You are on page 1of 197

DESIGN AND DEVELOPMENT OF PERVAPORATION

MEMBRANES FOR DEHYDRATION OF ALCOHOLS

XU YIMING
(B.Eng. Dalian University of Technology)

A THESIS SUBMITED
FOR THE DEGREE OF DOCTOR OF PHILOSOPHY
DEPARTMENT OF CHEMICAL AND
BIOMOLECULAR ENGINEERING
NATIONAL UNIVERSITY OF SINGAPORE

2017

Supervisor:
Professor Neal Tai-Shung Chung

Examiners:
Associate Professor Jianwen Jiang
Associate Professor Jianping Xie
Professor Xianshe Feng, University of Waterloo
DECLARATION

I hereby declare that this thesis is my original work and it has been written by

me in its entirety. I have duly acknowledged all the sources of information

which have been used in the thesis.

This thesis has also not been submitted for any degree in any university

previously.

Xu Yiming

21 August 2017

i
ACKNOWLEDGEMENTS

First of all, I would like express my sincere appreciation and token of thanks

to my supervisor, Prof. Chung Tai-Shung who provided me a great

opportunity and inspired me to work in the field of membrane science and

technology. His passion towards research gave me enthusiastic encouragement

and consistent support throughout the journey of my research tenure. His

trustful guidance and valuable suggestions provided me inspirations and

confidence to overcome the difficulties during my Doctoral study. I have

learned a lot from him not only in terms of research but also in terms of hard

working attitude, which will definitely lead me to a bright future in coming

days.

I would also like to gratefully acknowledge my mentor, Dr. Le Ngoc Lieu and

Dr. Zuo Jian who gave me their unselfish encouragements and detailed

guidance patiently. Special thanks are given to the other members of

pervaporation group including Dr. Ong Yee Kang, Dr. Shi Guimin, Dr. Tang

Yupan, Dr. Hua Dan and Mr. Peyman Salehian for sharing their research

knowledge and help in pervaporation. I also extend my sincere thanks to all

other research staffs and students in the research group who provided me their

kind help and shared with me all the joys and sorrows. Their company made

my 4-year stay at NUS memorable, colorful and stunning.

i
I would like to sincerely thank my thesis advisory committee (TAC) members,

Prof. Jiang Jianwen and Prof. Xie Jianping, who gave me their valuable

suggestions during my oral qualification exam.

I would also like extend my sincere gratitude to NUS for providing the

research scholarship for the whole duration of my PhD study. My heartiest

thanks go to all lab technologists and staffs in Department of Chemical and

Biomolecular Engineering, who helped me a lot in characterization techniques,

laboratory services and administrative related matters.

Last but not the least, I want to express my thanks to my parents and family

members from the core of my heart. Thanks for their love and care during the

past twenty-six years. No matter what happens, I know they are always

standing by my side with complete understanding and unwavering support.

Because of them, I could be able to put all my efforts in research without any

worries about other matters.

ii
Table of Contents

ACKNOWLEDGEMENTS ................................................................................ i

SUMMARY…………………………………………………………………viii

LIST OF TABLES ............................................................................................ xi

LIST OF FIGURES .........................................................................................xii

LIST OF SYMBOLS ...................................................................................... xvi

CHAPTER 1: INTRODUCTION ..................................................................... 1

1.1 Membrane pervaporation .................................................................... 1

1.2 Current applications of membrane pervaporation ............................... 2

1.2.1 Dehydration of organic solvents .................................................. 2

1.2.2 Removal of organics from aqueous solutions .............................. 4

1.2.3 Separation of organics mixtures................................................... 5

1.3 Main challenges of pervaporation ....................................................... 6

1.4 Comparison with other separation technologies ................................. 8

1.5 Research objectives and thesis organization ..................................... 10

1.6 References ......................................................................................... 14

CHAPTER 2: LITERATURE REVIEW ........................................................ 18

2.1 Mass transport mechanism ................................................................ 18

2.1.1 Solution-diffusion model ........................................................... 18

2.1.2 Pore flow model ......................................................................... 20

2.2 Evaluation of pervaporation membrane performance ....................... 21

2.3 Pervaporation membranes for alcohol dehydration .......................... 23

2.3.1 Membrane materials................................................................... 24

2.3.2 Membrane morphology and configuration ................................ 32

2.4 References ......................................................................................... 34

iii
CHAPTER 3: EXPERIMENTAL ................................................................... 42

3.1 Materials ............................................................................................ 42

3.1.1 Polymers .................................................................................... 42

3.1.2 Chemicals for polyimide syntheses ........................................... 42

3.1.3 Chemicals for UiO-66-Type MOFs syntheses ........................... 42

3.1.4 Organic solvents......................................................................... 43

3.2 Polyimide syntheses .......................................................................... 43

3.3 UiO-66-Type MOFs syntheses ......................................................... 44

3.4 Membrane fabrication ....................................................................... 45

3.4.1 Fabrication of polyarylether membranes by knife casting ......... 45

3.4.2 Fabrication of polyimide dense membranes by ring casting ..... 46

3.4.3 Fabrication of crosslinked polybenzoxazole membranes by

thermal treatment ..................................................................................... 46

3.4.4 Fabrication of mixed matrix membranes by ring casting .......... 47

3.5 Characterizations ............................................................................... 47

3.5.1 Fourier transform infrared spectrometer (FTIR)........................ 47

3.5.2 Wide angle X-ray diffraction (XRD) ......................................... 48

3.5.3 Themogravimetric analyses (TGA) ........................................... 48

3.5.4 Themogravimetric analyses-Fourier transform infrared

spectrometer (TGA-FTIR) ....................................................................... 49

3.5.5 Scanning electron microscope (SEM) ....................................... 49

3.5.6 Energy-dispersive X-ray spectroscopy (EDX) .......................... 49

3.5.7 Density measurements ............................................................... 50

3.5.8 Contact angle tests ..................................................................... 50

3.5.9 Solvent uptake analyses ............................................................. 51

iv
3.5.10 Positron annihilation lifetime spectroscopy (PALS) ................. 51

3.6 Pervaporation studies ........................................................................ 52

3.7 References ......................................................................................... 54

CHAPTER 4: POLYARYLERHER MEMBRANES FOR DEHYDRATION

OF ETHANOL AND METHANOL VIA PERVAPORATION ..................... 56

4.1 Introduction ....................................................................................... 56

4.2 Experimental ..................................................................................... 59

4.3 Results and discussion....................................................................... 60

4.3.1 Characterizations of membranes ................................................ 60

4.3.2 Solvent uptake analyses ............................................................. 63

4.3.3 PALS analyses ........................................................................... 64

4.3.4 Ethanol and methanol performance ........................................... 66

4.3.5 Long-term stability tests ............................................................ 73

4.4 Conclusions ....................................................................................... 74

4.5 References ......................................................................................... 75

CHAPTER 5: POLYIMDE AND CROSSLINKED THERMALLY

REARRANGED POLY(BENZOXAZOLE-CO-IMIDE) MEMBRANES FOR

ISOPROPANOL DEHYFRATION VIA PERVAPORATION ...................... 83

5.1 Introduction ....................................................................................... 83

5.2 Experimental ..................................................................................... 87

5.3 Results and discussion....................................................................... 88

5.3.1 Characterizations of PI and C-TR membranes .......................... 88

5.3.2 Solvent uptake analyses ............................................................. 92

5.3.3 Positron annihilation lifetime spectroscopy analyses ................ 93

5.3.4 Pervaporation performance of PI and C-TR membranes ........... 95

v
5.3.5 Performance benchmarking and stability of C-TR membranes . 98

5.4 Conclusions ..................................................................................... 100

5.5 References ....................................................................................... 101

CHAPTER 6: UiO-66/POLYIMIDE MIXED MATRIX MEMBRANES FOR

ETHANOL, ISOPROPANOL AND N-BUTANOL DEHYDRATION VIA

PERVAPORATION ...................................................................................... 108

6.1 Introduction ..................................................................................... 108

6.2 Experimental ................................................................................... 111

6.3 Results and discussion..................................................................... 112

6.3.1 Characterizations of UiO-66 and mixed matrix membranes ... 112

6.3.2 Solvent uptake analyses ........................................................... 117

6.3.3 Positron annihilation lifetime spectroscopy analyses .............. 118

6.3.4 Pervaporation performance ...................................................... 119

6.3.5 Pervaporation benchmarking ................................................... 126

6.4 Conclusions ..................................................................................... 129

6.5 References ....................................................................................... 130

CHAPTER 7: FUNCTIONALIZED UiO-66-TYPE MOFS FOR 6FDA-

POLYIMIDE BASED MIXED MATRIX MEMBRANES FOR

DEHYDRATION OF C1-C3 ALCOHOLS VIA PERVAPORATION ........ 137

7.1 Introduction ..................................................................................... 137

7.2 Experimental ................................................................................... 140

7.3 Results and discussion..................................................................... 140

7.3.1 Characterizations of particles and mixed matrix membranes .. 140

7.3.2 Solvent uptake analyses ........................................................... 145

7.3.3 Positron annihilation lifetime spectroscopy analyses .............. 147

vi
7.3.4 Pervaporation performance ...................................................... 148

7.3.5 Long stability of UiO-66-NH2 based MMMs .......................... 154

7.3.6 Performance benchmarking ..................................................... 155

7.4 Conclusions ..................................................................................... 157

7.5 References ....................................................................................... 159

CHAPTER 8: CONCLUSIONS AND ECOMMENDATIONS ................... 166

8.1 Conclusions ..................................................................................... 166

8.1.1 Polyarylether membranes for dehydration of ethanol and

methanol via pervaporation.................................................................... 166

8.1.2 Polyimide and Crosslinked Thermally Rearranged

Poly(benzoxazole-co-imide) Membranes for Isopropanol Dehydration via

Pervaporation ......................................................................................... 168

8.1.3 UiO-66/Polyimide Mixed Matrix Membranes for Ethanol,

Isopropanol and n-Butanol Dehydration via Pervaporation .................. 169

8.1.4 Functionalized UiO-66-Type MOFs for 6FDA-Polyimide based

Mixed Matrix Membranes for Dehydration of C1-C3 Alcohols via

Pervaporation ......................................................................................... 170

8.2 Recommendations ........................................................................... 171

BIBLIOGRAPHY .......................................................................................... 176

vii
SUMMARY

Pervaporation is a prospective candidate to separate azeotropic mixtures or

close boiling point mixtures due to its advantages of low energy consumption,

environmental benignity, small footprint and mild operating conditions.

Membrane is the heart of pervaporation separation. In this thesis, novel

polymeric and mixed matrix membranes (MMMs) with high separation

performance and good stability have been developed for alcohol dehydration

via pervaporation.

Firstly, pervaporation membranes made from four polyarylether (PAE)

polymers were developed for methanol and ethanol dehydrations.

Fundamental characteristics like free volume, d-spacing, water contact angle,

solvent uptakes were investigated. The hydrophilic polyethersulfone-co-

Pluronic (H-PESU) membrane shows the best ethanol dehydration

performance with the highest water permeability of 0.322 mg m-1 h-1 kPa-1,

which is 2-6 times higher than other three membranes and comparable mole-

based selectivity of 685. Long-term tests reveal the good operating stability of

H-PESU and polyethersulfone (PESU) membranes. Moreover, methanol

dehydration performance of these PAE membranes is superior compared with

other works. The water in the permeate can be concentrated to 86.8%.

Secondly, novel polyimide and crosslinked thermally rearranged

polybenzoxazole (C-TR-PBO) membranes were developed. Fundamental

chemical and thermal properties were characterized. Solvent uptake analyses

viii
and positron annihilation lifetime spectroscopy measures were also carried

out. The synergistic effects of crosslinking and thermal rearrangement on the

physicochemical properties of the membranes were studied. C-TR-PBO

membrane shows an obvious improvement in flux compared to their

polyimide precursors. C-TR-5-5 exhibits the best pervaporation performance

with a comparable water permeability of 0.15 mgm-1h-1kPa-1 and an

impressively high mole-based selectivity of 4019. Besides, long-term tests

show C-TR-PBO membrane has stable isopropanol dehydration performance.

Thirdly, MMMs with UiO-66 nanoparticles and 6FDA-HAB/DABA

polyimide were designed for dehydration of ethanol, isopropanol and n-

butanol. Uniform UiO-66 nanoparticles with a particle size of around 100nm

were synthesized and subsequently added into MMMs. UiO-66 nanoparticles

can be evenly dispersed in MMMs without visible agglomerations. Results

show that the incorporation of UiO-66 greatly improves the normalized flux of

MMMs. The MMMs show excellent separation performance for the

dehydration of isopropanol and n-butanol. At the highest loading of 30wt%,

the MMMs have the water permeability of 0.329 and 0.292 mg m-1 h-1 kPa-1

and mole-based selectivity of 2209 and 14214 respectively for

isopropanol/water and n-butanol/water systems. Furthermore, dehydration

performances for different alcohol/water systems were compared. The effects

of feed temperatures and compositions on pervaporation performance were

systematically studied.

ix
Last but not the least, MMMs consisting of two different functionalized UiO-

66-Type MOFs (UiO-66-NH2 and UiO-66-F4) and 6FDA-HAB/DABA

polyimide were further developed for methanol, ethanol, isopropanol

dehydration. UiO-66-NH2 (80-90nm) and UiO-66-F4 (90-130nm) were

successfully synthesized. The effects of functional groups of the MOFs

particles and the effects of particle loading on the dehydration performance

were systematically investigated. Results suggest UiO-66-NH2 based MMMs

have better separation performance than UiO-66-F4 based MMMs. Adding

UiO-66-NH2 nanoparticles can greatly improve normalized flux of the MMMs

and at the same time enhance separation factor within the particle loading

range of 0-20wt%. The UiO-66-NH2 based MMMs have stable long-term

isopropanol dehydration performance due to good compatibility between UiO-

66-NH2 and polyimide.

x
LIST OF TABLES

Table 1.1 Potential liquid separations by pervaporation process ............................ 2

Table 4.1 Densities and water contact angles of PESU, T-PESU, PPSU and H-

PESU membranes .................................................................................................. 62

Table 4.2 Solvent uptake results for PESU, T-PESU, PPSU and H-PESU

membranes ............................................................................................................. 63

Table 4.3 PALS data of dry PESU, T-PESU, PPSU and H-PESU membranes .... 64

Table 4.4 PALS data of wet PESU, T-PESU, PPSU and H-PESU membranes .... 65

Table 4.5 Ethanol dehydration performance of PESU, T-PESU, PPSU and H-

PESU membranes .................................................................................................. 66

Table 4.6 A comparison of pervaporation performance for ethanol dehydration.. 69

Table 4.7 Methanol dehydration performance of PESU, T-PESU, PPSU and H-

PESU membranes .................................................................................................. 70

Table 4.8 A comparison of pervaporation performance for methanol dehydration

................................................................................................................................ 71

Table 5.1 A comparison of pervaporation performance of dense membranes for

isopropanol dehydration......................................................................................... 99

Table 6.1 Solvent uptake analyses for the pristine PI membrane and MMMs .... 117

Table 6.2 PALS data for the pristine PI membrane and MMMs ......................... 118

Table 6.3 Thermodynamic properties of feed mixtures as a function of operating

temperature .......................................................................................................... 125

Table 6.4 Thermodynamic properties of feed mixtures as a function of isopropanol

concentration ........................................................................................................ 126

Table 6.5 A comparison of pervaporation performance of membranes for

isopropanol dehydration....................................................................................... 127

xi
Table 6.6 A comparison of pervaporation performance of membranes for n-

butanol dehydration ............................................................................................. 128

Table 6.7 A comparison of pervaporation performance of membranes for ethanol

dehydration .......................................................................................................... 128

Table 7.1 Solvent uptake ratios of the pristine PI membrane and UiO-66-NH2

based MMMs ....................................................................................................... 146

Table 7.2 Comparison of solvent uptake ratio among different UiO-66 MOF based

MMMs ................................................................................................................. 146

Table 7.3 PALS data of the pristine PI membrane and UiO-66-NH2 based MMMs

.............................................................................................................................. 147

Table 7.4 Comparison of PALS data between UiO-66-NH2 and UiO-66-F4 based

MMMs ................................................................................................................. 148

Table 7.5 Ethanol dehydration performance of UiO-66-NH2 and UiO-66-F4 based

MMMs ................................................................................................................. 149

Table 7.6 A comparison of pervaporation performance for methanol dehydration

.............................................................................................................................. 156

Table 7.7 A comparison of pervaporation performance of membranes for ethanol

dehydration .......................................................................................................... 156

Table 7.8 A comparison of pervaporation performance of membranes for

isopropanol dehydration....................................................................................... 157

Table 8.1 A comparison of pervaporation performance of dense membranes for

ethanol dehydration .............................................................................................. 172

LIST OF FIGURES

Figure 1.1 Schematic diagram of pervaporation process ......................................... 1

Figure 2.1 Schematic diagram of solution-diffusion transport model ................... 19

xii
Figure 2.2 Schematic diagram of pore flow model................................................ 21

Figure 2.3 Typical membrane morphologies ......................................................... 32

Figure 3.1 Synthesis scheme of polyimide and crosslinked poly(benzoxazole-co-

imide) ..................................................................................................................... 44

Figure 3.2 The ligand structures of UiO-66-Type MOFs ...................................... 45

Figure 4.1 Chemical structures of PESU, T-PESU, PPSU and H-PESU .............. 59

Figure 4.2 TGA curves of PESU, T-PESU, PPSU and H-PESU membranes ....... 61

Figure 4.3 XRD patterns of PESU, T-PESU, PPSU and H-PESU membranes..... 61

Figure 4.4 Comparison of FFVs at dry and wet states measured by PALS........... 66

Figure 4.5 (a) Water and (b) ethanol/methanol normalized flux of H-PESU

membranes at different feed concentrations at 60 ºC ............................................ 72

Figure 4.6 Ethanol dehydration performance of H-PESU membranes at different

feed concentrations at 60 ºC................................................................................... 72

Figure 4.7 Methanol dehydration performance of H-PESU membranes at different

feed concentrations at 60 ºC................................................................................... 73

Figure 4.8 Long-term stability of the PESU and H-PESU membranes for ethanol

dehydration at 60 ºC with the feed composition of 85wt%/15wt% ethanol/water 74

Figure 5.1 Crosslinking mechanism of the polyimide containing carboxylic groups

under thermal treatment ......................................................................................... 87

Figure 5.2 FTIR spectra of synthesized polyimides .............................................. 88

Figure 5.3 FTIR spectra of poly(benzoxazole-co-imide) ...................................... 89

Figure 5.4 TGA curves of polyimide precursors ................................................... 90

Figure 5.5 (a) Top-view 3D TGA-FTIR analyses of the polyimide precursor PI-

10-0 and (b) Heating time of TGA and weight percentage vs. temperature

correlation .............................................................................................................. 91

xiii
Figure 5.6 (a) Top-view 3D TGA-FTIR analyses of the polyimide precursor PI-9-

1 and (b) Heating time of TGA and weight percentage vs. temperature correlation

................................................................................................................................ 91

Figure 5.7 Solvent uptake analyses for PI and C-TR membranes ......................... 93

Figure 5.8 PALS data of PI membranes and C-TR membranes ............................ 94

Figure 5.9 Fractional free volumes of PI and C-TR membranes ........................... 95

Figure 5.10 Isopropanol dehydration performance of PI and C-TR membranes at

60 ºC....................................................................................................................... 96

Figure 5.11 Long-term stability of the C-TR-5-5 membrane for isopropanol

dehydration at 60 ºC ............................................................................................... 99

Figure 6.1 A comparison of XRD patterns of UiO-66 and MMMs..................... 113

Figure 6.2 FESEM images of (a) UiO-66 nanoparticles and cross-sectional

morphologies of (b) the pristine PI membrane and MMMs containing (c) 10wt%

UiO-66, (d) 20wt% UiO-66, (e) 30wt% UiO-66 ................................................. 113

Figure 6.3 (a) SEM image and (b) EDX mapping of Zr on the cross-section of

MMMs containing 30wt% UiO-66 ...................................................................... 114

Figure 6.4 (a) Top-view of 3D TGA-FTIR analyses of UiO-66 and (b) its

temperature vs. weight percentage and heating time ........................................... 115

Figure 6.5 (a) Top-view of 3D TGA-FTIR analyses of the MMM containing 20wt%

UiO-66 and (b) its temperature vs. weight percentage and heating time............. 115

Figure 6.6 TGA curves of the pristine PI membrane and MMMs ....................... 116

Figure 6.7 Ethanol dehydration performance of the pristine PI membrane and

MMMs at 60 ºC using ethanol/water (85/15wt%) as the feed ............................. 119

Figure 6.8 Isopropanol dehydration performance of the pristine PI membrane and

MMMs at 60 ºC using isopropanol/water (85/15wt%) as the feed ...................... 120

xiv
Figure 6.9 n-Butanol dehydration performance of the pristine PI membrane and

MMMs at 60 ºC using n-butanol/water (85/15wt%) as the feed ......................... 121

Figure 6.10 Dehydration performance of the pristine PI membrane and MMMs for

C2–C4 alcohols at 60 ºC using alcohol/water (85/15wt%) as the feed: (a) Total

normalized flux and (b) Water concentration in the permeate vs. carbon number

.............................................................................................................................. 123

Figure 6.11 Isopropanol dehydration performance of the MMM containing 20 wt%

UiO-66 using isopropanol /water (85/15wt%) as the feed................................... 124

Figure 6.12 Isopropanol dehydration of the MMM containing 20 wt% UiO-66 at

60 ºC..................................................................................................................... 126

Figure 7.1 XRD spectra of UiO-66-NH2 and UiO-66-F4 .................................... 141

Figure 7.2 FESEM images of (a) UiO-66-NH2 and (b) UiO-66-F4 nanoparticles

.............................................................................................................................. 142

Figure 7.3 FESEM images cross-sectional morphologies of (a) the pristine PI

membrane (b) MMM containing 10wt% UiO-66-F4 and MMMs containing (c)

10wt%, (d) 20wt%, (e) 30wt% UiO-66-NH2 ....................................................... 142

Figure 7.4 (a) SEM image and (b) EDX mapping of Zr on the cross-section of

MMMs containing 30wt% UiO-66-NH2 ............................................................. 143

Figure 7.5 (a) SEM image and (b) EDX mapping of Zr on the cross-section of

MMMs containing 10wt% UiO-66-F4 ................................................................. 143

Figure 7.6 TGA curves of UiO-66-NH2, pristine PI membrane and UiO-66-NH2

based MMMs under air atmosphere .................................................................... 144

Figure 7.7 TGA curves of UiO-66-F4, pristine PI membrane and UiO-66-F4 based

MMMs under air atmosphere............................................................................... 145

xv
Figure 7.8 Methanol dehydration performance of the pristine PI membrane and

MMMs at 60 ºC with feed composition of methanol/water (85/15wt%) ............. 150

Figure 7.9 Ethanol dehydration performance of the pristine PI membrane and

MMMs at 60 ºC with feed composition of ethanol/water (85/15wt%) ................ 150

Figure 7.10 Isopropanol dehydration performance of the pristine PI membrane and

MMMs at 60 ºC with feed composition of isopropanol/water (85/15wt%) ........ 151

Figure 7.11 Dehydration performance of the pristine PI membrane and UiO-66-

NH2 based MMMs for C1–C3 alcohols at 60 ºC with feed composition of alcohol

/water (85/15wt%): (a) Normalized flux and (b) Water concentration in the

permeate vs. carbon number ................................................................................ 153

Figure 7.12 Long-term stability performance of 10wt% UiO-66-NH2 MMMs for

isopropanol dehydration at 60 ºC with feed composition of isopropanol/water

(85/15wt%) .......................................................................................................... 154

Figure 8.1 Hydrophilic surface modification ....................................................... 173

Figure 8.2 Structures of four hydrophilic chemicals ........................................... 174

LIST OF SYMBOLS

 separation factor (water/alcohol)

 mole-based selectivity (water/alcohol)

i activity coefficient of component i

δ effective length of cylindrical pores

δ1 distance filled with liquid

δ2 distance filled with vapor

θ incident angle

𝜆 wavelength of radiation beam

xvi
𝜌𝑒 density of hexane

𝜌0 density of membrane

τ3 o-Ps lifetime

d d-spacing

l membrane thickness

𝑚0 mass of membrane in air

𝑚0′ mass of membrane in hexane

𝑛 any integer

t time interval during sample collections

xwi , xwj weight fraction of the component i or j in the feed

xi mole fraction of the component i in the feed

y wi , y wj weight fraction of the component i or j in the permeate

yi mole fraction of the component i in the permeate

A effective membrane area

I3 o-Ps intensity

J flux of a membrane

𝑀𝑤𝑒𝑡 weight of swollen membrane strip at the equilibrium condition

𝑀𝑑𝑟𝑦 weight of initial membrane strip at dry state

Pi permeability of component i

Pj permeability of component j

pP total pressure at the permeate side

pif partical vapor pressure of component i in the feed

piP partical vapor pressure of component i in the permeate

xvii
pisat saturated vapor pressure of component i

Q total weight of permeate

R mean free-volume radius

ΔR an empirical constant (1.66 Å)

ATR attenuated total reflectance

BDC 1,4-benzene-dicarboxylate

C-TR crosslinked thermal rearrangement

DABA 3,5-diaminobenzoicacid

DMAC N,N-dimethylacetamide

DMF N,N-dimethylformamide

DMSO N,N-dimethylsulfoxide

EDX energy-dispersive X-ray spectroscopy

FESEM field emission scanning electron microscopy

FFV fractional free volume

FTIR fourier transform infrared spectroscopy

GO graphene oxide

HAB 3,3’-dihydroxybenzidine diamine

H-PESU hydrophilic PESU-co-Pluronic

IPA isopropanol

MMMs mixed matrix membranes

MOFs metal organic frameworks

NMP N-methyl-2-pyrrolidone

PAA poly(acrylic acid)

PAE polyarylether

xviii
PALS positron annihilation lifetime spectroscopy

PBI polybenzimidazole

PBO polybenzoxazole

PBOZ polybenzoxazinone

PDMS polydimethylsiloxane

PESU polyethersulfone

PI polyimide

POSS polyhedral oligomeric silsesquioxane

PPSU polyphenylsulfone

PTFE polytetrafluoroethylene

PVA polyvinyl alcohol

PVDF poly(vinylidene fluoride)

SEM scanning electron microscopy

TGA thermogravimetric analyses

T-PESU polytrimethylphenylethersulfone

TR thermal rearrangement

UiO-66 Zr (IV) terephthalate metal-organic framework

UiO-66-F4 fluorine functionalized UiO-66

UiO-66-NH2 amine functionalized UiO-66

XRD X-ray diffraction

6FDA 4,4’-(hexafluoroisopropylidene) diphthalic anhydride

xix
CHAPTER 1: INTRODUCTION

1.1 Membrane pervaporation

A membrane is a semi-permeable barrier that preferentially allows certain

molecules to pass while retains others and therefore achieves separation [1].

Membrane technology becomes very important as an alternative to

conventional separation techniques, such as distillation and extraction because

of its advantages of environmental benignity, high energy efficiency and low

cost. Nowadays, membrane separation processes are widely applied in

different fields such as gas separation, waste water treatment, sea water

desalination, production of pharmaceuticals, separation of pharmaceutical

wastes, fuel cells applications etc.

Pervaporation, a kind of membrane technologies that combines the process of

permeation and vaporization of the permeate. Figure 1.1 shows the schematic

diagram of pervaporation process. In pervaporation process, liquid mixture to

be separated (feed) contacts with the membrane on one side and permeating

product (permeate) is removed in vapor state from the other side.

Figure 1.1 Schematic diagram of pervaporation process

1
1.2 Current applications of membrane pervaporation

The current applications of pervaporation can be classified into three major

categories: (i) dehydration of organic solvents, (ii) removal of organic

compounds from aqueous solutions, and (iii) separation of organic mixtures.

Table 1.1 lists some potential liquids separations that request technological

breakthroughs and also some existing pervaporation applications supported by

commercially available membranes based on the three categories [2].

Table 1.1 Potential liquid separations by pervaporation process

1.2.1 Dehydration of organic solvents

Among the three application categories, dehydration of organic solvents is the

main application in the industry. Some organics, such as alcohols with low

2
molecular weight, ethers, and ketones, can be completely miscible with water

at any ratio and easily form azeotropes. Traditional separation technologies

such as distillation usually need high energy thus are not economic.

Sometimes, the traditional separation technologies have difficulties or are

impossible to break the azeotropes. However, pervaporation can overcome these

problems because it is not restricted by the vapor-liquid equilibrium thus can

easily break the azeotropes. Pervaporation is a clean technology because it does

not need third component to break the azeotropes and can avoid cross-

contamination. Moreover, pervaporation process is energy-saving because only

the permeating component consumes the latent heat and operating condition of

pervaporation is mild with low pressures and temperatures. Therefore,

dehydration of organics by pervaporation has gained a considerable market.

One typical example is to remove water from concentrated alcohol solutions to

produce high-purity alcohols. The most important alcohols are ethanol and

isopropanol (IPA) due to their great demand in various industries. Ethanol is

considered as an alternative fuel to alleviate global warming and slow down

the depletion of fossil fuels and ethanol is also largely used in

biopharmaceutical and medical industries [3, 4]. IPA can be used as an

effective cleaner for electronic and optical devices because of its non-toxicity,

fast evaporation and ability to dissolve a wide range of non-polar compounds.

Moreover, IPA is also a chemical intermediate for the production of isopropyl

acetate, rubbing alcohol, vitamin B12 and a pharmaceutical additive used in

hand sanitizer and disinfecting pads [5, 6]. The production of ethanol and IPA

needs the removal of water, thus pervaporation dehydration attracts significant

3
attention. Besides, the dehydration of other solvents, such as glycerol, acetone

and acetic acid has also been gradually developed.

The first commercial pervaporation membrane to remove water from

concentrated alcohol solutions was developed by Gesellschaft für

Trenntechnik (GFT) Co., Germany in early 1980s, which is considered as a

milestone in dehydration of alcohols [7]. The membrane consists of a thin

layer of crosslinked polyvinyl alcohol (PVA) and a porous polyacrylonitrile

(PAN) substrate. Since then, more pervaporation systems for dehydration of

organic solvents were commercialized and installed worldwide [8].

1.2.2 Removal of organics from aqueous solutions

Removal or recovery of organic compounds from a dilute aqueous solution is

another application of pervaporation. It is applied in wastewater treatment,

recycle of expensive organics and pollution control. This application first

developed by Membrane Technology and Research (MTR) [9]. After that, the

process was further expanded to various areas: (1) removal of trace amount of

volatile organic compounds (VOCs) from water stream [10]; (2) recovery of

valuable organics from aqueous solutions [11]; (3) removal of certain product

from a fermentation broth, which is helpful to improve the conversion rate and

productivity and to maintain a continuous flow. This application has gained

great attention in recent years for bioethanol or biobutanol recovery [12].

4
In this application, membrane materials need to be organophilic and

hydrophobic which have preferential sorption of organic compounds. Therefore,

most of the commercial available membranes are made from elastomers or

rubbery materials such as polydimethylsiloxane (PDMS) and polybutadiene or

highly hydrophobic materials such as poly(vinylidene fluoride) (PVDF). So far,

this application still remains at the state of laboratory-scale-demonstration.

Further studies on novel materials are still in need.

1.2.3 Separation of organics mixtures

Organic/organic separation is another potential application of pervaporation in

chemical, petrochemical and pharmaceutical industries. In recent years, using

pervaporation technology for organic/organic mixtures separation has received

significant attention [13]. However, comparing the above mentioned two

applications, the development of organic/organic separation is more limited. To

date, there are no commercially available polymeric membranes for this

application with widespread industrial acceptance due to degradation of

membrane performance in organic mixtures which is caused by swelling and loss

of membrane integrity. Nowadays, pervaporation based separation of the

following four categories of organic mixtures including polar/non-polar

mixtures, aromatic/alicyclic mixtures, aromatic/aliphatic mixtures and

aromatic isomers, have been demonstrated [14-16]. Separation of organic

mixtures is based on the differences of steric effects, polarity, and affinity to

membrane materials of different organic molecules. Currently, typical

membrane materials for this application includes polyethylene,

5
poly(tetrafluoro ethylene) (PTFE), and cellulose acetate [13, 14, 17]. The main

limitation for its wide application is lack of suitable membranes and modules

with good long-term stability in organic solvents. Crosslinking or blending may

be effective ways to improve membrane stability. Another promising method

is to develop robust inorganic membranes with controlled pore size.

1.3 Main challenges of pervaporation

Based on the above review, it can be found that one of the major obstacles for

the development of pervaporation is the lack of membranes with high

productivity, high selectivity and good stability. Polymeric membranes are

worthy of study due to their advantages such as compactness, ease of

fabrication and scale-up, lower material costs. However, there is a trade-off

between permeability and selectivity which restricts their separation

performance. Inorganic materials usually exhibit higher separation

performance than polymeric membranes because of their superior thermal and

solvent stability. However, their brittleness, complicated production process

and high costs limit their applications. Researchers have tried to integrate

inorganic and organic components together to develop membranes with

combined advantages and superior separation performance. But compatibility

between polymeric and inorganic components is an issue to be considered and

addressed. Overall, the development of membrane material is critical for

pervaporation to compete with other separation techniques. In general,

desirable membrane materials should have outstanding productivity and

separation efficiency. Besides, satisfying physicochemical properties such as

6
good mechanical stability, thermal stability and chemical resistance are also

necessary to maintain good long-term stability and reduce solvent-induced

swelling in harsh environment [18]. The other challenge is due to the fact that

a certain material often can only show good separation performance for a

specific mixture but not for other types of mixtures. In other words, different

mixtures require different types of membranes. Thus, evaluation on material

selection and tests of membrane materials must be done for different mixture

pairs.

There are two strategies to develop suitable membranes for pervaporation.

