane

fabrication technology are research-intensive, aimed to produce high performance

membranes that can accomplish/conquer the separation challenges. Accordingly, an

investigation of membrane materials and fabrication based on commercially available

polyimides for variety of gas separation applications including but not limited to He/N2,

H2/N2, H2/CO2, H2/CH4, CO2/CH4, N2/CH4) was carried out in this research study

encompassing four diverse aspects listed below:

1. Development of nanocomposite membranes using MgO nanoparticles with

enhanced gas separation performance

2. Development of high performance membranes for hydrogen separation and

purification from the miscible polymer blends with interpenetration networks

3. Development of carbon molecular sieve membranes from the blends of high

performance materials N2/CH4 and CO2/CH4 separation and hydrogen

purification

270
4. Development of novel high-performnce gas separation membranes through

integration of polymer blending and dual-layer hollow fiber spinning process

The results obtained from these research studies provided further insights on the

important aspects of the gas separation membranes. The following discussions provide a

summary of conclusions derived from these studies.

7.1.1 Enhanced Gas Separation Performance of Nanocomposite Membranes Using

Mgo Nanoparticles

Nanocomposite membranes were fabricated for gas separation applications by the

introduction of magnesium oxide nanoparticles into the Matrimid polymer matrix. The

increase in particle content brought about increments in membrane’s permeability. This

was also accompanied by drops in gas selectivity attributed to the fact that the average

pore size of MgO (30 Å) was much larger than the kinetic diameters of gas molecules. It

was found that heat treatment of the membranes can influence their transport properties

differently. Heat treatment at temperatures below T g reduced the membranes’

permeability, whereas, heat treatment at temperatures exceeding the polymer T g resulted

in membranes with enhanced permeability. Further increments in membrane’s

permeability could be accomplished by quenching the membranes instead of natural

cooling. These phenomena were attributed mainly to the role of fractional free volumes in

the polymeric matrix of membrane.

271
Silver treatment was employed in order to tailor the performance of membranes.

Characterization of membranes revealed that in contrast to the polymeric membrane,

nanocomposite membranes had been impregnated through the adhesion of Ag cations

onto the surface of MgO nanoparticles. The adsorption process was confirmed by

different techniques and oxygen sites at the surface of MgO were recognized to be

responsible for this process. The amount of adsorption was monotonically increased with

the retention time. As a result, silver treated nanocomposite membranes exhibited

improved gas separation performances. Particularly, the H 2 /N 2 and CO 2 /CH 4 selectivity

of nanocomposite membranes after 10-day silver treatment improved to about 146.5 and

42.3, respectively; corresponding to 35% and 50% improvements compared to their

pristine sample.

The effective role of silver ions was recognized to be covering the porous

structure and control of the pore size of MgO nanoparticles. As a result, in addition to

the solution-diffusion mechanism, the size exclusion mechanism at MgO-Ag+ structures

as well as the presence of facilitated transport due to special interaction between CO 2 and

MgO were suggested by experimental analyses to be the governing mechanisms in the

superior separation performance of the newly developed Matrimid-MgO-Ag+

nanocomposite membranes. To our best knowledge, this is the first study that combines

the synergistic features of nanocomposite structure and facilitated transport mechanism

into one membrane by using unique characteristics of MgO for gas separation.

272
7.1.2 Hydrogen Separation and Purification in Membranes of Miscible Polymer

Blends with Interpenetration Networks

It was found in the course of this study that blend membranes prepared by mixing

PBI and Matrimid were completely miscible in molecular level for the wide range of

compositions. The miscibility of these components was largely attributed to the presence

of strong hydrogen bonding interactions between the functional groups. Analysis of the

transport properties of the blend membranes revealed that incorporation of PBI brought

about improvements in gas separation performance of membranes compared to the one

prepared from pure Matrimid. The effect was attributed to the increase in chain packing

density and hindrance in segmental mobility of polymer chains and the subsequent effects

on the diffusivity selectivity of the membranes. This was confirmed by characterization

of the microstructure of membranes.

