243
A thesis submitted for the degree of Doctor of Philosophy at The University of Queensland in 2014 School of Chemical Engineering Morphological, mechanical and gas transport properties of stainless steel and composite hollow fibres. Diego Ruben Schmeda Lopez Mechanical Engineer, MSc.

Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

  • Upload
    others

  • View
    3

  • Download
    0

Embed Size (px)

Citation preview

Page 1: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

A thesis submitted for the degree of Doctor of Philosophy at

The University of Queensland in 2014

School of Chemical Engineering

Morphological, mechanical and gas transport properties of stainless steel and

composite hollow fibres.

Diego Ruben Schmeda Lopez

Mechanical Engineer, MSc.

Page 2: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

ii

ABSTRACT

This work focuses on the development of stainless steel (SS) and SS composite hollow fibres (with

carbon (CSS) and alumina (CASS)) as novel, robust inorganic membranes prepared via dry-wet

phase inversion from a spinning dope containing a mixture of SS and/or alumina particles,

polymeric binders and solvents. The morphological features and mechanical properties of the

hollow fibres were evaluated by varying the spinning dope composition, including: different

binders; SS-to-alumina volume, binder-to-solvent, and binder-to-SS particle ratios; the spinning

dope viscosity; and the SS particle size (6-45m); as well as sintering parameters such as

temperature, atmosphere and time.

It was found that the addition of larger particles to the spinning dope favoured the creation of large

macrovoids in both the inner and outer shells of the hollow fibre. In contrast, small particles delayed

de-mixing, forming sponge-like regions at the outer shell, although the finger-like macrovoids were

retained at the lumen. Polyvinylpyrrolidone (PVP) was used as a viscosity enhancer, altering the

kinetics of the phase inversion process, leading to an increase in finger-like macrovoids with

increasing PVP addition. Polyetherimide (PEI) was preferred as the polymeric binder due to its

favourable phase inversion kinetics. Conversely, polyethersulfone promoted faster de-mixing

resulting in finger-like macrovoids at both surfaces.

The morphological structure of the sintered SS hollow fibres did mimic the morphology of the

green fibres. Sintering was controlled by SS mass diffusion limitations, lower densification was

achieved at 950°C with necks forming between particles in close contact. At 1000°C, surface

diffusion became important to densification. At 1050 and 1100°C, smaller pores in the sponge-like

region started closing whilst larger finger-like pores and macrovoids remained open as the inter-

particle space was too wide to be filled by surface diffusion. Densification, as function of mass

diffusion, was accelerated for smaller particles and retarded for larger particles, showing an inverse

impact in the total surface area available for diffusion. Fibres with SS particle loadings below

50wt% showed irregular geometries. However at 70wt% particle loading the required particle

packing condition was achieved to form inter-particle necks during sintering.

The mechanical resistance of the hollow fibres was closely related to morphology and densification.

Samples with high porosity showed low mechanical strength, especially for porosity related to

finger-like macrovoids, whereas smaller pores (sponge-like region) resulted in higher mechanical

Page 3: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

iii

strengths. Higher densification led to stronger hollow fibres, due to bold necks that could withstand

higher loads.

CSS hollow fibres were created by pyrolysing the polymeric binder during the sintering process,

resulting in inter-particle pore filling with the degraded and retained carbon. The morphology of the

CSS hollow fibres resembles the green fibres after the sintering/pyrolysis process, though some

differences appeared. The finger-like structure in the lumen disappeared due to carbon filling the

pores and/or densification. However, round macrovoids in the middle wall of the CSS hollow fibres

remained as they were too large to close. CSS hollow fibres made from the 6m SS particles

demonstrated a quasi-bi-modal pore size distribution dominated by sponge-like structures at both

surfaces. Increasing the particle size resulted in a multi-modal pore size distribution, with 45m SS

particles yielding a high porosity and larger macrovoids.

Incorporating alumina particles into the spinning dope resulted in the formation of a bi-modal pore

size distribution, independent of the alumina/SS particle ratio. The large pores were predominantly

finger-like structures at the lumen, indicating that the smaller alumina particles (0.5m) densified

less than the CSS fibres. Additionally, increasing the amount of alumina allowed the formation of a

closely packed carbon-alumina region predominantly within the sponge-like region of the hollow

fibre. The mechanical properties of the CSS and CASS were affected by the fact that carbon and

carbon/alumina inhibited sintering, thus a reduction in both strength and flexibility of the CSS and

CASS were obtained when compared to SS hollow fibres.

Finally, the hollow fibres were tested for single gas permeation (20-100°C), and binary gas

mixtures (75-150°C). Adsorption of nitrogen and carbon dioxide was investigated to understand the

synergistic effect of the composite material on the transport of gases. This work shows that CSS

hollow fibres did not adsorb N2. The isosteric heat of adsorption for CO2 was much higher for the

hollow fibres containing only SS as opposed to alumina particles. It was observed that the CSS

hollow fibres separated N2 from CO2 at low feed concentrations (less than 20/80 CO2 to N2). This

was attributed to the strong interaction of CO2 with the surface of the CSS hollow fibre, whilst the

non-adsorbable N2 flowed unimpeded through the pores. This behaviour is anomalous for

mesoporous or macroporous materials such as the CSS hollow fibres, and only possible due to the

synergistic effect of the SS and CO2 materials. This separation effect was not observed in the

composite fibres containing alumina, as N2 adsorption was noticeable resulting in high heat of

adsorption.

Page 4: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

iv

DECLARATION BY AUTHOR

This thesis is composed of my original work, and contains no material previously published or

written by another person except where due reference has been made in the text. I have clearly

stated the contribution by others to jointly-authored works that I have included in my thesis.

I have clearly stated the contribution of others to my thesis as a whole, including statistical

assistance, survey design, data analysis, significant technical procedures, professional editorial

advice, and any other original research work used or reported in my thesis. The content of my thesis

is the result of work I have carried out since the commencement of my research higher degree

candidature and does not include a substantial part of work that has been submitted to qualify for

the award of any other degree or diploma in any university or other tertiary institution. I have

clearly stated which parts of my thesis, if any, have been submitted to qualify for another award.

I acknowledge that an electronic copy of my thesis must be lodged with the University Library and,

subject to the General Award Rules of The University of Queensland, immediately made available

for research and study in accordance with the Copyright Act 1968.

I acknowledge that copyright of all material contained in my thesis resides with the copyright

holder(s) of that material. Where appropriate I have obtained copyright permission from the

copyright holder to reproduce material in this thesis.

Page 5: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

v

PUBLICATIONS DURING CANDIDATURE

Peer-reviewed papers

Zhang, Xiwang, David K. Wang, Diego Ruben Schmeda Lopez and João C. Diniz da Costa.

"Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for

Wastewater Treatment." Chemical Engineering Journal 236, no. 0 (2014): 314-322.

Conferences

Schmeda-Lopez, D.R.; Smart, S.K.; Meulenberg, W.; Diniz da Costa, J.C.; Stainless steel hollow

fibres. Early Career Researchers – Membrane Symposium Australia, 2011.

Schmeda-Lopez, D.R.; Smart, S.K.; Meulenberg, W.; Diniz da Costa, J.C.; Metallic hollow fibre

membranes. International Conference on Inorganic Membranes, 2012.

Schmeda-Lopez, D.R.; Smart, S.K.; Meulenberg, W.; Diniz da Costa, J.C.; Effect of particle size in

the production of stainless steel hollow fibres. Early Career Researchers – Membrane Society of

Australasia, 2012.

Zhang, X; Wang, D; Schmeda-Lopez, D.R.; Diniz da Costa, J.C.; Fabrication of nanostructured

TiO2 hollow fibre membrane and application for water purification. Early Career Researchers –

Membrane Society of Australasia, 2012.

PUBLICATIONS INCLUDED IN THIS THESIS

No publications included.