One is the molecular design of chemistry of membrane material with desired

physicochemical properties while the other is the macromolecular engineering

of membrane structure with desired morphology, ultrathin and defect free

selective layers [2]. The former method aims to tune membranes’

thermodynamic and/or kinetic responses to the permeate, while the latter

reduces defects and minimize transport resistance [2]. Based on the second

strategy, an asymmetric membrane structure with a very thin selective layer on

a porous supporting layer is desired to achieve high permeance. There are two

types of asymmetric pervaporation membranes: (1) asymmetric membranes

made of the same material [19] and (2) composite membranes consisting of

different materials for the dense selective layer and the porous substrate [20].

The second type membranes show great potential for commercialization due to

cost effective advantage. The composite membranes only need a small amount

of expensive material with high performance for the ultrathin selective layer

and the porous substrate can be made from cheap materials. However, the

7
compatibility of these different materials and the defect-free requirement for

the ultra-thin selective layer are the main challenges of composite membranes.

Besides, the mechanical properties also need be enhanced to maintain a longer

durability.

The success of membrane development not only depends on the improvements

of membrane materials and fabrication techniques but also needs a proper

understanding of the mass transport mechanism across the membrane.

Fundamental studies on mass transport mechanism can provide in-depth

knowledge in membrane structure design, process operation optimization, and

separation performance prediction.

1.4 Comparison with other separation technologies

In the industry, there are three major separation technologies for alcohol

dehydration i.e. distillation, adsorption, and pervaporation. The following

discussion compares these technologies including pros and cons.

• Distillation

As one of the most widely used technologies in the industry, distillation is

very effective at low alcohol concentrations and can provide a high separation

factor by multiple stages of vapor-liquid equilibrium (VLE). It performs well

at large-scale and is easier for mass and energy integration [5]. However, it

8
becomes difficult for azeotrope separation without adding in additional

separation steps or a third component. What’s more, it operates at high

temperature which consumes high energy and may cause deactivation of

thermally sensitive compounds such as proteins and enzymes. In addition, the

heat and mass integration becomes very difficult at small scales [5].

• Adsorption

In alcohol dehydration adsorption systems, a variety of hydrophilic adsorbents

have been developed including molecular sieves e.g. zeolites and bio-based

materials such as starch, corn cobs, cellulose and activated palm stone [5, 21].

Both liquid and vapor phase adsorptions are technically feasible, but the vapor

phase adsorption is usually practiced and it consumes lower energy than

distillation because only one-time vaporization is required. Zeolite molecular

sieves provide a high selectivity, but water is very strongly adsorbed thus high

temperature or low pressures are needed to regenerate them. What’s more,

molecular sieves are more expensive than bio-based adsorbents.

• Pervaporation

So far, pervaporation is the most significant competitor to distillation and

molecular sieve adsorption for alcohol dehydration. Compared with

conventional separation technologies, pervaporation offers several unique

advantages. First, pervaporation is not limited by the vapor-liquid equilibrium,

and thus expected to have surpassing separation efficiency, particularly in

9
separating azeotropic and close boiling point liquid mixtures [22]. Second, it is

a clean technology and can avoid cross-contamination because it does not

need third components, which are commonly employed in the so-called

“azeotropic distillations” [13]. Third, pervaporation is an energy-saving

process because only the permeating component consumes the latent heat [23,

24]. Moreover, pervaporation has a mild operation condition by using low

feed pressures and temperatures [22, 25]. Last, the compactness in module

fabrication as well as simplicity in process control are also advantageous

characteristics of pervaporation [18]. However, as discussed in previous part,

one of the major obstacles for the development of pervaporation is the lack of

membranes with high productivity, high selectivity and good stability. The

other challenge is due to the fact that a certain material often can only show

good separation performance for some specific mixtures but not for other

types of mixtures. In other words, different mixtures require different types of

membranes. Thus, evaluation on material selection and tests of membrane

materials must be done for different mixture pairs.

1.5 Research objectives and thesis organization

The objective of this thesis is to explore various new membrane materials with

high productivity, high selectivity and good stability, including polymers and

inorganic materials for alcohols dehydration via pervaporation. To achieve this

objective, novel membrane materials including commercial polymer

polyarylether and self-synthesized polymer polyimide were first developed

because they have good mechanical strength, excellent thermal and chemical

10
stability. A fundamental study of polymer properties and a systematic study

of various effects of polymer structures and operating conditions on separation

performance were conducted. After that, to further improve membrane

separation performance, thermal crosslinking and thermal rearrangement

reactions were carried out to convert polyimide into crosslinked

poly(benzoxazole-co-imide) and UiO-66-Type MOFs with different functional

groups were synthesized and incorporated into polyimide to make mixed

matrix membrane. The effects of particle loading and functional groups on

separation performance were investigated.

The detailed research objectives are as follows:

(1) To investigate the potential of polyarylether for dehydration of ethanol and

methanol and to study the influence of main-chain structure on mass transport

phenomena and pervaporation performance;

(2) To explore the potential of polyimide for isopropanol dehydration and

improve its separation performance by thermal crosslinking and thermal

rearrangement;

(3) To enhance separation performance for ethanol, isopropanol and n-butanol

dehydration by incorporating UiO-66 into polyimide to make mixed matrix

membranes;

(4) To further improve separation performance by developing mixed matrix

membranes containing polyimide and UiO-66-Type MOFs with different

functional groups and to study the effect of functional groups of UiO-66-Type

MOFs on membrane swelling and separation performance.

11
The dissertation is organized into eight chapters. Chapter one gives the

introduction of this thesis, including a brief introduction of pervaporation, the

major applications of pervaporation and the main challenges in the

development of pervaporation.

Chapter two presents the fundamentals of mass transport mechanisms, and the

parameters of performance evaluation for pervaporation membranes. Besides,

a comprehensive literature review on pervaporation membranes for alcohol

dehydration is also provided in this chapter. Membrane material, morphology

and configurations are also discussed in detail.

Chapter three provides the experimental details involved in the research,

including the materials used, the methods of fabricating membranes,

pervaporation tests and a detailed description of different characterization

techniques.

Chapter four shows pervaporation membranes made from polyethersulfone

(PESU), polyphenylsulfone (PPSU), polytrimethylphenylethersulfone (T-

PESU) and hydrophilic PESU-co-Pluronic (H-PESU) polymers for

dehydration of ethanol and methanol. Fundamental characteristics of pore

structure, free volume, d-spacing, water contact angle, water and ethanol

uptakes were investigated and correlated with their dehydration performance.

Long-term stability tests were carried out for both H-PESU and PESU

membranes.

12
Chapter five reports the synthesis of novel polyimide and the development of

crosslinked thermally rearranged polybenzoxazole (C-TR-PBO) membranes

for isopropanol dehydration. Fundamental chemical and thermal properties

were characterized by FTIR, TGA, TGA-IR. Solvent uptake analyses and

positron annihilation lifetime spectroscopy measures were also carried out to

know the solubility and permeability information. The synergistic effects of

crosslinking and thermal rearrangement on the physicochemical properties of

the membranes were studied. Long-term stability of C-TR membranes were

also studied.

Chapter six shows the design of novel mixed matrix membranes consisting of

UiO-66 nanoparticles and 6FDA-HAB/DABA polyimide for the dehydration

of ethanol, isopropanol and n-butanol via pervaporation. Uniform UiO-66

nanoparticles with a relatively small particle size were synthesized and

subsequently employed to prepare UiO-66/6FDA-HAB/DABA polyimide

MMMs with different particle loadings. Membranes and nanoparticle

morphologies, crystallographic structure, thermal properties were investigated.

Solvent uptakes and free volume properties were characterized to reveal the

actual mechanisms for the performance enhancement. The effects of UiO-66

loading, feed temperatures and feed compositions on pervaporation

performance were systematically studied.

Chapter seven shows a further development of mixed matrix membrane

consisting of two different functionalized UiO-66-type MOFs (UiO-66-NH2

and UiO-66-F4) and polyimide for alcohol dehydration via pervaporation.

13
UiO-66-NH2 and UiO-66-F4 nanoparticles with similar particle size were

synthesized and incorporated in to mixed matrix membranes. FESEM, SEM-

EDX, XRD, TGA characterizations were conducted to study morphologies,

crystallographic structure and thermal properties of nanoparticles and

membranes. Solvent uptake analyses and positron annihilation lifetime

spectroscopy test were carried out to reveal the actual mechanisms for the

improvements of separation performance. The effects of functional groups of

the MOFs particles and the effects of particle loading on the alcohol separation

performance were systematically investigated. Three different alcohols:

methanol, ethanol, isopropanol dehydration performances were studied and

compared. Long-term stability test of the UiO-66-NH2 based mixed matrix

membranes was also demonstrated.

Chapter eight draws the conclusions of the current research work and proposes

some recommendations for future work.

1.6 References

[1] R.W. Baker, Membrane technology and applications, J. Wiley, Chichester;

New York, 2004.

[2] L.Y. Jiang, Y. Wang, T.S. Chung, X.Y. Qiao, J.Y. Lai, Polyimides

membranes for pervaporation and biofuels separation, Prog. Polym. Sci. 34

(2009) 1135-1160.

[3] S. Atsumi, J.C. Liao, Metabolic engineering for advanced biofuels

production from Escherichia coli, Curr. Opin. Biotechnol. 19 (2008) 414-419.

14
[4] T. Jojima, M. Inui, H. Yukawa, Production of isopropanol by

metabolically engineered Escherichia coli, Appl. Microbiol. Biotechnol. 77

(2008) 1219-1224.

[5] L.M. Vane, Separation technologies for the recovery and dehydration of

alcohols from fermentation broths, Biofuels Bioprod. Biorefining, 2 (2008)

553-588.

[6] L.Y. Jiang, T.S. Chung, Homogeneous polyimide/cyclodextrin composite

membranes for pervaporation dehydration of isopropanol, J. Membr. Sci. 346

(2010) 45-58.

[7] G. Tusel, H. Brüschke, Use of pervaporation systems in the chemical

industry, Desalination, 53 (1985) 327-338.

[8] A. Jonquieres, R. Clément, P. Lochon, J. Néel, M. Dresch, B. Chrétien,

Industrial state-of-the-art of pervaporation and vapour permeation in the

western countries, J. Membr. Sci. 206 (2002) 87-117.

[9] I. Blume, J. Wijmans, R. Baker, The separation of dissolved organics from

water by pervaporation, J. Membr. Sci. 49 (1990) 253-286.

[10] L.M. Vane, F.R. Alvarez, Full-scale vibrating pervaporation membrane

unit: VOC removal from water and surfactant solutions, J. Membr. Sci. 202

(2002) 177-193.

[11] J. Willemsen, B. Dijkink, A. Togtema, Organophilic pervaporation for

aroma isolation—industrial and commercial prospects, Membr. Technol.

2004 (2004) 5-10.

[12] L.M. Vane, A review of pervaporation for product recovery from biomass

fermentation processes, J. Chem. Technol. Biot. 80 (2005) 603-629.

15
[13] B. Smitha, D. Suhanya, S. Sridhar, M. Ramakrishna, Separation of

organic–organic mixtures by pervaporation—a review, J. Membr. Sci. 241

(2004) 1-21.

[14] S. Sridhar, B. Smitha, A. Shaik, Pervaporation‐Based Separation of

Methanol/MTBE Mixtures—A Review, Sep. Purif. Rev. 34 (2005) 1-33.

[15] M. Yoshikawa, T. Yoshioka, J. Fujime, A. Murakami, Pervaporation

separation of MeOH/MTBE through agarose membranes, J. Membr. Sci. 178

(2000) 75-78.

[16] S.Y. Nam, Y.M. Lee, Pervaporation separation of methanol/methyl t-

butyl ether through chitosan composite membrane modified with surfactants, J.

Membr. Sci. 157 (1999) 63-71.

[17] J.G. Villaluenga, A. Tabe-Mohammadi, A review on the separation of

benzene/cyclohexane mixtures by pervaporation processes, J. Membr. Sci. 169

(2000) 159-174.

[18] X.S. Feng, R.Y.M. Huang, Liquid separation by membrane

pervaporation:  a review, Ind. Eng. Chem. Res. 36 (1997) 1048-1066.

[19] L.Y. Jiang, T.S. Chung, R. Rajagopalan, Dehydration of alcohols by

pervaporation through polyimide Matrimid® asymmetric hollow fibers with

various modifications, Chem. Eng. Sci. 63 (2008) 204-216.

[20] Y. Huang, J. Ly, D. Nguyen, R.W. Baker, Ethanol dehydration using

hydrophobic and hydrophilic polymer membranes, Ind. Eng. Chem. Res. 49

(2010) 12067-12073.

[21] M.R. LADISCH, K. DYCK, Dehydration of Ethanol: New Approach

Gives Positive Energy Balance, Science, 205 (1979) 898-900.

16
[22] P. Shao, R.Y.M. Huang, Polymeric membrane pervaporation, J. Membr.

Sci. 287 (2007) 162-179.

[23] Y. Huang, R.W. Baker, L.M. Vane, Low-energy distillation-membrane

separation process, Ind. Eng. Chem. Res. 49 (2010) 3760-3768.

[24] G. Liu, W.S. Hung, J. Shen, Q. Li, Y.H. Huang, W. Jin, K.R. Lee, J.Y.

Lai, Mixed matrix membranes with molecular-interaction-driven tunable free

volumes for efficient bio-fuel recovery, J. Mater. Chem. A, 3 (2015) 4510-

4521.

[25] F. Lipnizki, R.W. Field, P.-K. Ten, Pervaporation-based hybrid process: a

review of process design, applications and economics, J. Membr. Sci. 153

(1999) 183-210.

17
CHAPTER 2: LITERATURE REVIEW

2.1 Mass transport mechanism

A proper understanding of mass transport mechanism may give useful

information and guidelineson the development of appropriate membranes.

Pervaporation process involves a phase transition from liquid to vapor phase

during mass transport through the membrane. The driving force of mass transport

in pervaporation is chemical potential difference across the membrane. The

separation performance is affected by various factors, including (1) the

physicochemical properties of feed mixtures and their interactions, (2) the

affinities of feed components to the membrane material, and (3) the membrane

morphology and structure. Because of the complicated interactions between

permeant and permeant as well as penetrants and membranes, it is hard to

formulate one single explanation for the sophisticated transport process. There

are principally two widespread approaches to describe mass transport in

pervaporation: (i) the solution-diffusion model and (ii) the pore flow model.

2.1.1 Solution-diffusion model

The solution-diffusion model is accepted by the majority of membrane

researchers describe the mass transport through a non-porous membrane [1]. The

solution-diffusion model assumes that the pressure within a membrane is uniform

while a concentration gradient across the membrane leads to the chemical

potential difference, which provides the driving force. According to this

mechanism, pervaporation consists of three consecutive steps: (i) sorption of

18
the permeant from the feed liquid onto the membrane surface, (ii) diffusion of

the permeant in the membrane, and (iii) desorption of the permeant in a vapor

phase on the downstream side of the membrane [2]. This process is illustrated

in Figure 2.1. The first two steps are the rate-controlling steps and the difference

in solubility and diffusivity of each species in the mixtures determines the final

separation performance. Thus, to pursue high performance, a desirable

membrane should have higher solubility and diffusivity towards the permeate

components.

Generally, sorption step prefers more condensable molecules or molecules with

special interaction and/or high affinity with membrane materials, which could be

estimated by solubility parameter [3]. On the other hand, the diffusivity is

significantly affected by the molecular size and shape of the permeate, the

mobility of polymer chains, the interstitial space between polymer chains or the

fractional free volume properties of the membrane, as well as the interactions

between the permeate components and membrane [4].

Figure 2.1 Schematic diagram of solution-diffusion transport model

19
2.1.2 Pore flow model

Pore flow model is proposed more recently by Okada and Matsuura [5].

Figure 2.2 illustrates the schematic diagram of the pore-flow mechanism.

Different from solution-diffusion model which is applied for non-porous

membranes, pore flow model assumes there are straight and cylindrical pores

with an effective length of δ, penetrating in the selective layer of membranes.

The pores are perpendicular to the membrane surface and all the pores are

under an isothermal condition. Moreover, it also hypothesizes that the pores

are filled with liquid with a distance of δ1 while the rest part of the pores is

filled with vapor with a length of δ2. In other words, there is a liquid/vapor

phase boundary somewhere in the middle of the membrane pores. In this

model, the mass transport process also includes three steps: (i) the permeates

transport through the liquid-filled portion of the pores from feed solution, (ii) a

liquid-to-vapor phase transition happens at the phase interface and (iii) the

permeates transport through the vapor-filled portion of the pores to

downstream. In other words, this model suggests that the mass transport can

be viewed as a combination of liquid-phase and vapor-phase transport in series.

Thus, this model simplifies the membrane structure. Pore flow model has been

studied and applied for various membranes and different separating mixtures

[6-9]. However, pore flow model may neglect the membrane-permeant

interaction and might not be suitable if the membrane is severely swelled by

the feed solution.

20
Figure 2.2 Schematic diagram of pore flow model

Solution-diffusion model and pore flow model employ different approaches to

describe pervaporation process and they can be used to explain and predict

membrane performance in different conditions.

2.2 Evaluation of pervaporation membrane performance

The selection of a pervaporation membrane is generally based on its capability

to separate mixtures. There are typically two sets of inter-connected systems to

define membrane performance, which are flux-separation factor system and

permeability-selectivity system. The flux-separation factor system is used to

evaluate the apparent membrane performance obtained under a certain

operation condition, while the permeability-selectivity system is employed to

describe the intrinsic membrane performance which is independent of the

operation conditions.

The flux (J) was determined by the total mass of permeate (Q) divided by the

product of the effective membrane area (A) and interval time (t).

21
Q
J (2.1)
A t

The separation factor α is defined by the equation below:

ywi / ywj
 (2.2)
xwi / xwj

Wherein, y w and x w are the weight fractions of the component in the

permeate and feed, respectively; the subscripts i and j refer to components of

water and alcohol in this work, respectively.

Flux and separation factor are usually reported in most studies, and are widely

used in the comparison with literature date. However, flux and separation

factor lie on both the intrinsic properties of membrane material and operation

parameters (e.g. film thickness and temperature, feed compositions). Hence, to

get intrinsic properties of membranes, flux and separation factor could be

converted to permeability (P) and selectivity (β) based on the solution-

diffusion model [10]. The permeability Pi can be calculated by:

Ji  l
Pi  (2.3)
pif  pip

Wherein, l is the membrane thickness. J i is the flux of component i . While

pif and pip are the partial vapor pressure of component i in feed and permeate,

respectively. Feed solutions in pevaporation process is in liquid state.

f
Therefore, pi represents the partial vapor pressure of component i in a

hypothetical vapor phase in equilibrium with the feed liquid, which can be

determined using the Raoult’s Law:

pif  xi i pisat (2.4)

22
Wherein, xi is the mole fraction of component i in the feed.  i and pi are the
sat

activity coefficient and the saturated vapor pressure of component i ,

respectively. Both  i and pisat can be determined using the AspenTech

Process Modelling software (version 7.2) based on the Wilson equation and

the Antoine equation. The partial vapor pressure of component i in the

permeate pip is calculated by:

pip  yi p p (2.5)

Wherein, yi represents the mole fraction of component i in the permeate. p p

is the total pressure at permeate side, was assumed to be zero due to the

vacuum condition at permeate side. As a result, the value of pip is also

determined to be zero. Therefore, permeability could be defined by the

following equation:

Ji  l
Pi  (2.6)
xi i pisat

Ideal selectivity of the membrane (β) was expressed as the ratio of the

permeability of components i and j:


𝑃
β = 𝑃𝑖 (2.7)
𝑗

2.3 Pervaporation membranes for alcohol dehydration

Since dehydration of organics, particularly alcohol dehydration is the most

significant application of pervaporation in the industry as aforementioned, a

thorough review of pervaporation membranes is focused on this application.

23
The two important aspects of pervaporation membrane development are

included: (1) membrane materials selection and (2) membrane morphology,

structure and configurations.

2.3.1 Membrane materials

To design a pervaporation membrane with good separation performance, two

aspects should be considered. One is affinity between the feed components

and membrane material. For alcohol dehydration, hydrophilic materials are

usually used due to their preferential sorption for water molecules. However,

higher hydrophilicity may cause higher degree of membrane swelling, which

results in an increase of permeation flux but a deterioration of separation

factor. Therefore, a balanced hydrophilicity and hydrophobicity is needed. The

other is the free-volume property of the membrane material. Normally a higher

fractional free volume gives rise to a higher diffusivity but a lower diffusivity

selectivity. Besides, the general evaluation criteria for pervaporation

membrane materials also includes high chemical stability and good

mechanical strength.

The materials available for pervaporation mainly include polymeric and

inorganic materials. Generally, polymeric membranes are easy to fabricate and

relatively cheap but often encounter the trade-off between permeability and

selectivity. While inorganic membranes have good mechanical stability and

high performance, but are limited by the expensive costs and fragility. In the

24
following part, both polymeric and inorganic materials as well as mixed

matrix membrane are review.

• Polymeric materials

In the early development stage, highly hydrophilic polymers, such as

polyvinyl alcohol (PVA) [11], cellulose, chitosan [12] and alginate [13], have

been widely used to dehydrate alcohols. The high hydrophilicity of these

polymers would enhance water transport and improve solubility selectivity of

water through hydrogen bonding interactions. However, these membranes are

susceptible to swell, which leads to both unstable long-term performance and

low mechanical strengths. To stabilize the membranes, various modifications

have been applied, such as chemical cross-linking, UV irradiation crosslinking

and plasma grafting [1, 14].

In recent years, research focus of pervaporation membranes for solvent

dehydration has transferred to the exploration of new materials with good

chemical and thermal stable materials to dehydrate aggressive solvents at high

operating temperatures. Therefore, polymers with stiff and rigid chains have

been subsequently developed in this application [15]. These polymers have

excellent solvent resistance, good thermal stability and high mechanical

strength due to the rigid polymer chains. Moreover, the glassy characteristic of

these polymers could help to improve the diffusivity selectivity of the

membranes.

25
Polyimides have been viewed as a kind of promising polymers due to its

excellent thermal stability, high chemical resistance and good mechanical

strength, as well as good film forming property [15]. There are some

commercially available polyimide materials such as P84, Matrimid, Ultem and

Torlon, which have been used widely in recent years. Usually, polyimides are

synthesized by a two-step approach including polycondensation between

amine monomer and anhydride monomer to form poly(amic acid) and

imidization of poly(amic acid) to form polyimide. The easy synthesis

procedures and plenty choices of amine and anhydride monomers as well as

simple modification techniques make it versatile and flexibile to tailor the

structures and physiochemical properties of polyimides. Recently, a lot of

work have been done by Chung’s research group using commercial

polyimides or self-synthesized polyimides for alcohol dehydration [16-19].

Liu et al. successfully applied P84 co-polyimide hollow fiber membranes for

IPA dehydration [16] and modifications including silicon rubber coating and

post heat treatment are used to enhance membrane performance. Jiang and

Chung developed polyimide/cyclodextrin composite membranes using a

commercial polyimide Matrimid® for IPA dehydration [17]. Le et al.

synthesized co-polyimide using 1,5-naphthalene, 3,5-benzoicacid and 2,2′-

bis(3,4-dicarboxyphenyl) hexafluoro-propanedimide and fabricated dual-layer

hollow fiber membranes using this co-polyimide as the selective layer [18].

Xu et al. successfully synthesized co-polyimide using three monomers

including 4,4’-(hexafluoroisopropylidene) diphthalic anhydride, 3,3’-

dihydroxybenzidine diamine and 3,5-diaminobenzoic acid and the polyimide

26
membrane shows a good separation efficiency with water concentration in

permeate of 99.9% [19].

Polyamides known as Nylon are another kind of heat resistance materials with

good chemical and mechanical stability [20-22]. Polyamides can be

synthesized by step-growth polymerization, such as interfacial polymerization.

Usually, polyamide dense membranes have extremely low flux. Therefore,

polyamides are often used as a thin selective layer in composite membranes to

minimize the transport resistance and improve separation performance.

Interfacial polymerization has been demonstrated to be an effective way to

form a thin layer of polyamide on a substrate.

Amorphous perfluoropolymer such as Teflon AF, Hyflon AD and Cytop is

another class of promising membrane materials with high free-volume. They

have extraordinary thermal and chemical stability thus can effectively resist

aggressive chemicals including acids, bases, organic solvents, oils and strong

oxidizers. Despite of their intrinsic hydrophobicity, the perfluoropolymer-

based membranes have been applied for dehydration of butanol, isopropanol,

ethanol, N,N-dimethylformamide (DMF), N,N-dimethylsulfoxide (DMSO),

N,N-dimethylacetamide (DMAc) and hydrogen peroxide (H2O2) [23-26] based

on size exclusion mechanism. Studies reveal that perfluoropolymer-based

membranes have a very low sorption of alcohol/water mixtures. They have

high permeability due to the large free volumes but relatively low selectivity.

27
In addition, aromatic polymers such as polybenzoxazole (PBO),

polybenzoxazinone (PBOZ) and polybenzimidazole (PBI) have also been

developed with a great potential in solvent dehydration due to their excellent

chemical and thermal stability. Both PBO and PBOZ membranes were

synthesized by the thermal rearrangement process from their respective

precursors and were applied in dehydration of alcohols [19, 27, 28]. Ong et al.

and Xu et al. found that the thermally rearranged PBO membranes show a

stable performance in dehydration of isopropanol throughout long-term period

of 200-250h [19, 27] whereas Pulyalin A et al. found that water diffusivity

was greatly improved in the PBOZ membrane compared to its precursor [28].

On the other hand, PBI-based pervaporation membranes were pioneered by

Chung and co-workers [29-32]. They have developed PBI membranes both in

flat-sheet and hollow fiber configurations for dehydration of different solvents

such as alcohols, glycols and acetone. In addition, PBI was also used as a filler

to improve the performance of Matrimid® membranes by polymer blending

[33].

• Inorganic materials

Inorganic materials possess some unique advantages of outstanding over

polymeric materials such as outstanding chemical resistance and high

temperature stability. They usually do not have solvent-induced swelling issue,

and can thus be applied in harsh chemical environment. Because of the ability to

be used at high temperature, inorganic membranes show high flux with a stable

long-term performance. However, inorganic membranes have weaknesses of

28
brittleness, high cost and complicated processbility, which restrict their wide

applications in industry. The separation of inorganic membranes is mainly

achieved by size exclusion mechanism. Inorganic membranes developed for

alcohol dehydration include zeolite, silica, ceramic, carbon, graphene oxide

and metal organic frameworks (MOFs) etc. [34-39]. Currently, most of the

commercialized inorganic membranes are made of zeolites and silica. Other

inorganic materials are still at the development stage.

Zeolites membranes have high permeability and selectivity due to the highly

ordered and well-defined structure. Zeolites have various structures with different

pore sizes thus can have molecular sieving property for different pairs of mixtures.

Recently developed zeolite membranes, such as zeolite A, zeolite T and zeolite X,

have remarkable separation performances, better than traditional polymeric

membranes. Moreover, zeolite membranes can overcome the problem of swelling

and have stable performance under harsh environments. Morigami et al. first

reported the industrial application of zeolite membranes in an ethanol production

plant [40]. This shows the feasibility of large-scale use of zeolite membranes in

industry.

Carbon membranes made from polymeric precursors have shown promising

performance in dehydration of alcohols. However, the fragility and poor

mechanical strength limit their further development. Graphene oxide (GO)

membranes, which can be viewed as a novel type of carbon-based membrane,

recently have attracted a lot of attention since it is reported to have ultra-high

water permeability but almost no permeation of organic vapor [41].

Furthermore, GO membranes possess exceptionally good mechanical

29
properties and high stability in water due to the hydrogen-bonding interactions

between GO nanosheets. Recently, Yeh et al. and Hung et al. applied the GO

membranes for the dehydration of ethanol and isopropanol, respectively [42, 43].

These membranes show remarkable permeability and high selectivity. However,

there are still some issues that need to be addressed to further develop GO

membranes such as fabrication of an ultra-thin and defect-free membrane to

achieve higher separation performance and stability in feed solutions,

particularly at flowing mode or high temperature.

Besides, metal organic frameworks (MOFs) are also an emerging family of

inorganic materials due to their unique characteristics such as large porosity,

controllable window size, high adsorption capacity and tuneable chemical

properties [44]. It has been a hot topic for gas separation [45]. In recent years,

MOFs membranes have also been applied in pervaporation. However, some

MOFs are not stable in water, therefore very few research works develop pure

MOFs membranes for alcohol-water separation. Nonetheless, MOFs are still

very promising to be used as fillers to enhance the separation performance of

polymeric membranes. In 2008, Cavka et al. successfully synthesized UiO-66

by ZrCl4 and 1,4-benzene-dicarboxylate (BDC) [46]. UiO-66-Type MOFs

have received significant attention because of the high stability in various

solvents, which makes it very hopeful to be used in solvent dehydration via

pervaporation.

30
• Mixed matrix membranes

In order to combine the strengths of both polymeric and inorganic materials,

mixed matrix membranes (MMMs) were first developed by Kulprathipanja et

al. in 1988 [47]. Afterwards, numerous inorganic fillers including zeolites,

carbon molecular sieves, silica, metal oxides, carbon nanotubes, silicate, metal

organic frameworks (MOFs), graphene oxide (GO), porous organic cages,

covalent organic frameworks and many others have been investigated and

employed in mixed matrix membranes to improve pervaporation performance

[14, 44, 45, 48]. Among them, MOFs, have been demonstrated as promising

materials for MMMs due to their advantages such as large surface areas,

tuneable porosity, high adsorption capacity and good affinity with polymer

chains, as well as the controllable chemical properties [44, 45, 48].

The incorporation of inorganic fillers has different effects on the performance

of MMMs: some works reported an enhancement of flux but a sacrifice of

separation factor [49], while others observed improvements of both flux and

separation factor [50]. The different results can be ascribed to several factors,

including the compatibility between inorganic fillers and the polymeric matrix,

the hydrophilicity of inorganic fillers, the pore size of inorganic fillers and the

rigidity of polymer chains etc.

One challenge faced by MMMs is the agglomeration of inorganic fillers in the

polymeric matrix which generates defects and results in a decrease in

selectivity [14]. To improve the compatibility between the polymeric and the

31
inorganic phases and achieve an even dispersion of inorganic fillers, following

approaches have been developed recently: (1) modifying inorganic fillers with

coupling agents which have strong bondings with polymeric phase [51]; (2)

coating inorganic fillers with a thin polymeric layer [52, 53]; (3) employing

hybrid particles with both organic and inorganic components, such as

polyhedral oligomeric silsesquioxane (POSS) [18, 54] or MOFs [49, 50, 55];

and (4) developing new MMMs fabrication methods [56].

2.3.2 Membrane morphology and configuration

Besides membrane materials, membrane morphology and configuration also

greatly affect the separation performance of a membrane. Therefore, it is

important to design a suitable membrane structure with desired morphology

and configuration according to different separation requirements. Based on

morphology, membranes can be classified into two categories: (1) symmetric

and (2) asymmetric membranes, as illustrated in Figure 2.3.

Figure 2.3 Typical membrane morphologies

32
Asymmetric membranes are developed to reduce thickness of selective layer

of membranes thus to achieve a high flux. Generally, an asymmetric

membrane possesses an ultra-thin dense selective layer and a porous substrate

as mechanical support. Based on materials used, there are two types of

asymmetric membranes: (1) wholly integral asymmetric membranes made from

the same material for selective layer and porous substrate and (2) composite

membranes with different materials for selective layer and porous substrate.

Usually, the wholly integral asymmetric membranes are prepared by non-

solvent phase inversion method. In this method, a homogeneous polymer

solution is precipitated in a non-solvent coagulation bath then a continuous

porous solid phase and a dense skin layer are formed. This phase inversion

process is quite complicated therefore fabricating conditions needs to be well

designed to obtain a desired structure. Composite membranes have more

freedom in designing and fabricating. They can be fabricated by direct coating,

casting on the substrate, layer-by-layer assembly, in-situ surface polymerization

such as interfacial polymerization and plasma polymerization [57, 58]. A small

amount of expensive materials with outstanding separation performance can

be used for the ultra-thin selective layer, while the porous substrate can be

made from cheap materials. Therefore, composite membranes are cost-

efficient and promising for industrialization.

Based on configuration, membranes can be divided into flat sheet, hollow

fiber or tubular membranes. Among them, hollow fiber membranes have been

widely used because hollow fiber membranes have superior advantages such

as high surface area, high packing density, self-support structure, excellent

33
flexibility and ease of fabrication as well as scale-up [1, 59]. Conventional

hollow fiber membranes have single-bore geometry. Recently, multi-bore

hollow fiber membranes have been widely developed [60, 61]. Multi-bore

hollow fiber membranes show higher mechanical properties compared with

traditional single-bore hollow fiber membranes thus can reduce the chances of

fiber breakage and increase operation stability. Currently, multi-bore hollow

fiber membranes are still in the developing stage and are only commercialized

for ultrafiltration application.

2.4 References

[1] X.S. Feng, R.Y.M. Huang, Liquid separation by membrane pervaporation: 

a review, Ind. Eng. Chem. Res. 36 (1997) 1048-1066.

[2] J.G. Wijmans, R.W. Baker, The solution-diffusion model: a review, J.

Membr. Sci. 107 (1995) 1-21.

[3] F. Xianshe, R.Y.M. Huang, Estimation of activation energy for permeation

in pervaporation processes, J. Membr. Sci. 118 (1996) 127-131.

[4] J. Mulder, Basic principles of membrane technology, Springer Science &

Business Media, 2012.

[5] T. Okada, T. Matsuura, A new transport model for pervaporation, J.

Membr. Sci. 59 (1991) 133-149.