The separation performance of the membranes was further ameliorated through

chemical modification of blend constituents. Modification of PBI phase with p-

xylenedichloride brings about slight improvements in selectivity performance, especially

for H 2 /CO 2 and H 2 /N 2 . In contrast, the selectivity of membranes was improved

significantly after cross-linking Matrimid with p-xylenediamine. The Matrimid/PBI

(25/75 wt.%) blend membrane cross-linked with p-xylenediamine exhibited the best

separation performance for hydrogen with H 2 /CO 2 selectivity of about 26. The higher

tendency of Matrimid toward cross-linking reaction contributes more in controlling the

transport properties of membranes through diffusion coefficient by increase in chain

packing density and diminishing of excess free volumes.

273
In overall, this study demonstrated successful implications of blending technique

cum chemical modification for the fabrication of high performance polymeric membranes

for gas separation applications. The effect of variation in composition on miscibility and

microstructure, gas permeability and selectivity of blend membranes were investigated.

The results obtained in this study revealed the promising features of developed

membranes for gas separation applications with great potential for hydrogen separation

and purification on industrial scale. To our best knowledge, this is the first study focusing

the exploitation of unique properties of blending high performance polymers of Matrimid

and PBI for the development of specialty membranes for hydrogen separation and

purification.

7.1.3 Carbon Membranes from Blends of PBI and Polyimides For N 2 /CH 4 and

CO 2 /CH 4 Separation and Hydrogen Purification

Carbon molecular sieving membranes developed from the blend of PBI and

various polyimides were analyzed on their potentials for various gas separation

applications. Polymeric films fabricated from blending Matrimid, P84 and Torlon with

PBI were found homogenous in nature and exhibiting desirable gas transport properties.

Therefore, these membranes could be considered as viable precursors for further

development and derivation of carbon membranes. It was realized by comparing the

performance of the carbon membranes prepared at different temperatures that the final

temperature can greatly influence the efficiency of the membranes for their gas

separation. The results suggested achieving a better performance at elevated

temperatures. This was mainly ascribed to the extent of progress in carbonization which

274
could bring about microstructures with higher carbon concentrations and more effective

pore sizes with fine distributions.

All derived carbon membranes demonstrated higher gas permeability compared to

their respective precursors. More interestingly, precursors with tighter polymeric

microstructures where found to offer carbon membranes with higher permeability. It was

found that carbon membranes derived from PBI/Matrimid (50/50 wt. %) exhibited more

attractive performance than the carbon membrane derived from other blends. Further

investigations on this promising blend system revealed the higher Matrimid content could

bring about improved permeability and selectivity particularly for separation of N 2 and

CO 2 from CH 4 .

Chemical modification of PBI/Matrimid precursors prior to carbonization

produced membranes with enhanced performance for separation of H 2 molecules from

N 2 and CO 2 . The results suggested that carbon molecular sieve membranes developed in

this study surpass various trade-off lines and are of great potentials for industrial

applications.

Developed membranes are amenable to surpass several separation performance trade-offs

with great potentials for various industrial applications including CO 2 /CH 4 (α=203.95),

H 2 /CO 2 (α=33.44) and particularly N 2 /CH 4 separation with unprecedented high

permeability (PN2 =2.78 Barrer) and selectivity (α=7.99) for this gas pair. To our best

knowledge, this is the highest ever reported selectivity for this gas pair. In addition,

tuning the PBI content could diversify the application and offer attractive membranes for

275
separation of other important gas pairs.

The results of this research study accentuated the necessity and importance of

adopting a proper strategy in exploiting synergistic beneficial features of advanced

materials (PBI), technology (polymer blending), process (carbonization) and modification

techniques (cross-linking) toward achieving membranes with desired performance.