Page 6: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

vi

CONTRIBUTIONS BY OTHERS TO THIS THESIS

Contributions were made by Prof. Joe da Costa and Dr. Simon Smart in experiment design, concept,

analysis, interpretation, drafting, and writing in the advisory capacity.

STATEMENT OF PARTS OF THE THESIS SUBMITTED TO QUALIFY

FOR THE AWARD OF ANOTHER DEGREE

None

Page 7: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

vii

ACKNOWLEDGEMENTS

I would like to acknowledge and thank all those that in some way helped and assisted me through

my PhD, in special to:

My wife Ana, for her support, partnership, understanding, love and patience.

My family for the continued love, support and understanding even through the big distance.

My advisors Professor Joe Diniz da Costa and Dr Simon Smart for their guidance, support

and patience during the course of the project.

All other past and present members of FIMLabs for their assistance, friendship and support,

in special to: Adi Darmawan, Aida Zubir, Ben Ballinger, Christelle Yacou, Dana Martens,

David Wang, Gianni Olguin, Guozhao Ji, Ingrid Song, Julius Motuzas, Liang Liu, Patrick

Haworth, Shengnan (Alice) Wang, Xiwang Zhang, Xuechao Gao and Yen Chua.

Dr Wilhelm Meulenberg, Mr Martin Bitzer, Dr Martin Bram, Dr. Sebold and all the crew at

the IEK1 in the Forschungszentrum Julich that helped me with the experimental and

discussion during my stay in Germany.

Dr Grant Edwards for his help getting me access to the mechanical measurements setup.

Mr David Page for his help and discussion over mercury porosimetry.

Dr Wade Martens for his help understanding experimental results.

The staff in the workshops for their high quality workmanship and willingness to help.

The administrative, building and IT staff in Chemical Engineering for their assistance.

The Queensland Government for the financial aid through the NIRAP program.

The University of Queensland for the scholarship and the travel grant.

Page 8: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

viii

KEYWORDS

stainless steel hollow fibres, porous metallic membrane, composite membrane, dry-wet phase

inversion, gas separation

AUSTRALIAN AND NEW ZEALAND STANDARD RESEARCH

CLASSIFICATIONS (ANZSRC)

ANZSRC code: 090404, Membrane and Separation Technologies, 55%.

ANZSRC code: 091202 Composite and Hybrid Material, 15%.

ANZSRC code: 091207 Metals and Alloy Materials, 30%.

FIELDS OF RESEARCH (FOR) CLASSIFICATION

FoR code: 0904, Chemical Engineering, 70%.

FoR code: 0912 Materials Engineering, 30%.

Page 9: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

ix

TABLE OF CONTENTS

Abstract ............................................................................................................................................... ii

Introduction ...................................................................................................................................... 22

1.1. Background ......................................................................................................................... 22

1.2. Scope and research contributions ........................................................................................ 24

1.3. Structure of the thesis .......................................................................................................... 25

1.4. References ........................................................................................................................... 27

Literature Review ............................................................................................................................ 32

2.1. Abstract ............................................................................................................................... 32

2.2. Membrane supports ............................................................................................................. 32

2.3. Polymeric membrane supports through phase inversion ..................................................... 35

2.4. Polymeric hollow fibre production process......................................................................... 41

2.5. Effect of solid particles on the hollow fibre production process ......................................... 44

2.6. Sintering of porous solids .................................................................................................... 45

2.6.1. The Raw Powder .......................................................................................................... 45

2.6.2. Sintering Mechanism ................................................................................................... 46

2.6.3. Effect of sintering parameters ...................................................................................... 52

2.7. Concluding Remarks ........................................................................................................... 54

2.8. References ........................................................................................................................... 55

Experimental .................................................................................................................................... 64

3.1. Introduction ......................................................................................................................... 64

3.2. Materials .............................................................................................................................. 64

3.3. Viscosity measurements ...................................................................................................... 65

Page 10: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

x

3.4. Preparation of membrane precursors ................................................................................... 65

3.5. Sintering .............................................................................................................................. 67

3.6. Thermogravimetric Analysis (TGA) ................................................................................... 69

3.7. Helium pycnometry ............................................................................................................. 69

3.8. Nitrogen adsorption ............................................................................................................. 69

3.9. Mercury porosimetery ......................................................................................................... 70

3.10. Scanning Electron Microcopy (SEM) ................................................................................. 70

3.11. Computerized X-ray microtomography (µ-CT) .................................................................. 71

3.12. Mechanical Strength Tests. ................................................................................................. 71

3.13. Inductively-Coupled Plasma – Optical Emission Spectrometry Analysis .......................... 72

3.14. Permeation tests ................................................................................................................... 73

3.14.1. Single gas permeation tests ...................................................................................... 73

3.14.2. Binary gas permeation tests ...................................................................................... 74

3.15. Uncertainty within experiments .......................................................................................... 75

3.16. errors in concentricity of the surfaces in Hollow Fibre Production .................................... 78

3.17. References ........................................................................................................................... 80

Formation of asymmetric hollow fibre produced via phase inversion of a stainless steel /

polymer / solvent / non-solvent system. .......................................................................................... 83

4.1. Abstract ............................................................................................................................... 83

4.2. Introduction ......................................................................................................................... 85

4.3. Choice of binder .................................................................................................................. 87

4.4. Effect of the viscosity of the spinning dope ........................................................................ 92

4.4.1. Polymer to Solvent Ratio ............................................................................................. 92

Page 11: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

xi

4.4.2. Viscosity Modifier ....................................................................................................... 96

4.5. Effect of the addition of particles to the spinning dope .................................................... 100

4.5.1. Particle Loading ......................................................................................................... 101

4.5.2. Particle Size................................................................................................................ 105

4.6. General Discussion ............................................................................................................ 108

4.7. Conclusion ......................................................................................................................... 114

4.8. References ......................................................................................................................... 115

Characterisation of the stainless steel hollow fibres produced via phase inversion and the

effect of sintering parameters ....................................................................................................... 122

5.1. Abstract ............................................................................................................................. 122

5.2. Introduction ....................................................................................................................... 124

5.3. Sintering Temperature ....................................................................................................... 125

5.4. Sintering Atmosphere ........................................................................................................ 130

5.5. Dwelling Time ................................................................................................................... 136

5.6. Stainless Steel Particle Size ............................................................................................... 138

5.7. Polyetherimide to Solvent Ratio........................................................................................ 142

5.8. Viscosity Modification with Polyvinylpyrrolidone ........................................................... 144

5.9. Polyetherimide vs Polyethersulfone .................................................................................. 147

5.10. Particle Loading ................................................................................................................ 152

5.11. Discussion of Morphology Effects .................................................................................... 155

5.12. Discussion of Mechanical Effects ..................................................................................... 162

5.13. Conclusions ....................................................................................................................... 172

5.14. References ......................................................................................................................... 173

Page 12: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

xii

Production and characterisation of composite carbon stainless steel, composite carbon

alumina hollow fibres and carbon alumina hollow fibres. ......................................................... 180

6.1. Abstract ............................................................................................................................. 180

6.2. Introduction ....................................................................................................................... 182

6.3. Carbon Stainless Steel (CSS) Hollow Fibres .................................................................... 183

6.4. Carbon Alumina Stainless Steel (CASS) Hollow Fibres .................................................. 189

6.5. Discussion ......................................................................................................................... 195

6.5.1. 6.4.1 Morphological effects of pore filling ................................................................ 195

6.5.2. Mechanical properties of pore filling ......................................................................... 201

6.6. Conclusion ......................................................................................................................... 205

6.7. References ......................................................................................................................... 207

Gas testing of the composite hollow fibres ................................................................................... 211

7.1. Abstract ............................................................................................................................. 211

7.2. Introduction ....................................................................................................................... 212

7.3. Transport Phenomena in Porous Media ............................................................................ 212

7.3.1. Adsorption .................................................................................................................. 216