[6] S. Deng, B. Shiyao, S. Sourirajan, T. Matsuura, A study of the

pervaporation of isopropyl alcohol/water mixtures by cellulose acetate

membranes, J. Colloid. Interf. Sci. 136 (1990) 283-291.

34
[7] T. Okada, T. Matsuura, Predictability of transport equations for

pervaporation on the basis of pore-flow mechanism, J. Membr. Sci. 70 (1992)

163-175.

[8] T. Okada, M. Yoshikawa, T. Matsuura, A study on the pervaporation of

ethanol/water mixtures on the basis of pore flow model, J. Membr. Sci. 59

(1991) 151-168.

[9] P. Sukitpaneenit, T.S. Chung, Molecular design of the morphology and

pore size of PVDF hollow fiber membranes for ethanol–water separation

employing the modified pore-flow concept, J. Membr. Sci. 374 (2011) 67-82.

[10] R.W. Baker, J.G. Wijmans, Y. Huang, Permeability, permeance and

selectivity: A preferred way of reporting pervaporation performance data, J.

Membr. Sci. 348 (2010) 346-352.

[11] B. Bolto, T. Tran, M. Hoang, Z. Xie, Crosslinked poly (vinyl alcohol)

membranes, Prog. Polym. Sci. 34 (2009) 969-981.

[12] M. Goto, A. Shiosaki, T. Hirose, Separation of water/ethanol vapor

mixtures through chitosan and crosslinked chitosan membranes, Sep. Sci.

Technol. 29 (1994) 1915-1923.

[13] R.Y.M. Huang, R. Pal, G.Y. Moon, Characteristics of sodium alginate

membranes for the pervaporation dehydration of ethanol-water and

isopropanol-water mixtures, J. Membr. Sci. 160 (1999) 101-113.

[14] Y.K. Ong, G.M. Shi, N.L. Le, Y.P. Tang, J. Zuo, S.P. Nunes, T.-S. Chung,

Recent membrane development for pervaporation processes, Prog. Polym. Sci.

57 (2016) 1-31.

35
[15] L.Y. Jiang, Y. Wang, T.S. Chung, X.Y. Qiao, J.-Y. Lai, Polyimides

membranes for pervaporation and biofuels separation, Prog. Polym. Sci. 34

(2009) 1135-1160.

[16] R. Liu, X. Qiao, T.S. Chung, The development of high performance P84

co-polyimide hollow fibers for pervaporation dehydration of isopropanol,

Chem. Eng. Sci. 60 (2005) 6674-6686.

[17] L.Y. Jiang, T.S. Chung, Homogeneous polyimide/cyclodextrin composite

membranes for pervaporation dehydration of isopropanol, J. Membr. Sci. 346

(2010) 45-58.

[18] N.L. Le, Y.P. Tang, T.S. Chung, The development of high-performance

6FDA-NDA/DABA/POSS/Ultem® dual-layer hollow fibers for ethanol

dehydration via pervaporation, J. Membr. Sci. 447 (2013) 163-176.

[19] Y.M. Xu, N.L. Le, J. Zuo, T.S. Chung, Aromatic polyimide and

crosslinked thermally rearranged poly (benzoxazole-co-imide) membranes for

isopropanol dehydration via pervaporation, J. Membr. Sci. 499 (2016) 317-

325.

[20] X. Tang, R. Wang, Z. Xiao, E. Shi, J. Yang, Preparation and

pervaporation performances of fumed-silica-filled polydimethylsiloxane–

polyamide (PA) composite membranes, J. Appl. Polym. Sci. 105 (2007) 3132-

3137.

[21] Y. Wang, S.H. Goh, T.S. Chung, P. Na, Polyamide-imide/polyetherimide

dual-layer hollow fiber membranes for pervaporation dehydration of C1–C4

alcohols, J. Membr. Sci. 326 (2009) 222-233.

[22] Y. Wang, L.Y. Jiang, T. Matsuura, T.S. Chung, S.H. Goh, Investigation

of the fundamental differences between polyamide-imide (PAI) and

36
polyetherimide (PEI) membranes for isopropanol dehydration via

pervaporation, J. Membr. Sci. 318 (2008) 217-226.

[23] J. Tang, K.K. Sirkar, Perfluoropolymer membrane behaves like a zeolite

membrane in dehydration of aprotic solvents, J. Membr. Sci. 421 (2012) 211-

216.

[24] S. Roy, A. Thongsukmak, J. Tang, K.K. Sirkar, Concentration of aqueous

hydrogen peroxide solution by pervaporation, J. Membr. Sci. 389 (2012) 17-

24.

[25] V. Smuleac, J. Wu, S. Nemser, S. Majumdar, D. Bhattacharyya, Novel

perfluorinated polymer-based pervaporation membranes for the separation of

solvent/water mixtures, J. Membr. Sci. 352 (2010) 41-49.

[26] Y. Huang, R.W. Baker, J. Wijmans, Perfluoro–coated hydrophilic

membranes with improved selectivity, Ind. Eng. Chem. Res. 52 (2012) 1141-

1149.

[27] Y.K. Ong, H. Wang, T.S. Chung, A prospective study on the application

of thermally rearranged acetate-containing polyimide membranes in

dehydration of biofuels via pervaporation, Chem. Eng. Sci. 79 (2012) 41-53.

[28] A. Pulyalina, G. Polotskaya, M. Goikhman, I. Podeshvo, L. Kalyuzhnaya,

M. Chislov, A. Toikka, Study on polybenzoxazinone membrane in

pervaporation processes, J. Appl. Polym. Sci. 130 (2013) 4024-4031.

[29] K.Y. Wang, T.S. Chung, R. Rajagopalan, Dehydration of

tetrafluoropropanol (TFP) by pervaporation via novel PBI/BTDA-TDI/MDI

co-polyimide (P84) dual-layer hollow fiber membranes, J. Membr. Sci. 287

(2007) 60-66.

37
[30] Y. Wang, M. Gruender, T.S. Chung, Pervaporation dehydration of

ethylene glycol through polybenzimidazole (PBI)-based membranes. 1.

Membrane fabrication, J. Membr. Sci. 363 (2010) 149-159.

[31] Y. Wang, T.S. Chung, B.W. Neo, M. Gruender, Processing and

engineering of pervaporation dehydration of ethylene glycol via dual-layer

polybenzimidazole (PBI)/polyetherimide (PEI) membranes, J. Membr. Sci.

378 (2011) 339-350.

[32] G.M. Shi, Y. Wang, T.S. Chung, Dual‐layer PBI/P84 hollow fibers for

pervaporation dehydration of acetone, AIChE J.58 (2012) 1133-1145.

[33] T.S. Chung, W.F. Guo, Y. Liu, Enhanced Matrimid membranes for

pervaporation by homogenous blends with polybenzimidazole (PBI), J.

Membr. Sci. 271 (2006) 221-231.

[34] N. Itoh, J. Ishida, Y. Kikuchi, T. Sato, Y. Hasegawa, Continuous

dehydration of IPA–water mixture by vapor permeation using Y type zeolite

membrane in a recycling system, Sep. Purif. Technol.147 (2015) 346-352.

[35] A. Verkerk, P. Van Male, M. Vorstman, J. Keurentjes, Description of

dehydration performance of amorphous silica pervaporation membranes, J.

Membr. Sci. 193 (2001) 227-238.

[36] A. Verkerk, P. Van Male, M. Vorstman, J. Keurentjes, Properties of high

flux ceramic pervaporation membranes for dehydration of alcohol/water

mixtures, Sep. Purif. Technol. 22 (2001) 689-695.

[37] P.S. Tin, H.Y. Lin, R.C. Ong, T.S. Chung, Carbon molecular sieve

membranes for biofuel separation, Carbon, 49 (2011) 369-375.

38
[38] Y.P. Tang, D.R. Paul, T.S. Chung, Free-standing graphene oxide thin

films assembled by a pressurized ultrafiltration method for dehydration of

ethanol, J. Membr. Sci. 458 (2014) 199-208.

[39] A. Kasik, Y. Lin, Organic solvent pervaporation properties of MOF-5

membranes, Sep. Purif. Technol. 121 (2014) 38-45.

[40] Y. Morigami, M. Kondo, J. Abe, H. Kita, K. Okamoto, The first large-

scale pervaporation plant using tubular-type module with zeolite NaA

membrane, Sep. Purif. Technol.25 (2001) 251-260.

[41] D.R. Dreyer, S. Park, C.W. Bielawski, R.S. Ruoff, The chemistry of

graphene oxide, Chem. Soc. Rev. 39 (2010) 228-240.

[42] T.M. Yeh, Z. Wang, D. Mahajan, B.S. Hsiao, B. Chu, High flux ethanol

dehydration using nanofibrous membranes containing graphene oxide barrier

layers, J. Mater. Chem. A, 1 (2013) 12998-13003.

[43] W.S. Hung, Q.F. An, M. De Guzman, H.Y. Lin, S.H. Huang, W.R. Liu,

C.C. Hu, K.R. Lee, J.Y. Lai, Pressure-assisted self-assembly technique for

fabricating composite membranes consisting of highly ordered selective

laminate layers of amphiphilic graphene oxide, Carbon, 68 (2014) 670-677.

[44] Z. Jia, G. Wu, Metal-organic frameworks based mixed matrix membranes

for pervaporation, Microporous Mesoporous Mat.235 (2016) 151-159.

[45] Y. Zhang, X. Feng, S. Yuan, J. Zhou, B. Wang, Challenges and recent

advances in MOF-polymer composite membranes for gas separation, Inorg.

Chem. Front. 3 (2016) 896-909.

[46] J.H. Cavka, S. Jakobsen, U. Olsbye, N. Guillou, C. Lamberti, S. Bordiga,

K.P. Lillerud, A new zirconium inorganic building brick forming metal

39
organic frameworks with exceptional stability, J. Am. Chem. Soc. 130 (2008)

13850-13851.

[47] S. Kulprathipanja, R.W. Neuzil, N.N. Li, Separation of fluids by means of

mixed matrix membranes, Google Patents, 1988.

[48] J. Caro, Are MOF membranes better in gas separation than those made of

zeolites?, Curr. Opin. Chem. Eng. 1 (2011) 77-83.

[49] G.M. Shi, T.X. Yang, T.S. Chung, Polybenzimidazole (PBI)/zeolitic

imidazolate frameworks (ZIF-8) mixed matrix membranes for pervaporation

dehydration of alcohols, J. Membr. Sci. 415-416 (2012) 577-586.

[50] D. Hua, Y.K. Ong, Y. Wang, T.X. Yang, T.S. Chung, ZIF-90/P84 mixed

matrix membranes for pervaporation dehydration of isopropanol, J. Membr.

Sci. 453 (2014) 155-167.

[51] I.F. Vankelecom, S. Van den broeck, E. Merckx, H. Geerts, P. Grobet,

J.B. Uytterhoeven, Silylation to improve incorporation of zeolites in polyimide

films, J. Phys. Chem. 100 (1996) 3753-3758.

[52] M.D. Jia, K.V. Pleinemann, R.D. Behling, Preparation and

characterization of thin-film zeolite–PDMS composite membranes, J. Membr.

Sci. 73 (1992) 119-128.

[53] Y. Huang, P. Zhang, J. Fu, Y. Zhou, X. Huang, X. Tang, Pervaporation of

ethanol aqueous solution by polydimethylsiloxane/polyphosphazene nanotube

nanocomposite membranes, J. Membr. Sci. 339 (2009) 85-92.

[54] D. Xu, L.S. Loo, K. Wang, Pervaporation performance of novel

chitosan‐POSS hybrid membranes: effects of POSS and operating conditions,

J. Polym. Sci. Pol. Phys. 48 (2010) 2185-2192.

40
[55] Y.M. Xu, T.S. Chung, High-performance UiO-66/polyimide mixed

matrix membranes for ethanol, isopropanol and n-butanol dehydration via

pervaporation, J. Membr. Sci. 531 (2017) 16-26.

[56] Q. Zhao, J. Qian, C. Zhu, Q. An, T. Xu, Q. Zheng, Y. Song, A novel

method for fabricating polyelectrolyte complex/inorganic nanohybrid

membranes with high isopropanol dehydration performance, J. Membr. Sci.

345 (2009) 233-241.

[57] W.S. Hung, M. De Guzman, S.H. Huang, K.R. Lee, Y. Jean, J.Y. Lai,

Characterizing free volumes and layer structures in asymmetric thin-film

polymeric membranes in the wet condition using the variable monoenergy

slow positron beam, Macromolecules, 43 (2010) 6127-6134.

[58] T. Yamaguchi, S. Nakao, S. Kimura, Plasma-graft filling polymerization:

preparation of a new type of pervaporation membrane for organic liquid

mixtures, Macromolecules, 24 (1991) 5522-5527.

[59] F. Lipnizki, R.W. Field, P.-K. Ten, Pervaporation-based hybrid process: a

review of process design, applications and economics, J. Membr. Sci. 153

(1999) 183-210.

[60] P. Sukitpaneenit, T.S. Chung, Fabrication and use of hollow fiber thin

film composite membranes for ethanol dehydration, J. Membr. Sci. 450 (2014)

124-137.

[61] N. Peng, N. Widjojo, P. Sukitpaneenit, M.M. Teoh, G.G. Lipscomb, T.S.

Chung, J.Y. Lai, Evolution of polymeric hollow fibers as sustainable

technologies: past, present, and future, Prog. Polym. Sci. 37 (2012) 1401-1424.

41
CHAPTER 3: EXPERIMENTAL

3.1 Materials

3.1.1 Polymers

Four kinds of polyarylether polymers: PESU (Ultrason® E6020P), PPSU

(Ultrason® P3010), T-PESU (GM1324-0016/0059) and H-PESU (GM584/P-

0302/1) were obtained from BASF SE, Germany.

3.1.2 Chemicals for polyimide syntheses

The 6FDA-HAB/DABA co-polyimides were synthesized from monomers

6FDA, HAB and DABA with four various diamine HAB to DABA ratios of

10:0, 9:1, 7:3 and 5:5 (abbreviated as PI-10-0, PI-9-1, PI-7-3 and PI-5-5) via

the imidization reaction. The monomers 4,4’-(hexafluoroisopropylidene)

diphthalic anhydride (6FDA, Clariant) and 3,5-diaminobenzoicacid (DABA,

Aldrich) were purified by vacuum sublimation while 3,3’-dihydroxybenzidine

diamine (HAB, Tokyo Chemical Industry) was used without further treatments.

Acetic anhydride and pyridine from Aldrich of the reagent grade or higher,

were used as received.

3.1.3 Chemicals for UiO-66-Type MOFs syntheses

For UiO-66-Type MOFs synthesis, zirconium (IV) chloride from Merck

(Germany), terephthalic acid and 2-aminoterephthalic acid from Aldrich

42
(Singapore), tetrafluorobenzene-1,4-dicarboxylic acid from Matrix Scientific

and acetic acid (glacial) from Merck (Germany) were obtained and used as

received.

3.1.4 Organic solvents

N-methyl-2-pyrrolidone (NMP, analytical grade) from Merck (Germany) was

purified by vacuum distillation as solvent for polyimide syntheses.

Dimethylformamide (DMF) of the reagent grade or higher, from Fisher

Scientific (UK) was used as solvents for UiO-66-Type MOFs syntheses and

for membrane casting. Methanol, ethanol, isopropanol and n-butanol from

Fisher Scientific (UK) of the reagent grade or higher, were used to prepare the

aqueous feed solutions for pervaporation studies. Hexane (HPLC grade) from

Merck (Germany) was used for density measurements.

3.2 Polyimide syntheses

HAB and DABA at ratios of 10:0, 9:1, 7:3 and 5:5 were accurately weighed

and dissolved in freshly distilled NMP under a nitrogen environment. The

mixture was then cooled to 0 °C and an equimolar amount of 6FDA

dianhydride was added. The reaction mixture was stirred for half an hour at

0 °C followed by one day at room temperature to form a viscous poly(amic

acid) solution. Subsequently, in order to transform the poly(amic acid) to a

polyimide, acetic anhydride and pyridine were poured into the solution.

Further stirring in one day was conducted to achieve the complete imidization.

43
The synthesis scheme is shown in Figure 3.1. The polyimide solution was then

precipitated in methanol, washed several times with methanol and dried in an

oven at 120 °C for 12 h.

Figure 3.1 Synthesis scheme of polyimide and crosslinked poly(benzoxazole-


co-imide)

3.3 UiO-66-Type MOFs syntheses

To synthesize UiO-66-Type MOFs, 0.080g ZrCl4, 0.057g terephthalic acid

(for UiO-66) or 0.062g 2-aminoterephthalic acid (for UiO-66-NH2) or 0.082g

tetrafluorobenzene-1,4-dicarboxylic acid (for UiO-66-F4) and 0.6mL acetic

acid (glacial) were added into 20mL DMF in a well-sealed blue cap bottle.

The mixture was stirred at 100°C (for UiO-66 and UiO-66-NH2) or at room

temperature (for UiO-66-F4) for 24h. After cooling to room temperature, the

44
UiO-66-Type MOFs particles were obtained by centrifugation and further

purified by washing with fresh DMF for several times. After that, the UiO-66-

Type MOFs particles were immersed in fresh DMF prior to use. The ligand

structures of UiO-66-Type MOFs are presented in Figure 3.2.

Figure 3.2 The ligand structures of UiO-66-Type MOFs

3.4 Membrane fabrication

3.4.1 Fabrication of polyarylether membranes by knife casting

The flat dense membranes were prepared by the knife casting and solvent

evaporation method. Prior to use, all polymers were dried overnight in a

vacuum oven at 120°C. Firstly, the polymer was dissolved in NMP to prepare

a 20 wt% dope solution. After completely dissolving and degassing, the

solution was cast on a glass plate using a casting knife with a gap thickness of

200μm. The as-cast membrane was dried in a vacuum oven at 55°C to

guarantee a slow solvent evaporation. After peeled off from the glass plate,

the membrane was further dried in a vacuum oven at 250°C for 15h to remove

the residual solvent and then naturally cooled down.

45
3.4.2 Fabrication of polyimide dense membranes by ring casting

The synthesized polyimide was dissolved in DMF to prepare a 3wt% polymer

solution. The solution was then filtered through a 1µm polytetrafluoroethylene

(PTFE) syringe filter and poured onto silicon wafers. The solution was left to

dry at 80 °C with a monitored condition to ensure slow evaporation of the

solvent. The formed membranes were then removed carefully. In order to

totally remove the residual solvent, the membranes were further dried under

vacuum using a gradient heating procedure: kept at 75 °C for 1h, raised to

200 °C at a rate of 25 °C/h, kept at 200 °C for 12h and naturally cooled down

to room temperature. The thickness of the resultant membrane was about 20-

30 mm measured by a Mitutoyo micrometer.

3.4.3 Fabrication of crosslinked polybenzoxazole membranes by thermal

treatment

In order to obtain a crosslinked polybenzoxazole membrane, the polyimide

dense membrane was thermally treated in a Lenton horizontal vacuum tube

furnace (Model VTF12/50/550) at a vacuum level of ~10 -6 mbar. The heating

protocol was to increase temperature to 450 °C at a rate of 5 °C/min and then

to keep isothermally for half an hour in a high vacuum level. After thermal

treatment, the membrane was naturally cooled down to room temperature in

the furnace before tests. The resultant membranes are abbreviated as C-TR-10-

0, C-TR-9-1, C-TR-7-3 and C-TR-5-5. It should be noted that C-TR-10-0

means there is no crosslinking due to the absence of DABA moiety.

46
3.4.4 Fabrication of mixed matrix membranes by ring casting

The mixed matrix membranes in this study were made by the solution casting

and solvent evaporation method. Firstly, the synthesized polyimide was

dissolved in DMF to prepare a 5 wt% polymer solution. Then, the solution

was filtered using a 5 µm PTFE syringe filter before mixing with the UiO-66-

Type nanoparticles. After centrifugation and washing, the wet-state UiO-66-

Type nanoparticles were re-dispersed in fresh DMF and sonicated for 1h to

achieve good dispersion. After that, the UiO-66-Type nanoparticles were

directly mixed with the polymer solution. The mixture was stirred for another

30min and cast onto a silicon wafer. The casting solution was put in an oven at

80°C to ensure slow evaporation of the solvent. After one to two days, the

mixed matrix membranes were peeled off from the silicon wafer. To totally

remove the residual solvent, the membranes were further dried in a vacuum

oven at 200°C for 12h and naturally cooled down.

3.5 Characterizations

3.5.1 Fourier transform infrared spectrometer (FTIR)

The chemical structures of the newly synthesized polyimide and

poly(benzoxazole-co-imide) were analyzed by Fourier transform infrared

spectroscopy (Bio-Rad FTS 135) (FTIR) in the range of 500-4000 cm-1 in both

attenuated total reflectance (ATR) and transmission modes.

47
3.5.2 Wide angle X-ray diffraction (XRD)

Wide-angle X-ray diffraction (XRD) measurements were carried out by a

Bruker D8 Advance X-ray diffractometer using Cu Kα with a wavelength of

1.54 Å as the radiation source. The d-spacing (𝑑), representing the average

intersegmental distance of polymer chains, was determined by the Bragg's law:

2𝑑𝑠𝑖𝑛𝜃 = 𝑛𝜆 (3.1)

where, 𝜃 is the incident angle, 𝑛 is the any integer and 𝜆 is the wavelength of

the beam.

3.5.3 Themogravimetric analyses (TGA)

To examine the membranes’ thermal stability, thermogravimetric analyses

(TGA) were employed with a Shimadzu Thermal Analyzer (DTG-60A/TA-

60WS/FC-60A) with a heating rate of 5 °C/min from 50 to 800 °C under

nitrogen atmosphere (for polyarylether membranes, PI membranes, C-TR

membranes) or air atmosphere (for PI membranes and MMMs) with a flow

rate of 30 mL min-1. The MOFs loading in the mixed matrix membranes was

calculated by the following equation:

𝑃𝑎𝑡𝑖𝑐𝑙𝑒 𝑤𝑒𝑖𝑔ℎ𝑡
𝑃𝑎𝑟𝑡𝑖𝑐𝑙𝑒 𝑙𝑜𝑎𝑑𝑖𝑛𝑔 = ∗ 100% (3.2)
𝑃𝑎𝑟𝑡𝑖𝑐𝑙𝑒 𝑤𝑒𝑖𝑔ℎ𝑡+𝑝𝑜𝑙𝑦𝑚𝑒𝑟 𝑤𝑒𝑖𝑔ℎ𝑡

48
3.5.4 Themogravimetric analyses-Fourier transform infrared spectrometer

(TGA-FTIR)

In order to investigate the reactions during the thermal treatment, TGA-IR was

also carried out, which consisted of a Shimadzu Thermal Analyzer (DTG-

60A/TA-60WS/FC-60A) and a Shimadzu IRPrestige-21 under nitrogen

atmosphere (for PI membranes, C-TR membranes) or air atmosphere (for PI

membranes and MMMs) with a flow rate of 50 mL min-1. All samples were

pretreated from room temperature to 50 °C to remove moisture; after which,

the temperature was increased to 800 °C with a rate of 10 °C.min-1. The

released vapor was online analyzed by a scanning interval of 60s.

3.5.5 Scanning electron microscope (SEM)

The morphologies of the hollow fiber substrates and TFC membranes were

observed by scanning electron microscopy (SEM JEOL JSM-5600LV) and field

emission scanning electron microscopy (FESEM JEOL JSM-6700F). For cross-

section observation, all samples were made by cryogenically cracking in liquid

N2 to obtain proper cross-sections and coated with platinum using a JEOL JFC-

1200 ion sputtering device.

3.5.6 Energy-dispersive X-ray spectroscopy (EDX)

Elemental mapping of MMMs was carried out by X-ray energy dispersive

spectrometry (EDX) to examine the distribution of zirconium element in

49
MMMs using an Oxford INCA energy dispersion of X-ray system together

with a super ultrathin window (SUTW) connected to a scanning electron

microscopy (SEM JEOL JSM-5600LV) operating at 15 kV.

3.5.7 Density measurements

Membrane density was measured by a micro balance (ML204, Mettler Toledo)

and a density kit (ML-DNY-43, Zurich) provided by Mettler-Toledo (S.E.A.)

Pte Ltd. The membrane was first measured in air and then in hexane, and the

density was finally calculated by the following equation:


𝑚0
(𝑚0 − 𝑚0′ ) = 𝜌𝑒 (3.3)
𝜌0

where, 𝑚0 and 𝑚0′ represent the masses of the membrane in air and in hexane,

respectively. 𝜌0 and 𝜌𝑒 are the densities of the membrane and hexane,

respectively. At least three samples were tested for each kind of membranes to

obtain the average values.

3.5.8 Contact angle tests

A Rame´- Hart Contact Angle Goniometer (model 100-22) was used to

measure the contact angle at room temperature. Deionized water droplets were

placed onto the membrane surface by a Gilmont microsyringe and the contact

angles were automatically measured. For each membrane, at least 10

measurements were conducted and the average was reported.

50
3.5.9 Solvent uptake analyses

Before conducting the experiments, the membrane strips were first dried in a

vacuum oven overnight to remove any moisture. The pre-weighed dry strips

were subsequently immersed in pure solvents and water separately at 60°C,

the same temperature of pervaporation testing. At different time intervals, the

swollen strips were removed from the solvent, blotted with tissue papers to

remove the excess solvent on their surfaces and their weights were determined

by a digital microbalance in a closed condition. The solvent uptake data at

different time intervals were recorded until there was no further notable

change in their weights which indicates that the swollen membranes have

reached sorption equilibrium. The solvent uptake ratio for each sample was

determined by the weight difference between the swollen strip (𝑀𝑤𝑒𝑡 ) at the

equilibrium condition and the initial strip (𝑀𝑑𝑟𝑦 ) as shown below

𝑀𝑤𝑒𝑡 −𝑀𝑑𝑟𝑦
𝑆𝑜𝑙𝑣𝑒𝑛𝑡 𝑢𝑝𝑡𝑎𝑘𝑒 𝑟𝑎𝑡𝑖𝑜 = (3.4)
𝑀𝑑𝑟𝑦

3.5.10 Positron annihilation lifetime spectroscopy (PALS)

The microstructures of membranes were measured by bulk positron

annihilation lifetime spectroscopy (PALS). A fast-fast coincident PALS

spectrometer in our laboratory with a system channel width of 50.53

picoseconds per channel was used to measure the lifetime and intensity of
22
positron species. Na was selected as the positron source and the counting

rate was around 165 counts s-1. A detailed description of the experimental

procedures can be found elsewhere [1, 2]. To perform the wet sample tests for

51
four polyarylether membranes, the membranes were firstly saturated in an

85/15 ethanol/water feed at 60 °C for 12 hours, the same duration as

pervaporation tests, and then hermetically sealed in a plastic bag with the 22Na

source sandwiched in between these samples. Details can be found elsewhere

[3]. The PATFIT program was employed to analyze the raw data and acquire

the three lifetime components (τ1, τ2 and τ3). The mean free-volume radius R

(Å) was correlated with τ3 and calculated by the following semi-empirical

correlation equation [4-7]:

𝑅 1 2𝜋𝑅
𝜏3−1 = 2[1 − ∆𝑅 + 2𝜋 𝑠𝑖𝑛 ( ∆𝑅 )] (3.5)

where τ3 is the o-Ps lifetime. ΔR represents an empirical constant (1.66 Å). The

relative fractional free volume (FFV) can be obtained by the following

equation [6-8]:
4
𝐹𝐹𝑉 = 0.0018𝐼3 (3 𝜋𝑅33 ) (3.6)

where I3 is the o-Ps intensity (%).

3.6 Pervaporation studies

The pervaporation system was a lab-scale static cell as depicted elsewhere [9].

Aqueous alcohol solutions were used as the feed solutions. The downstream

pressure was maintained at less than 1 mbar by a vacuum pump. The test

temperatures were controlled at 40°C, 60°C and 80°C. The system was

conditioned for 2h before any sample collection. Thereafter, the permeate

mixtures were collected in a cold trap immersed in liquid nitrogen. The

permeate samples were weighted, and their compositions were analyzed by a

52
gas chromatography (HP-INNOWAX column, Hewlett-Packard GC 7890)

with a TCD detector. Throughout the experiment, the feed concentration

could be considered as constant because the quantity of the permeate sample

was much smaller than that of the feed solution and the variation (less than 1

wt%) could be neglected. In order to make sure the good reproducibility, all

the data were repeated at least three times.

The total flux J was determined by the following equation:

𝑄
𝐽= (3.7)
𝐴𝑡

where A, Q and t are the effective membrane area, total mass of the permeate

and operating time, respectively. Separation factor was defined as below:

𝑦 /𝑦
1/2 = 𝑥𝑤,1/𝑥𝑤,2 (3.8)
𝑤,1 𝑤,2

where subscripts 1 and 2 represent water and alcohol, respectively; xw and yw

indicate the components’ weight fractions in the feed and the permeate.

To study the intrinsic properties of the MMMs, the permeability and mole-

based selectivity were calculated by the following equation:

𝑙
𝑃𝑖 = 𝐽𝑖 × (3.9)
𝑥𝑖 𝛾𝑖 𝑃𝑖𝑠𝑎𝑡 −𝑦𝑖 𝑃𝑝

where 𝑃𝑖 is the permeability of component i, l is the membrane thickness,

𝐽𝑖 ,𝑃𝑖𝑠𝑎𝑡 and 𝛾𝑖 are the flux, the saturated vapor pressure of component 𝑖 and the

activity coefficient, respectively. 𝑥𝑖 and 𝑦𝑖 are the mole fractions of

component 𝑖 in the feed and the permeate, respectively. 𝑃𝑝 is the total pressure

at the permeate side, which could be viewed as zero because the pressure in

the permeate side is much smaller than the saturated vapor pressure in the feed

53
side due to the vacuum condition. By using the AspenTech Process Modelling

software, both 𝛾𝑖 and 𝑃𝑖𝑠𝑎𝑡 were calculated by the Wilson and the Antoine

equations, respectively. Mole-based selectivity was calculated by the

following equation:

𝑀 𝑃
𝛽𝑚𝑜𝑙𝑒 = ( 𝑀𝑗 ) (𝑃 𝑖 ) (3.10)
𝑖 𝑗

where 𝑀𝑖 and 𝑀𝑗 are the molecular weights of water and alcohol, respectively.

3.7 References

[1] N.L. Le, Y.P. Tang, T.S. Chung, The development of high-performance

6FDA-NDA/DABA/POSS/Ultem® dual-layer hollow fibers for ethanol

dehydration via pervaporation, J. Membr. Sci. 447 (2013) 163-176.

[2] Y.P. Tang, H. Wang, T.S. Chung, Towards high water permeability in

triazine-framework-based microporous membranes for dehydration of ethanol,

ChemSusChem, 8 (2015) 138-147.

[3] Y.P. Tang, D.R. Paul, T.S. Chung, Free-standing graphene oxide thin films

assembled by a pressurized ultrafiltration method for dehydration of ethanol, J.

Membr. Sci. 458 (2014) 199-208.

[4] S.J. Tao, Positronium annihilation in molecular substances, J. Chem. Phys.

56 (1972) 5499.

[5] S.J. Lue, D.T. Lee, J.Y. Chen, C.H. Chiu, C.C. Hu, Y.C. Jean, J.Y. Lai,

Diffusivity enhancement of water vapor in poly(vinyl alcohol)–fumed silica

nano-composite membranes: Correlation with polymer crystallinity and free-

volume properties, J. Membr. Sci. 325 (2008) 831-839.

54
[6] V.P. Shantarovich, I.B. Kevdina, Y.P. Yampolskii, A.Y. Alentiev, Positron

annihilation lifetime study of high and low free volume glassy polymers: 

dffects of free volume sizes on the permeability and permselectivity,

Macromolecules, 33 (2000) 7453-7466.

[7] H.M. Chen, W.S. Hung, C.H. Lo, S.H. Huang, M.L. Cheng, G. Liu, K.R.

Lee, J.Y. Lai, Y.M. Sun, C.C. Hu, R. Suzuki, T. Ohdaira , N. Oshima, Y.C.

Jean, Free-volume depth profile of polymeric membranes studied by positron

annihilation spectroscopy: layer structure from interfacial polymerization,

Macromolecules, 40 (2007) 7542-7557.

[8] M.L. Williams, R.F. Landel, J.D. Ferry, The temperature dependence of

relaxation mechanisms in amorphous polymers and other glass-forming

liquids, J. Am. Chem. Soc. 77 (1955) 3701-3707.

[9] Y. Wang, M. Gruender, T.S. Chung, Pervaporation dehydration of

ethylene glycol through polybenzimidazole (PBI)-based membranes. 1.

Membrane fabrication, J. Membr. Sci. 363 (2010) 149-159.

55
CHAPTER 4: POLYARYLERHER MEMBRANES FOR

DEHYDRATION OF ETHANOL AND

METHANOL VIA PERVAPORATION

4.1 Introduction

The idea of sustainability has stimulated global interests toward new

technologies that can conserve energy, minimize waste discharge and reduce

costs. In this regard, membrane technology is emerging as a promising

technology and has been widely used in various industries because it has the

advantages of ‘high selectivity, low energy consumption, moderate cost to

performance ratio and compact modular design’ [1].

Pervaporation, one of membrane-based processes, provides the distinct

advantages of high energy efficiency, environmental friendliness, low cost,

small footprint and ease of operation [2, 3]. In addition, this process is not

limited by the thermodynamic vapor-liquid equilibrium [3]. Therefore,

pervaporation is a promising candidate to compete with conventional

separation techniques such as distillation and absorption processes, especially

in the separation of azeotropic mixtures and thermally sensitive compounds [4,

5]. Among different applications of pervaporation, dehydration of alcohols is

now very popular in the industry [6-8]. In solvent dehydration applications,

hydrophilic polymers such as poly(vinyl alcohol) (PVA), poly(acrylic acid)

(PAA), chitosan, cellulose and alginate have been employed in early

generation pervaporation membranes due to their preferential affinity to water

56
and the enhanced solubility selectivity of water to organic solvents [9-11].