7.1.4 Development of Novel High-performance Gas Separation Membranes

Through Integration of Polymer Blending Technology and Dual-layer

Hollow Fiber Spinning Process

Dual-layer hollow fiber membranes were successfully fabricated by suing a

polymer blend as a functional material as the outer layer. The effect of various spinning

parameters including the air gap distance, elongational drawing outer dope flow rate on

the morphological evolution, transport properties and separation performance of the

hollow fiber membranes were investigated.

The emergence of these two distinct morphologies for inner and outer layers was

attributed to various factors including the dope constituents, their composition and the

effect of coagulants and the chemistry of materials. A general increasing trend in the gas

permeance of the hollow fiber membranes could be established as a result of increase in

the air gap distance. Similar effects were caused by the application of elongational

drawing. Neither spinning at high air gap distance and application of elongational

276
drawing during the spinning could affect the state of macrovoids in inner layer, most

probably due to the low viscosity of the inner dope. However, the microstructure of the

outer layer was greatly affected and brought about significant improvements separation

performance of the hollow fiber membranes. The silicone rubber coated hollow fiber

membrane spun at the air gap distance 3 cm exhibited the H 2 /CO 2 selectivity of 11.11.

Similarly, silicone rubber coated hollow fiber membrane spun with elongational drawing

exhibited the CO 2 /CH 2 selectivity of 41.81.

Increase in outer dope flow rate was found to result in membranes with

diminished gas permeance accompanied with improvements in CO 2 /CH 4 separation

performance of the membranes. Majority of the hollow fiber membranes exhibited a very

good resistance against the CO 2 -induced plasticization. Additionally, chemical cross-

linking for short period of time was found effective in boosting the selectivity of

membranes for H 2 /CO 2 separation, though at the expense of the gas permeance.

7.2 RECOMMENDATIONS FOR FUTURE WORK

Based on the obtained experimental results, presented discussions and drawn

conclusions from this research study, the following recommendations may provide

further insights for future investigations related to the development of membrane

materials with potentially high-separation characteristics and further advancement of

membrane fabrication technology.

277
7.2.1 Nanocomposite Membranes

The research work presented here on nanocomposite membranes was originated

by the idea of exploiting the unexplored microstructural and surface properties of metal

oxide nanoparticles for enhanced gas separation applications. The development of

successful nanocomposite membranes from such materials in fact opens new doors for

further research on the use of various metal oxides with special inherent and surfacial

properties.

It was revealed in the course of this study that further post-modifications were

required in order revitalize the separation performance of the as-fabricated

nanocomposite membranes. This was mainly attributed to the relatively large pore size of

the magnesium oxides compared to that of gas molecules. The future study may focus on

the careful selection of metal oxide nanoparticles with more suitable pore sizes that fall

within the range the kinetic diameter of gas molecules. This may even eliminate the need

for further modification (e.g., heat treatment, silver ion treatment) of nanocomposite

membranes. Another interesting are of investigation may include the use of other high

performance polymers as the matrix for inclusion of nanoparticles. In this case, special

affinities and interactions between the polymers and nanoparticles may come

advantageous. It can also be interesting to apply surface modification or functionalization

techniques in order to tailor the properties of the nanoparticles.

Considering the silver ion treatment, effect of variation in the concentration of the

metal ion solution may also need to be explored. Besides the heat treatment and silver ion

278
modifications used in this research study, it would also be valuable to use other

modification techniques to tailor the transport and separation performance of

nanocomposite membranes. For examples, chemical cross-linking modification may be

used in order to modify the properties of nanocomposite membranes. The presence of the

nanoparticles inside the polymeric matrix may enhance the intrusion of cross-linking

agents into the membrane structure and facilitate the chance for internal modification in

addition to the surfacial modification.

It also would be worthwhile to examine the long term performance of the

membranes for gas separation especially in terms of their stability and also interaction

between the magnesium oxides and the metal ions used for their modification. Further

investigations may be required in order to examine and identify the role of individual

governing transport mechanisms hypothesized for the performance of the developed

nanocomposite membranes.