7.4. Single gas permeation for the CSS, CASS and CA Hollow Fibres .................................. 219

7.4.1. Modelling ................................................................................................................... 219

7.4.2. Results ........................................................................................................................ 223

7.5. Binary gas test for the CSS Hollow Fibres........................................................................ 227

7.5.1. Long term testing ....................................................................................................... 231

7.6. Discussion ......................................................................................................................... 231

7.7. Conclusion ......................................................................................................................... 237

7.8. References ......................................................................................................................... 238

Page 13: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

xiii

Conclusions and recommendations .............................................................................................. 241

8.1. Conclusions ....................................................................................................................... 241

8.2. Recommendations ............................................................................................................. 242

Page 14: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

14

LIST OF FIGURES

Figure 2.1 - Different shapes of membrane supports (a) Tubular porous alumina substrates, (b)

tubular porous metallic substrates, (c) porous ceramic hollow fibres, (d) porous metallic discs, (e)

porous metallic platelets, from [2]. .................................................................................................... 33

Figure 2.2 – Schematic view of three membrane modules [3]. ......................................................... 34

Figure 2.3 – Composition paths of a cast film immediately after immersion demonstrating

instantaneous demixing (a); and delayed demixing (b), adapted from [14]. ..................................... 36

Figure 2.4 – SEM image of a typical sponge-like structure (a) and a typical finger-like structure (b),

adapted from [17]. .............................................................................................................................. 37

Figure 2.5 – Different membrane morphologies obtained by different demixing paths. ................... 39

Figure 2.6 – Effect of parameters in the kinetics of the phase inversion process. ............................. 41

Figure 2.7 - Cross-section of the tube-in-orifice spinneret (a); Scheme of the process of extrusion of

the Hollow fibre (b)............................................................................................................................ 42

Figure 2.8 - Coagulation phenomena in the hollow fibre production process. .................................. 43

Figure 2.9 - Scheme of various types of sintering, adapted from [67]. ............................................. 47

Figure 2.10 - Basic phenomena occurring during sintering. .............................................................. 48

Figure 2.11 - Scheme of densification stages by sintering process [68-70]. ..................................... 49

Figure 2.12 - Sintering model of two particles: (a) shows a schematic representation of bonded

particles; x, y, a and ρ indicates neck radius, interpenetrated depth, particle radius, and radius of

neck surface, respectively. (b) shows the different diffusion paths during the initial stage of

sintering. ............................................................................................................................................. 49

Figure 2.13 - Effect of sintering parameters on densification. ........................................................... 52

Figure 3.1 – Schematics of the hollow fibre production rig. ............................................................. 66

Figure 3.2 – Schematic of sintering furnace ...................................................................................... 67

Figure 3.3 – Three point bending test ................................................................................................ 71

Figure 3.4 – Gas permeation rig setup ............................................................................................... 74

Figure 3.5 – Binary gas permeation rig.............................................................................................. 75

Figure 3.6 – Schematics of the spinning process used to produce hollow fibres. ............................. 79

Figure 3.7 – Schematics showing the cause of errors of concentricity in the extrusion process of a

hollow fibre, (a) caused by tilting the spinneret and (b) caused misalignment between the non-

solvent injection needle and the spinneret hole. ................................................................................ 80

Figure 4.1 – Viscosity results as a function of rotational speed for PESf (circle) and PEI (triangle)

dissolved in NMP in the ratio of 1:3 (closed symbols) and the same mixtures including stainless

steel particles (open symbols). ........................................................................................................... 88

Figure 4.2 – Morphology resulting after sintering of the stainless steel hollow fibre prepared with

PESf (A) and PEI (B). ........................................................................................................................ 89

Figure 4.3 – Mercury porosimetry results for the fibres prepared using PESf and PEI polymers. ... 91

Figure 4.4 – Viscosity of the spinning dope with PEI / NMP ratios of 1:3 (solid circles) and 1:4

(hollow circles) as a function of stainless steel loading (6μm sized particles) .................................. 93

Page 15: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

15

Figure 4.5 – SEM images of green fibres produced using spinning dopes containing stainless steel

and polymer to solvent ratios of 1:3 (a) and 1:4 (b) ........................................................................... 94

Figure 4.6 – Mercury porosimetry (a), porosity distribution (b) and maximum porosity (c) of the

hollow fibres containing 6μm stainless steel particles prepared using polymer to solvent ratios of

1:3 and 1:4. ......................................................................................................................................... 95

Figure 4.7 – Spinning dope viscosity as a function of PVP content. ................................................. 97

Figure 4.8 – SEM images of the morphology resulting of the phase inversion process of samples

produced using polymer to solvent ratio of 1:3 and contain 0% PVP (A), 0.5% PVP (B),, 1% PVP

(C), and 1.5% PVP (D). ..................................................................................................................... 97

Figure 4.9 – SEM images of the morphology resulting of the phase inversion process of samples

produced using polymer to solvent ratio of 1:4 and contain 0% PVP (A), 0.5% PVP (B),, 1% PVP

(C), and 1.5% PVP (D). ..................................................................................................................... 98

Figure 4.10 – Total porosity as determined by mercury porosimetry (a) porosity distribution of

nascent hollow fibres with 0.5% (b) 1% (c) and 1.5% (d) PVP added as a viscosity enhancer. ....... 99

Figure 4.11 – Viscosity as a function of particle loading for particles of different sizes ................ 102

Figure 4.12 – Morphology resulting from spinning hollow fibres using slurry with particle size of

6µm and loading of 10% (A), 50% (B) and 70% (C) ..................................................................... 103

Figure 4.13 – Mercury porosimetry (a), porosity distribution (b), maximum porosity (c) and average

pore radius (d) of samples produced using different particle loading. ............................................. 104

Figure 4.14 – Viscosity of solutions containing stainless steel particles of 6, 16 and 45 μm particle

diameter. ........................................................................................................................................... 105

Figure 4.15 – Green hollow fibres produced with stainless steel particles of 6µm (A), 10µm (B),

16µm(C) and 45µm(D). ................................................................................................................... 106

Figure 4.16 – Mercury porosimetry (a), porosity distribution (b) and maximum porosity (c) of green

hollow fibres made using particles of different sizes. ...................................................................... 107

Figure 4.17 – Schematic of the phase inversion process of low(A, B) and high (C, D) viscosity

polymer. ........................................................................................................................................... 110

Figure 4.18 – Relative effect of particle size in the solvent – non-solvent exchange velocity. ....... 114

Figure 5.1 - Surface morphology of the hollow fibres sintered at temperatures various temperatures

(a) 950°C, (b)1000°C, (c) 1050°C, and (d) 1100°C for 1 hour. ....................................................... 126

Figure 5.2 – (a) Porosity and (b) pore size distribution. .................................................................. 127

Figure 5.3 - Nitrogen Permeation results ......................................................................................... 128

Figure 5.4 - Effect of sintering temperature on the ultimate flexural stress (a) and flexural strain(b).