However, these membranes are prone to swell which results in a reduction of

separation efficiency and a loss of mechanical strength. Therefore, these

membranes need further modifications such as crosslinking, blending, grafting,

copolymerization as well as addition of adsorbent materials to improve their

stability and selectivity [6, 12-14]. Meanwhile, glassy polymers with rigid and

stiff chains such as polyimide (PI), polyamide, polybenzoxazole (PBO),

polybenzoxazinone (PBOZ) and polybenzimidazole (PBI) have emerged as

potential membrane materials for this application because their rigid and stiff

structures could not only improve the diffusivity selectivity but also enhance

the anti-swelling property and long-term stability [15-19]. However, the

applications of these polymers in solvent dehydration are frustrated by their

relative hydrophobicity and poor solubility selectivity.

As one of the most attractive membrane materials, polyarylether (PAE)

polymer has been widely used owing to its good mechanical properties,

excellent thermal and hydrolytic stability [20]. To alter their hydrophobic

properties and enhance their performance [20-22], substantial modifications

have been carried out: namely, (1) bulk modification of the PAE material, e.g.

sulfonation of PAE [23] and grafting hydrophilic groups to its chains [24, 25];

(2) surface modifications of PAE membranes via coating [26, 27] and

interfacial polymerization [28, 29]; and (3) blending directly with hydrophilic

polymers such as polyvinylpyrrolidone (PVP) [30], polyethyleneglycol (PEG)

[31]. These modifications not only improve membrane hydrophilicity but also

have positive effects on performance enhancement. For example, Li et al.

57
modified PESU membranes using a copolymer Pluronic F127 as a surface

modifier via surface segregation and applied the membranes to separate

thiophene from n-octane via pervaporation [32]. Both the permeation flux and

enrichment factor were effectively improved.

Different from previous work which just modified the membrane surface or

grafted hydrophilic segments onto the side-chains of the PAE polymers, we

recently synthesized a new hydrophilic PESU-co-Pluronic (H-PESU)

copolymer by directly incorporating a Pluronic copolymer into the main chain

of PESU. To our best knowledge, no exploration has been conducted on this

polymer for pervaporation. Therefore, the first objective of this work is to

investigate the potential of this new H-PESU for dehydration of alcohols via

pervaporation. The second objective of this work aims to study the influence

of main-chain structure on mass transport phenomena and pervaporation

performance. Therefore, another three different PAE polymers have also

been synthesized for comparison. They are polyethersulfone (PESU),

polyphenylsulfone (PPSU), polytrimethylphenylethersulfone (T-PESU). The

structures of these four PAE polymers are depicted in Figure 4.1. To meet our

objectives, flat dense membranes were prepared and fundamental

characteristics of these membranes were investigated. Dehydration of ethanol

and methanol aqueous solutions is our focus because they are smaller

molecules and difficult to be dehydrated compared to isopropanol and butanol

[33]. Besides, solvent uptake and fractional free volume were characterized to

reveal their solubility and diffusivity preferences and to explain their different

swelling phenomena as well as pervaporation performances. Finally, long-

58
term stability tests have also been carried out to give a comprehensive

evaluation about these membranes. This work may provide useful insights to

understand the correlations between polymer microstructure and mass

transport phenomena and to design membrane materials for alcohol

dehydration.

Figure 4.1 Chemical structures of PESU, T-PESU, PPSU and H-PESU

4.2 Experimental

The PAE membranes were prepared by knife casting as introduced in Chapter

3. Thermal stability of the membranes was studied by TGA under nitrogen

atmosphere. Density measurements were carried out as shown in Chapter 3.

D-spacing was obtained by XRD measurements. Hydrophilicity was

determined by contact angle tests. Solvent uptake tests were conducted at 60℃

to study the affinities of alcohol and water with the membranes. PALS tests

59
were carried out for membranes at both dry state and wet state to study the

influence of swelling on free-volume.

4.3 Results and discussion

4.3.1 Characterizations of membranes

A comparison of thermal behaviors of the four membranes is displayed in

Figure 4.2. PESU and PPSU have better thermal stability than T-PESU and H-

PESU. The former two show no obvious weight loss (less than 1 wt%) until

490°C when the polymer backbones start to decompose. In contrast, the T-

PESU membrane starts to degrade from 400°C, while the H-PESU membrane

undergoes an earlier weight loss at a lower temperature range from 200 to 400

°C probably due to the decomposition of the Pluronic moiety, which is in good

agreement with previous studies [34, 35]. After that, the decomposition of the

PESU moiety in H-PESU leads to a more weight loss at a higher temperature

range. Nevertheless, these four membranes have sufficient thermal stability for

pervaporation applications because most operating temperatures of

pervaporation are lower than 100 oC.

60
PESU
100
PPSU
Weight percentage (%) T-PESU
80 H-PESU

60

40

20

0
0 200 400 600 800
Temperature (ºC)

Figure 4.2 TGA curves of PESU, T-PESU, PPSU and H-PESU membranes
Intensity (a.u.)

PPSU

T-PESU

H-PESU

PESU

5 10 15 20 25 30 35 40 45 50
2 theta (º)

Figure 4.3 XRD patterns of PESU, T-PESU, PPSU and H-PESU membranes

61
Figure 4.3 displays the XRD patterns of the four membranes. It is found that

the peak of T-PESU appears at the smallest angle while the other three show

insignificant differences. Based on the Bragg's law, a polymer with a smaller

peak angle should have a larger d-spacing value. Therefore, among these four

kinds of membranes, T-PESU possesses the largest d-spacing value. This is

because the three -CH3 groups in T-PESU can prevent polymer chains from

close packing due to the steric hindrance effect. Table 4.1 shows that the

density of these four membranes follows the trend of T-PESU < PPSU < H-

PESU < PESU. The lowest density of T-PESU agrees well with its highest d-

spacing. This phenomenon was also reported in a previous study [36].

Table 4.1 Densities and water contact angles of PESU, T-PESU, PPSU and H-
PESU membranes
Membrane Density (g/cm3) Water Contact Angle (º)
PESU 1.354±0.021 88.8±1.4
PPSU 1.266±0.005 85.8±0.9
T-PESU 1.213±0.017 85.4±0.9
H-PESU 1.294±0.013 41.8±1.4

Table 4.1 also summarizes the water contact angles of these four membranes.

PESU, PPSU and T-PESU membranes have close and comparable water

contact angles, while the water contact angle of the H-PESU membrane is

much smaller than those of the other three. Therefore, it can be concluded that

among these four membranes, the H-PESU membrane has the largest

hydrophilicity which can be attributed to the existence of hydrophilic Pluronic

moieties in its polymer chains.

62
4.3.2 Solvent uptake analyses

In order to investigate the affinity of water and ethanol towards the

membranes, the solvent uptake tests were carried out and characterized by the

membrane weight changes due to the sorption of different solvents. Table 4.2

compares the solvent uptake ratios of the membranes in water and ethanol and

methanol. The water uptake ratios follow the trend of PESU < PPSU < T-

PESU<H-PESU while the ethanol and methanol uptake ratios present an

inverse trend. This phenomenon can be explained by the contact angle results.

Since the water contact angles follow the trend of PESU < PPSU < T-PESU <

H-PESU as aforementioned, so the hydrophilicities follow the same trend and

the membranes can absorb more water and less ethanol and methanol

according to this trend. Therefore, the H-PESU membrane provides the

highest water solubility and lowest ethanol and methanol solubilities, followed

by T-PESU, PPSU and PESU. Correspondingly, the solubility selectivity

towards water of these four membranes obeys the trend of PESU < PPSU < T-

PESU < H-PESU.

Table 4.2 Solvent uptake results for PESU, T-PESU, PPSU and H-PESU
membranes
Water Ethanol Methanol Uptake ratio Uptake ratio
Membrane uptake uptake uptake ratio of water to of water to
ratio (%) ratio (%) (%) ethanol methanol
PESU 3.1±0.7 9.9±1.6 5.2±1.1 0.31 0.60
PPSU 3.9±0.8 8.1±1.5 4.8±0.7 0.48 0.81
T-PESU 4.1±0.7 7.5±1.5 4.6±1.1 0.55 0.89
H-PESU 7.9±0.8 5.7±1.5 3.3±0.7 1.39 2.39

63
4.3.3 PALS analyses

To further quantitatively analyze the free volume size of the four membranes,

PALS measurements were carried out. The cavity size of free volume and its

concentration are correlated with two important parameters; namely, o-Ps

lifetime τ3 and its intensity I3, respectively [37, 38]. Table 4.3 shows PALS

raw data, free volume sizes and FFVs of the four membranes at the dry state.

The FFVs follow the trend of PESU<H-PESU<T-PESU< PPSU. To further

investigate the effects of penetrants on membrane swelling and polymer

microstructure, PALS measurements were carried out for the membranes at

the wet state by immersing membrane samples in an 85/15 wt% ethanol/water

solution for 12 hr before tests. Table 4.4 tabulates the results. Comparing the

PALS data at dry and wet states, increases in τ3 and R are observed for all four

membranes, suggesting that swelling results in a larger free volume. This

clearly accounts for the significant flux increment after swelling. On the

contrary, I3 decreases for all the membranes, indicating that adjacent small

cavities at the dry state may coalesce to form large cavities at the wet state or

some cavities are filled with the penetrants at the wet state [39].

Table 4.3 PALS data of dry PESU, T-PESU, PPSU and H-PESU membranes
Membrane τ3 (ns) I3 (%) R (Å) FFV (%)
PESU 1.779±0.009 19.52±0.15 2.644±0.008 2.720±0.046
PPSU 1.922±0.008 23.83±0.14 2.782±0.006 3.871±0.049
T-PESU 1.928±0.009 20.19±0.13 2.788±0.007 3.299±0.048
H-PESU 1.839±0.009 18.91±0.14 2.703±0.008 2.816±0.045

64
Table 4.4 PALS data of wet PESU, T-PESU, PPSU and H-PESU membranes
Membrane τ3 (ns) I3 (%) R (Å) FFV (%)
PESU 1.867±0.010 18.88±0.14 2.731±0.008 2.898±0.047
PPSU 2.044±0.009 22.16±0.12 2.896±0.006 4.057±0.050
T-PESU 1.967±0.010 20.12±0.13 2.824±0.007 3.418±0.049
H-PESU 2.117±0.010 18.07±0.11 2.961±0.007 3.535±0.048

A comparison of FFVs at dry and wet states is shown in Figure 4.4. For PESU,

T-PESU and PPSU, the percentage of FFV increment follows the order of

PESU>PPSU>T-PESU which is consistent with the trend of ethanol uptake

ratio. This is because the FFV increment is caused by swelling and for these

three membranes, swelling is mainly induced by ethanol since their ethanol

uptake prevails over the water uptake as shown in Table 4.2. Therefore, the

FFV incremental trend is in good agreement with the ethanol uptake ratio

trend for these three membranes. On the contrary, swelling of the H-PESU

membrane is mainly caused by water due to its preferential affinity to water as

implied in Table 4.2. Figure 4.4 shows that the FFV increment for the H-

PESU membrane is much more severe than the other three. As a result, FFV of

the H-PESU membrane becomes the second largest after swelling. In addition,

by comparing the structures of H-PESU and PESU, it can be concluded that

the big difference between their percentages of FFV increment is caused by

the incorporation of the Pluronic moiety. This suggests that the swelling of H-

PESU membrane mainly occurs in the hydrophilic Pluronic regions.

65
5 30
4.5
3.8714.057 25

FFV increment percentage (%)


4 3.535
3.5 3.2993.418
20
3 2.72 2.898 2.816
FFV(%)

2.5 15
2
1.5 10
1 5
0.5
0 0
PESU PPSU T-PESU H-PESU

Dry state Wet state FFV increment percentage %

Figure 4.4 Comparison of FFVs at dry and wet states measured by PALS

4.3.4 Ethanol and methanol performance

Table 4.5 summarizes the ethanol dehydration performance of the four

membranes at 60 oC. Normalized flux is used in this study, which is the

product of flux and membrane thickness. This is to eliminate the impact of the

slight inconsistency in membrane thickness.

Table 4.5 Ethanol dehydration performance of PESU, T-PESU, PPSU and H-


PESU membranes
Total Water Ethanol Water
normalized normalized normalized Conc. in Separation
Membrane flux* flux* flux* permeate factor
2 2
(gμm/m h) (gμm/m h) (gμm/m h) (wt%)
2

PESU 661.8±82.6 657.2±82.0 4.6±0.6 99.3±0.4 784.6


T-PESU 815.8±5.0 810.9±4.9 4.9±0.1 99.4±0.2 936.4
PPSU 1333.1±25.4 1314.4±25.0 18.7±0.4 98.6±0.5 389.6
H-PESU 4352.1±124.7 4299.9±123.2 52.2±1.5 98.8±0.2 483.9
o
* Feed composition: 85% ethanol+15% water; Test temperature: 60 C

66
For these four PAE membranes, the water normalized flux is much higher than

the ethanol normalized flux and accounts for most of the total normalized flux,

suggesting that these PAE membranes are highly water selective. In

pervaporation, the commonly cited separation mechanism is the solution-

diffusion model. Based on this model, permeability depends on solubility and

diffusivity. Generally, higher hydrophilicity and water uptake may signify a

higher water solubility while a higher FFV is conducive to the fast diffusion of

the permeate components across the membranes. Thus, membranes with

higher hydrophilicity and higher FFV are favorable to attain a high permeation

flux. In this regard, the PESU membrane shows the smallest water normalized

flux due to its smallest water solubility and diffusivity since it has the lowest

water uptake ratio and the smallest FFV as aforementioned in Table 4.2 and

Table 4.4, respectively. The increases of water normalized flux for T-PESU

and PPSU are because of the increases in both water solubility and diffusivity.

It's worth noting that the H-PESU membrane shows the largest water

normalized flux, which is 2-6 times higher than those of the other three

membranes. This can be explained by the following reasons: (1) the H-PESU

membrane has the largest water solubility as shown in Table 4.2 due to the

hydrophilic nature of the Pluronic moiety and (2) its FFV value is moderate as

shown in Table 4.4, and its hydrophilic Pluronic moiety may form special

water transport channels for fast water diffusion, which lead to a high water

diffusivity. At the same time, the H-PESU membrane also has the highest

ethanol normalized flux among these four PAE membranes which may be due

to the swelling effect.

67
In addition, the PPSU membrane shows the lowest separation factor which is

mainly ascribed to its smallest diffusivity selectivity since it has the largest

FFV. It is interesting to observe that the H-PESU membrane exhibits a

comparable separation factor although its degree of swelling is much higher

than the others. This is due to the constraint of the hydrophobic and rigid

PESU moiety. In other words, the swelling mainly occurs in the Pluronic

regions, which have the highest solubility selectivity and most likely promote

the transport of water molecules rather than ethanol molecules. These Pluronic

channels are preferentially occupied by water molecules and this possibly adds

the resistance for ethanol transport. Therefore, the water molecules have

priority to diffuse through these channels. On the other hand, the rigid PESU

moiety prevents the membrane from uncontrolled swelling and ensures good

separation efficiency. Clearly, the combination of PESU and Pluronic moieties

results in a synergistic microstructure that endows the membrane with a

superior flux and a remarkable separation factor.

Table 4.6 shows a benchmarking comparison with literature data [40-45] . The

four membranes developed in this study not only exhibit comparable water

permeability but also higher mole-based selectivity than most of the rest in the

table. Especially, the H-PESU membrane displays the most promising

performance and may have great potential for industrialization.

68
Table 4.6 A comparison of pervaporation performance for ethanol
dehydration
Water
Membrane Water Mole-
Temp. permeability
Membrane thickness concentration o based Ref.
( C) (mg m-1 h-1
(μm) in feed selectivity
kPa-1)
Crosslinked
20 10% 60 1.136 417 [40]
Chitosan
Crosslinked
29 10% 60 0.823 456 [41]
PVA
P84®
--- 15% 60 0.034 1874 [42]
copolyimide
Pervap® 2201 --- 10% 60 0.022 309 [43]
TR PBO 20-25 10% 25 1.254 186 [44]
Matrimid
45-55 15% 60 0.136 481 [45]
/hPIM-1
This
PESU 21 15% 60 0.049 1111
study
This
T-PESU 20 15% 60 0.061 1326
study
This
PPSU 23 15% 60 0.098 552
study
This
H-PESU 36 15% 60 0.322 685
study

To date, methanol dehydration by polymer membranes via pervaporation has

not been widely studied, which may be attributed to the lack of suitable and

affordable membrane materials with excellent anti-methanol swelling

properties [33]. Inorganic membranes are dominant in methanol dehydration

since they don’t have the solvent-induced swelling problem. However, their

extensive applications are restricted by the fragility and high costs of

fabrication. In contrast, polymeric membranes are easy to fabricate and scale-

up with relatively low costs. However, polymeric membranes for methanol

dehydration usually suffer from poor separation efficiency because of swelling

69
[4, 6, 17]. Therefore, polymeric membranes are still not popular for methanol

dehydration.

In view of the good performance for ethanol dehydration, the four membranes

developed in this study were employed for methanol dehydration. Table 4.7

summarizes the results. PESU, T-PESU and PPSU membranes have

comparable water and methanol normalized fluxes. While similar to ethanol

dehydration, the H-PESU membrane shows both the highest water and

methanol normalized fluxes. On the other hand, all these four PAE membranes

show separation factor higher than 28 and mole-based selectivity more than 82

which is far better than most of the other polymeric membranes listed in

Table 4.8 [33, 46, 47]. Especially, the T-PESU membrane has the highest

separation factor of 39.8 which corresponds to 86.8% water in the permeate.

In addition, compared to other polymeric membranes, the membranes

developed in this work have the more balanced separation performance in

terms of superior separation factors and moderate normalized fluxes.

Table 4.7 Methanol dehydration performance of PESU, T-PESU, PPSU and


H-PESU membranes
Total Water Methanol Water
normalized normalized normalized Conc. in Separation
Membrane flux* flux* flux* permeate factor
2 2
(gμm/m h) (gμm/m h) (gμm/m h) (wt%)
2

PESU 1066.2±32.3 886.0±26.8 180.2±5.5 83.1±1.8 29.7


T-PESU 1072.7±64.9 931.1±56.3 141.6±8.6 86.8±1.5 39.8
PPSU 1133.5±21.6 936.3±17.8 197.2±3.8 82.6±1.3 28.7
H-PESU 2134.7±254.1 1788.9±212.9 345.8±41.2 83.8±1.8 31.3
* Feed composition: 85% methanol+15% water; Test temperature: 60 oC

70
Table 4.8 A comparison of pervaporation performance for methanol
dehydration
Water Mole-
Membrane Water
Temp. permeability based
Membrane thickness concentration Ref.
(oC) (mg m-1 h-1 selectivity
(μm) in feed
kPa-1)
5%-sPPSU 16 15% 60 0.033 33.8 [33]
2.5%-sPPSU 16 15% 60 0.030 32.1 [33]
Agarose 14-20 14% 30 0.925-1.321 8.3 [46]
Polyamide-
30 10% 20 0.268-0.415 9.2-99.9 [47]
sulfonamide
This
PESU 21 15% 60 0.085 85.1
study
This
T-PESU 20 15% 60 0.089 114.0
study
This
PPSU 23 15% 60 0.089 82.0
study
This
H-PESU 36 15% 60 0.171 89.7
study

The H-PESU membrane is chosen for further studies of the effect of feed

concentration on pervaporation performace since its has the best separation

performance with the highest normalized flux and comparable separation

factor. Aqueous ethanol and methanol systems with alcohol concentrations of

85wt%, 90wt% and 95wt% are investigated. Figure 4.5 shows that the water

normalized flux decreases while the ethanol and methanol normalized fluxes

increase with an increase in alcohol concentration in ethanol/water and

methanol/water systems. These phenomena are mainly because a higher

alcohol concentration results in a lower driving force for water transport but a

higher driving force for alcohol transport [44]. As a result, both total

normalized flux and separation factor decrease with an increase in alcohol

concentration as shown in Figure 4.6 and Figure 4.7. Similar trends have been

reported elsewhere [44, 48].

71
a 6000

Water normalized flux


Ethanol/water system
5000 Methanol/water system

(gμm/m2h)
4000
3000
2000
1000
0
80 85 90 95 100
Alcohol Concentration (%)

1000
b Ethanol/water system
900
Alcohol normalized flux

800 Methanol/water system


700
(gμm/m2h)

600
500
400
300
200
100
0
80 85 90 95 100
Alcohol Concentration (%)
Figure 4.5 (a) Water and (b) ethanol/methanol normalized flux of H-PESU
membranes at different feed concentrations at 60 ºC

6000
Total normalized flux

500
Separation factor

5000
400
(gμm/m2h)

4000
3000 300

2000 200

1000 100
0 0
80 85 90 95 100
Ethanol Concentration (%)

Figure 4.6 Ethanol dehydration performance of H-PESU membranes at


different feed concentrations at 60 ºC

72
6000 40

Total normalized flux


5000 35

Separation factor
30

(gμm/m2h)
4000 25
3000 20
2000 15
10
1000 5
0 0
80 85 90 95 100
Methanol Concentration (%)

Figure 4.7 Methanol dehydration performance of H-PESU membranes at


different feed concentrations at 60 ºC

4.3.5 Long-term stability tests

Long-term operating stability is another important factor to evaluate the

potential of a membrane material. Therefore, the long-term stability tests of

the H-PESU membrane were conducted for ethanol dehydration at 60 oC with

the feed composition of 85wt%/15wt% ethanol/water. The PESU membrane

was selected for comparison. Their fluxes and separation factors were

continuously monitored for 250h. Figure 4.8 shows their results. It can be

found that for H-PESU membrane, the water concentration in the permeate

decreases slightly from 98.7 wt% to 96.5 wt% at the early stage, which is

caused by swelling. Subsequently, it becomes stable at around 96.3%, which is

still acceptable for ethanol dehydration. Overall, the long-term performance of

the H-PESU membrane is quite reliable. This is again due to the combination

of the strengths from both rigid hydrophobic PESU moiety and flexible

hydrophilic Pluronic moiety in the H-PESU polymer. In contrast, the PESU

membrane shows quite stable performance with little fluctuations in both

73
normalized flux and separation factor during the entire testing duration of

250h. This is because the PESU membrane has higher anti-swelling properties

than the H-PESU membrane because the latter contains the hydrophilic

Pluronic moiety.

100.0
14000
Normalized flux (gμm/m2h)

Water concentration in the


98.0
12000 96.0

permeate (wt%)
10000 94.0
92.0
8000 90.0
6000 88.0
4000 86.0
84.0
2000 82.0
0 80.0
100 0 50
150 200 250
Test Duration (h)
PESU H-PESU
Figure 4.8 Long-term stability of the PESU and H-PESU membranes for
ethanol dehydration at 60 ºC with the feed composition of 85wt%/15wt%
ethanol/water

4.4 Conclusions

In this study, pervaporation membranes made from four polyarylether (PAE)

materials: PESU, PPSU, T-PESU and H-PESU have been explored for

pervaporation dehydration of ethanol and methanol. These PAE membranes

showed impressive separation performance for both dehydration cases. Some

interesting correlations between mass transport phenomena and structural

characteristics were also unveiled.

74
PALS measurements reveal that swelling results in a larger free volume radius

and a bigger FFV. In addition, the adjacent small cavities in the membranes at

the dry state may coalesce to form large cavities at the wet state. Among these

four PAE membranes, the H-PESU membrane shows the best ethanol

dehydration performance with the highest water permeability of 0.322 mg m-1

h-1 kPa-1 and a comparable mole-based selectivity of 685 due to its unique

structure consisting of both hydrophilic and hydrophobic moieties. The

hydrophilic Pluronic moiety may form special channels which are

preferentially occupied by water molecules and promote the fast transport of

water molecules. Therefore, the H-PESU membrane has the highest water

solubility and solubility selectivity. On the other hand, its hydrophobic and

rigid PESU moiety can prevent the membrane from uncontrolled swelling and

contribute to good separation efficiency. The long-term operating stabilities of

PESU and HPESU have been demonstrated. In addition, the separation

performance of these PAE membranes are better than most of reported

polymeric membranes for both ethanol and methanol dehydration, implying

their great potential for pervaporation applications.

4.5 References

[1] F. Lipnizki, R.W. Field, P.K. Ten, Pervaporation-based hybrid process: a

review of process design, applications and economics, J. Membr. Sci. 153

(1999) 183-210.

[2] L.M. Vane, A review of pervaporation for product recovery from biomass

fermentation processes, J. Chem. Technol. Biot. 80 (2005) 603-629.

75
[3] Y.M. Xu, T.S. Chung, High-performance UiO-66/polyimide mixed matrix

membranes for ethanol, isopropanol and n-butanol dehydration via

pervaporation, J. Membr. Sci. 531 (2017) 16-26.

[4] Y.K. Ong, G.M. Shi, N.L. Le, Y.P. Tang, J. Zuo, S.P. Nunes, T.S. Chung,

Recent membrane development for pervaporation processes, Prog. Polym. Sci.

57 (2016) 1-31.

[5] D. Figueroa, D. S. Laoretani, J. Zelin, R. Vargas, A. R. Vecchietti, J.

Espinosa, Screening of pervaporation membranes for the separation of

methanol-methyl acetate mixtures: an approach based on the conceptual

design of the pervaporation-distillation hybrid process, Sep. Purif. Technol.

189 (2017) 296-309.

[6] X.S. Feng, R.Y.M. Huang, Liquid separation by membrane pervaporation: 

a review, Ind. Eng. Chem. Res. 36 (1997) 1048-1066.

[7] Z.W. Song, J.M. Zhu, L.Y. Jiang, Novel

polysiloxaneimide/polyetherimide/non-woven fabric composite membranes

for organophilic pervaporation, J. Membr. Sci. 472 (2014) 77-90.

[8] S. Xu, L. Liu, Y. Wang, Network cross-linking of polyimide membranes

for pervaporation dehydration, Sep. Purif. Technol. 185 (2017) 215-226.

[9] Y. Huang, J. Ly, D. Nguyen, R.W. Baker, Ethanol dehydration using

hydrophobic and hydrophilic polymer membranes, Ind. Eng. Chem. Res. 49

(2010) 12067-12073.

[10] M.E. Dmitrenko, A.V. Penkova, A.B. Missyul, A.I. Kuzminova, D.A.

Markelov, S.S. Ermakov, D. Roizard, Development and investigation of

mixed-matrix PVA-fullerenol membranes for acetic acid dehydration by

pervaporation, Sep. Purif. Technol. 187 (2017) 285-293.

76
[11] A.J. Toth, A. Andre, E. Haaz, P. Mizsey, New horizon for the membrane

separation: Combination of organophilic and hydrophilic pervaporations, Sep.

Purif. Technol. 156 (2015) 432-443.

[12] W.S. Hung, S.M. Chang, R.L.G. Lecaros, Y.L. Ji, Q.F. An, C.C. Hu, K.R.

Lee, J.Y. Lai, Fabrication of hydrothermally reduced graphene oxide/chitosan

composite membranes with a lamellar structure on methanol dehydration,

Carbon, 117 (2017) 112–119.

[13] H. Zhang, Y. Wang, Poly (vinyl alcohol)/ZIF-8-NH2 mixed matrix

membranes for ethanol dehydration via pervaporation, AICHE J. 62 (2016)

1728-1739.

[14] L.L. Xia, C.L. Li, Y. Wang, In-situ crosslinked PVA/organosilica hybrid

membranes for pervaporation separations, J. Membr. Sci. 498 (2016) 263-275.

[15] Y.M. Xu, N.L. Le, J. Zuo, T.S. Chung, Aromatic polyimide and

crosslinked thermally rearranged poly (benzoxazole-co-imide) membranes for

isopropanol dehydration via pervaporation, J. Membr. Sci. 499 (2016) 317-

325.

[16] N.L. Le, Y. Wang, T.S. Chung, Synthesis, cross-linking modifications of

6FDA-NDA/DABA polyimide membranes for ethanol dehydration via

pervaporation, J. Membr. Sci. 415-416 (2012) 109-121.

[17] L.Y. Jiang, Y. Wang, T.S. Chung, X.Y. Qiao, J.Y. Lai, Polyimides

membranes for pervaporation and biofuels separation, Prog. Polym. Sci. 34

(2009) 1135-1160.

[18] M.M. Teoh, T.S. Chung, K.Y. Wang, M.D. Guiver, Exploring Torlon/P84

co-polyamide-imide blended hollow fibers and their chemical cross-linking

77
modifications for pervaporation dehydration of isopropanol, Sep. Purif.

Technol. 61 (2008) 404-413.

[19] A. Pulyalina, G. Polotskaya, M. Goikhman, I. Podeshvo, L. Kalyuzhnaya,

M. Chislov, A. Toikka, Study on polybenzoxazinone membrane in

pervaporation processes, J. Appl. Polym. Sci. 130 (2013) 4024-4031.

[20] C. Zhao, J. Xue, F. Ran, S. Sun, Modification of polyethersulfone

membranes – A review of methods, Prog. Mater. Sci. 58 (2013) 76-150.

[21] B. Van der Bruggen, Chemical modification of polyethersulfone

nanofiltration membranes: a review, J. Appl. Polym. Sci. 114 (2009) 630-642.

[22] K. Khulbe, C. Feng, T. Matsuura, The art of surface modification of

synthetic polymeric membranes, J. Appl. Polym. Sci. 115 (2010) 855-895.

[23] D. Lu, H. Zou, R. Guan, H. Dai, L. Lu, Sulfonation of polyethersulfone

by chlorosulfonic acid, Polym. Bull. 54 (2005) 21-28.

[24] B. Deng, X. Yang, L. Xie, J. Li, Z. Hou, S. Yao, G. Liang, K. Sheng, Q.

Huang, Microfiltration membranes with pH dependent property prepared from

poly (methacrylic acid) grafted polyethersulfone powder, J. Membr. Sci. 330

(2009) 363-368.

[25] L. Li, Y. Wang, Sulfonated polyethersulfone Cardo membranes for direct

methanol fuel cell, J. Membr. Sci. 246 (2005) 167-172.

[26] A. Razmjou, J. Mansouri, V. Chen, M. Lim, R. Amal, Titania

nanocomposite polyethersulfone ultrafiltration membranes fabricated using a

low temperature hydrothermal coating process, J. Membr. Sci. 380 (2011) 98-

113.

78
[27] R.X. Liu, X.Y. Qiao, T.S. Chung, Dual-layer P84/polyethersulfone

hollow fibers for pervaporation dehydration of isopropanol, J. Membr. Sci.

294 (2007) 103-114.

[28] Y. Mansourpanah, S. Madaeni, A. Rahimpour, Preparation and

investigation of separation properties of polyethersulfone supported poly

(piperazineamide) nanofiltration membrane using microwave-assisted

polymerization, Sep. Purif. Technol. 69 (2009) 234-242.

[29] M.A. Seman, M. Khayet, N. Hilal, Development of antifouling properties

and performance of nanofiltration membranes modified by interfacial

polymerisation, Desalination, 273 (2011) 36-47.

[30] D.B. Mosqueda ‐ Jimenez, R.M. Narbaitz, T. Matsuura, Effects of

preparation conditions on the surface modification and performance of

polyethersulfone ultrafiltration membranes, J. Appl. Polym. Sci. 99 (2006)

2978-2988.

[31] Y.Q. Wang, T. Wang, Y.L. Su, F.B. Peng, H. Wu, Z.Y. Jiang, Protein-

adsorption-resistance and permeation property of polyethersulfone and

soybean phosphatidylcholine blend ultrafiltration membranes, J. Membr. Sci.

270 (2006) 108-114.

[32] B. Li, W. Zhao, Y. Su, Z. Jiang, X. Dong, W. Liu, Enhanced

desulfurization performance and swelling resistance of asymmetric

hydrophilic pervaporation membrane prepared through surface segregation

technique, J. Membr. Sci. 326 (2009) 556-563.

[33] Y. Tang, N. Widjojo, G.M. Shi, T.S. Chung, M. Weber, C. Maletzko,

Development of flat-sheet membranes for C1–C4 alcohols dehydration via

79
pervaporation from sulfonated polyphenylsulfone (sPPSU), J. Membr. Sci.

415-416 (2012) 686-695.

[34] N.R. Hendricks, J.J. Watkins, K.R. Carter, Formation of hierarchical

silica nanochannels through nanoimprint lithography, J. Mater. Chem. 21

(2011) 14213-14218.

[35] S. Tanaka, A. Doi, N. Nakatani, Y. Katayama, Y. Miyake, Synthesis of

ordered mesoporous carbon films, powders, and fibers by direct triblock-

copolymer-templating method using an ethanol/water system, Carbon, 47

(2009) 2688-2698.

[36] M. Calle, C.M. Doherty, A.J. Hill, Y.M. Lee, Cross-linked thermally

rearranged poly(benzoxazole-co-imide) membranes for gas separation,

Macromolecules, 46 (2013) 8179-8189.

[37] H.M. Chen, W.S. Hung, C.H. Lo, S.H. Huang, M.L. Cheng, G. Liu, K.R.

Lee, J.Y. Lai, Y.M. Sun, C.C. Hu, R. Suzuki, T. Ohdaira , N. Oshima, Y.C.

Jean, Free-volume depth profile of polymeric membranes studied by positron

annihilation spectroscopy: layer structure from interfacial polymerization,

Macromolecules, 40 (2007) 7542-7557.

[38] J. Zuo, Y. Wang, T.-S. Chung, Novel organic–inorganic thin film

composite membranes with separation performance surpassing ceramic

membranes for isopropanol dehydration, J. Membr. Sci. 433 (2013) 60-71.

[39] S. Zhang, R. Zhang, Y.C. Jean, D.R. Paul, T. S. Chung, Cellulose esters

for forward osmosis: Characterization of water and salt transport properties

and free volume, Polymer, 53 (2012) 2664-2672.