7.2.2 Polymeric Blend Membranes

The promising results presented here based on the membranes developed from the

blend of high polymeric material may highlight great potentials for research in this area.

Therefore, future studies may consider investigation for identification of potential

polymeric pairs from high performance materials that can form miscible blends in

molecular level. This also may include trials of novel synthesized polymers. Other

techniques such as enhancement in miscibility through addition of compatiblizers may

apply to those blends in the level of partial miscibility.

279
Due to the promising features of the PBI, it is highly recommended that future

studies focus on further exploitation of this high performance polymer. For example,

other successful candidates for gas separation such as 6FDA-based polyimides, P84 etc,

may also be explored for blending with PBI.

Specifically on the blends of PBI and Matrimid, one area of investigation may

include further amelioration of performance and separation efficiency of the membranes

through the use of other modification. For example, other types of chemical cross-linking

agents varying in chemical structure (aliphatic- or aromatic-based), functional groups,

size, pendants, etc may be used. In addition, effort on the improvements of cross-linking

modifications process would be valuable. In this respect, the primary areas of focus may

include reduction in reaction time and enhancement in reaction efficiency and effect of

concentration. Research on the possible ways to accelerate the cross-linking reaction may

enhance its applicability for industrial practice. In addition, it would be worthwhile to

investigate the possible means to further extend the cross-linking domain from the

surface. Advanced modification approaches may be envisaged to promote cross-linking

to the entire membrane structure.

The developed polymeric blend of PBI/Matrimid at different compositions, can

itself be considered as a material of choice for development of other types of membranes

such as nanocomposite membranes. The areas of application may be extended for

separation of other gases or examined for other membranes-based separation processes.

280
7.2.3 Carbon Membranes from Polymer Blends

The promising results of our research on development of carbon membranes from

miscible blends of polymeric materials provide a viable platform for further studies in

this area. In general, yet more fundamental research studies are required for further

understanding on the preparation of carbon molecular sieve membranes through

carbonization since it involves many complexities. Some of the parameters such as the

effect of final temperature, and profile were studied in the course of this research, but

other areas are suggested to be considered for further investigations. For example,

knowledge about the material degradation, carbon yields and emergence of the

microstructure are worthwhile to be further explored.

The influence of operating temperature and pressure on the gas transport

properties of the carbon molecular sieve membranes is recommended for further studies.

Efforts also worth on getting better insights on the governing transport mechanisms in

carbon molecular sieve membranes. Especial attentions may need to be paid to

differentiation of the roles of entropic and energetic selectivities.

Considering the polymer blend used in this study as the precursor, use of other

modification methods is suggested in order to investigate their effects on the gas

separation performance of the resultant carbon membranes. Further developments would

be valuable on transforming the carbon membranes into the hollow fiber geometry which

is more favorable for industrial applications.

281
7.2.4 Dual-layer Hollow Fiber Membranes from Polymeric Blends

Hollow fiber membranes are more advantages for industrial applications largely

due to their high surface area per unit volume of membrane module. Therefore, this area

has attracted many attentions for further research. One of the key developments made in

the course of this research was the pioneering integration of polymer blending technology

into dual-layer hollow fiber spinning. Therefore, it is highly recommended for future

research to further develop the technique and try on the other materials.

One of the important areas of research may be optimization of the geometry of the

hollow fibers through using different spinneret. Improvement in the mechanical stability

of the membranes can make great impacts on the long term performance of the

membranes and also their resistance to high feed pressures. In addition, further research

may be required to improve the morphological characteristics of the hollow fibers

especially in their inner layer. Various methods such as using alternative materials,

changes in rheological properties of the polymer, changes in bore fluid characteristics, etc

can be examined in this respect in order to reduce or eliminate the macrovoids in the

inner layer.

It is suggested that future studies on the developed membranes explore the

application of these membranes for separation of other gases. Another interesting area is

investigation on the potential use of dual-layer hollow fiber membranes as precursors for

fabrication of carbon molecular sieve membranes.

282

You might also like