.......................................................................................................................................................... 130

Figure 5.5 – Light Microscope pictures of samples produced with AISI316L 10µm and sintered at

1050°C for 1 hour in a) Nitrogen atmosphere b) Argon atmosphere. ............................................. 130

Figure 5.6 – Cumulative and frequency porosimetry distributions for samples sintered at 950°C (a,

b) and 1050°C (c, d) in Argon and Nitrogen.................................................................................... 131

Figure 5.7 –SEM image of a 316L 10m sample sintered in (a) Argon and (d) Nitrogen for two

hours. Respective EDX spectra for Argon (b)spectrum 1 and (c) spectrum 2; Nitrogen (e) spectrum

1 and (f) spectrum 2. ........................................................................................................................ 133

Figure 5.8 - Flexural Strength (a) and strain (b) of AISI 316L hollow fibres product of different

sintering atmospheres. ...................................................................................................................... 135

Page 16: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

16

Figure 5.9 - SEM image of a hollow fibre sintered at 1050°C under nitrogen for (a) 1 hour and (b) 4

hours. ................................................................................................................................................ 136

Figure 5.10 – Porosimetry of samples made of particles of 6µm (a), 16µm (b) and 45µm(c) results

as a function of dwelling time for hollow fibres sintered at 1050oC. .............................................. 137

Figure 5.11 - Effect of dwelling time on mechanical strength (a) and strain (b). ............................ 138

Figure 5.12 – CT scans of hollow fibres produced with stainless steel of 6µm (a), 10µm (b), 16µm

(c) and 45µm (d) particle sizes sintered at 1050°C for one hour. .................................................... 139

Figure 5.13 – Comparison of the dimensions of green and sintered (1050°C) hollow fibres, in terms

of diameters (a) and wall thickness (b) ............................................................................................ 140

Figure 5.14 - Mercury porosimetry results comparing hollow fibres produced in function of porosity

(a) and pore distribution (b). ............................................................................................................ 141

Figure 5.15 – Effect of particle size on the ultimate flexural stress (a) and Flexural strain (b). ..... 141

Figure 5.16 - Hollow fibres sintered at 1050°C produced using polymer to solvent ratio of 1:4 (left)

and 1:3 (right) in the spinning dope ................................................................................................. 143

Figure 5.17 - Mercury porosimetry of sintered samples produced using polymer to solvent ratio of

1:4 and 1:3. ....................................................................................................................................... 143

Figure 5.18 - Result of 3 point bending test carried out on fibres produced using polymer ratio of

1:4 and 1:3 as function of maximum bending strength (a) and strain (b) ........................................ 144

Figure 5.19 - SEM images of the hollow fibres produced using PVP in the spinning dope. .......... 145

Figure 5.20 - Mercury porosimetry of samples produced using green fibres containing polymer to

solvent ratios of 1:3 and 1:4 and PVP content from 0.5% (a, d), 1% (b, e) and 1.5% (c, f). ........... 146

Figure 5.21 - Mechanical strength (a) and strain (b) of the hollow fibres produced using 1:3 and 1:4

PEI to NMP ratios and adding PVP as viscosity modifier. .............................................................. 147

Figure 5.22 - TGA results for hollow fibres produced with PEI and PESf carried out under argon

atmosphere. ...................................................................................................................................... 149

Figure 5.23 - Mercury porosimetry results for the fibres prepared using PESf and PEI polymers. 150

Figure 5.24 - Results of 3 point bending test carried out in the hollow fibres. (a) Maximum Flexural

Stress, (b) Maximum bending strain. ............................................................................................... 151

Figure 5.25 - SEM images of the hollow fibres produced using solid loads of 50% (a) and 70% (b).

.......................................................................................................................................................... 153

Figure 5.26 - Mercury porosimetry of samples producing with 50% and 70% particle loading. (a)

Porosity function of pore radius and (b) pore size distribution........................................................ 154

Figure 5.27 - Three point bending test result for sintered hollow fibres containing 50% and 70%

solid loads. (a) Maximum Flexural strength (b) Maximum flexural strain. .................................... 154

Figure 5.28 - Evolution of the stainless steel hollow fibre during sintering process. ...................... 156

Figure 5.29 – Schematic effect of sintering temperature ................................................................. 157

Figure 5.30 – Comparison of neck radius measured with prediction of Eq. 5.2 as a function of

sintering temperature (particle size 6µm). ....................................................................................... 159

Figure 5.31 – Average diffusion coefficient for neck radius showed in Figure 5.30 in function of

small pores (a) and total porosity (b) ............................................................................................... 160

Figure 5.32 – Effect of particle size and sintering time in the wall thickness of the hollow fibres as

predicted by Eq. 5.4. ........................................................................................................................ 162

Figure 5.33 – Global maximal flexural stress of the hollow fibres produced versus total porosity (a)

and small porosity (under 0.5µm)(b) for all parameters studied. .................................................... 163

Page 17: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

17

Figure 5.34 – Maximum flexural stress as function of total porosity and porosity of pores under

0.5µm for: particle size and sintering temperature (a, d); dwelling time, particle loading and

sintering atmosphere (b, e); polymers(c, f) ...................................................................................... 164

Figure 5.35 – Position of the sponge-like region for samples made from 6µm SS particles and

sintered for 1 hour (a) and made from 6µm SS particles and 1.5% PVP and sintered for 1 hour (b)

.......................................................................................................................................................... 167

Figure 5.36 - Global maximal flexural strain of the hollow fibres produced versus total porosity (a)

and small porosity (under 0.5µm) (b) for all parameters studied. ................................................... 169

Figure 5.37 – Maximum flexural strain as function of porosity of pores under 0.5µm for: particle

size and sintering temperature (a); dwelling time, particle loading and sintering atmosphere (b);

polymers(c) ...................................................................................................................................... 170

Figure 5.38 – Schematic on the effect of macrovoids in strain and crack initiation. ....................... 171

Figure 6.1 – Example of the process of production of the carbon composite hollow fibre ............. 183

Figure 6.2 – Picture of the carbon stainless steel produced of 6µm (a), 10µm (b), 16µm (c) and

45µm (d). .......................................................................................................................................... 184

Figure 6.3 – Carbon Stainless steel hollow fibre made with particles of 6µm. ............................... 184

Figure 6.4 – Magnification of SEM images for hollow fibres made of 6µm (a) and 45µm (b)

particles showing the presence of carbon within the matrix. ........................................................... 185

Figure 6.5 – TGA analysis of the carbon stainless steel hollow fibres ............................................ 186

Figure 6.6 – Porosity and pore size distribution of carbon stainless steel hollow fibres produced

with particles of 6µm (a), 10µm (b), 16µm (c) and 45µm(d). ......................................................... 188

Figure 6.7 – Maximum flexural stress(a) and strain (b) that the carbon stainless steel hollow fibres

can stand before breaking as function of particle size. .................................................................... 188

Figure 6.8 – Picture of the hollow fibres obtained with 25 vol% (a), 50 vol% (b), 75 vol% (c) and

100 vol% (d) alumina loading sintered at 1050oC. .......................................................................... 190

Figure 6.9 – Morphology of hollow fibres produced using alumina to SS particle ratios of 25 vol%

(a), 50 vol% (b), 75 vol% (c) and 100 vol% (d) and sintered at 1050°C for 1 hour ........................ 191

Figure 6.10 – TGA results for hollow fibres produced using alumina loadings between 25% and

100% loading in volume. ................................................................................................................. 191

Figure 6.11 – Pore size distribution for samples made using 25 vol% (a), 50 vol% (b), 75 vol% (c)

and 100 vol% (d) alumina. ............................................................................................................... 192

Figure 6.12 - Nitrogen adsorption results for hollow fibres containing alumina particles varying

from 0 to 100 vol% .......................................................................................................................... 193

Figure 6.13 – Surface area as function of the content of alumina. .................................................. 194

Figure 6.14 – Mechanical properties of hollow fibres (a) flexural stress and (b) flexural strain. ... 195

Figure 6.15 – Schematic of the carbon stainless steel hollow fibre sintering / pyrolysis process. .. 196

Figure 6.16 – Overall morphological features of the CSS hollow fibres. ........................................ 198

Figure 6.17 – Morphological features of CASS .............................................................................. 199

Figure 6.18 – Effect of particle loading in the nitrogen adsorption. ................................................ 201

Figure 6.19 – Comparison between (a) flexural stress and (b) flexural strain for the CSS and SS

hollow fibres. ................................................................................................................................... 202

Figure 6.20 – Interaction between alumina, SS and carbon within the matrix of the hollow fibres 204

Page 18: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

18

Figure 7.1 – Adsorption isotherms for CSS (a, d), CASS (b, e) and CA(c, f) for CO2 ................... 217

Figure 7.2 – Arrhenius plot of the Henry’s law constant for CSS, CASS and CA hollow fibres for