80
[40] W. Zhang, G. Li, Y. Fang, X. Wang, Maleic anhydride surface-

modification of crosslinked chitosan membrane and its pervaporation

performance, J. Membr. Sci. 295 (2007) 130-138.

[41] M.L. Gimenes, L. Liu, X. Feng, Sericin/poly (vinyl alcohol) blend

membranes for pervaporation separation of ethanol/water mixtures, J. Membr.

Sci. 295 (2007) 71-79.

[42] X. Qiao, T.S. Chung, Fundamental characteristics of sorption, swelling,

and permeation of P84 co-polyimide membranes for pervaporation

dehydration of alcohols, Ind.Eng.Chem.Res. 44 (2005) 8938-8943.

[43] D. Van Baelen, B. Van der Bruggen, K. Van den Dungen, J. Degrève, C.

Vandecasteele, Pervaporation of water–alcohol mixtures and acetic acid–water

mixtures, Chem.Eng.Sci. 60 (2005) 1583-1590.

[44] Y.K. Ong, H. Wang, T.S. Chung, A prospective study on the application

of thermally rearranged acetate-containing polyimide membranes in

dehydration of biofuels via pervaporation, Chem. Eng. Sci. 79 (2012) 41-53.

[45] W.F. Yong, P. Salehian, L. Zhang, T.-S. Chung, Effects of hydrolyzed

PIM-1 in polyimide-based membranes on C2–C4 alcohols dehydration via

pervaporation, J. Membr. Sci. 523 (2017) 430-438.

[46] M. Yoshikawa, K. Masaki, M. Ishikawa, Pervaporation separation of

aqueous organic mixtures through agarose membranes, J. Membr. Sci. 205

(2002) 293-300.

[47] W.H. Chan, C.F. Ng, S.Y. Lam-Leung, X. He, Water–alcohol separation

by pervaporation through chemically modified poly (amidesulfonamide)s, J.

Membr. Sci. 160 (1999) 77-86.

81
[48] Y. Wang, T.S. Chung, B.W. Neo, M. Gruender, Processing and

engineering of pervaporation dehydration of ethylene glycol via dual-layer

polybenzimidazole (PBI)/polyetherimide (PEI) membranes, J. Membr. Sci.

378 (2011) 339-350.

This chapter has been accepted as a journal paper:

Y.M. Xu, Y.P. Tang, T.S. Chung, M. Weber, C. Maletzkoc, Polyarylether

membranes for dehydration of ethanol and methanol via pervaporation, Sep.

Purif. Technol. 193 (2018) 165-174.

82
CHAPTER 5: POLYIMDE AND CROSSLINKED

THERMALLY REARRANGED

POLY(BENZOXAZOLE-CO-IMIDE)

MEMBRANES FOR ISOPROPANOL

DEHYFRATION VIA PERVAPORATION

5.1 Introduction

Isopropanol (IPA) is an important alcohol which is used for a variety of

industrial and consumer applications. For example, it can be used as an

effective cleaner for electronic and optical devices because of its non-toxicity,

fast evaporation and ability to dissolve a wide range of non-polar compounds.

Moreover, it is also a chemical intermediate for the production of isopropyl

acetate, rubbing alcohol, vitamin B12 and a pharmaceutical additive used in

hand sanitizer and disinfecting pads [1, 2]. In addition, it can be considered as

an alternative fuel to alleviate global warming and slow down the depletion of

fossil fuels [3, 4]. Isopropanol is produced either by the chemical reaction

between propene and water or by microorganism’s fermentation [4]. However,

both processes require the separation of water from IPA to produce high purity

IPA. Hence, the successful production and applications of isopropanol

strongly depends on the development of effective dehydration technologies

[5]. Currently, distillation is the dominant separation technology employed for

the dehydration of IPA. However, its high energy consumption and

ineffectiveness in separating azeotropic mixtures pave the way for the

development of other separation technologies [6].

83
Pervaporation, a membrane-based separation technology, may be regarded as

a promising dehydration process [7] because it offers several unique

advantages over distillation. First, pervaporation is not limited by the vapor-

liquid equilibrium, and thus expected to have surpassing separation efficiency,

particularly in separating azeotropic and close boiling point liquid mixtures [8].

Second, it is a clean technology and can avoid cross-contamination because it

does not need third components, which are commonly employed in the so-

called “azeotropic distillations” [9]. Third, pervaporation is an energy-saving

process because only the permeating component consumes the latent heat [10,

11]. Moreover, pervaporation has a mild operation condition by using low

feed pressures and temperatures [8, 12]. Last, the compactness in module

fabrication as well as simplicity in process control are also advantageous

characteristics of pervaporation [13].

In pervaporation processes, membrane is vitally important. Polymeric

membranes are currently widely used in pervaporation because of their

advantages in ease of fabrication and scale-up, low capital costs, and low

footprint [14]. Polybenzoxazole (PBO) is a kind of glassy polymers which has

a unique rigid-rod structure and possesses advantageous characteristics such as

excellent solvent resistance, good thermal stability and high mechanical

strength [15-21]. However, its high solvent resistance reduces its solubility in

common solvents, and this brings about obstructions in membrane preparation

and constrains their applications [15-17]. To solve this problem, researchers

have developed an alternative PBO membrane preparation method via the in-

84
situ thermal conversion of hydroxyl-containing polyimide precursors. In this

method, an aromatic polyimide, which contains hydroxyl groups ortho to the

imide nitrogen, can be converted to a polybenzoxazole through thermal

rearrangement (TR) upon heating between 350 and 500 °C in an inert

atmosphere or under vacuum [15-17]. This method makes the fabrication of

PBO membranes more feasible.

Numerous works have focused on the application of thermally rearranged

polybenzoxazole (TR-PBO) membranes for gas separation. TR-PBO

membranes have shown superior gas separation properties especially in

separating CO2/CH4 mixtures, surpassing the 2008 upper bound [22-25]. The

existence of micropores and high free volume in TR polymers is the reason for

the high permeability, while the rigid-rod benzoxazole structure contributes to

the relatively stable selectivity [22]. Researchers have also investigated the

effects of different polyimide precursor structures [26-28], imidization

methods [29], TR protocols [27, 30-32] and TR-PBO membrane thickness [33]

on the free volume distribution and gas separation performance.

Although TR-PBO membranes have been widely explored for gas separation,

there are limited studies on their applications for pervaporation [34, 35].

Among the limited studies, only one of them was related to alcohol

dehydration where Ong et al. studied the effects of dwell duration and thermal

rearrangement temperature on alcohol dehydration performance [34].

Experimental results indicated that the thermal rearrangement temperature had

more significant impact on pervaporation performance than the dwell time. In

85
addition, the TR-PBO membranes which were treated at 450°C for half an

hour exhibited the highest flux with a reasonable separation factor for ethanol

dehydration.

In this study, a set of crosslinked thermally rearranged polybenzoxazole (C-

TR-PBO) membranes are prepared via in-situ thermal conversion of hydroxy-

containing polyimide precursors and are utilized for isopropanol dehydration

via pervaporation. The polyimide precursors are synthesized from the

polycondensation of 4,4’-(hexafluoroisopropylidene) diphthalic anhydride

(6FDA), 3,3’-dihydroxybenzidine diamine (HAB) and 3,5-diaminobenzoic

acid (DABA). The resultant polyimide precursors are first cast into

membranes, and then converted to C-TR-PBO through the thermal treatment.

The advantages of these three chosen monomers are: (1) 6FDA contains

fluorine groups –CF3, which are expected to prevent polymer chains from

compact packing and increase the free volume of the resultant polymer; (2)

HAB has a hydroxyl group ortho to the amine group, which meets the

essential requirement for the TR process; and (3) DABA possesses a

carboxylic group in the structure, which can undergo thermal crosslinking

during the thermal treatment. The thermally induced crosslinking mechanism

for the carboxylic-group containing polyimide was proposed by Qiu et al. [36]

as shown in Figure 5.1. Under vacuum or an inert condition and at high

temperatures, the carboxylic groups of two neighboring polymer chains react

to form anhydride groups. Subsequently, two phenyl free radicals are

generated from the decarboxylation of the anhydride along with the release of

a CO and CO2 molecule. After that, these adjacent phenyl radicals combine to

86
form linkages that yield biphenyl crosslinking. This would further enhance the

stability and swelling resistance of resultant membranes.

The present research aims to (1) develop new polybenzoxazole membrane

materials with combination of crosslinking and thermal rearrangement for

pervaporation applications (2) study the synergistic effects of crosslinking and

thermal rearrangement on the physicochemical properties of the membranes.

To our best knowledge, this is the first time that this material is synthesized

and studied for pervaporation dehydration.

Figure 5.1 Crosslinking mechanism of the polyimide containing carboxylic


groups under thermal treatment

5.2 Experimental

6FDA-HAB/DABA polyimides were synthesized as the method shown in

Chapter 3. Crosslinked polybenzoxazole membranes were obtained by

thermal rearrangement of polyimide precursors. The chemical structures of the

synthesized polyimide and poly(benzoxazole-co-imide) were analyzed by

FTIR. Membranes’ thermal stability was examined by TGA. In order to

investigate the reactions during the thermal treatment, TGA-IR was carried

out. PALS measurements and solvent uptake analyses were also conducted to

know the diffusivity and solubility information.

87
5.3 Results and discussion

5.3.1 Characterizations of PI and C-TR membranes

In order to verify the success in synthesizing the 6FDA-HAB/DABA

polyimide, ATR-FTIR analyses are carried out and the results are shown in

Figure 5.2. The synthesized polymers possess the characteristic peaks at 1774

and 1716 cm-1 which are mainly attributed to typical double bands stretching

of the C=O group in the imide ring [37-40]. In addition, the C–N stretching at

1365 cm-1, the transverse stretching of C–N–C groups at 1084 cm-1, and the

out-of-plane bending of C–N–C groups at 717 cm-1 also characterize the

presence of imide groups [34, 37, 38]. This confirms the successful fabrication

of the polyimide membrane.

Figure 5.2 FTIR spectra of synthesized polyimides

In contrast, for the C-TR-PBO membranes, it should be noted that the ATR-

FTIR spectra show weak peak intensities due to the dark brown colour of the

88
membranes. Therefore, a transmission mode of FTIR is used for the C-TR-

PBO membranes. According to Figure 5.3, three new peaks for C-TR-PBO

membranes can be observed at around 1477, 1554 and 1620 cm−1, which

characterize the vibration of benzoxazole rings [32, 41, 42]. Peaks at 1477 and

1554 cm−1 are the absorbing bands of benzoxazoles whereas the characteristic

peak at 1620 cm−1 is due to C=N stretching [32, 42]. These results validate the

successful formation of the polybenzoxazole structure.

Figure 5.3 FTIR spectra of poly(benzoxazole-co-imide)

Figure 5.4 shows the thermal behavior of the synthesized polyimides. The

polyimides undergo a three-stage weight loss as the temperature increases

from 100 to 700 oC. The initial weight loss (less than 1%) between 100oC to

250oC is ascribed to the elimination of residual solvents or moisture. The

second stage weight loss at 250-500oC is due to the release of gas molecules

89
produced from crosslinking and thermal rearrangement reactions. From 500 ºC

onwards, polymer chains start to decompose which results in the last stage

weight loss.

100
PI-5-5
90 PI-7-3
Weight percentage (%)

PI-9-1
80 PI-10-0

70

60

50

40
100 150 200 250 300 350 400 450 500 550 600 650 700
Temperature (ºC)

Figure 5.4 TGA curves of polyimide precursors

Figure 5.5 and Figure 5.6 show the TGA curve and the simultaneous IR

spectra of gases released from polyimide membranes PI-10-0 and PI-9-1

during pyrolysis, respectively. In TGA-IR spectra, yellow represents the

lowest intensity of a peak, red signifies the medium intensity, and turquoise

denotes the highest intensity. After about 48 min of heating corresponding to a

temperature of about 420oC, the CO2 characteristic IR peak (2250-2400 cm−1)

and the -C=O characteristic peak (1750 to 1850 cm−1) show up in the TGA-IR

spectra. The detected CO2 comes from the thermal rearrangement process.

When the temperature increases to about 520 oC, additional IR peaks including

H2O (3600-3750 cm−1), CFx (1150-1200 cm−1) and -C-O (1000-1050 cm−1)

groups are evolved. This suggests that the polymer backbone starts to

decompose at this high temperature.

90
Compared with the IR spectrum of PI-10-0, Figure 5.6 reveals that PI-9-1

experiences an earlier appearance of H2O peak, -C=O peak, -C-O peak and

CO2 peak at a lower temperature and illustrates the distinct peak of CO

molecules.

Figure 5.5 (a) Top-view 3D TGA-FTIR analyses of the polyimide precursor


PI-10-0 and (b) Heating time of TGA and weight percentage vs. temperature
correlation

Figure 5.6 (a) Top-view 3D TGA-FTIR analyses of the polyimide precursor


PI-9-1 and (b) Heating time of TGA and weight percentage vs. temperature
correlation

91
According to Figure 5.1, water, CO2 and CO molecules are released during the

crosslinking reaction. Therefore, this phenomenon implies that the thermally

induced crosslinking reaction occurs at lower temperatures than thermal

rearrangement. As temperature further increases, thermal rearrangement starts

to occur, while the crosslinking process is continuing. When temperature

reaches about 500 oC, decomposition of polymer chains starts, similar to the

case of PI-10-0.

5.3.2 Solvent uptake analyses

In order to understand the affinity between the permeate molecules and the

membrane, solvent uptake analyses were carried out at 60 oC. This temperature

is kept the same as the pervaporation test in order to have a fair comparison.

Figure 5.7 shows a comparison of solvent uptakes for the three polyimide

precursors and their respective C-TR membranes in isopropanol and water.

Both the polyimide precursors and C-TR membranes have higher affinity

towards isopropanol than water, which implies that the membranes remain

relatively hydrophobic. However, as the amount of carboxylic group increases,

the water uptake increases while isopropanol uptake decreases for both

polyimide precursors and C-TR membranes, indicating an improvement in

solubility selectivity towards water. This phenomenon is attributed to the

hydrophilic nature of carboxylic groups, which can play as absorption sites of

water. Comparing the polyimide and its respective C-TR membrane, it can be

observed that the latter absorbs less water and more isopropanol, which

implies C-TR membranes have lower solubility selectivity towards water than

92
polyimide membranes. This phenomenon corresponds well with the

observation reported by Ong et al. and Wang et al. [34, 43].

HAB-DABA

Figure 5.7 Solvent uptake analyses for PI and C-TR membranes

5.3.3 Positron annihilation lifetime spectroscopy analyses

Positron annihilation lifetime spectroscopy (PALS) was applied to

quantitatively analyze the free volume size and distribution for the polyimide

and C-TR membrane samples. In PALS measurements, the o-Ps lifetime τ3

and its intensity I3 are correlated to the size and concentration of the free

volume, respectively [44-46]. Figure 5.8 presents the PALS results. It can be

observed that the polyimide membranes have τ3 data in the range of 2.19 to

2.32 ns that corresponds to a mean cavity radius of 3.02 to 3.13Å. After

thermal rearrangement, τ3 increases to 2.47-2.69 ns, indicating an increase in

cavity sizes to 3.28-3.41Å. This increase is attributed to the formation of

93
cavities during the thermal rearrangement and thermal crosslinking reactions.

At the same time, compared with polyimide precursors, the intensities I3 of C-

TR membranes also increase, signifying that more pores are created after the

thermal treatment.

6 8
5.5 C-TR
5
PI
6
4.5
τ3(ns)

I3 (%)
4 4
3.5
3 C-TR 2
2.5
PI
2 0
10-0 9-1 7-3 5-5
HAB-DABA

Figure 5.8 PALS data of PI membranes and C-TR membranes

On the other hand, Figure 5.8 also reveals the effect of DABA content on

membrane free volume. For the PI membranes, τ3 and I3 do not show

significant change (4.59-4.72%) as the DABA to HAB ratio increases. This is

probably because the carboxylic side chain in DABA is too small to affect

polymer free volume. In contrast, the intensity I3 of C-TR membranes

decreases from C-TR-10-0 to C-TR-9-1 and subsequently increases from C-

TR-9-1 to C-TR-5-5. This decrease-and-increase trend of intensity for C-TR

membranes may be due to the combined effects of cavity formation during

thermal treatment and crosslinking’s inhibitory influence on TR [23, 24]. A

similar trend was reported by Calle et al. [24].

94
Figure 5.9 indicates the calculated fractional free volume as a function of

DABA to HAB ratio. Given that the lifetime τ3 for PI and C-TR membranes

remains relatively constant, it can be concluded that their fractional free

volumes are mainly affected by the change of intensity. The fractional free

volumes of PI membranes do not change significantly (0.96 to 1.09%) while

those of C-TR membranes exhibit a V-shape in the range from 1.44 to 2.20 %.

2.5
Fractional free volume (%)

2 C-TR

1.5

1 PI

0.5

0
10-0 9-1 7-3 5-5
HAB-DABA

Figure 5.9 Fractional free volumes of PI and C-TR membranes

5.3.4 Pervaporation performance of PI and C-TR membranes

Figure 5.10 summarizes the pervaporation performance of PI membranes and

C-TR membranes for dehydration of 85 wt% isopropanol aqueous solutions.

In order to eliminate the influence of membrane thickness, normalized flux is

used in this study for comparison.

95
4000 100.0
C-TR

Water concentration in permeate (wt%)


99.5
Total normalized flux (gμm/m h)

3500
2

PI 99.0
3000 98.5
98.0
2500
C-TR 97.5
2000
97.0
PI
1500 96.5
96.0
1000
95.5
500 95.0
10-0 9-1 7-3 5-5
HAB-DABA

Figure 5.10 Isopropanol dehydration performance of PI and C-TR membranes


at 60 ºC

As illustrated in Figure 5.10, as the DABA to HAB ratio increases, the

normalized flux of PI membranes also increases. However, their separation

factor initially increases from PI-10-0 to PI-7-3 but subsequently decreases for

PI-5-5. The increase in normalized flux is attributed to two factors. First, the

high DABA to HAB ratio increases the hydrophilicity of the membranes, as

discussed in section 3.2, which allows more water to be absorbed onto the

membranes. This, in turn, results in the loosening of polymeric chains (in the

wet state) and the increase in flux. Second, the membranes with a higher

DABA to HAB ratio have higher fraction free volumes (PI-7-3 and PI-5-5) as

indicated in Figure 5.9, which may lead to a higher flux in the wet state. On

the other hand, when the DABA to HAB ratio increases from 10:0 to 7:3, the

96
increase in separation factor is mainly due to the enhanced solubility

selectivity as discussed in section 3.2. For the membrane PI-5-5, the slight

decrease in separation factor may be attributed to the water-induced swelling.

Nonetheless, the water concentrations in permeate of all studied PI membranes

remain higher than 99.0%, indicating good separation performance.

Although the separation factors of PI membranes are high, their normalized

fluxes are relatively low. In order to further enhance the pervaporation

performance, the PI membranes are thermally treated to enable thermal

rearrangement and crosslinking. Figure 5.10 shows that the normalized fluxes

of C-TR membranes are higher than those of PI membranes for all DABA to

HAB ratios. The increase in flux of C-TR membranes is attributed to their

larger free volume sizes and higher fractional free volumes. However, when

compared with the PI membranes, the C-TR-10-0, C-TR-9-1 and C-TR-7-3

membranes have slightly lower separation factors. This is due to the decreased

diffusivity selectivity when their free volume sizes are larger and decreased

solubility selectivity as discussed in section 3.2. When the DABA to HAB

ratio increases to 5:5; however, the separation factor of C-TR-5-5 membranes

increases while that of PI membranes decreases, leading to the result that the

former is higher than the latter. This interesting result is attributed to anti-

swelling capacity of C-TR-5-5 membranes produced by the thermal

crosslinking reaction among carboxylic groups. Among C-TR membranes, C-

TR-5-5 displays the best separation performance with a normalized flux of

2138 gμm/m2h and a water concentration in permeate of approximate 99.8

wt%, as shown in Figure 5.10.

97
5.3.5 Performance benchmarking and stability of C-TR membranes

Table 5.1 compares the pervaporation performance in terms of permeability

and selectivity for the membranes in this study with various pervaporation

membranes available in literatures [34, 47-52]. Both the PI-5-5 and C-TR-5-5

membranes show comparable water permeability with other membranes. In

terms of selectivity, C-TR-5-5 exhibits outstanding water selectivity, which is

much higher than others. This superior performance opens up the opportunity

of using the newly designed polybenzoxazole-based membranes for biofuel

dehydration via pervaporation.

Besides having satisfactory performance, their long-term stability also needs

to be addressed. Therefore, in order to evaluate the long-term stability of the

C-TR membrane, the pervaporation performance is monitored continuously

for 200 h at 60 ºC. Figure 5.11 shows that the C-TR membrane displays quite

stable separation performance and no obvious reduction of flux and separation

factor can be observed during the entire testing duration of 200 h. The stable

long-term performance has verified the benefits of crosslinking and signifies

the feasibility of the C-TR membranes for pervaporation processes.

98
Table 5.1 A comparison of pervaporation performance of dense membranes
for isopropanol dehydration
Water Mole-
Water permeability
Temp. based
Membrane concentration -1 -1 Ref.
(ºC) (mg m h selectivity
in feed -1
kPa )
Matrimid 18% 100 0.490 50 [47]
Torlon 15% 60 0.011 3302 [48]
Ultem 15% 60 0.012 683 [48]
BPADA-ODA-
20% 60 0.145 9530 [49]
DABA polyimide
Crosslinked
10% 60 0.835 609 [50]
Chitosan
Matrimid/MgO
18% 100 0.067 1513 [47]
MMMs
PBI/ZIF-8 MMMs 15% 60 0.358 1969 [51]
P84/ZIF-90 MMMs 15% 60 0.166 449 [52]
TR PBO 10% 80 0.299 670 [34]
C-TR-5-5 15% 60 0.150 4019 This study
PI-5-5 15% 60 0.129 882 This study

200 100 Water concentration in permeate


90
180
80
160 70
Total flux (g/m2h)

140 60
(%)

50
120 40
100 30
20
80
10
60 0
0 50 100 150 200
Operating time (h)

Figure 5.11 Long-term stability of the C-TR-5-5 membrane for isopropanol


dehydration at 60 ºC

99
5.4 Conclusions

In this study, 6FDA-HAB/DABA copolyimides with four different HAB to

DABA molar ratios were synthesized and subsequently thermally treated to

form C-TR membranes. The science of both PI and C-TR membranes has

been studied and their isopropanol dehydration performance of isopropanol

has been evaluated. The following conclusions can be drawn:

(1) Thermal treatment of the 6FDA-HAB/DABA polyimides at 450 C for

half an hour induces thermal rearrangement and crosslinking reaction though

decarboxylation. A higher fractional free volume is achieved during thermal

rearrangement and crosslinking processes.

(2) As the DABA to HAB ratio increases, both PI and C-TR membranes

absorb more water and less isopropanol, which suggests an increase in

solubility selectivity towards water. On the other hand, the C-TR membrane

shows a higher affinity towards isopropanol than PI precursors, which

indicates that C-TR membranes have lower solubility selectivity towards

water than PI membranes.

(3) C-TR membranes show a higher normalized flux than their PI precursor

membranes. As the DABA to HAB ratio increases, both normalized flux and

separation factor of C-TR membranes increase simultaneously. C-TR-5-5

exhibits the best pervaporation performance.

(4) The C-TR-5-5 membrane displays stable isopropanol dehydration

performance at 60 C throughout the continuous duration of 200 h, indicating

great potential of the C-TR material for pervaporation application.

100
5.5 References

[1] L.M. Vane, Separation technologies for the recovery and dehydration of

alcohols from fermentation broths, Biofuels Bioprod. Biorefining, 2 (2008)

553-588.

[2] L.Y. Jiang, T.S. Chung, Homogeneous polyimide/cyclodextrin composite

membranes for pervaporation dehydration of isopropanol, J. Membr. Sci. 346

(2010) 45-58.

[3] S. Atsumi, J.C. Liao, Metabolic engineering for advanced biofuels

production from Escherichia coli, Curr. Opin. Biotechnol. 19 (2008) 414-419.

[4] T. Jojima, M. Inui, H. Yukawa, Production of isopropanol by

metabolically engineered Escherichia coli, Appl. Microbiol. Biotechnol. 77

(2008) 1219-1224.

[5] A.J. Ragauskas, C.K. Williams, B.H. Davison, G. Britovsek, J. Cairney,

C.A. Eckert, W.J. Frederick, J.P. Hallett, D.J. Leak, C.L. Liotta, J.R. Mielenz,

R. Murphy, R. Templer, T. Tschaplinski, The path forward for biofuels and

biomaterials, Science, 311 (2006) 484-489.

[6] L.M. Vane, A review of pervaporation for product recovery from biomass

fermentation processes, J. Chem. Technol. Biot. 80 (2005) 603-629.

[7] S.P. Nunes, K.V. Peinemann, Membrane technology: in the chemical

industry, John Wiley & Sons, 2006.

[8] P. Shao, R.Y.M. Huang, Polymeric membrane pervaporation, J. Membr.

Sci. 287 (2007) 162-179.

101
[9] B. Smitha, D. Suhanya, S. Sridhar, M. Ramakrishna, Separation of

organic–organic mixtures by pervaporation—a review, J. Membr. Sci. 241

(2004) 1-21.

[10] Y. Huang, R.W. Baker, L.M. Vane, Low-energy distillation-membrane

separation process, Ind. Eng. Chem. Res. 49 (2010) 3760-3768.

[11] G. Liu, W.S. Hung, J. Shen, Q. Li, Y.H. Huang, W. Jin, K.R. Lee, J.Y.

Lai, Mixed matrix membranes with molecular-interaction-driven tunable free

volumes for efficient bio-fuel recovery, J. Mater. Chem. A, 3 (2015) 4510-

4521.

[12] F. Lipnizki, R.W. Field, P.K. Ten, Pervaporation-based hybrid process: a

review of process design, applications and economics, J. Membr. Sci. 153

(1999) 183-210.

[13] X.S. Feng, R.Y.M. Huang, Liquid separation by membrane

pervaporation:  a review, Ind. Eng. Chem. Res. 36 (1997) 1048-1066.

[14] L.Y. Jiang, Y. Wang, T.S. Chung, X.Y. Qiao, J.Y. Lai, Polyimides

membranes for pervaporation and biofuels separation, Prog. Polym. Sci. 34

(2009) 1135-1160.

[15] G. Tullos, L. Mathias, Unexpected thermal conversion of hydroxy-

containing polyimides to polybenzoxazoles, Polymer, 40 (1999) 3463-3468.

[16] G.L. Tullos, J.M. Powers, S.J. Jeskey, L.J. Mathias, Thermal Conversion

of Hydroxy-Containing Imides to Benzoxazoles:  Polymer and Model

Compound Study, Macromolecules, 32 (1999) 3598-3612.

[17] X.D. Hu, S.E. Jenkins, B.G. Min, M.B. Polk, S. Kumar, Rigid-Rod

Polymers: Synthesis, Processing, Simulation, Structure, and Properties,

Macromol. Mater. Eng. 288 (2003) 823-843.

102
[18] K.I. Fukukawa, M. Ueda, Recent Development of Photosensitive

Polybenzoxazoles, Polym. J. 38 (2006) 405-418.

[19] Y. Imai, Recent advances in synthesis of high-temperature aromatic

polymers, React. Funct. Polym. 30 (1996) 3-15.

[20] Y. Imai, K. Itoya, M.A. Kakimoto, Synthesis of aromatic

polybenzoxazoles by silylation method and their thermal and mechanical

properties, Macromol. Chem. Phys. 201 (2000) 2251-2256.

[21] Y.H. So, J.P. Heeschen, Mechanism of Polyphosphoric Acid and

Phosphorus Pentoxide−Methanesulfonic Acid as Synthetic Reagents for

Benzoxazole Formation, J. Org. Chem. 62 (1997) 3552-3561.

[22] H.B. Park, C.H. Jung, Y.M. Lee, A.J. Hill, S.J. Pas, S.T. Mudie, E. Van

Wagner, B.D. Freeman, D.J. Cookson, Polymers with Cavities Tuned for Fast

Selective Transport of Small Molecules and Ions, Science, 318 (2007) 254-

258.

[23] M. Calle, C.M. Doherty, A.J. Hill, Y.M. Lee, Cross-Linked Thermally

Rearranged Poly(benzoxazole-co-imide) Membranes for Gas Separation,

Macromolecules, 46 (2013) 8179-8189.

[24] M. Calle, H.J. Jo, C.M. Doherty, A.J. Hill, Y.M. Lee, Cross-Linked

Thermally Rearranged Poly(benzoxazole-co-imide) Membranes Prepared

fromortho-Hydroxycopolyimides Containing Pendant Carboxyl Groups and

Gas Separation Properties, Macromolecules, 48 (2015) 2603-2613.

[25] H.B. Park, S.H. Han, C.H. Jung, Y.M. Lee, A.J. Hill, Thermally

rearranged (TR) polymer membranes for CO2 separation, J. Membr. Sci. 359

(2010) 11-24.

103
[26] C.H. Jung, J.E. Lee, S.H. Han, H.B. Park, Y.M. Lee, Highly permeable

and selective poly(benzoxazole-co-imide) membranes for gas separation, J.

Membr. Sci. 350 (2010) 301-309.

[27] M. Calle, Y.M. Lee, Thermally Rearranged (TR)

Poly(ether−benzoxazole) Membranes for Gas Separation, Macromolecules, 44

(2011) 1156-1165.

[28] S.H. Han, J.E. Lee, K.J. Lee, H.B. Park, Y.M. Lee, Highly gas permeable

and microporous polybenzimidazole membrane by thermal rearrangement, J.

Membr. Sci. 357 (2010) 143-151.

[29] S.H. Han, N. Misdan, S. Kim, C.M. Doherty, A.J. Hill, Y.M. Lee,

Thermally Rearranged (TR) Polybenzoxazole: Effects of Diverse Imidization

Routes on Physical Properties and Gas Transport Behaviors, Macromolecules,

43 (2010) 7657-7667.

[30] D.F. Sanders, Z.P. Smith, C.P. Ribeiro, R. Guo, J.E. McGrath, D.R. Paul,

B.D. Freeman, Gas permeability, diffusivity, and free volume of thermally

rearranged polymers based on 3,3′-dihydroxy-4,4′-diamino-biphenyl (HAB)

and 2,2′-bis-(3,4-dicarboxyphenyl) hexafluoropropane dianhydride (6FDA), J.

Membr. Sci. 409-410 (2012) 232-241.

[31] Z.P. Smith, D.F. Sanders, C.P. Ribeiro, R. Guo, B.D. Freeman, D.R. Paul,

J.E. McGrath, S. Swinnea, Gas sorption and characterization of thermally

rearranged polyimides based on 3,3′-dihydroxy-4,4′-diamino-biphenyl (HAB)

and 2,2′-bis-(3,4-dicarboxyphenyl) hexafluoropropane dianhydride (6FDA), J.

Membr. Sci. 415-416 (2012) 558-567.

104
[32] H. Wang, T.S. Chung, The evolution of physicochemical and gas

transport properties of thermally rearranged polyhydroxyamide (PHA), J.

Membr. Sci. 385-386 (2011) 86-95.

[33] H. Wang, T.S. Chung, D.R. Paul, Thickness dependent thermal

rearrangement of an ortho-functional polyimide, J. Membr. Sci. 450 (2014)

308-312.

[34] Y.K. Ong, H. Wang, T.S. Chung, A prospective study on the application

of thermally rearranged acetate-containing polyimide membranes in

dehydration of biofuels via pervaporation, Chem. Eng. Sci. 79 (2012) 41-53.

[35] C.P. Ribeiro, B.D. Freeman, D.S. Kalika, S. Kalakkunnath, Aromatic

polyimide and polybenzoxazole membranes for the fractionation of

aromatic/aliphatic hydrocarbons by pervaporation, J. Membr. Sci. 390-391

(2012) 182-193.

[36] W.L. Qiu, C.C. Chen, L.R. Xu, L.L. Cui, D.R. Paul, W.J. Koros, Sub-Tg

cross-linking of a polyimide membrane for enhanced CO2 plasticization

resistance for natural gas separation, Macromolecules, 44 (2011) 6046-6056.

[37] N.L. Le, Y. Wang, T.S. Chung, Synthesis, cross-linking modifications of

6FDA-NDA/DABA polyimide membranes for ethanol dehydration via

pervaporation, J. Membr. Sci. 415-416 (2012) 109-121.

[38] X. Qiao, T.S. Chung, Diamine modification of P84 polyimide membranes

for pervaporation dehydration of isopropanol, AIChE J. 52 (2006) 3462-3472.

[39] H. Chen, Y. Xiao, T.S. Chung, Synthesis and characterization of poly

(ethylene oxide) containing copolyimides for hydrogen purification, Polymer,

51 (2010) 4077-4086.

105
[40] W. Albrecht, B. Seifert, T. Weigel, M. Schossig, A. Holländer, T. Groth,

R. Hilke, Amination of Poly(ether imide) Membranes Using Di- and

Multivalent Amines, Macromol. Chem. Phys. 204 (2003) 510-521.

[41] S.Y. Wu, S.M. Yuen, C.C.M. Ma, Y.L. Huang, Synthesis and properties

of aromatic polyimide, poly(benzoxazole imide), and poly(benzoxazole amide

imide), J. Appl. Polym. Sci. 113 (2009) 2301-2312.

[42] D. Likhatchev, C. Gutierrez-Wing, I. Kardash, R. Vera-Graziano, Soluble

aromatic polyimides based on 2,2-bis(3-amino-4-hydroxyphenyl)

hexafluoropropane: Synthesis and properties, J. Appl. Polym. Sci. 59 (1996)

725-735.

[43] Y. Wang, S.H. Goh, T.S. Chung, P. Na, Polyamide-imide/polyetherimide

dual-layer hollow fiber membranes for pervaporation dehydration of C1–C4

alcohols, J. Membr. Sci. 326 (2009) 222-233.

[44] S.J. Lue, D.T. Lee, J.Y. Chen, C.H. Chiu, C.C. Hu, Y.C. Jean, J.Y. Lai,

Diffusivity enhancement of water vapor in poly(vinyl alcohol)–fumed silica

nano-composite membranes: Correlation with polymer crystallinity and free-

volume properties, J. Membr. Sci. 325 (2008) 831-839.

[45] H.M. Chen, W.S. Hung, C.H. Lo, S.H. Huang, M.L. Cheng, G. Liu, K.R.

Lee, J.Y. Lai, Y.M. Sun, C.C. Hu, R. Suzuki, T. Ohdaira , N. Oshima, Y.C.

Jean, Free-volume depth profile of polymeric membranes studied by positron

annihilation spectroscopy: layer structure from interfacial polymerization,

Macromolecules, 40 (2007) 7542-7557.

[46] J. Zuo, Y. Wang, T.S. Chung, Novel organic–inorganic thin film

composite membranes with separation performance surpassing ceramic

membranes for isopropanol dehydration, J. Membr. Sci. 433 (2013) 60-71.

106
[47] L.Y. Jiang, T.S. Chung, R. Rajagopalan, Matrimid®/MgO mixed matrix

membranes for pervaporation, AIChE J. 53 (2007) 1745-1757.

[48] Y. Wang, L.Y. Jiang, T. Matsuura, T.S. Chung, S.H. Goh, Investigation

of the fundamental differences between polyamide-imide (PAI) and

polyetherimide (PEI) membranes for isopropanol dehydration via

pervaporation, J. Membr. Sci. 318 (2008) 217-226.

[49] S. Xiao, X.S. Feng, R.Y.M. Huang, 2,2-Bis[4-(3,4-dicarboxyphenoxy)

phenyl]propane dianhydride (BPADA)-based polyimide membranes for

pervaporation dehydration of isopropanol: cCharacterization and comparison

with 4,4′-(hexafluoroisopropylidene) diphthalic anhydride (6FDA)-based

polyimide membranes, J. Appl. Polym. Sci. 110 (2008) 283-296.

[50] W. Zhang, G. Li, Y. Fang, X. Wang, Maleic anhydride surface-

modification of crosslinked chitosan membrane and its pervaporation

performance, J. Membr. Sci. 295 (2007) 130-138.

[51] G.M. Shi, T.X. Yang, T.S. Chung, Polybenzimidazole (PBI)/zeolitic

imidazolate frameworks (ZIF-8) mixed matrix membranes for pervaporation

dehydration of alcohols, J. Membr. Sci. 415-416 (2012) 577-586.

[52] D. Hua, Y.K. Ong, Y. Wang, T.X. Yang, T.S. Chung, ZIF-90/P84 mixed

matrix membranes for pervaporation dehydration of isopropanol, J. Membr.

Sci. 453 (2014) 155-167.

This chapter has been accepted as a journal paper:

Y.M. Xu, N.L. Le, J. Zuo, T.S. Chung, Aromatic polyimide and crosslinked

thermally rearranged poly (benzoxazole-co-imide) membranes for isopropanol

dehydration via pervaporation, J. Membr. Sci. 499 (2016) 317-325.

107
CHAPTER 6: UiO-66/POLYIMIDE MIXED MATRIX

MEMBRANES FOR ETHANOL,

ISOPROPANOL AND N-BUTANOL

DEHYDRATION VIA PERVAPORATION

6.1 Introduction

The global warming resulting from the heavy utilization of fossil fuels has

raised the urgency to explore renewable and sustainable energy so that

mankind can alleviate the reliance on fossil fuels. Biofuels, produced from

biomass by biological processes, have been viewed as a promising candidate

to substitute fossil fuels [1]. In the production of biofuels, the separation of

biofuels from fermentation broth is the most expensive process, which can

account for 60 to 80% of the total cost [1]. Therefore, the development of an

economic and effective separation technology is essential to produce cost-

effective biofuels. Pervaporation has been considered as a promising

candidate to replace conventional separation technologies such as distillation

and adsorption in chemical and energy industries as it has advantages of low

energy consumption, environmental friendliness, small footprint and mild

operation conditions [2-5]. What’s more, pervaporation is powerful to

separate azeotropic mixtures because it is not restricted by the vapor-liquid

equilibrium [6].

108
Based on membrane materials, pervaporation membranes can be classified

into inorganic membranes and polymeric membranes. Inorganic membranes

have superior thermal, chemical and mechanical stabilities. They are not

subjected to any solvent-induced swelling problem. Therefore, inorganic

membranes have superior separation performance especially in harsh

environments with strong solvents, low pH or high temperatures [7]. However,

their fragility and high costs of fabrication limit their wide applications.

Compared with inorganic membranes, polymeric membranes are more widely

utilized for pervaporation due to the advantages such as ease of fabrication and

scale-up, lower capital costs, better utilization of space and reasonable

separation performance [6, 8-10]. However, these membranes usually

encounter the trade-off relationship between permeability and selectivity.

They also suffer from the problem of long-term stability due to the solvent-

induced swelling effect.

In order to combine the strengths of inorganic and polymeric membranes,

mixed matrix membranes (MMMs) consisting of inorganic fillers dispersed in

organic polymers were first developed by Kulprathipanja et al. in 1988 [11].

The incorporation of inherently high permeability and selectivity inorganic

particles into polymeric matrices may result in MMMs that not only have

better separation performance than the original polymeric membranes but also

minimize the fragility issue in inorganic membranes [12, 13]. Numerous

inorganic fillers including zeolites, carbon molecular sieves, silica, metal

oxides, carbon nanotubes, silicate, metal organic frameworks (MOFs),

graphene oxide (GO), porous organic cages, covalent organic frameworks

109
have been investigated and employed in mixed matrix membranes [8, 14-16].

Among them, metal organic frameworks (MOFs), have been demonstrated as

promising materials for MMMs due to their unique characteristics such as

large surface areas, controllable porosity, high adsorption capacity for certain

gases and good affinity with polymer chains, as well as the tuneable chemical

properties [14-16].

Among various MOFs, UiO-66, synthesized by ZrCl4 and 1,4-benzene-

dicarboxylate (BDC), possessing 12-coordinated zirconium-oxo clusters, has

received significant attention since the pioneering work by Cavka et al. in

2008 [17]. UiO-66 based membranes have shown excellent gas separation

performance especially in CO2/CH4 separation with high selectivity and good

adsorption capacity [18-23]. Nik et al. reported CO2 permeability of the UiO-

66 based MMMs increased significantly compared to the neat polymer

membrane without compromising the selectivity [18]. Besides, UiO-66 based

membranes also show great potential in water reuse and desalination [24, 25].

The pure-phase UiO-66 polycrystalline membranes fabricated on alumina

hollow fibres exhibited excellent multivalent ion rejections with good

permeability and outstanding water stability up to 170h tests in a wide range

of saline solutions [24, 25]. The excellent chemical stability was attributed to

the strong coordination covalent bonds between the hard-acid−hard-base

interactions of the Zr atoms and carboxylate oxygens. Yin et al. combined

amino-functionalized UiO-66 (UiO-66-NH2) with ceramic membranes for

adsorptive removal of Pb (II) from wastewater [25]. The resultant membrane

110
was able to effectively remove most of Pb (II) (61.4%) and exhibited

remarkably high adsorption capacity (1795.3 mg·g-1).

Although UiO-66 based MMMs have a wide range of applications as

aforementioned, to our best knowledge, there is no study of UiO-66 based

MMMs for pervaporation. Since UiO-66 possesses many important

characteristics suitable for pervaporation such as high porosity, excellent

chemical and thermal stability and tunable chemical properties [26-29], this

work aims to explore the feasibility of applying UiO-66 based MMMs for

pervaporation and investigate the fundamental science behind the performance

enhancement. To meet the objectives, uniform UiO-66 nanoparticles with a

relatively small particle size would be firstly synthesized and subsequently

employed to prepare UiO-66/6FDA-HAB/DABA polyimide (PI) MMMs with

different particle loadings for dehydration of ethanol, isopropanol and n-

butanol via pervaporation. In addition, basic research on water and alcohol

uptakes and free volume size would be characterized in order to reveal the

actual mechanisms for the performance enhancement. This study may provide

useful insights to develop high performance UiO-66 based MMMs for alcohol

dehydration via pervaporation.

6.2 Experimental

The 6FDA-HAB/DABA polyimide was synthesized by polycondensation of

three monomers 6FDA, HAB and DABA with HAB to DABA of 9:1. UiO-66

was synthesized and MMMs were prepared as the methods in Chapter 3. The

111
crystallographic structure of the UiO-66 nanoparticles was explored by XRD

and cross-sectional morphologies of MMMs were observed by FESEM. EDX

was used to study the distribution of zirconium element. Thermal stability and

UiO-66 loading of MMMs were investigated by TGA. TGA-IR was

conducted to investigate the reactions during the heat treatment. PALS

measurements and solvent uptake tests were carried out to reveal the

mechanism of performance enhancement.

6.3 Results and discussion

6.3.1 Characterizations of UiO-66 and mixed matrix membranes

The crystallographic structure of the synthesized UiO-66 was measured by

XRD. Figure 6.1 compares the tested XRD result with the simulated one. All

characteristic diffraction peaks match very well with the simulated UiO-66

pattern in the database [30], showing the existence of a highly crystalline

structure of UiO-66 without other impurities. The peaks of the original MMM

are in accordance with the peaks of UiO-66 nanoparticle, indicating the

successful incorporation of UiO-66. In order to investigate the water stability

of UiO-66 based MMMs, a MMM with 20wt% loading was soaked in pure

water at 60°C for one week followed by XRD testing. Remarkably, the peaks

of the MMM after soaking in pure water at 60°C for one week are the same as

the original MMM without any treatment, which means that UiO-66 can

maintain its structure in pure water at 60°C for one week. This finding

confirms the good water stability of UiO-66 based MMMs reported by

previous works [24, 25]. Figure 6.2 shows the FESEM images of UiO-66

112
particles after removing the solvent by drying them in a vacuum oven. The

particles are quite uniform with a particle size of around 100nm.


Intensity (AU)

MMM, in water, 60 ºC, 1 week

MMM
UiO-66 Simulated result

UiO-66 Experimental result

5 15 25 35 45
2 Theta (º)

Figure 6.1 A comparison of XRD patterns of UiO-66 and MMMs

Figure 6.2 FESEM images of (a) UiO-66 nanoparticles and cross-sectional


morphologies of (b) the pristine PI membrane and MMMs containing (c) 10wt%
UiO-66, (d) 20wt% UiO-66, (e) 30wt% UiO-66

Figure 6.2 also displays the cross-sectional morphologies of MMMs consisting

of different UiO-66 loadings. The UiO-66 nanoparticles are dispersed evenly

113
in the polyimide matrix. There is no apparent particle agglomeration even at

the highest Uio-66 loading (30 wt%). These observations demonstrate good

compatibility between UiO-66 nanoparticles and the 6FDA-HAB/DABA

polyimide matrix. Since Zr element only exists in UiO-66 nanoparticles, EDX

mappings of Zr element can be utilized to investigate the dispersion of UiO-66

nanoparticles across the membrane. The result shown in Figure 6.3 further

confirms the even distribution of UiO-66 nanoparticles across the MMM.

Figure 6.3 (a) SEM image and (b) EDX mapping of Zr on the cross-section of
MMMs containing 30wt% UiO-66

TGA-IR was conducted to investigate the thermal stability of the UiO-66

nanoparticles and MMMs as well as the reactions during heat treatment.

Figure 6.4 and Figure 6.5 display the TGA curves and the simultaneous IR

spectra of gases released from UiO-66 nanoparticles and the MMM containing

20% UiO-66 during pyrolysis, respectively. For UiO-66, characteristic peaks

of CO2 (2250-2400 cm−1) and H2O (3600-3750 cm−1) are evolved when the

temperature increases to 530°C, which indicates the onset of decomposition of

the UiO-66 framework. These results confirm the superior thermal stability of

UiO-66 nanoparticles.

114
Figure 6.4 (a) Top-view of 3D TGA-FTIR analyses of UiO-66 and (b) its
temperature vs. weight percentage and heating time

Figure 6.5 (a) Top-view of 3D TGA-FTIR analyses of the MMM containing


20wt% UiO-66 and (b) its temperature vs. weight percentage and heating time

For MMMs, the situation is more complicated compared with pure UiO-66

nanoparticles because the 6FDA-HAB/DABA polyimide itself can go through

decarboxylation, thermal rearrangement and decomposition processes during

the heat treatment as explained in our previous study [31]. Figure 6.5(b) shows

that the MMM undergoes two stages of weight loss during the heat treatment.

The first stage of weight loss starts from 300°C to 520°C due to the

decarboxylation reaction in the DABA moieties and the thermal

rearrangement of the pristine polyimide. Figure 6.5(a) reveals that -C=O peak

(1750 to 1850 cm−1) and -C-O peak (1100 to 1200 cm−1) firstly show up at

115
around 300°C, which suggests the beginning of the decarboxylation reaction

[31, 32]. The evolved CO2 peak (2250-2400 cm−1) between 300°C and 520°C

comes from the decarboxylation reaction and the thermal rearrangement. As

temperature further increases to about 520oC, decompositions of UiO-66 and

polyimide chains begin, which results in the second stage of weight loss.

Along with these decomposition processes, -C-O, CFx, CO, CO2 and H2O are

released and their characteristic peaks are detected.

Besides the thermal stability, TGA curves can be used to determine the actual

loadings of UiO-66 nanoparticles in MMMs. As illustrated in Figure 6.6, the

residual ash at 800°C is ZrO2 because of the complete oxidation of UiO-66

under air atmosphere. Based on the remaining weight of ZrO2, the actual

loadings of UiO-66 nanoparticles in the UiO-66/PI MMM-10wt%, UiO-66/PI

MMM-20wt% UiO-66/PI MMM-30wt% are calculated to be 13.15wt%,

18.17wt% and 28.23wt%, respectively.

100
PI (0%)
90 MMM (10%)
Weight percentage (%)

80 MMM (20%)
70 MMM (30%)
60
50
40
30
20
10
0
0 100 200 300 400 500 600 700 800
Temperature (ºC)

Figure 6.6 TGA curves of the pristine PI membrane and MMMs

116
6.3.2 Solvent uptake analyses

The solvent uptake ratio gives the information about the affinity of water and

organic species towards the membranes. The solvent uptake tests were

conducted at 60 oC and characterized by the percentages of weight gain caused

by the sorption of tested solvents. Table 6.1 presents the solvent uptake ratios

of the membranes in water, ethanol, isopropanol and n-butanol. As the UiO-66

particle loading increases, the water uptake ratio increases from 3.0% to

14.0%. This suggests that the addition of UiO-66 nanoparticles into MMMs

would provide a higher water solubility. On the other hand, all ethanol,

isopropanol and n-butanol uptake ratios decrease, implying that the

membrane’s solubility selectivity towards water increases with an increase in

UiO-66 nanoparticle loading. This phenomenon may be because the ligand of

UiO-66, terephthalic acid, contains the hydrophilic functional groups that may

favour the sorption of water rather than alcohols. As a result, the water uptake

ratio increases while the organic species uptake ratio decreases with an

increase in UiO-66 nanoparticle loading.

Table 6.1 Solvent uptake analyses for the pristine PI membrane and MMMs
UiO-66 Water Ethanol Isopropanol n-Butanol
Membrane loading uptake uptake uptake uptake
wt% ratio% ratio% ratio% ratio%
PI 0 3.0±0.8 11.9±0.2 20.0±1.2 14.6±0.3
MMM 10 4.9±0.5 9.5±0.4 15.1±1.0 14.2±0.5
MMM 20 8.4±1.2 8.7±0.4 11.3±0.2 14.0±0.5
MMM 30 14.0±1.5 7.5±0.5 8.5±0.7 13.2±0.1
Test temperature: 60 ºC

117
6.3.3 Positron annihilation lifetime spectroscopy analyses

Positron annihilation lifetime spectroscopy (PALS) measurements were

carried out to investigate the free volume size and its distribution for the

membranes. In PALS, the o-Ps lifetime τ3 and its intensity I3 are related to the

size and concentration of the free volume, respectively [33-35]. Based on the

results of τ3 and I3, fractional free volume (FFV) of the MMMs can be

calculated from equations 2 and 3. Table 6.2 shows that τ3, I3, R and FFV all

increase with an increase in UiO-66 nanoparticle loading. Clearly, the

incorporation of the highly porous UiO-66 nanoparticles into the polyimide

matrix can create more and larger free volume, thus leads to an increase in

FFV. Remarkably, the FFV value is almost doubled when embedding 30wt%

UiO-66 nanoparticles into the polyimide matrix. Since permeability is a

product of diffusivity and solubility, and a larger FFV is beneficial for higher

diffusivity in alcohol dehydration, the addition of UiO-66 nanoparticles into

the polyimide would likely increase the overall permeability for alcohol

dehydration.

Table 6.2 PALS data for the pristine PI membrane and MMMs
UiO-
66
Membrane τ3 (ns) I3 (%) R (Å) FFV (%)
loading
wt%
PI 0 1.928±0.029 4.416±0.084 2.788±0.023 0.722±0.032
MMM 10 2.015±0.027 5.041±0.080 2.870±0.020 0.898±0.033
MMM 20 2.164±0.025 5.767±0.074 3.001±0.017 1.175±0.035
MMM 30 2.304±0.023 6.298±0.069 3.119±0.015 1.441±0.037

118
6.3.4 Pervaporation performance

• The effect of UiO-66 loading on dehydration performance of

ethanol, isopropanol and n-butanol

Figure 6.7, Figure 6.8 and Figure 6.9 present the ethanol, isopropanol and n-

butanol dehydration performance of the pristine polyimide membrane and

MMMs as a function of UiO-66 loading at 60 ºC, respectively. In order to

exclude the influence of membrane thickness, the normalized flux (i.e., the

product of flux and membrane thickness) is used in this study.

10000 100.0
Total normalized flux (gμm/m2h)

9000 90.0

Water concentration in the


8000 80.0
7000 70.0

permeate (wt%)
6000 60.0
5000 50.0
4000 40.0
3000 30.0
2000 Normalized Flux 20.0
1000 10.0
Water Concentration in Permeate
0 0.0
0 10 20 30
UiO-66 loading (%)
Figure 6.7 Ethanol dehydration performance of the pristine PI membrane
and MMMs at 60 ºC using ethanol/water (85/15wt%) as the feed

In terms of normalized flux, ethanol/water, isopropanol/water and n-

butanol/water systems have similar trends with an increase in UiO-66 loading.

The normalized flux increases rapidly with the increase of UiO-66 content in

119
MMMs. For example, the highest normalized flux for isopropanol dehydration

is 4724.5 gμm/m2h when the MMM contains 30wt% UiO-66. This value is

more than two times higher than that of the pristine polyimide membrane. The

improvements of normalized flux are attributed to two factors: higher

solubility and diffusivity. As shown in solvent uptake analyses in Section 3.2,

membranes with higher UiO-66 loadings display higher water uptake ratios

and higher water solubilities, which can allow more water molecules to be

absorbed by the membranes. The higher diffusivity can be proved by the

PALS analyses. Table 6.2 demonstrates that membranes with higher UiO-66

loadings exhibit higher FFV values, which can help penetrants to diffuse faster

across the membranes. As a consequence, the resultant MMMs exhibit a

higher normalized flux.

5000 100.0
Total normalized flux (gμm/m2h)

4500 99.0 Water concentration in the


4000 98.0
3500 97.0
permeate (wt%)

3000 96.0
2500 95.0
2000 94.0
1500 93.0
1000 Normalized Flux 92.0
500 Water Concentration in Permeate 91.0
0 90.0
0 10 20 30
UiO-66 loading (%)

Figure 6.8 Isopropanol dehydration performance of the pristine PI membrane


and MMMs at 60 ºC using isopropanol/water (85/15wt%) as the feed

120
5000 100.0
4500 99.0
Total normalized flux (gμm/m2h)

Water concentration in the


4000 98.0
3500 97.0

permeate (wt%)
3000 96.0
2500 95.0
2000 94.0
1500 93.0
1000 Normalized Flux 92.0
500 Water Concentration in Permeate 91.0
0 90.0
0 10 20 30
UiO-66 loading (%)

Figure 6.9 n-Butanol dehydration performance of the pristine PI membrane


and MMMs at 60 ºC using n-butanol/water (85/15wt%) as the feed

In terms of separation efficiency, the increase of UiO-66 loading may induce

two opposite effects. On one hand, an increase in UiO-66 loading would

enlarge the free volume size and the FFV value of the membranes, as shown

by the PALS data, resulting in the decrease of diffusivity selectivity which is

not favorable for alcohol and water separation. On the other hand, a higher

UiO-66 loading may result in MMMs that possess higher solubility selectivity,

as evidenced by the solvent uptake data, which is beneficial for higher

separation efficiency. These two effects compete with each other and may lead

to different results for different pairs of alcohol/water systems. For ethanol

dehydration, an increase in UiO-66 loading results in a lower water

concentration in permeate, as illustrated in Figure 6.7. However, for

isopropanol dehydration (Figure 6.8), the water concentration in the permeate

is kept at 99.9% if the MMMs contain UiO-66 less than 20wt%, but drops

slightly to 99.7% if the particle loading increases to 30wt%. For n-butanol

121
dehydration (Figure 6.9), the water concentration in the permeate stays

impressively high at 99.9% for all MMMs studied. The fundamental reasons

behind different separation behaviors for these three alcohol/water systems

will be further discussed in the following section.

• Comparison of dehydration performance for ethanol, isopropanol

and n-butanol

Figure 6.10 plots and compares the normalized fluxes and water

concentrations in permeate of the 6FDA-HAB/DABA polyimide and the

newly designed MMMs as a function of carbon number in alcohols. The

normalized flux follows the trend of ethanol/water system > isopropanol/water

system > n-butanol/water system. This trend is consistent with the reverse

order of alcohols’ molecular sizes because the alcohol with a smaller

molecular size can diffuse through the membrane easier and faster. Therefore,

the ethanol/water system shows the highest normalized flux, while the n-

butanol/water system has the lowest normalized flux. Meanwhile, the

separation efficiency in terms of water concentration in the permeate follows

the trend of n-butanol/water system≥ isopropanol/water system> ethanol/water

system. This trend obeys the order of alcohols’ molecular sizes because the

alcohol with a larger molecular size is easier to be separated. In addition, the

swelling effect induced by the ethanol/water system may lead to the loosening

of polymer chains and thus deteriorate the membrane’s ability to dehydrate

ethanol [36]. Furthermore, ethanol can form clusters with water and then pass

through the membranes together [6, 36]. This coupling effect also reduces the

122
separation efficiency of MMMs for ethanol dehydration. In summary, except

for the ethanol/water system, the newly developed MMMs are very effective

for isopropanol and n-butanol dehydration because their water concentrations

in the permeate are more than 99.7% for all MMMs studied.

a 10000
PI (0%)
9000
MMM (10%)
8000
Total normalized flux

MMM (20%)
7000 MMM (30%)
(gμm/m2h)

6000
5000
4000
3000
2000
1000
0
1 2 3 4 5
Carbon number of alcohols

b 100
98
Water concentration in the

96
94
permeate (wt%)

92
90
88
86 PI (0%)
84 MMM (10%)
MMM (20%)
82
MMM (30%)
80
1 2 3 4 5
Carbon number of alcohols

Figure 6.10 Dehydration performance of the pristine PI membrane and


MMMs for C2–C4 alcohols at 60 ºC using alcohol/water (85/15wt%) as the
feed: (a) Total normalized flux and (b) Water concentration in the permeate vs.
carbon number

123
• The effect of feed temperature and feed composition on

dehydration performance of isopropanol

A MMM containing 20% UiO-66 and an isopropanol/water system were

chosen to study the temperature effect. Pervaporation performance at

temperatures of 40, 60 and 80°C are shown in Figure 6.11. The normalized

flux increases with an increase in operating temperature from 40 to 80°C. The

flux enhancement is due to a significant increase in fugacity at the feed side

for the isopropanol/water system as shown in Table 6.3. This finding is

consistent with previous studies [36, 37]. On the other hand, the water

concentration in permeate keeps high and stable throughout the entire

temperature testing range.

5000 100.0
Water concentration in permeate
Total normalized flux (gμm/m2h)

4500 95.0
4000 90.0
3500 85.0
3000 80.0
(wt%)

2500 75.0
2000 70.0
1500 65.0
Normalized Flux
1000 60.0
Water Concentration in
500 55.0
Permeate
0 50.0
20 40 60 80 100
Feed Temperature (ºC)

Figure 6.11 Isopropanol dehydration performance of the MMM containing 20


wt% UiO-66 using isopropanol /water (85/15wt%) as the feed

124
Table 6.3 Thermodynamic properties of feed mixtures as a function of
operating temperature
Saturated Saturated
vapor vapor Fugacity Fugacity
Feed Activity Activity
pressure pressure of of
Feed temp. coefficient coefficient
of of solvent water
(℃) of solvent of water
solvent water (kPa) (kPa)
(kPa) (kPa)

Isopropanol
/Water 40 1.164 1.901 14.047 7.387 13.898 2.106

60 1.161 1.942 38.949 19.940 38.437 5.809


80 1.158 1.969 93.316 47.368 91.851 13.990

To study the effect of feed concentration, pervaporation performance of a

MMM with 20wt% UiO-66 loading was tested under three different feed

concentrations: 75/25 wt%, 85/15 wt% and 95/5 wt% IPA/water. Figure 6.12

shows the results. As the feed IPA concentration increases, the normalized

flux shows a decreasing trend. The reduction of the normalized flux is mainly

due to the decrease of driving force as the feed IPA concentration increases.

Since the vacuum condition at the permeate side is maintained, the fugacity at

the permeate side can be neglected. Therefore, the normalized flux mainly

depends on the fugacity at the feed side which is determined by concentration,

saturated vapour pressure and activity coefficient of species in the feed side.

As shown in Table 6.4, the fugacities of IPA and water at the permeate side

decrease as the IPA concentration increases. Therefore, the normalized flux

decreases. In terms of separation efficiency, water concentration in the

permeate can keep more than 99% when the feed concentration is lower than

85% but decrease to around 96% when the feed concentration increases to

95%. The drop of separation efficiency may be due to the swelling of the

MMM at this high feed concentration.

125
6000 100.0

Water concentration in permeate


Total ormalized flux (gμm/m2h)
95.0
5000
90.0
85.0
4000
80.0

(wt%)
3000 75.0
70.0
2000
65.0
Normalized Flux
60.0
1000
Water Concentration in Permeate 55.0
0 50.0
70 75 80 85 90 95 100
Feed Concentration(%)

Figure 6.12 Isopropanol dehydration of the MMM containing 20 wt% UiO-66


at 60 ºC

Table 6.4 Thermodynamic properties of feed mixtures as a function of


isopropanol concentration
Fugacity
Solvent Activity Activity Fugacity of
of
Feed concentration coefficient coefficient solvent
water
wt% of solvent of water (kPa)
(kPa)
Isopropanol/Water 75 1.385 1.566 40.458 7.807
85 1.161 1.942 38.437 5.809
95 1.023 2.773 37.853 2.765

6.3.5 Pervaporation benchmarking

A benchmarking of the 30/70 (wt/wt) UiO-66/6FDA-HAB/DABA polyimide

MMM with other pervaporation membranes for isopropanol and n-butanol,

ethanol dehydration is presented in Table 6.5, Table 6.6 and Table 6.7 [31, 36-

49]. The newly developed MMM containing 30wt% UiO-66 not only displays

a comparable mole-based selectivity of 2209 for isopropanol dehydration but

also possesses a much higher water permeability of 0.329 mg m-1 h-1 kPa-1

126
than most of others. For n-butanol dehydration, the 30/70 UiO-66/6FDA-

HAB/DABA polyimide MMM exhibits a high water permeability of 0.292 mg

m-1 h-1 kPa-1 and an extremely high mole-based selectivity of 14214. For

ethanol dehydration, the 30/70 UiO-66/6FDA-HAB/DABA polyimide MMM

shows a comparable water permeability but a low mole-based selectivity.

Since permeance is the ratio of permeability to the active layer thickness [56],

Table 6.5 and Table 6.6 also indicate the newly developed membranes have

comparable or superior dehydration performance of isopropanol and n-butanol

to other membranes. Therefore, the newly developed UiO-66/6FDA-

HAB/DABA MMMs have great potential for isopropanol and n-butanol

dehydration via pervaporation.

Table 6.5 A comparison of pervaporation performance of membranes for


isopropanol dehydration
Water
Membrane Water
Temp. permeability Mole-based
Membrane thickness concentration in Ref.
(ºC) (mg m-1 h-1 selectivity
(μm) feed
kPa-1)

®
Matrimid 30-60 18% 100 0.490 50 [38]
®
Torlon 23 15% 60 0.011 3302 [39]
®
Ultem 24 15% 60 0.012 683 [39]

BPADA-ODA-
DABA 15-20 20% 60 0.145 9530 [40]
polyimide
Crosslinked
20 10% 60 0.835 609 [41]
Chitosan
®
Matrimid /MgO 30-60 18% 100 0.067 1513 [38]
MMM
P84/ZIF-90
19-22 15% 60 0.166 449 [37]
MMM
TR PBO 20-25 10% 80 0.299 670 [36]
C-TR-5-5 27 15% 60 0.150 4019 [31]
Polyimide/UiO- This
30 15% 60 0.329 2209
66 MMM study

127
Table 6.6 A comparison of pervaporation performance of membranes for n-
butanol dehydration
Water
Membrane Water
Temp. permeability Mole-based
Membrane thickness concentration Ref.
(ºC) (mg m-1 h-1 selectivity
(μm) in feed
kPa-1)

Crosslinked
10 10% 30 0.916 87 [42]
PVA
SYMPLEX 2 10% 25 3.881 153 [43]
®
Pervap 2510 0.5-2 15% 60 0.08 17 [44]

TR PBO 20-25 10% 80 0.122 390 [36]


PBI 35-65 15% 60 0.029 2766 [45]
PBI/ZIF-8
35-65 15% 60 0.276 690 [45]
MMM
Polyimide/UiO- This
25 15% 60 0.292 14214
66 MMM study

Table 6.7 A comparison of pervaporation performance of membranes for


ethanol dehydration
Water
Membrane Water
Temp. permeability Mole-based
Membrane thickness concentration Ref.
(ºC) (mg m-1 h-1 selectivity
(μm) in feed
kPa-1)

Crosslinked
20 10% 60 1.136 417 [41]
Chitosan
Crosslinked
29 10% 60 0.823 456 [46]
PVA
®
P84 --- 15% 60 0.034 1874 [47]
copolyimide
®
Pervap 2201 --- 10% 60 0.022 309 [48]
TR PBO 20-25 10% 25 1.254 186 [36]
Polyimide/UiO- This
30 15% 60 0.572 34.5
66 MMM study

128
6.4 Conclusions

In this study, novel MMMs consisting of UiO-66 nanoparticles and 6FDA-

HAB/DABA polyimide have been designed for dehydration of ethanol,

isopropanol and n-butanol via pervaporation. The MMMs show superior

separation performance, signifying great potential for isopropanol and n-

butanol dehydration. The following conclusions can be also drawn:

(1) Uniform UiO-66 nanoparticles of around 100nm have been

successfully synthesized and confirmed with XRD. They disperse

evenly in the 6FDA-HAB/DABA polyimide matrix without any visible

agglomeration even at the highest 30wt% loading.

(2) An increase in UiO-66 loading results in an increase in water uptake

ratio but a decrease in alcohol uptake ratio. Thus, the MMMs’ solubility

selectivity towards water increases with the increase of UiO-66 loading

in MMMs.

(3) PALS data have confirmed that both free-volume radius R and

fractional free volume (FFV) increase with an increase in UiO-66

loading due to the high porosity nature of the UiO-66 nanoparticles.

(4) The incorporation of UiO-66 nanoparticles significantly improves

MMMs’ normalized flux and permeability because of enhanced

diffusivity and solubility. The water concentration in the permeate

decreases with the increase of UiO-66 loading for ethanol dehydration.

However, water concentration in the permeate for isopropanol

dehydration stays high at 99.9% for MMMs containing UiO-66

loadings less than 20wt%. For n-butanol dehydration, water

129
concentration in the permeate remains always high at 99.9% for all

MMMs.

(5) The normalized flux follows the trend of ethanol/water system >

isopropanol/water system > n-butanol/water system. However, the

separation efficiency in terms of water concentration in the permeate

obeys the opposite trend.

6.5 References

[1] A.J. Ragauskas, C.K. Williams, B.H. Davison, G. Britovsek, J. Cairney,

C.A. Eckert, W.J. Frederick, J.P. Hallett, D.J. Leak, C.L. Liotta, J.R. Mielenz,

R. Murphy, R. Templer, T. Tschaplinski, The path forward for biofuels and

biomaterials, Science, 311 (2006) 484-489.

[2] B. Smitha, D. Suhanya, S. Sridhar, M. Ramakrishna, Separation of

organic–organic mixtures by pervaporation—a review, J. Membr. Sci. 241

(2004) 1-21.

[3] Y. Huang, R.W. Baker, L.M. Vane, Low-energy distillation-membrane

separation process, Ind. Eng. Chem. Res. 49 (2010) 3760-3768.

[4] G. Liu, W.S. Hung, J. Shen, Q. Li, Y.H. Huang, W. Jin, K.-R. Lee, J.-Y.

Lai, Mixed matrix membranes with molecular-interaction-driven tunable free

volumes for efficient bio-fuel recovery, J. Mater. Chem. A, 3 (2015) 4510-

4521.