CO2. .................................................................................................................................................. 218

Figure 7.3 – CA hollow fibres (a) nitrogen sorption and (b) Arrhenius plot. .................................. 218

Figure 7.4 – Effect of carbon content (a), alumina content (b) and BET surface area (c) on isosteric

heat of adsorption of CO2 ................................................................................................................. 219

Figure 7.5 – Schematic representation of the flux (J) across a membrane of thickness l. ............... 220

Figure 7.6 – Single gas permeation (± 10%) results for CSS hollow fibre for He (a), N2 (b) and CO2

(c) at transmembrane pressures of 1, 1.5, 2 and 2.5 bar. .................................................................. 224

Figure 7.7 – Single gas permeation (± 10%) results for CASS hollow fibre for He (a), N2 (b) and

CO2 (c) at transmembrane pressures of 1, 1.5, 2 and 2.5 bar. .......................................................... 225

Figure 7.8 – Single gas permeation (± 10%) results for CA hollow fibre for He (a), N2 (b) and CO2

(c) at transmembrane pressures of 1, 1.5, 2 and 2.5 bar. .................................................................. 225

Figure 7.9 – Single gas selectivities (± 22%) for the CSS hollow fibres for (a) N2/CO2, (b)He/CO2

and (c) He/N2. .................................................................................................................................. 226

Figure 7.10 – Single gas selectivities (± 22%) for CASS for N2/CO2 (a), He/CO2 (b), He/N2 (c) for

different pressures. ........................................................................................................................... 226

Figure 7.11 – Single gas selectivities (± 22%) for CA for N2/CO2 (a), He/CO2 (b), He/N2 (c) for

different pressures. ........................................................................................................................... 227

Figure 7.12 – Binary gas permeance (±15%) for CO2 (bottom) and N2 (top) as a function of the flue

gas composition for a transmembrane pressure difference of 150 kPa. ........................................... 229

Figure 7.13 – (a) Separation factor for CO2 and N2 and (b) Ratio between permeance of N2 and CO2

as function of feed composition, and feed stream temperature ........................................................ 229

Figure 7.14 – Binary gas separation factor (±20%) for (a) He and N2 and (b) He and CO2 ........... 230

Figure 7.15 – Long-time testing results for Helium permeance (±20%). ........................................ 231

Figure 7.16 – Effect of the (a) total porosity and (b) sponge-like porosity for the all hollow fibre for

gas permeation at 50°C and 150 kPa. .............................................................................................. 233

Figure 7.17 – Purity of N2 in the permeate stream (±20%) ............................................................. 234

Figure 7.18 – Permeance (±10%) of He, CO2 and N2 as a function of time at 100oC. .................... 235

Figure 7.19 – Schematics showing the effect that adsorption has on selectivity. ............................ 236

Page 19: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

19

LIST OF TABLES

Table 2.1- Influences that powder properties has on porous materials [64]. ..................................... 46

Table 2.2 - Path routes for diffusion during sintering [77, 78]. ......................................................... 50

Table 2.3 – Plausible values for the constants in Eq. 2.7 and 2. 8 for the initial stage of sintering.

Ds,Dl,Dgb are diffusion coefficients for surface, lattice, and grain boundary respectively. δs, δgb

are thickness for surface and grain boundary diffusion. γsv, specific surface energy, p0, vapour

pressure, m atomic mass, k Boltzmann constant, T absolute temperature, η viscosity and Ω is the

atomic volume [73, 74, 79-81]. .......................................................................................................... 51

Table 3.1 – Fabrication parameters used in the production of Green Hollow Fibres ........................ 66

Table 3.2 – Fabrication parameters used in the production of hollow fibres in this work ................ 68

Table 3.3– Particle size distribution as reported by the manufacturer ............................................... 76

Table 4.1– Flory interaction parameters for water (1) / solvent (2) / polymer (3) systems[40] ........ 90

Table 5.1 – Chemical Analysis of samples sintered under different atmospheres. ......................... 132

Table 5.2 – Percentage of surface area associated with each region of EDX analysis relative to

sintering atmosphere ........................................................................................................................ 134

Table 5.3 – Chemical composition of the original 316L SS particles and the post-sintered PEI+SS

and PESf+SS hollow fibres. ............................................................................................................. 152

Table 5.4 – Comparison of mechanical strength with literature values ........................................... 167

Table 6.1 – TGA mass loss of CSS hollow fibres. .......................................................................... 186

Table 6.2 – Composition of the green fibre obtained after spinning in vol% (left) and wt% (right)

.......................................................................................................................................................... 190

Table 6.3 – Comparison of mechanical strength values available in literature. .............................. 205

Table 7.1 – Hollow fibre dimensions. .............................................................................................. 223

Page 20: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

20

LIST OF ABBREVIATIONS

Ea : Apparent activation energy

µ : Absolute viscosity

m : Atomic mass

Ω : Atomic volume

σ : Bending stress

k : Boltzmann constant

CA : Carbon alumina

CASS : Carbon alumina stainless steel

CSS : Carbon stainless steel

α : Coefficient that depends on the sintering mechanism

dq/dl : Concentration gradient

x/a : Densification rate

Ρ : Density

Dgb : Diffusion coefficients for grain boundary

Dl : Diffusion coefficients for lattice

Ds : Diffusion coefficients for surface

EDX : Energy dispersive X-Ray difraction

J : Flux

xi : Fraction of component i

KH : Henry's constant

H : hours

ICP/OES : Inductively-Coupled Plasma – Optical Emission Spectrometry

Di : Inner diameter of the hollow fibre

Χ : Interaction parameter

Y : Interpenetrating depth

Qst : Isosteric heat of adsorption

Sx/y : Knudsen selectivity

F : Load

m : mass flow

δm : Membrane thickness

Em : Mobility energy

MW : Molecular weight

λ : Molecule mean free path

x : Neck radius

NMP : N-Methyl-2-pyrrolidinone

n : Number of moles

m : Numerical exponent dependent on the mechanism of sintering

n : Numerical exponent dependent on the mechanism of sintering

Do : Outer diameter of the hollow fibre

a : Particle radius

Page 21: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

21

(P/L) : Permeance

FP,0 : Poiseuille flow

PEI : Polyetherimide

PESf :

Polyethilene sulfone

PVP : Polyvinylpryyolidone

dp : Pore diameter

r :

Pore radius

εp; ε : Porosity

X : Quantity adsorbed

ρ : Radius of neck surface

η : Relative viscosity

λ : Scaling constant

SEM : Scanning Electron Microscopy

α : Separation factor

K : Span

γsv : Specific surface energy

γ : Specific surface energy

SS : Stainless steel

A : Surface area

Ds : Surface diffusion coefficient

T : Temperature

TGA : Thermogravimetric Analysis

δgb : Thickness for grain boundary diffusion

δs : Thickness for surface diffusion

H : Thickness of bed of spheres

L : Thickness of the porous layer

t : Time

τ : Tortuosity

ΔP : Transmembrane pressure

R : Universal gas constant

H : Value of a function depending on geometrical and material parameters of the powder and sintering mechanism

p0 : Vapour pressure

U : Velocity of fluid across the bed of spheres

V : Volume

φ : Volume fraction

ν : Volume fraction in the coagulation bath

σys : Yield stress of the original material

σpl : Yield stress of the porous solid

Page 22: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

22

CHAPTER 1

INTRODUCTION

1.1. BACKGROUND

Membranes are semi-permeable barriers that allow passage only to selected components. Nowadays

they are commonplace and widely used in the medical, desalination, water treatment, food and oil

and gas industries [1-4]. The most common material used to produce membranes are polymers [5].

The first polymeric membranes consisted of a thick, dense layer of polymer which gave extremely

low fluxes and were of limited use [6]. Loeb and Sourirajan [2] proposed a novel solution to this

issue, by developing an asymmetric membrane structure consisting of a thin selective layer on a

thick supportive sub-layer. By adopting this strategy the productivity of the membrane was largely

increased. In addition, the supportive sub-layer provided adequate mechanical strength without

interfering with the flow. The process developed by Loeb and Sourirajan [2], known as dry-wet

phase inversion, was largely dependent on the thermodynamic instabilities that exist in a ternary

system composed of solvent – polymer – non-solvent. Studied extensively by Strathmann and

Kock [7] and by Koenhen et al. [8], they conclude that the formation of the asymmetric structure

was due mainly to differences in the kinetics of the phase inversion process as mediated by the

presence of the non-solvent [9]. A wide range of polymers have been used to produce

commercially-relevant, selective membranes, including cellulose acetate, polysulfone,

polyethersulfone, poly(4-methyl-1-pentene) and polyetherimide, while employing solvents such as

dioxane, acetic acid, triethyl phosphate, N-N-dimethylacetamide, and N-Methyl-2-pyrrolidinone [9-

23]. However, these polymeric membranes tend to fail when exposed to high temperatures,

corrosive environments and high mechanical stresses [24-26]. Inorganic membranes, first developed

in the 1940s to separate isotopes of uranium using gaseous diffusion [27, 28], offer a potential

solution to these stability issues; though it was not until 1980s and 1990s that inorganic membranes,

particularly ceramics and dense metallic membranes, were heavily studied for other applications

[27, 29].

Page 23: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

23

A natural evolution of the work of Loeb and Sourirajan [2] was achieved when Tan et al. [30]

mixed particles to a solution of polymer and solvent, inducing phase inversion by exposing to a

non-solvent; a procedure similar to how asymmetric polymeric membranes are produced. This

process resulted in an asymmetric polymeric hollow fibre membrane that contained high loads of

solid ceramic particles. The polymer was burnt off and the particles were sintered at high

temperature resulting in a ceramic hollow fibre where most of the asymmetric characteristics were

preserved. Alumina, titania and perovskites are the most common materials used to produce

ceramic membranes by this combined phase inversion and sintering technique [31-41]. However,

the use of ceramic membranes is not without pitfalls, including relatively low mechanical strength

and brittleness. With these drawbacks in mind, porous metal membranes have enjoyed moderate

recent attention due to their high mechanical strength. A few initial works have been reported for

porous nickel membranes [42], flat stainless steel substrates [43-45], symmetric [46] and

asymmetric stainless steel hollow fibres [47, 48].

Current evidence suggests that despite the inclusion of particles to the spinning dope (mixture of

polymer, solvent and particles), the resultant morphology is very similar to that which would have

been observed in a polymer-only system. Indeed particle addition induces changes in the kinetics of

the phase inversion process [49]. Previous works have primarily studied the effects of the air-gap,

viscosity and non-solvent concentration on the hollow fibre morphology, where the viscosity of the

spinning dope was found to be the dominating factor for ceramic hollow fibre morphology. While

some of these studies suggest that altering parameters such as particle size, particle loading,

sintering temperature and time affect the final hollow fibre morphology, no definitive relationship

or in-depth study has been conducted to determine the relations between those parameters and the

characteristics of the final hollow fibre.

In this thesis the development of asymmetric stainless steel hollow fibres derived from

polyetherimide (PEI) produced by combined phase inversion and various sintering methodologies

were explored with the aim of obtaining a strong metallic membrane or membrane support. To this

end, a mixture of stainless steel (SS) containing a polymeric solution of PEI and N-Methyl-2-

pyrrolidone (NMP) was spun via phase inversion using water as the non-solvent. Controlling and

predicting the resultant morphology was of high importance, thus the initial part of this work

focused on the preparation of the spinning dope. Specifically, the effect of particle size, particle

loading, polymer / solvent ratio, different polymers and additives on the morphology obtained after

phase inversion was investigated. A definitive correlation between these parameters and the ratio

between large macrovoids and sponge-like region was achieved.

Page 24: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

24

The second part of this work focused on the impact of sintering parameters on the final properties of

the stainless steel hollow fibre. Properties like tailored pore sizes and high mechanical strength are

desirable in the membrane industry. With this in mind the hollow fibres produced in the first part of

this work were sintered under different sintering atmospheres, at different sintering times and

temperatures, in addition to the variables used to produce the hollow fibre precursor. The sintered

hollow fibres were then tested for the subsequent morphological and mechanical characteristics.

The result was a panorama of ways that sintering can modify the hollow fibre morphology produced

during the phase inversion stage, and relations were postulated for these parameters.

In the final part of this work a novel composite hollow fibre was produced and tested. In this case

the green stainless steel hollow fibres produced earlier in the thesis were thermally altered in a

combined pyrolysis / sintering process in an inert atmosphere. This allowed for large macrovoids

between SS particles of the hollow fibre matrix to be filled by porous carbon derived from the

polymeric binder. This carbonisation process was adapted to metallic-ceramic composites through

the addition of small alumina particles (i.e. 12 times smaller than SS particles) to the spinning dope,

which resulted in a tri-component composite hollow fibre to explore further the concept of pore

filling. Both the carbon stainless steel and the carbon alumina stainless steel hollow fibres were

tested for morphological and mechanical features. The effect of the carbonisation process on hollow

fibre performance was compared to the pure stainless steel hollow fibres produced in this work. As

an extension of this work, single gas permeation tests were carried out for these novel hollow fibres,

and even binary gas test were carried out for composite membranes made of carbon and stainless

steel.

1.2. SCOPE AND RESEARCH CONTRIBUTIONS

This thesis focuses on the preparation of SS hollow fibres by a combined phase inversion, forming

green fibres. Subsequently, these green fibres were sintered to study the morphological and

mechanical properties, where preparation and sintering methods were analysed to determine

relationships between porosity, pore sizes and mechanical strength. The concept of macrovoid pore

filling was also investigated by the pyrolysis of the polymeric binder to produce a binary carbon-SS

hollow fibre, or by adding small alumina particles to form a ternary carbon-SS-alumina hollow

fibre. The hollow fibres were tested for gas separation and adsorption.

Page 25: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

25

The key research contributions of this thesis are summarised as follows:

The addition of SS particles to the spinning dope altered the kinetics of the phase inversion

process by impeding the solvent / non-solvent exchange when compared to the base

polymer. This yields significant differences in the resultant morphology despite maintaining

a constant volume fraction. In general, smaller particles generate a greater resistance which

results in delayed demixing.

The SS particle size influenced the mechanical properties attributed to mass transfer

limitations during the sintering process.

The sintered hollow fibres did mimic most morphological characteristics obtained during the

phase inversion process, since the densification of the hollow fibre was limited by the

surface diffusion kinetics.

Superior morphological control was achieved by filling inter-particle pores formed by SS

particles with porous carbon retained after the pyrolysis process and by adding small

alumina particles.

It is shown for the first time that N2 / CO2 separation can be achieved by the carbon-SS

hollow fibres, directly attributed the unique property of this material. As a result, the slow

surface transport was preferential by CO2 adsorption whilst the N2 non-adsorptive permeated

unimpeded via fast viscous/Knudsen flow through the carbon-SS hollow fibres.

1.3. STRUCTURE OF THE THESIS

Chapter 1 – Introduction

This chapter provides background information to the thesis work along with the scope and research

contributions to the relevant field.

Chapter 2 – Literature Review

This chapter presents an overview of the formation mechanisms of hollow fibres via phase

inversion and discusses the current state of art for producing asymmetric hollow fibres. The second

half of the chapter focuses on the sintering mechanisms relevant to the production of porous

metallic bodies.

Chapter 3 – Experimental procedures

Page 26: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

26

This chapter details the experimental procedures used to produce, characterise and test the hollow

fibres. The particulars in which the various hollow fibres were spun and later sintered are presented

alongside characterisation techniques like scanning electron microscopy (SEM), helium pycnometry

and mercury porosimetry. Three point bending tests were performed to measure the mechanical

resistance and flexibility of the fibres whilst single and binary gas permeation tests were carried out

to measure the performance of the hollow fibres.