[5] L.M. Vane, A review of pervaporation for product recovery from biomass

fermentation processes, J. Chem. Technol. Biot. 80 (2005) 603-629.

130
[6] P. Shao, R.Y.M. Huang, Polymeric membrane pervaporation, J. Membr.

Sci. 287 (2007) 162-179.

[7] T.C. Bowen, R.D. Noble, J.L. Falconer, Fundamentals and applications of

pervaporation through zeolite membranes, J. Membr. Sci. 245 (2004) 1-33.

[8] Y.K. Ong, G.M. Shi, N.L. Le, Y.P. Tang, J. Zuo, S.P. Nunes, T.-S. Chung,

Recent membrane development for pervaporation processes, Prog. Polym. Sci.

57 (2016) 1-31.

[9] X.S. Feng, R.Y.M. Huang, Liquid separation by membrane pervaporation: 

a review, Ind. Eng. Chem. Res. 36 (1997) 1048-1066.

[10] L.Y. Jiang, Y. Wang, T.S. Chung, X.Y. Qiao, J.Y. Lai, Polyimides

membranes for pervaporation and biofuels separation, Prog. Polym. Sci. 34

(2009) 1135-1160.

[11] S. Kulprathipanja, R.W. Neuzil, N.N. Li, Separation of fluids by means of

mixed matrix membranes, US patent 4740219, 1988.

[12] T.S. Chung, L.Y. Jiang, Y. Li, S. Kulprathipanja, Mixed matrix

membranes (MMMs) comprising organic polymers with dispersed inorganic

fillers for gas separation, Prog. Polym. Sci. 32 (2007) 483-507.

[13] R. Mahajan, W.J. Koros, Factors Controlling Successful Formation of

Mixed-Matrix Gas Separation Materials, Ind. Eng. Chem. Res. 39 (2000)

2692-2696.

[14] Z. Jia, G. Wu, Metal-organic frameworks based mixed matrix membranes

for pervaporation, Microporous Mesoporous Mat. 235 (2016) 151-159.

[15] Y. Zhang, X. Feng, S. Yuan, J. Zhou, B. Wang, Challenges and recent

advances in MOF-polymer composite membranes for gas separation, Inorg.

Chem. Front. 3 (2016) 896-909.

131
[16] J. Caro, Are MOF membranes better in gas separation than those made of

zeolites?, Curr. Opin. Chem. Eng. 1 (2011) 77-83.

[17] J.H. Cavka, S. Jakobsen, U. Olsbye, N. Guillou, C. Lamberti, S. Bordiga,

K.P. Lillerud, A new zirconium inorganic building brick forming metal

organic frameworks with exceptional stability, J. Am. Chem. Soc. 130 (2008)

13850-13851.

[18] O.G. Nik, X.Y. Chen, S. Kaliaguine, Functionalized metal organic

framework-polyimide mixed matrix membranes for CO2/CH4 separation, J.

Membr. Sci. 413–414 (2012) 48-61.

[19] M.W. Anjum, F. Vermoortele, A.L. Khan, B. Bueken, D.E. De Vos, I.F.J.

Vankelecom, Modulated UiO-66-based mixed-matrix membranes for CO2

separation, ACS Appl. Mater. Inter. 7 (2015) 25193-25201.

[20] Q. Yang, H. Jobic, F. Salles, D. Kolokolov, V. Guillerm, C. Serre, G.

Maurin, Probing the dynamics of CO2 and CH4 within the porous zirconium

terephthalate UiO-66 (Zr): a synergic combination of neutron scattering

measurements and molecular simulations, Chem. Eur. J. 17 (2011) 8882-8889.

[21] Q.Y. Yang, A.D. Wiersum, H. Jobic, V. Guillerm, C. Serre, P.L.

Llewellyn, G. Maurin, Understanding the thermodynamic and kinetic behavior

of the CO2/CH4 gas mixture within the porous zirconium terephthalate UiO-66

(Zr): a joint experimental and modeling approach, J. Phys. Chem. C, 115

(2011) 13768-13774.

[22] A.D. Wiersum, E. Soubeyrand ‐ Lenoir, Q. Yang, B. Moulin, V.

Guillerm, M.B. Yahia, S. Bourrelly, A. Vimont, S. Miller, C. Vagner, M.

Daturi, G. Clet, C. Serre, G. Maurin, P.L. Llewellyn, An evaluation of UiO‐

66 for gas‐based applications, Chem. Asian J. 6 (2011) 3270-3280.

132
[23] S.J.D. Smith, B.P. Ladewig, A.J. Hill, C.H. Lau, M.R. Hill, Post-synthetic

Ti exchanged UiO-66 metal-organic frameworks that deliver exceptional gas

permeability in mixed matrix membranes, Sci. Rep. 5 (2015).

[24] X. Liu, N.K. Demir, Z. Wu, K. Li, Highly water-stable zirconium metal–

organic framework UiO-66 membranes supported on alumina hollow fibers

for desalination, J. Am. Chem. Soc. 137 (2015) 6999-7002.

[25] N. Yin, K. Wang, L. Wang, Z. Li, Amino-functionalized MOFs

combining ceramic membrane ultrafiltration for Pb (II) removal, Chem. Eng. J.

306 (2016) 619-628.

[26] M. Kandiah, M.H. Nilsen, S. Usseglio, S. Jakobsen, U. Olsbye, M. Tilset,

C. Larabi, E.A. Quadrelli, F. Bonino, K.P. Lillerud, Synthesis and stability of

tagged UiO-66 Zr-MOFs, Chem. Mat. 22 (2010) 6632-6640.

[27] J. Liu, N. Canfield, W. Liu, Preparation and characterization of a

hydrophobic metal–organic framework membrane supported on a thin porous

metal sheet, Ind. Eng. Chem. Res. 55 (2016) 3823-3832.

[28] F. Vermoortele, M. Vandichel, B. Van de Voorde, R. Ameloot, M.

Waroquier, V. Van Speybroeck, D.E. De Vos, Electronic effects of linker

substitution on Lewis acid catalysis with metal-organic frameworks, Angew.

Chem. Int. Ed. 51 (2012) 4887-4890.

[29] M.S. Denny, S.M. Cohen, In situ modification of metal – organic

frameworks in mixed‐matrix membranes, Angew. Chem. Int. Ed. 54 (2015)

9029-9032.

[30] Crystallography Open Database, Information card for 4512072.

〈http://www.crystallography.net/cod/4512072.html〉, 2016 (Accessed 08

September 2016).

133
[31] Y.M. Xu, N.L. Le, J. Zuo, T.S. Chung, Aromatic polyimide and

crosslinked thermally rearranged poly (benzoxazole-co-imide) membranes for

isopropanol dehydration via pervaporation, J. Membr. Sci. 499 (2016) 317-

325.

[32] W.L. Qiu, C.C. Chen, L.R. Xu, L.L. Cui, D.R. Paul, W.J. Koros, Sub-Tg

cross-linking of a polyimide membrane for enhanced CO2 plasticization

resistance for natural gas separation, Macromolecules, 44 (2011) 6046-6056.

[33] S.J. Lue, D.T. Lee, J.Y. Chen, C.H. Chiu, C.C. Hu, Y.C. Jean, J.Y. Lai,

Diffusivity enhancement of water vapor in poly(vinyl alcohol)–fumed silica

nano-composite membranes: Correlation with polymer crystallinity and free-

volume properties, J. Membr. Sci. 325 (2008) 831-839.

[34] H.M. Chen, W.S. Hung, C.H. Lo, S.H. Huang, M.L. Cheng, G. Liu, K.R.

Lee, J.Y. Lai, Y.M. Sun, C.C. Hu, R. Suzuki, T. Ohdaira , N. Oshima, Y.C.

Jean, Free-volume depth profile of polymeric membranes studied by positron

annihilation spectroscopy: layer structure from interfacial polymerization,

Macromolecules, 40 (2007) 7542-7557.

[35] J. Zuo, Y. Wang, T.S. Chung, Novel organic–inorganic thin film

composite membranes with separation performance surpassing ceramic

membranes for isopropanol dehydration, J. Membr. Sci. 433 (2013) 60-71.

[36] Y.K. Ong, H. Wang, T.S. Chung, A prospective study on the application

of thermally rearranged acetate-containing polyimide membranes in

dehydration of biofuels via pervaporation, Chem. Eng. Sci. 79 (2012) 41-53.

[37] D. Hua, Y.K. Ong, Y. Wang, T.X. Yang, T.S. Chung, ZIF-90/P84 mixed

matrix membranes for pervaporation dehydration of isopropanol, J. Membr.

Sci. 453 (2014) 155-167.

134
[38] L.Y. Jiang, T.S. Chung, R. Rajagopalan, Matrimid®/MgO mixed matrix

membranes for pervaporation, AIChE J. 53 (2007) 1745-1757.

[39] Y. Wang, L.Y. Jiang, T. Matsuura, T.S. Chung, S.H. Goh, Investigation

of the fundamental differences between polyamide-imide (PAI) and

polyetherimide (PEI) membranes for isopropanol dehydration via

pervaporation, J. Membr. Sci. 318 (2008) 217-226.

[40] S. Xiao, X.S. Feng, R.Y.M. Huang, 2,2-Bis[4-(3,4-dicarboxyphenoxy)

phenyl]propane dianhydride (BPADA)-based polyimide membranes for

pervaporation dehydration of isopropanol: cCharacterization and comparison

with 4,4 ′ -(hexafluoroisopropylidene) diphthalic anhydride (6FDA)-based

polyimide membranes, J. Appl. Polym. Sci. 110 (2008) 283-296.

[41] W. Zhang, G. Li, Y. Fang, X. Wang, Maleic anhydride surface-

modification of crosslinked chitosan membrane and its pervaporation

performance, J. Membr. Sci. 295 (2007) 130-138.

[42] M.C. Burshe, S.B. Sawant, J.B. Joshi, V.G. Pangarkar, Sorption and

permeation of binary water-alcohol systems through PVA membranes

crosslinked with multifunctional crosslinking agents, Sep. Purif. Technol. 12

(1997) 145-156.

[43] N. Scharnagl, K.V. Peinemann, A. Wenzlaff, H.H. Schwarz, R.D.

Behling, Dehydration of organic compounds with SYMPLEX composite

membranes, J. Membr. Sci. 113 (1996) 1-5.

[44] W.F. Guo, T.S. Chung, T. Matsuura, Pervaporation study on the

dehydration of aqueous butanol solutions: a comparison of flux vs. permeance,

separation factor vs. selectivity, J. Membr. Sci. 245 (2004) 199-210.

135
[45] G.M. Shi, T.X. Yang, T.S. Chung, Polybenzimidazole (PBI)/zeolitic

imidazolate frameworks (ZIF-8) mixed matrix membranes for pervaporation

dehydration of alcohols, J. Membr. Sci. 415-416 (2012) 577-586.

[46] M.L. Gimenes, L. Liu, X. Feng, Sericin/poly (vinyl alcohol) blend

membranes for pervaporation separation of ethanol/water mixtures, J. Membr.

Sci. 295 (2007) 71-79.

[47] X. Qiao, T.S. Chung, Fundamental characteristics of sorption, swelling,

and permeation of P84 co-polyimide membranes for pervaporation

dehydration of alcohols, Ind.Eng.Chem.Res. 44 (2005) 8938-8943.

[48] D. Van Baelen, B. Van der Bruggen, K. Van den Dungen, J. Degrève, C.

Vandecasteele, Pervaporation of water–alcohol mixtures and acetic acid–water

mixtures, Chem.Eng.Sci. 60 (2005) 1583-1590.

[49] R.W. Baker, J.G. Wijmans, Y. Huang, Permeability, permeance and

selectivity: A preferred way of reporting pervaporation performance data, J.

Membr. Sci. 348 (2010) 346-352.

This chapter has been accepted as a journal paper:

Y.M. Xu, T.S. Chung, High-Performance UiO-66/Polyimide Mixed Matrix

Membranes for Ethanol, Isopropanol and n-Butanol Dehydration via

Pervaporation, J. Membr. Sci. 531 (2017) 16-26.

136
CHAPTER 7: FUNCTIONALIZED UiO-66-TYPE MOFS

FOR 6FDA-POLYIMIDE BASED MIXED

MATRIX MEMBRANES FOR

DEHYDRATION OF C1-C3 ALCOHOLS VIA

PERVAPORATION

7.1 Introduction

The development of renewable and sustainable energy has received worldwide

attention due to the highly fluctuated oil price and global warming.

Bioalcohols are one of promising renewable and sustainable energy sources to

substitute fossil fuels [1]. For bioalcohol dehydration, pervaporation is a

promising technology due to its advantages of high energy efficiency,

environmental benignity, mild operation conditions and small footprint [2-5].

More importantly, pervaporation can be used to separate azeotropic mixtures

because it is not restricted by the thermodynamic vapor-liquid equilibrium [6].

Membrane is the heart of pervaporation processes. Substantial works have

been done to develop pervaporation membranes [7-10]. Polymeric

membranes are widely used for pervaporation because they have reasonable

separation performance. In addition, they can be fabricated easily and scaled

up cost-effectively [6, 11-13]. However, there is a trade-off relationship

between permeability and selectivity for polymeric membranes. They are also

vulnerable to the solvent-induced swelling issue that may deteriorate their

137
separation performance. Compared with polymeric membranes, inorganic

membranes don’t suffer from the swelling problem because of their superior

chemical, thermal and mechanical stability [14]. Nonetheless, their wide

applications are limited by the fragility and high fabrication cost. In order to

integrate the advantages of both polymeric and inorganic membranes, mixed

matrix membranes (MMMs) consisting of inorganic fillers and polymeric

matrices were invented by Kulprathipanja et al. in 1988 [15]. The inorganic

particles with an inherently high permeability and selectivity can help increase

the separation performance, while the polymeric matrices can minimize the

fragility issue and reduce the overall fabrication cost for MMMs [16,17].

Among numerous inorganic fillers [11, 18-20], metal organic frameworks

(MOFs) have been regarded as a promising material due to their unique

characteristics of ultrahigh porosities, tuneable chemical functionalities,

controllable structures, good affinity with polymers and high adsorption

capacity [18-20]. Among them, UiO-66 MOFs have been extensively studied

due to their noticeably high thermal, chemical and mechanical stability [21].

By altering ligands with various functional groups, one can easily manipulate

the pore size and chemical properties of UiO-66-Type MOFs [22,23]. UiO-66

functionalized with amine groups, UiO-66-NH2 MOFs, have been widely

applied in a variety of applications [24-28]. For example, Anjum et al.

reported that both selectivity and permeability were significantly improved by

incorporating UiO-66-NH2 into Matrimid because the amine groups inside

MOF pores facilitated the CO2 transport [24]. Nik et al. found that the

presence of amine-functional groups in UiO-66-NH2 MOFs increased both the

138
ideal selectivity and CO2 permeability for CO2/CH4 separation [25]. The study

of Peterson et al. suggested that UiO-66-NH2 MOFs could effectively remove

nitrogen dioxide from air with the aid of amine groups [26]. In addition, UiO-

66-NH2 based membranes also show great potential for water treatment and

dye removal [27-29].

Although UiO-66-NH2 based membranes have a wide range of applications as

aforementioned, there are limited studies for pervaporation [30, 31]. However,

none of them is related to alcohol dehydration. Therefore, the first objective of

this work is to explore the feasibility of applying UiO-66-NH2 based MMMs

for pervaporation. 6FDA-HAB/DABA polyimide is chosen as the polymer in

the MMMs because the original polyimide membrane has high separation

factor [32]. The incorporation of UiO-66-NH2 nanoparticles aims to improve

the overall flux due to the high porosity and the positive influence of

hydrophilic amino-functional groups on water transport. The other objectives

of this work are to (1) explore the effect of functional groups of UiO-66-Type

MOF on alcohol dehydration and (2) investigate and compare their separation

performances in three alcohol dehydration systems including isopropanol

/water, ethanol/water and methanol /water as a function of MOF particle

loading.

To meet the objectives, uniform UiO-66-NH2 nanoparticles would be firstly

synthesized for the fabrication of MMMs. To explain the effects of functional

groups on dehydration performance, fluorine functionalized UiO-66-Type

MOFs, UiO-66-F4 nanoparticles would be synthesized and incorporated into

139
MMMs for comparison. In addition, solvent uptake and positron annihilation

lifetime spectroscopy (PALS) analyses would be carried out to reveal the

fundamental science and mechanisms for the performance enhancement. 200-

h stability tests would be carried out to study their long-term performance.

This study suggests UiO-66-NH2 based MMMs are very promising for

dehydration of alcohols.

7.2 Experimental

The 6FDA-HAB/DABA polyimide, UiO-66-NH2 and UiO-66-F4 were

synthesized as the methods described in Chapter 3. The crystallographic

structure of the MOFs particles was explored by XRD. Cross-sectional

morphologies of MMMs were observed by FESEM. EDX was conducted to

study the distribution of particles by zirconium element. Thermal stabilities of

the MOFs particles and MMMs were examined by TGA. Solvent uptake and

positron PALS analyses as well as 200-hour stability tests were carried out to

reveal the fundamental science and underline mechanisms for the performance

enhancement.

7.3 Results and discussion

7.3.1 Characterizations of particles and mixed matrix membranes

XRD characterizations were carried to study the crystalline structure of the

synthesized particles. Figure 7.1 shows the XRD spectra of UiO-66-NH2 and

140
UiO-66-F4. All the characteristic diffraction peaks agree quite well with the

simulated result in the database [33], indicating that both UiO-66-NH2 and

UiO-66-F4 have a highly crystalline structure and an isostructural framework

similar to UiO-66.

40000
UiO-66-Simulated
35000 UiO-66-NH2
30000 UiO-66-F4
Intensity (a.u.)

25000
20000
15000
10000
5000
0
5 15 25 35 45
2 theta (º)

Figure 7.1 XRD spectra of UiO-66-NH2 and UiO-66-F4

Figure 7.2 shows the morphology of the synthesized UiO-66-NH2 and UiO-66-

F4 observed by FESEM. The synthesized UiO-66-NH2 particles are quite

uniform with a particle size of 80-90nm while the size of UiO-66-F4 particles

is around 90-130nm. Figure 7.3 shows the cross-sectional morphologies of the

pristine polyimide membrane and MMMs with various UiO-66-NH2 and UiO-

66-F4 loadings. The images reveal that UiO-66-NH2 and UiO-66-F4

nanoparticles can be dispersed homogenously in the polyimide matrix. The

nanoparticles don’t show any obvious agglomeration even at the highest

loading (30 wt%). These observations indicate good compatibility between

polyimide matrix and the synthesized nanoparticles. Figure 7.4 and Figure 7.5

141
show the EDX mapping of Zr across the MMMs and further confirms the

uniform distributions of UiO-66-NH2 and UiO-66-F4 in the MMMs since Zr

element only exists in the nanoparticles.

Figure 7.2 FESEM images of (a) UiO-66-NH2 and (b) UiO-66-F4


nanoparticles

Figure 7.3 FESEM images cross-sectional morphologies of (a) the pristine PI


membrane (b) MMM containing 10wt% UiO-66-F4 and MMMs containing (c)
10wt%, (d) 20wt%, (e) 30wt% UiO-66-NH2

142
Figure 7.4 (a) SEM image and (b) EDX mapping of Zr on the cross-section of
MMMs containing 30wt% UiO-66-NH2

Figure 7.5 (a) SEM image and (b) EDX mapping of Zr on the cross-section of
MMMs containing 10wt% UiO-66-F4

TGA was carried out to study the thermal stability of the UiO-66-NH2

nanoparticles, pristine PI membrane and MMMs. Figure 7.6 and Figure 7.7

depict the TGA curves of the UiO-66-NH2, UiO-66-F4 nanoparticles, pristine

PI membrane and MMMs under air atmosphere. The TGA curves of UiO-66-

NH2 and UiO-66-F4 agree well with literatures [22, 34]. The weight loss from

200 °C is attributed to the removal of the residual DMF solvent in the particles.

After that, the nanoparticles start to decompose. The detailed explanation of

the thermal behaviour of the pristine polyimide can be found in our previous

study [32]. The weight loss at 250–500 °C is because of decarboxylation

crosslinking and thermal rearrangement reactions. The weight loss after

500 °C is owing to the decomposition of polyimide chains. For MMMs, the

143
weight loss is mainly because of (1) the decomposition of UiO-66-type

nanoparticles, (2) decarboxylation crosslinking of the pristine polyimide, (3)

thermal rearrangement of the pristine polyimide and (4) degradation of the

pristine polyimide.

100
UiO-66-NH2
90 PI
MMM-10%
Weight percentage (%)

80
MMM-20%
70 MMM-30%
60
50
40
30
20
10
0
0 200 400 600 800
Temperature (℃)

Figure 7.6 TGA curves of UiO-66-NH2, pristine PI membrane and UiO-66-


NH2 based MMMs under air atmosphere

In addition, the TGA results can be applied to calculate the actual MOFs

loadings in MMMs. Assuming all nanoparticles have been completely

oxidized at 800°C under air atmosphere and the final ash is ZrO2. the actual

MOFs loading in MMMs can be determined based on the amount of ZrO2.

Therefore, the actual loadings of UiO-66-NH2 nanoparticles in MMM-NH2-

10wt%, MMM-NH2-20wt% and MMM-NH2-30wt% are calculated to be

8.8wt%, 19.7wt% and 27.5wt%, respectively while the actual loading of UiO-

66- F4 nanoparticles in MMM-F4-10wt% is 12.2wt%.

144
100
UiO-66-F4
90
PI
Weight percentage (%)
80
70 MMM-F4-10%

60
50
40
30
20
10
0
0 200 400 600 800
Temperature (℃)

Figure 7.7 TGA curves of UiO-66-F4, pristine PI membrane and UiO-66-F4


based MMMs under air atmosphere.

7.3.2 Solvent uptake analyses

To understand the interaction and affinity of water and different organic

solvents towards the membranes, solvent uptake tests were carried out at 60oC,

the same temperature used in the pervaporation tests. Table 7.1 shows the

effects of UiO-66-NH2 loading on solvent uptake ratios of its MMMs in water,

methanol, ethanol and isopropanol, while Table 7.2 tabulates the solvent

uptake ratios of MMMs containing different type MOFs. With an increase in

UiO-66-NH2 loading, the water uptake ratio increases but the uptake ratios of

methanol, ethanol and isopropanol decrease. This means that the adding of

UiO-66-NH2 nanoparticles into MMMs results in a higher water solubility but

a lower alcohol solubility. As a result, the solubility selectivity of UiO-66-NH2

based MMMs towards water increases as the UiO-66-NH2 loading increases.

145
Clearly, the hydrophilic amine groups in UiO-66-NH2 nanoparticles provide

more water sorption sites and work as water selective fillers.

Table 7.1 Solvent uptake ratios of the pristine PI membrane and UiO-66-NH2
based MMMs
UiO-66- Water Methanol Ethanol
Isopropanol
NH2 uptake uptake uptake
Membrane uptake ratio
loading ratio ratio ratio
%
wt % % % %
PI 0 3.3±0.3 9.2±0.3 11.5±0.7 18.4±0.6
MMM 10 5.6±0.5 8.0±0.6 8.3±0.3 10.4±0.6
MMM 20 9.4±0.5 6.1±0.9 7.1±0.3 9.8±0.8
MMM 30 16.9±0.7 5.0±0.2 6.5±0.6 8.2±0.8

Table 7.2 Comparison of solvent uptake ratio among different UiO-66 MOF
based MMMs
Particle Water Ethanol
Membrane loading uptake ratio uptake ratio
wt % % %
MMM-UiO-66-NH2 10 5.6±0.5 8.3±0.3
MMM-UiO-66-F4 10 4.4±0.2 12.1±1.0

The effect of hydrophilic amine groups in UiO-66-NH2 nanoparticles on water

uptake ratio can be further confirmed by comparing the solvent uptake ratios

of MMMs containing MOFs with different functional groups at the same 10

wt% loading. As shown in Table 7.2, the water uptake ratio follows the order

of UiO-66-NH2 > UiO-66-F4, which is mainly due to the different

hydrophilicities of the UiO-66-type particles resulted from the hydrophilic

amine and hydrophobic fluorine functional groups. While the ethanol uptake

ratio follows the opposite order. These orders indicate that UiO-66-NH2 based

MMMs should have greater water solubility and higher solubility selectivity

towards water.

146
7.3.3 Positron annihilation lifetime spectroscopy analyses

Positron annihilation lifetime spectroscopy (PALS) was employed to

investigate the free volume of the MMMs. The free volume size and its

concentration were characterized by the o-Ps lifetime τ3 and its intensity I3,

respectively [35-37]. Based on τ3 and I3, fractional free volume (FFV) can also

calculated. Table 7.3 summarizes the PALS data for the pristine polyimide

membrane and UiO-66-NH2 based MMMs. τ3, I3, R and FFV all increase with

an increase in UiO-66-NH2 loading because of the highly porous structure of

UiO-66-NH2 nanoparticles.

Table 7.3 PALS data of the pristine PI membrane and UiO-66-NH2 based
MMMs
UiO-
66-NH2
Membrane τ3 (ns) I3 (%) R (Å) FFV (%)
loading
wt%
PI 0 1.650±0.038 3.020±0.100 2.510±0.036 0.360±0.028
MMM 10 1.769±0.034 3.449±0.089 2.633±0.030 0.475±0.028
MMM 20 1.773±0.029 4.187±0.092 2.637±0.025 0.579±0.029
MMM 30 1.913±0.029 4.439±0.084 2.774±0.023 0.715±0.032

For comparison, the PALS data of UiO-66-F4 based MMMs at 10wt% particle

loading were also measured. Table 7.4 shows that at the same particle loading,

UiO-66-HN2 based MMMs have a larger free volume size (R) and a higher

FFV than UiO-66-F4 based MMMs because UiO-66-HN2 has a larger pore

volume (0.763cm3g−1) than that of UiO-66-F4 (0.616cm3g−1) [22].

147
Table 7.4 Comparison of PALS data between UiO-66-NH2 and UiO-66-F4
based MMMs
Particle
Membrane loading τ3 (ns) I3 (%) R (Å) FFV (%)
wt%
MMM-
UiO-66- 10 1.769±0.034 3.449±0.089 2.633±0.030 0.475±0.028
NH2
MMM-
10 1.660±0.030 3.639±0.105 2.520±0.029 0.420±0.028
UiO-66-F4

7.3.4 Pervaporation performance

• The effect of functional groups of MOFs particles on ethanol

dehydration

Table 7.5 compares the ethanol dehydration performance of UiO-66-NH2 and

UiO-66-F4 based MMMs using an ethanol/water mixture of 85/15 wt% as the

feed. All MMMs have a particle loading of 10wt%. To exclude the influence

of variation in membrane thickness, normalized flux is employed in this work,

which is defined as the product of membrane thickness and flux. It is found

that at the same particle loading, UiO-66-NH2 based MMMs have both higher

normalized flux and higher separation factor than that of UiO-66-F4 based

MMMs. Clearly, the functional groups of MOF particles play a determining

role on the dehydration performance of these MMMs. Not only the high FFV

of the UiO-66-NH2 MMMs facilitates the water transport, but also the

hydrophilic amino-functional groups on the MOF surface or inside the MOF

pores provide high water solubility and help water diffuse across the

membranes. Since the UiO-66-NH2 based MMMs have both the higher

148
normalized flux and separation factor, the UiO-66-NH2 based MMMs are the

focus of the following studies.

Table 7.5 Ethanol dehydration performance of UiO-66-NH2 and UiO-66-F4


based MMMs
Particle Normalized Water
Separation
Membrane loading flux Concentration in
factor
wt% (gμm/m2h) permeate (wt%)
MMM-UiO-
10 3560.4 96.2 142
66-NH2
MMM-UiO-
10 3057.1 93.7 84
66-F4

• The effect of UiO-66-NH2 loading on dehydration performance of

alcohols

Because UiO-66-NH2 based MMMs have better separation performance, their

methanol, ethanol and isopropanol dehydration performance as a function of

UiO-66-NH2 loading at 60 ºC were studied and shown in Figure 7.8, Figure

7.9 and Figure 7.10.

In terms of normalized flux, all methanol/water, ethanol/water and

isopropanol/water systems have similar trends. The normalized flux increases

as the UiO-66-NH2 loading increases. The flux enhancement can be attributed

to two factors: (1) higher water solubility due to the presence of hydrophilic

hydrophilic amino-functional groups and (2) higher diffusivity due to the

existence of higher FFV values. As a result, the MMMs with a 30wt% UiO-

66-NH2 loading have the highest normalized flux of 8878 gμm/m2h for ethanol

149
dehydration which is more than three times of the normalized flux of the

pristine polyimide membrane.

12000 100

Water concentration in permeate


Total normalized flux (gμm/m2h)

90
10000
80
70
8000
60

(wt%)
6000 50
40
4000
30
Normalized Flux 20
2000
10
Water Concentration in Permeate
0 0
0 10 20 30
UiO-66-NH2 loading (%)

Figure 7.8 Methanol dehydration performance of the pristine PI membrane


and MMMs at 60 ºC with feed composition of methanol/water (85/15wt%)

10000 100
Total normalized flux (gμm/m2h)

9000 90
Water concentration in permeate

8000 80
7000 70
6000 60
(wt%)

5000 50
4000 40
3000 30
2000 Normalized Flux 20
1000 Water Concentration in Permeate 10
0 0
0 10 20 30
UiO-66-NH2 loading (%)

Figure 7.9 Ethanol dehydration performance of the pristine PI membrane and


MMMs at 60 ºC with feed composition of ethanol/water (85/15wt%)

150
4500 100

Total normalized flux (gμm/m2h)

Water concentration in permeate


4000 90

3500 80
70
3000
60

(wt%)
2500
50
2000
40
1500
30
1000 Normalized Flux 20
500 10
Water Concentration in Permeate
0 0
0 10 20 30
UiO-66-NH2 loading (%)

Figure 7.10 Isopropanol dehydration performance of the pristine PI membrane


and MMMs at 60 ºC with feed composition of isopropanol/water (85/15wt%)

In terms of separation efficiency, the water concentration in permeate firstly

increases when the UiO-66-NH2 loading is increased from 0 to 10wt% for

methanol/water and ethanol/water systems, but it decreases if the MOF

loading is further increased to 30wt%. As shown by solvent uptake and PALS

analyses, with an increase in UiO-66-NH2 loading, solubility selectivity

increases while diffusivity selectivity decreases. These two contradictory

effects lead to different results at various particle loadings. At a lower particle

loading range of 0–10 wt%, the FFV increment degree is relatively small and

the decrease of diffusivity selectivity is not severe. Therefore, the

enhancement of solubility selectivity is the main reason for the increase of

water concentration in permeate. However, when the particle loading is further

increased, the FFV value becomes much higher especially at the highest

particle loading of 30 wt% thus the decrease of diffusivity selectivity becomes

the dominant factor for the decrease of water concentration in permeate at

151
higher particle loadings. In addition, the up-and-down convex trend between

water concentration and UiO-66-NH2 loading is much severer for the

methanol/water system than the ethanol/water system because methanol has a

much small kinetic diameter than ethanol (i.e., 0.38 vs. 0.43 nm).

Nevertheless, the water concentrations in permeate of MMMs-10% and

MMMs-20wt% are higher than that of the pristine polyimide membrane for

methanol/water and ethanol/water systems.

The isopropanol/water system shows the best separation performance and the

water concentration in permeate can keep impressively high at 99.9% for all

the membranes studied. In addition, there is no an up-and-down convex trend

between water concentration and UiO-66-NH2 loading because IPA has a large

kinetic diameter and molecular size.

• Comparison of dehydration performance for methanol, ethanol

and isopropanol

Figure 7.11 replots the normalized flux and water concentration in permeate as

a function of carbon number in alcohols. The normalized flux follows the

trend of methanol/water system > ethanol/water system > isopropanol/water

system. This order is consistent with the reverse order of alcohol molecular

sizes because a small alcohol size tends to diffuse faster across the membrane.

While the water concentration in permeate follows the order of alcohol

molecular sizes: isopropanol/water system > ethanol/water system >

methanol/water system. This is because (1) an alcohol with a larger molecular

152
size can be easily intercepted and separated by the membrane and (2) an

alcohol with a higher linearity can pass through the membrane network more

easily [38, 39]. Since isopropanol has the largest molecular size and lowest

linearity among these three alcohols, the isopropanol/water system shows the

highest separation factor.

a
12000
PI (0%)
MMM (10%)
Total normalized flux

10000
MMM (20%)
8000 MMM (30%)
(gμm/m2h)

6000
4000
2000
0
0 1 2 3 4
Carbon number of alcohols
b

100
Water concentration in

80
permeate (wt%)

60

40 PI (0%)
MMM (10%)
20 MMM (20%)
MMM (30%)
0
0 1 2 3 4
Carbon number of alcohols

Figure 7.11 Dehydration performance of the pristine PI membrane and UiO-


66-NH2 based MMMs for C1–C3 alcohols at 60 ºC with feed composition of
alcohol /water (85/15wt%): (a) Normalized flux and (b) Water concentration
in the permeate vs. carbon number

153
7.3.5 Long stability of UiO-66-NH2 based MMMs

Besides separation performance, long-term stability is another significant

factor to evaluate a membrane. Therefore, a 200-h test was carried out for IPA

dehydration at 60oC using MMMs consisting of 10wt% UiO-66-NH2. Figure

7.12 presents the results. The MMMs display very stable performance without

obvious changes of flux and separation factor in the entire testing duration.

The good stability of UiO-66-NH2 based MMMs is mainly because the amine

groups in UiO-66-NH2 can improve the compatibility between the MOF

particles and the 6FDA-HAB/DABA polyimide through intermolecular

bonding with imide groups, thus effectively reduce membrane swelling [24].

5000

Water concentration in the permeate


4500 100
Total normalized flux (gμm/m2h)

4000
3500 80

3000
(wt%)

60
2500
2000
40
1500
1000 20
500
0 0
0 50 100 150 200
Test Duration (h)

Figure 7.12 Long-term stability performance of 10wt% UiO-66-NH2 MMMs


for isopropanol dehydration at 60 ºC with feed composition of
isopropanol/water (85/15wt%)

154
7.3.6 Performance benchmarking

Table 7.6, Table 7.7 and Table 7.8 compare the dehydration performance of

methanol, ethanol and isopropanol of the newly developed UiO-66-NH2 based

MMMs with different pervaporation membranes in literatures [39-51]. Up to

now, methanol dehydration via pervaporation has not been widely used due to

the lack of suitable materials with both affordable price and good anti-swelling

properties [39]. However, compared with other polymeric membranes, Table

7.6 shows that the MMM comprising 10 wt% UiO-66-NH2 has better

methanol dehydration performance than most of other membranes with the

most balanced performance of normalized flux and separation factor. For

ethanol dehydration, the MMMs have comparable performance with the

literature data. For isopropanol dehydration, the MMMs with 30wt% UiO-66-

NH2 not only show a comparable water permeability of 0.295 mg m-1 h-1 kPa-1

but also have an outstanding mole-based selectivity of 66250, which is much

higher than most of others. Overall, the newly developed UiO-66-NH2 based

MMMs possess great potential for alcohol dehydration via pervaporation.

155
Table 7.6 A comparison of pervaporation performance for methanol
dehydration
Water Mole-
Membrane Water
Temp. permeability based
Membrane thickness concentration in Ref.
(ºC) (mg m-1 h-1 selectivity
(μm) feed
kPa-1)

5%-sPPSU 16 15% 60 0.033 33.8 [39]


2.5%-sPPSU 16 15% 60 0.030 32.1 [39]
Agarose 14-20 14% 30 0.925-1.321 8.3 [47]
Polyamidesulfon-
30 10% 20 0.268-0.415 9.2-99.9 [48]
amide
Polyimide/UiO-66- This
NH MMM-10% 34 15% 60 0.413 43.0
2 study

Polyimide/UiO-66- This
NH MMM-20% 40 15% 60 0.455 28.8
2 study
Polyimide/UiO-66- This
NH MMM-30% 22 15% 60 0.599 19.8
2 study

Table 7.7 A comparison of pervaporation performance of membranes for


ethanol dehydration
Water
Membrane Water Mole-
Temp. permeability
Membrane thickness concentration based Ref.
(ºC) (mg m-1 h-1
(μm) in feed selectivity
kPa-1)

Crosslinked
20 10% 60 1.136 417 [44]
Chitosan
Crosslinked PVA 29 10% 60 0.823 456 [49]
®
P84 copolyimide --- 15% 60 0.034 1874 [50]
®
Pervap 2201 --- 10% 60 0.022 309 [51]

TR PBO 20-25 10% 25 1.254 186 [41]


Polyimide/UiO-66- This
NH2 MMM-10% 33 15% 60 0.256 202.5
study
Polyimide/UiO-66- This
NH2 MMM-20% 44 15% 60 0.345 124.5
study
Polyimide/UiO-66- This
NH2 MMM-30% 21 15% 60 0.602 80.3
study

156
Table 7.8 A comparison of pervaporation performance of membranes for
isopropanol dehydration
Water
Membrane Water Mole-
Temp. permeability
Membrane thickness concentration based Ref.
(ºC) (mg m-1 h-1
(μm) in feed selectivity
kPa-1)
®
Matrimid 30-60 18% 100 0.490 50 [40]
®
Torlon 23 15% 60 0.011 3302 [42]
®
Ultem 24 15% 60 0.012 683 [42]

BPADA-ODA-
15-20 20% 60 0.145 9530 [43]
DABA polyimide
Crosslinked
20 10% 60 0.835 609 [44]
Chitosan
P84/ZIF-90 MMM 19-22 15% 60 0.166 449 [46]
TR PBO 20-25 10% 80 0.299 670 [41]
C-TR-5-5 27 15% 60 0.150 4019 [32]
Polyimide/UiO-66- This
NH2 MMM-10% 34 15% 60 0.205 66250
study
Polyimide/UiO-66- This
NH2 MMM-10% 47 15% 60 0.265 66250
study
Polyimide/UiO-66- This
NH2 MMM-10% 19 15% 60 0.295 66250
study

7.4 Conclusions

In this study, making mixed matrix membrane (MMMs) containing porous

UiO-66-NH2 particles and 6FDA-HAB/DABA polyimide was demonstrated

as an effective method of improving membrane’s normalized flux and

separation factor simultaneously. The UiO-66-NH2 based MMMs have

excellent separation performance, showing great potential for alcohol

dehydration. The following conclusions can be drawn:

(1) UiO-66-NH2 nanoparticles with particle size of 80-90nm and UiO-66-

F4 nanoparticles with particle size of 90-130nm have been successfully

synthesized. The nanoparticles can disperse uniformly in the polyimide

157
matrix without visible agglomerations even for the highest loading of

30wt%.

(2) The hydrophilic amine groups in UiO-66-NH2 nanoparticles can

provide more water sorption sites and work as water selective fillers. A

higher UiO-66-NH2 loading results in higher water solubility, lower

alcohol solubility, thus higher solubility selectivity towards water. At

the same particle loading, UiO-66-NH2 based MMMs have higher

water solubility, lower ethanol solubility and higher solubility

selectivity towards water than UiO-66-F4 based MMMs.

(3) FFV increases with an increase in UiO-66-NH2 loading increases due

to the highly porous structure of UiO-66-NH2. At the same particle

loading of 10wt%, UiO-66-HN2 based MMMs have larger free volume

size (R) and higher FFV than that of UiO-66-F4 based MMMs.

(4) UiO-66-NH2 based MMMs have better separation performance than

UiO-66-F4 based MMMs. Incorporation of UiO-66-NH2 nanoparticles

can greatly improve both the normalized flux and separation factor of

MMMs simultaneously if the particle loading is within the range of 0–

20 wt%.

(5) The normalized flux shows the trend of methanol/water system >

ethanol/water system > isopropanol/water system and water

concentration in permeate follows the reverse trend.

(6) The UiO-66-NH2 based MMMs have stable isopropanol dehydration

performance throughout the uninterrupted 200-h testing due to the good

compatibility of UiO-66-NH2 particle and the polyimide.

158
7.5 References

[1] A.J. Ragauskas, C.K. Williams, B.H. Davison, G. Britovsek, J. Cairney,

C.A. Eckert, W.J. Frederick, J.P. Hallett, D.J. Leak, C.L. Liotta, J.R. Mielenz,

R. Murphy, R. Templer, T. Tschaplinski, The path forward for biofuels and

biomaterials, Science, 311 (2006) 484-489.

[2] B. Smitha, D. Suhanya, S. Sridhar, M. Ramakrishna, Separation of

organic–organic mixtures by pervaporation—a review, J. Membr. Sci. 241

(2004) 1-21.

[3] Y. Huang, R.W. Baker, L.M. Vane, Low-energy distillation-membrane

separation process, Ind. Eng. Chem. Res. 49 (2010) 3760-3768.

[4] G. Liu, W.S. Hung, J. Shen, Q. Li, Y.H. Huang, W. Jin, K.-R. Lee, J.-Y.

Lai, Mixed matrix membranes with molecular-interaction-driven tunable free

volumes for efficient bio-fuel recovery, J. Mater. Chem. A, 3 (2015) 4510-

4521.

[5] L.M. Vane, A review of pervaporation for product recovery from biomass

fermentation processes, J. Chem. Technol. Biotechnol. 80, (2005) 603-629.

[6] P. Shao, R.Y.M. Huang, Polymeric membrane pervaporation, J. Membr.

Sci. 287 (2007) 162-179.

[7] A.A. Alomair, S.M. Al-Jubouri, S.M. Holmes, A novel approach to

fabricate zeolite membranes for pervaporation processes, J. Mater. Chem. A, 3

(2015) 9799-9806.

[8] L.L. Xia, C.L. Li, Y. Wang, In-situ crosslinked PVA/organosilica hybrid

membranes for pervaporation separations, J. Membr. Sci. 498 (2016) 263-275.

159
[9] A. Svang-Ariyaskul, R.Y.M. Huang, P.L. Douglas, R. Pal, X. Feng, P.

Chen, L. Liu, Blended chitosan and polyvinyl alcohol membranes for the

pervaporation dehydration of isopropanol, J. Membr. Sci. 280 (2006) 815-823.

[10] H.A. Tsai, W.H. Chen, C.Y. Kuo, K.R. Lee, J.Y. Lai, Study on the

pervaporation performance and long-term stability of aqueous iso-propanol

solution through chitosan/polyacrylonitrile hollow fiber membrane, J. Membr.

Sci. 309 (2008) 146-155.

[11] Y.K. Ong, G.M. Shi, N.L. Le, Y.P. Tang, J. Zuo, S.P. Nunes, T.S. Chung,

Recent membrane development for pervaporation processes, Prog. Polym. Sci.

57 (2016) 1-31.

[12] X.S. Feng, R.Y.M. Huang, Liquid separation by membrane

pervaporation:  a review, Ind. Eng. Chem. Res. 36 (1997) 1048-1066.

[13] L.Y. Jiang, Y. Wang, T.S. Chung, X.Y. Qiao, J.Y. Lai, Polyimides

membranes for pervaporation and biofuels separation, Prog. Polym. Sci. 34

(2009) 1135-1160.

[14] T.C. Bowen, R.D. Noble, J.L. Falconer, Fundamentals and applications of

pervaporation through zeolite membranes, J. Membr. Sci. 245 (2004) 1-33.

[15] S. Kulprathipanja, R.W. Neuzil, N.N. Li, Separation of fluids by means of

mixed matrix membranes, US patent 4740219, 1988.

[16] T.S. Chung, L.Y. Jiang, Y. Li, S. Kulprathipanja, Mixed matrix

membranes (MMMs) comprising organic polymers with dispersed inorganic

fillers for gas separation, Prog. Polym. Sci. 32 (2007) 483-507.

[17] R. Mahajan, W.J. Koros, Factors Controlling Successful Formation of

Mixed-Matrix Gas Separation Materials, Ind. Eng. Chem. Res. 39 (2000)

2692-2696.

160
[18] Z. Jia, G. Wu, Metal-organic frameworks based mixed matrix membranes

for pervaporation, Microporous Mesoporous Mater. 235 (2016) 151-159.

[19] Y. Zhang, X. Feng, S. Yuan, J. Zhou, B. Wang, Challenges and recent

advances in MOF-polymer composite membranes for gas separation, Inorg.

Chem. Front. 3 (2016) 896-909.

[20] J. Caro, Are MOF membranes better in gas separation than those made of

zeolites? Curr. Opin. Chem. Eng. 1 (2011) 77-83.

[21] J.H. Cavka, S. Jakobsen, U. Olsbye, N. Guillou, C. Lamberti, S. Bordiga,

K.P. Lillerud, A new zirconium inorganic building brick forming metal

organic frameworks with exceptional stability, J. Am. Chem. Soc. 130 (2008)

13850-13851.

[22] Z. Hu, Y. Peng, Z. Kang, Y. Qian, D. Zhao, A Modulated Hydrothermal

(MHT) Approach for the Facile Synthesis of UiO-66-Type MOFs, Inorg.

Chem. 54 (2015) 4862-4868.

[23] M. Kandiah, M.H. Nilsen, S. Usseglio, S. Jakobsen, U. Olsbye, M. Tilset,

C. Larabi, E.A. Quadrelli, F. Bonino, K.P. Lillerud, Synthesis and stability of

tagged UiO-66 Zr-MOFs, Chem. Mater. 22 (2010) 6632-6640.

[24] M.W. Anjum, F. Vermoortele, A.L. Khan, B. Bueken, D.E. De Vos, I.F.J.

Vankelecom, Modulated UiO-66-based mixed-matrix membranes for CO2

separation, ACS Appl. Mater. Inter. 7 (2015) 25193-25201.

[25] O.G. Nik, X.Y. Chen, S. Kaliaguine, Functionalized metal organic

framework-polyimide mixed matrix membranes for CO2/CH4 separation, J.

Membr. Sci. 413–414 (2012) 48-61.

161
[26] G.W. Peterson, J.J. Mahle, J.B. DeCoste, W.O. Gordon, J.A. Rossin,

Extraordinary NO2 Removal by the Metal-Organic Framework UiO-66-NH2,

Angew. Chem. Int. Ed. 55 (2016) 6235–6238.

[27] N. Yin, K. Wang, L. Wang, Z. Li, Amino-functionalized MOFs

combining ceramic membrane ultrafiltration for Pb (II) removal, Chem. Eng. J.

306 (2016) 619-628.

[28] B.J. Yao, W.L. Jiang, Y. Dong, Z.X. Liu, Y.B. Dong, Post‐Synthetic

polymerization of UiO-66-NH2 nanoparticles and polyurethane oligomer

toward stand-alone membranes for dye removal and separation, Chem. Eur.J.

22 (2016) 10565-10571.

[29] H. Sun, B. Tang, P. Wu, Development of hybrid ultrafiltration

membranes with improved water separation properties using modified

superhydrophilic metal-organic framework nanoparticles, ACS Appl. Mater.

Interfaces, 9 (2017) 21473-21484.

[30] N. Wang, G. Zhang, L. Wang, J. Li, Q. An, S. Ji, Pervaporation

dehydration of acetic acid using NH2-UiO-66/PEI mixed matrix membranes,

Sep. Purif. Technol. 186 (2017) 20-27.

[31] L. Wan, C. Zhou, K. Xu, B. Feng, A. Huang, Synthesis of highly stable

UiO-66-NH2 membranes with high ions rejection for seawater desalination,

Microporous Mesoporous Mater, 252 (2017) 207-213.

[32] Y.M. Xu, N.L. Le, J. Zuo, T.S. Chung, Aromatic polyimide and

crosslinked thermally rearranged poly (benzoxazole-co-imide) membranes for

isopropanol dehydration via pervaporation, J. Membr. Sci. 499 (2016) 317-

325.

162
[33] Crystallography Open Database, Information card for 4512072.

〈http://www.crystallography.net/cod/4512072.html〉, 2016 (Accessed 08

September 2016).

[34] J.B. DeCoste, G.W. Peterson, H. Jasuja, T.G. Glover, Y.G. Huang, K.S.

Walton, Stability and degradation mechanisms of metal–organic frameworks

containing the Zr6O4(OH)4 secondary building unit, J. Mater. Chem. A, 1

(2013) 5642.

[35] S.J. Lue, D.T. Lee, J.Y. Chen, C.H. Chiu, C.C. Hu, Y.C. Jean, J.Y. Lai,

Diffusivity enhancement of water vapor in poly(vinyl alcohol)–fumed silica

nano-composite membranes: Correlation with polymer crystallinity and free-

volume properties, J. Membr. Sci. 325 (2008) 831-839.

[36] H.M. Chen, W.S. Hung, C.H. Lo, S.H. Huang, M.L. Cheng, G. Liu, K.R.

Lee, J.Y. Lai, Y.M. Sun, C.C. Hu, R. Suzuki, T. Ohdaira , N. Oshima, Y.C.

Jean, Free-volume depth profile of polymeric membranes studied by positron

annihilation spectroscopy: layer structure from interfacial polymerization,

Macromolecules, 40 (2007) 7542-7557.

[37] J. Zuo, Y. Wang, T.S. Chung, Novel organic–inorganic thin film

composite membranes with separation performance surpassing ceramic

membranes for isopropanol dehydration, J. Membr. Sci. 433 (2013) 60-71.

[38] L.Y. Jiang, T.S. Chung, R. Rajagopalan, Dehydration of alcohols by

pervaporation through polyimide Matrimid® asymmetric hollow fibers with

various modifications, Chem. Eng. Sci. 63 (2008) 204-216.

[39] Y. Tang, N. Widjojo, G.M. Shi, T.S. Chung, M. Weber, C. Maletzko,

Development of flat-sheet membranes for C1–C4 alcohols dehydration via

163
pervaporation from sulfonated polyphenylsulfone (sPPSU), J. Membr. Sci.

415-416 (2012) 686-695.

[40] L.Y. Jiang, T.S. Chung, R. Rajagopalan, Matrimid®/MgO mixed matrix

membranes for pervaporation, AIChE J. 53 (2007) 1745-1757.

[41] Y.K. Ong, H. Wang, T.S. Chung, A prospective study on the application

of thermally rearranged acetate-containing polyimide membranes in

dehydration of biofuels via pervaporation, Chem. Eng. Sci. 79 (2012) 41-53.

[42] Y. Wang, L.Y. Jiang, T. Matsuura, T.S. Chung, S.H. Goh, Investigation

of the fundamental differences between polyamide-imide (PAI) and

polyetherimide (PEI) membranes for isopropanol dehydration via

pervaporation, J. Membr. Sci. 318 (2008) 217-226.

[43] S. Xiao, X.S. Feng, R.Y.M. Huang, 2,2-Bis[4-(3,4-dicarboxyphenoxy)

phenyl]propane dianhydride (BPADA)-based polyimide membranes for

pervaporation dehydration of isopropanol: characterization and comparison

with 4,4 ′ -(hexafluoroisopropylidene) diphthalic anhydride (6FDA)-based

polyimide membranes, J. Appl. Polym. Sci. 110 (2008) 283-296.

[44] W. Zhang, G. Li, Y. Fang, X. Wang, Maleic anhydride surface-

modification of crosslinked chitosan membrane and its pervaporation

performance, J. Membr. Sci. 295 (2007) 130-138.

[45] G.M. Shi, T.X. Yang, T.S. Chung, Polybenzimidazole (PBI)/zeolitic

imidazolate frameworks (ZIF-8) mixed matrix membranes for pervaporation

dehydration of alcohols, J. Membr. Sci. 415-416 (2012) 577-586.

[46] D. Hua, Y.K. Ong, Y. Wang, T.X. Yang, T.-S. Chung, ZIF-90/P84 mixed

matrix membranes for pervaporation dehydration of isopropanol, J. Membr.

Sci. 453 (2014) 155-167.

164
[47] M. Yoshikawa, K. Masaki, M. Ishikawa, Pervaporation separation of

aqueous organic mixtures through agarose membranes, J. Membr. Sci. 205

(2002) 293-300.

[48] W.H. Chan, C.F. Ng, S.Y. Lam-Leung, X. He, Water–alcohol separation

by pervaporation through chemically modified poly (amidesulfonamide)s, J.

Membr. Sci. 160 (1999) 77-86.

[49] M.L. Gimenes, L. Liu, X. Feng, Sericin/poly (vinyl alcohol) blend

membranes for pervaporation separation of ethanol/water mixtures, J. Membr.

Sci. 295 (2007) 71-79.

[50] X. Qiao, T.S. Chung, Fundamental characteristics of sorption, swelling,

and permeation of P84 co-polyimide membranes for pervaporation

dehydration of alcohols, Ind.Eng.Chem.Res., 44 (2005) 8938-8943.

[51] D. Van Baelen, B. Van der Bruggen, K. Van den Dungen, J. Degrève, C.

Vandecasteele, Pervaporation of water–alcohol mixtures and acetic acid–water

mixtures, Chem.Eng.Sci. 60 (2005) 1583-1590.

This chapter has been accepted as a journal paper:

Y.M. Xu, S. Japip, T.S. Chung, Mixed Matrix Membranes with nano-sized

functional UiO-66-type MOFs embedded in 6FDA-HAB/DABA Polyimide for

Dehydration of C1-C3 Alcohols via Pervaporation, J. Membr. Sci. (2018) In

Press.

165
CHAPTER 8: CONCLUSIONS AND ECOMMENDATIONS

8.1 Conclusions

Understanding the limitations and challenges of current pervaporation

technology, design and development of novel polymeric and mixed matrix

membranes with high separation performance and good stability have been

conducted in this PhD study. Polyarylether (PESU, PPSU, T-PESU and H-

PESU), polyimide and crosslinked thermally rearranged poly(benzoxazole-co-

imide) were explored as membrane materials. Three UiO-66-type MOFs

(UiO-66, UiO-66-NH2 and UiO-66-F4) were synthesized as fillers of mixed

matrix membranes. Different alcohol/water systems including methanol/water,

ethanol/water, isopropanol/water, and n-butanol/water were investigated.

Fundamental chemical and thermal properties, mass transport behaviour,

dehydration performance and long-term stability were systematically studied.

The following conclusions can be drawn from the aforementioned study.

8.1.1 Polyarylether membranes for dehydration of ethanol and methanol

via pervaporation

In this work, pervaporation membranes made from four kinds of polyarylether

(PAE) polymers including polyethersulfone (PESU), polyphenylsulfone

(PPSU), polytrimethylphenylethersulfone (T-PESU) and hydrophilic PESU-

co-Pluronic (H-PESU) were developed for dehydration of ethanol and

methanol. These PAE membranes exhibited excellent separation performance

166
for both alcohol dehydration cases. Fundamental characteristics of pore

structure, free volume, d-spacing, water contact angle, water and ethanol

uptakes were investigated and correlated with their dehydration performance.

PALS measurements show that swelling gives rise to a larger free volume

radius and a bigger FFV. Moreover, the adjacent small cavities in the

membranes at the dry state may coalesce to form large cavities at the wet state.

The H-PESU membrane shows the best ethanol dehydration performance due

to its synergetic combination of the strengths from both rigid hydrophobic

PESU and flexible hydrophilic Pluronic moieties. It has the highest water

permeability of 0.322 mg m-1 h-1 kPa-1 which is 2-6 times higher than other

three membranes because the hydrophilic Pluronic moiety may form special

water transport channels. It also has a comparable mole-based selectivity of

685 due to the rigid hydrophobic PESU moiety that prevents the membrane

from uncontrolled swelling. Long-term tests evidence the operating stability of

both H-PESU and PESU membranes. In addition, the methanol dehydration

performance of these membranes is superior to the other reported polymeric

membranes. The water in the permeate can be concentrated to 86.8%. These

promising preliminary results suggest that these polyarylether (PAE)

membranes have great potential for ethanol and methanol dehydration.

167
8.1.2 Polyimide and Crosslinked Thermally Rearranged Poly(benzoxazole-

co-imide) Membranes for Isopropanol Dehydration via

Pervaporation

In this study, successful synthesis of PI precursors by the polycondensation of

three monomers; namely, 4,4’-(hexafluoroisopropylidene) diphthalic

anhydride (6FDA), 3,3’-dihydroxybenzidine diamine (HAB) and 3,5-

diaminobenzoic acid (DABA) was demonstrated. Novel crosslinked thermally

rearranged polybenzoxazole (C-TR-PBO) membranes, which show impressive

results for isopropanol dehydration, have been obtained via in-situ thermal

conversion of hydroxyl-containing polyimide precursors. Due to the

incorporation of the carboxylic-group containing diamine DABA into an

ortho-hydroxypolyimide precursor, the thermal induced crosslinking reaction

can be achieved together with the thermal rearrangement process.

Consequently, a synergistic effect of high permeability and high selectivity

can be realized in one step. The resultant C-TR-PBO membrane exhibits an

unambiguous enhancement in permeation flux compared to their PI precursors

due to the higher fractional free volume of C-TR-PBO membrane achieved in

thermal rearrangement and crosslinking processes. As the DABA to HAB

ratio increases, both normalized flux and separation factor of C-TR

membranes increase simultaneously. C-TR-5-5 exhibits the best pervaporation

performance with a comparable water permeability of 0.15 mgm-1h-1kPa-1 and

an impressively high mole-based selectivity of 4019. Moreover, the newly

developed C-TR-PBO membrane displays stable isopropanol dehydration

performance at 60 C throughout the continuous 200 hours. The promising

168
preliminary results achieved in this study may offer useful insights for the

selection of membrane materials for pervaporation and new methods to

molecularly design next-generation pervaporation membranes.

8.1.3 UiO-66/Polyimide Mixed Matrix Membranes for Ethanol,

Isopropanol and n-Butanol Dehydration via Pervaporation

In this study, novel MMMs consisting of UiO-66 nanoparticles and 6FDA-

HAB/DABA polyimide have been developed for dehydration of ethanol,

isopropanol and n-butanol via pervaporation. Uniform UiO-66 nanoparticles

with a particle size of around 100nm were successfully synthesized and they

can be evenly dispersed in the 6FDA-HAB/DABA polyimide matrix without

visible agglomeration even at the highest 30wt% loading. Adding UiO-66 into

the 6FDA-HAB/DABA polyimide not only significantly enhances both free-

volume radius and fractional free volume but also water solubility and

solubility selectivity towards water of the MMMs. Therefore, the

incorporation of UiO-66 remarkably improves the normalized flux of MMMs

for the dehydration of ethanol/water, isopropanol/water and n-butanol/water

systems. The MMMs show excellent separation performance for the

dehydration of isopropanol and n-butanol. At the highest UiO-66 loading of

30wt%, the MMMs have the water permeability of 0.329 and 0.292 mg m-1 h-1

kPa-1 and mole-based selectivity of 2209 and 14214 respectively for

isopropanol/water and n-butanol/water systems, outperforming most literature

data. These experimental results strongly suggest the newly developed UiO-

169
66/polyimide MMMs have great potential for isopropanol and n-butanol

dehydration via pervaporation.

Moreover, dehydration performances for different alcohol/water systems were

compared. Normalized flux of the MMMs for different alcohol/water systems

follows the trend of ethanol/water system > isopropanol/water system > n-

butanol/water system. However, the separation efficiency in terms of water

concentration in the permeate obeys the opposite trend. In addition, the study

of feed temperature effects shows that with the increase of feed temperature,

normalized flux increases and water concentration in permeate keeps high and

stable at 99.9%. The study of feed composition effects reveals that with the

increase of feed concentration, both normalized flux and separation factor

show decreasing trends.

8.1.4 Functionalized UiO-66-Type MOFs for 6FDA-Polyimide based

Mixed Matrix Membranes for Dehydration of C1-C3 Alcohols via

Pervaporation

This work further developed mixed matrix membranes consisting of two

different functionalized UiO-66-type MOFs (UiO-66-NH2 and UiO-66-F4) and

6FDA-HAB/DABA polyimide for alcohol dehydration via pervaporation.

UiO-66-NH2 with particle size of 80-90nm and UiO-66-F4 with particle size

of 90-130nm were successfully synthesized. The nanoparticles can disperse

uniformly in the polyimide matrix without visible agglomerations even for the

highest loading of 30wt%. Three different alcohols: methanol, ethanol,

170
isopropanol dehydration performances were studied and compared.

Normalized flux shows the trend of methanol/water system > ethanol/water

system > isopropanol/water system and water concentration in permeate

follows the reverse trend. The effects of functional groups of the MOFs

particles and the effects of particle loading on the alcohol separation

performance were systematically investigated. Results show the UiO-66-NH2

based MMMs have better separation performance than UiO-66-F4 based

MMMs due to the positive influence of hydrophilic amino-functional groups

on water transport. Incorporation of UiO-66-NH2 nanoparticles can greatly

improve normalized flux of the MMMs due to the enhanced water solubility

and diffusivity, and at the same time enhance separation factor within the

particle loading range of 0-20wt% due to the higher selectivity towards water.

The UiO-66-NH2 based MMMs have stable isopropanol dehydration

performance throughout the uninterrupted 200-hour testing due to the good

compatibility of UiO-66-NH2 particle and polyimide resulting from the

intermolecular bonding among their amine and imide groups. The UiO-66-

NH2 based MMMs show excellent alcohol dehydration performance. The

encouraging results suggest a promising future of the UiO-66-NH2 based

MMMs for alcohol dehydration via pervaporation.

8.2 Recommendations

In this thesis, various membranes including polymeric and mixed matrix

membranes have been developed. The performances for ethanol dehydration

of these membranes developed are summarized and compared in Table 8.1.

171
Table 8.1 A comparison of pervaporation performance of dense membranes
for ethanol dehydration
Water
Membrane Water Water Mole-based
Temp. permeability
Membrane thickness concentration concentration selectivity
(℃) (mg m-1 h-1
(μm) in feed in permeate (water/ethanol)
kPa-1)

PESU 21 15% 60 0.049 99.3% 1111


T-PESU 20 15% 60 0.061 99.4% 1326
PPSU 23 15% 60 0.098 98.6% 552
H-PESU 36 15% 60 0.322 98.8% 685
PI-9-1 19 15% 60 0.161 94.0% 123.9
Polyimide/UiO-
66-NH2 MMM- 33 15% 60 0.256 96.2% 202.5
10%
Polyimide/UiO-
66-NH2 MMM- 44 15% 60 0.345 94.0% 124.5
20%
Polyimide/UiO-
66-NH2 MMM- 21 15% 60 0.602 91.0% 80.3
30%

Based on this table, it can be concluded that:

H-PESU membrane is the most promising candidates for commercialization

with the most balanced separation performance with highest water

permeability and good selectivity. For commercialization, asymmetric

membranes are usually preferred to get a higher flux. But in the asymmetric

membrane fabrication, because of the hydrophilic nature, H-PESU may have

difficulty in forming the defect-free selective layer due to its slow non-solvent

induced phase inversion process. Adjusting dope formula such as blending

with other hydrophobic polymers e.g. PESU and adding non-solvent in dope

may be helpful for fast phase inversion and formation of defect-free selective

layer.

172
The other three PAE membranes i.e. PESU, PPSU and T-PESU are high

barrier materials and show high selectivity but low permeability. Therefore,

there is an urgent need to improve the permeability of these membranes.

Hydrophilic surface modification would be a worthwhile method to enhance

the performance. Based on solution-diffusion model, permeability depends on

solubility and diffusivity. One method to improve solubility of the membranes

is to increase the hydrophilicity of the membrane surface. One can incorporate

hydrophilic polymer chains like poly(4-vinylbenzoic acid) on the top of the

highly water selective membrane as shown in Figure 8.1. The membrane is

first treated with ozone for a certain time then surface initiated radical

polymerization will happen after adding 4-vinylbenzoic acid. Finally, a

hydrophilic layer will be formed on the top layer. The hydrophilic polymer

chains will increase the sorption of the water on the membrane surface and the

PAE layer can provide high selectivity so as to increase its water permeability

without sacrificing membrane selectivity. Besides hydrophilic surface

modification method, coating PAE as an ultrathin selective layer on top of

porous support to make composite membrane is also an effective way to

improve membrane flux. The porous support and the ultrathin selective layer

can greatly reduce the transport resistance and contribute to high flux while

the highly water-selective PAE layer would make sure the good separation

factor.

Figure 8.1 Hydrophilic surface modification

173
PI membrane shows moderate permeability and selectivity. On one hand, to

improve its permeability, the same hydrophilic surface modification method

presented in Figure 8.1 is also applicable for PI membrane. Besides this way,

hydrophilic surface modification of PI membranes can also be achieved by

directly reacting with some chemicals with both hydrophilic functional groups

and amine groups like β-alanine and ethanolamine. Structures of these

chemicals are shown in Figure 8.2. Polyimide can react with –NH2 in these

chemicals to form more hydrophilic polyamide layer. And the hydrophilic

functional groups in these chemicals can help further increase the water

solubility on the membrane surface, thus improve the permeability. On the

other hand, to enhance the selectivity of PI membrane, crosslinking is still

needed.

Figure 8.2 Structures of four hydrophilic chemicals

UiO-66-NH2 based MMMs also show a great potential for further

development with high permeability and moderate selectivity. However, at

the condition of very high particle loading, particles may agglomerate and

voids may form at the particle-polymer interface. The MMMs become brittle

and difficult for handling and testing. Adding modulator with amine groups

e.g. 4-aminobenzoic acid, may be a good approach to solve this problem. The

174
presence of amine groups on the modulators may create a chemical reaction

between the MOF outer surface and the imide group of the polymer, thus

increase the compatibility between MOF particle and polyimide and result in a

stable membrane.

Last, all the developed membranes in this these are used for dehydration of

alcohols via pervaporation. In the future, one can also extend the application

of the developed membranes to dehydration of other solvent or other

separation systems. And also, besides pervaporation, these membranes may

also be applied for vapor permeation because vapor permeation is very similar

to pervaporation but it has higher the operation temperature than pervaporation

and the feed is in vapor phase. A higher temperature is beneficial for a higher

flux. Therefore, vapor permeation may reduce the required membrane size for

a certain separation. Membranes with good thermal-stability may be suitable

for vapor permeation.

175
BIBLIOGRAPHY

PUBLICATIONS

[1] Y.M. Xu, N.L. Le, J. Zuo, T.S. Chung, Aromatic polyimide and

crosslinked thermally rearranged poly (benzoxazole-co-imide) membranes for

isopropanol dehydration via pervaporation, J. Membr. Sci. 499 (2016) 317-

325.

[2] Y.M. Xu, T.S. Chung, High-Performance UiO-66/Polyimide Mixed

Matrix Membranes for Ethanol, Isopropanol and n-Butanol Dehydration via

Pervaporation, J. Membr. Sci. 531 (2017) 16-26.

[3] Y.M. Xu, Y.P. Tang, T.S. Chung, M. Weber, C. Maletzkoc, Polyarylether

membranes for dehydration of ethanol and methanol via pervaporation, Sep.

Purif. Technol. 193 (2018) 165-174.

[4] Y.M. Xu, S. Japip, T.S. Chung, Mixed Matrix Membranes with nano-sized

functional UiO-66-type MOFs embedded in 6FDA-HAB/DABA Polyimide

for Dehydration of C1-C3 Alcohols via Pervaporation, J. Membr. Sci. (2018)

In Press.

CONFERENCE PROCEEDINGS

[1] Y.M. Xu, N.L. Le, J. Zuo, T.S. Chung, Aromatic polyimide and

crosslinked thermally rearranged poly (benzoxazole-co-imide) membranes for

isopropanol dehydration via pervaporation, 26th NAMS Annual Meeting,

Bellevue, WA USA, May 21-25, 2016.

176

You might also like