Chapter 4 – Formation of asymmetric hollow fibre produced via phase inversion of a

stainless steel / polymer / solvent / non-solvent system.

In this chapter the effect that SS particle size, particle loading, polymer / solvent ratio, viscosity

modifiers and different polymers have on the formation mechanism and morphology of the green

hollow fibres is elucidated. The various ways and means of producing sponge-like regions or

finger-like macrovoids is discussed from the point of view of the ternary or quaternary phase

equilibria and the solvent / non-solvent exchange kinetics.

Chapter 5 – Characterisation of the stainless steel hollow fibres produced via phase

inversion and the effect of sintering parameters

In this chapter the effect of sintering variables on the morphology and performance of the resultant

hollow fibres is presented. In particular it is shown that the morphology formed in the green fibre

during phase inversion is largely preserved during the de-binding and sintering stages. Densification

was observed to varying degrees and its effect on the mechanical properties of the hollow fibres is

discussed.

Chapter 6 – Production and characterisation of composite carbon stainless steel,

composite carbon alumina hollow fibres and carbon alumina hollow fibres.

In this chapter the production of a novel composite hollow fibre is introduced. A combined sintering

/ pyrolysis process is used to produce hollow fibre that retains significant amounts of amorphous

carbon within the matrix. The technique is adapted to produce a three component (carbon-alumina-

SS) hollow fibre by mixing SS and alumina particles together into the spinning dope. The impact of

incorporating varying amounts of alumina (including a carbon-alumina fibre without SS) on the

mechanical and morphological properties of the hollow fibre is presented.

Chapter 7 – Gas testing of the composite hollow fibres

Page 27: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

27

This chapter explores the gas separation performance of the hollow fibres produced in Chapter 6.

Initially single gas permeation tests were carried followed by binary gas mixtures of CO2/He, N2/He

and CO2/N2. The permeation and adsorption properties of all hollow fibres (SS, carbon-SS, and

carbon-SS-alumina) are analysed in the context of gas separation.

Chapter 8 – Conclusions and recommendations

The concluding findings of the thesis and recommendations for future work are given in this

chapter.

1.4. REFERENCES

[1] Baker RW. Medical Applications of Membranes. Membrane Technology and Applications:

John Wiley & Sons, Ltd; 2004. p. 465-90.

[2] Loeb S, Sourirajan S. Sea Water Demineralization by Means of an Osmotic Membrane. Saline

Water Conversion II: AMERICAN CHEMICAL SOCIETY; 1963. p. 117-32.

[3] Cho J, Amy G, Pellegrino J. Membrane filtration of natural organic matter: factors and

mechanisms affecting rejection and flux decline with charged ultrafiltration (UF) membrane.

Journal of Membrane Science 2000; 164:89-110.

[4] Baker RW, Lokhandwala K. Natural gas processing with membranes: an overview. Industrial &

Engineering Chemistry Research 2008; 47:2109-21.

[5] Bose AC. Inorganic Membranes for Energy and Environmental Applications. New York, NY:

Springer New York; 2009.

[6] Koros WJ, Fleming GK. Membrane-based gas separation. Journal of Membrane Science 1993;

83:1-80.

[7] Strathmann H, Kock K. The formation mechanism of phase inversion membranes. Desalination

1977; 21:241-55.

Page 28: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

28

[8] Koenhen DM, Mulder MHV, Smolders CA. Phase separation phenomena during the formation

of asymmetric membranes. Journal of Applied Polymer Science 1977; 21:199-215.

[9] Altena FW, Smolders CA. Calculation of liquid-liquid phase separation in a ternary system of a

polymer in a mixture of a solvent and a nonsolvent. Macromolecules 1982; 15:1491-7.

[10] Cabasso I, Klein E, Smith JK. Polysulfone hollow fibers. II. Morphology. Journal of Applied

Polymer Science 1977; 21:165-80.

[11] Cabasso I, Klein E, Smith JK. Polysulfone hollow fibers. I. Spinning and properties. Journal of

Applied Polymer Science 1976; 20:2377-94.

[12] Lafreniere LY, Talbot FDF, Matsuura T, Sourirajan S. Effect of poly(vinylpyrrolidone)

additive on the performance of poly(ether sulfone) ultrafiltration membranes. Industrial &

Engineering Chemistry Research 1987; 26:2385-9.

[13] Broens L, Altena FW, Smolders CA, Koenhen DM. Asymmetric membrane structures as a

result of phase separation phenomena. Desalination 1980; 32:33-45.

[14] Wang D, Li K, Sourirajan S, Teo WK. Phase separation phenomena of

polysulfone/solvent/organic nonsolvent and polyethersulfone/solvent/organic nonsolvent systems.

Journal of Applied Polymer Science 1993; 50:1693-700.

[15] Miyano T, Matsuura T, Sourirajan S. EFFECT OF POLYVINYLPYRROLIDONE

ADDITIVE ON THE PORE SIZE AND THE PORE SIZE DISTRIBUTION OF

POLYETHERSULFONE (VICTREX) MEMBRANES. Chemical Engineering Communications

1993; 119:23-39.

[16] Liu T, Zhang D, Xu S, Sourirajan S. Solution-Spun Hollow Fiber Polysulfone and

Polyethersulfone Ultrafiltration Membranes. Separation Science and Technology 1992; 27:161-72.

[17] Miyano T, Matsuura T, Sourirajan S. EFFECT OF POLYMER MOLECULAR WEIGHT,

SOLVENT AND CASTING SOLUTION COMPOSITION ON THE PORE SIZE AND THE

PORE SIZE DISTRIBUTION OF POLYETHERSULFONE (VICTREX) MEMBRANE. Chemical

Engineering Communications 1990; 95:11-26.

Page 29: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

29

[18] Li DF, Chung T-S, Wang R, Liu Y. Fabrication of fluoropolyimide/polyethersulfone (PES)

dual-layer asymmetric hollow fiber membranes for gas separation. Journal of Membrane Science

2002; 198:211-23.

[19] Wang D, Li K, Teo WK. Preparation of Poly(ether sulfone) and Poly(ether imide) Hollow

Fiber Membranes for Gas Separation: Effect of Internal Coagulant. Membrane Formation and

Modification: American Chemical Society; 1999. p. 96-109.

[20] Wang D, Li K, Teo WK. Preparation and characterization of polyetherimide asymmetric

hollow fiber membranes for gas separation. Journal of Membrane Science 1998; 138:193-201.

[21] Wang D, Li K, Teo WK. Polyethersulfone hollow fiber gas separation membranes prepared

from NMP/alcohol solvent systems. Journal of Membrane Science 1996; 115:85-108.

[22] Lai J-Y, Lin F-C, Wang C-C, Wang D-M. Effect of nonsolvent additives on the porosity and

morphology of asymmetric TPX membranes. Journal of Membrane Science 1996; 118:49-61.

[23] Mok S, Worsfold DJ, Fouda AE, Matsuura T, Wang S, Chan K. Study on the effect of spinning

conditions and surface treatment on the geometry and performance of polymeric hollow-fibre

membranes. Journal of Membrane Science 1995; 100:183-92.

[24] Jones CW, Koros WJ. Carbon molecular sieve gas separation membranes-II. Regeneration

following organic exposure. Carbon 1994; 32:1427-32.

[25] Raman NK, Brinker CJ. Organic “template” approach to molecular sieving silica membranes.

Journal of Membrane Science 1995; 105:273-9.

[26] Porter MC. Handbook of Industrial Membrane Technology. William Andrew

Publishing/Noyes; 1990. p. 559-91.

[27] Gillot J. The developing use of inorganic membranes: A historical perspective. Inorganic

Membranes Synthesis, Characteristics and Applications: Springer; 1991. p. 1-9.

[28] Keizer K, Verweij H. Progress in inorganic membranes. Chemtech 1996; 26:37-41.

[29] Uhlhorn RJR, Burggraaf AJ. Gas Separations with Inorganic Membranes. Inorganic

Membranes Synthesis, Characteristics and Applications: Springer Netherlands; 1991. p. 155-76.

Page 30: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

30

[30] Tan X, Liu S, Li K. Preparation and characterization of inorganic hollow fiber membranes.

Journal of Membrane Science 2001; 188:87-95.

[31] Liu S, Tan X, Li K, Hughes R. Preparation and characterisation of SrCe0.95Yb0.05O2.975

hollow fibre membranes. Journal of Membrane Science 2001; 193:249-60.

[32] Liu S, Li K. Preparation TiO2/Al2O3 composite hollow fibre membranes. Journal of

Membrane Science 2003; 218:269-77.

[33] Liu S, Li K, Hughes R. Preparation of porous aluminium oxide (Al2O3) hollow fibre

membranes by a combined phase-inversion and sintering method. Ceramics International 2003;

29:875-81.

[34] Liu S, Gavalas GR. Oxygen selective ceramic hollow fiber membranes. Journal of Membrane

Science 2005; 246:103-8.

[35] Wang H, Werth S, Schiestel T, Caro J. Perovskite Hollow-Fiber Membranes for the Production

of Oxygen-Enriched Air. Angewandte Chemie International Edition 2005; 44:6906-9.

[36] Li K, Tan X, Liu Y. Single-step fabrication of ceramic hollow fibers for oxygen permeation.

Journal of Membrane Science 2006; 272:1-5.

[37] Wang YH, Cheng JG, Liu XQ, Meng GY, Ding YW. Preparation and Sintering of

Macroporous Ceramic Membrane Support from Titania Sol-Coated Alumina Powder. Journal of the

American Ceramic Society 2008; 91:825-30.

[38] García-García FR, Rahman MA, Kingsbury BFK, Li K. Asymmetric Ceramic Hollow fibres:

New micro-supports for gas-phase catalytic reactions. Applied Catalysis A: General 2010; In Press,

Accepted Manuscript.

[39] Leo A, Smart S, Liu S, Diniz da Costa JC. High performance perovskite hollow fibres for

oxygen separation. Journal of Membrane Science 2011; 368:64-8.

[40] Tan X, Liu N, Meng B, Liu S. Morphology control of the perovskite hollow fibre membranes

for oxygen separation using different bore fluids. Journal of Membrane Science 2011; 378:308-18.

Page 31: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

31

[41] Zhang X, Wang DK, Lopez DRS, Diniz da Costa JC. Fabrication of nanostructured TiO2

hollow fiber photocatalytic membrane and application for wastewater treatment. Chemical

Engineering Journal 2014; 236:314-22.

[42] Meng B, Tan X, Meng X, Qiao S, Liu S. Porous and dense Ni hollow fibre membranes.

Journal of Alloys and Compounds 2009; 470:461-4.

[43] Brands K, Uhlmann D, Smart S, Bram M, da Costa JCD. Long-term flue gas exposure effects

of silica membranes on porous steel substrate. Journal of Membrane Science 2010; 359:110-4.

[44] Van Gestel T, Sebold D, Meulenberg WA, Bram M, Buchkremer H-P. Manufacturing of new

nano-structured ceramic-metallic composite microporous membranes consisting of ZrO2, Al2O3,

TiO2 and stainless steel. Solid State Ionics 2008; 179:1360-6.

[45] W. Schafbauer FS-K, S. Baumann, W.A. Meulenberg, N.H. Menzler, H.P. Buchkremer and D.

Stöver. Tape Casting as a Multi Purpose Shaping Technology for Different Applications in Energy

Issues. Materials Science Forum 2012; 706-709:1035-40.

[46] Michielsen B, Chen H, Jacobs M, Middelkoop V, Mullens S, Thijs I, et al. Preparation of

porous stainless steel hollow fibers by robotic fiber deposition. Journal of Membrane Science 2013;

437:17-24.

[47] Luiten-Olieman MWJ, Raaijmakers MJT, Winnubst L, Wessling M, Nijmeijer A, Benes NE.

Porous stainless steel hollow fibers with shrinkage-controlled small radial dimensions. 2011.

[48] Luiten-Olieman MWJ, Winnubst L, Nijmeijer A, Wessling M, Benes NE. Porous stainless

steel hollow fiber membranes via dry–wet spinning. Journal of Membrane Science 2011; 370:124-

30.

[49] Kingsbury BFK, Li K. A morphological study of ceramic hollow fibre membranes. Journal of

Membrane Science 2009; 328:134-40.

Page 32: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

32

CHAPTER 2

LITERATURE REVIEW

2.1. ABSTRACT

This literature review examines the published academic literature related to the production of

hollow fibre membranes and supports. Special focus has been given to the use of the wet spinning,

phase inversion technique as this forms the basis for the hollow fibre production for the thesis.

Studies on the phenomenon of coagulation, phase equilibria and the resulting hollow fibre

morphology are of particular relevance and have been discussed in detail. Finally, the process of

sintering stainless steel porous bodies and related parameters including the effect of particle size,

sintering temperature, time and atmosphere are discussed.

2.2. MEMBRANE SUPPORTS

Membranes are semi-permeable barriers that allow the preferential passage of one or more selected

components. The two key parameters that define membrane performance are flux of components

through the membrane and its selectivity. As membrane thickness is directly proportional to flux,

membrane manufacturers and researchers strive to reduce the thickness of the selective membrane

layer in order to maximise the flow of desired components through it. However, reducing

membrane thickness dramatically increases the risk of mechanical failure during operation.

Membrane supports, alternatively called substrates, allow the thin membrane layer to overcome its

lack of mechanical resistance without interfering significantly with the transport process [1]. This is

particularly true for inorganic membranes where the thin film membrane layer can be brittle and

Page 33: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

33

prone to cracking. These supports can be produced in a variety of shapes and sizes from simple

platelets and discs to porous tubes and hollow fibres [2], as can be seen in Figure 2.1.

Figure 2.1 - Different shapes of membrane supports (a) Tubular porous alumina substrates, (b)

tubular porous metallic substrates, (c) porous ceramic hollow fibres, (d) porous metallic discs, (e)

porous metallic platelets, from [2].

For industrial operation, membrane supports are packed into larger modules so as to maximise the

membrane surface area to module volume ratio and thus production for a given plant footprint. The

geometry of the support influences the module design and size as shown in Figure 2.2. Planar

membranes can be mounted in spiral wound or plate-and-frame modules; these designs are simple

and have been developed for a large number of industrial applications, in particular water treatment

with polymeric membranes. However for inorganic membranes, only the plate and frame option is

suitable and as gaskets are required for each plate, sealing is a costly problem.

Tubular membrane supports are generally limited to inorganic membranes employed in filtration

applications, for which the benefit of resistance to membrane fouling due to good fluid

hydrodynamics outweighs their high cost. In a typical tubular membrane system a large number of

tubes are manifolded in series. The permeate (i.e. the material that has passed through the

membrane) is then removed from each tube and sent to a permeate collection header. Hollow fibre

membrane modules are a subset of tubular membranes, but they are so commonly employed in

industry (in polymeric form) they are often spoken of separately. They are easy to make and allow

very large membrane areas to be contained in small, cost effective modules. Indeed, the greatest

single advantage of hollow fibre modules is the ability to pack a very large membrane area into a

single module.

(a)

(a)

(b)

(c)

(d)

(e)

Page 34: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

34

Figure 2.2 – Schematic view of three membrane modules [3].

However, regardless of geometry, membrane supports must have the capacity to support the

membrane layer without interacting chemically or significantly affecting the transport

characteristics. For inorganic membranes this manifests as a robust support with a very smooth top

layer for depositing uniform thin films. A controlled pore size and pore size distribution are highly

desired to make membrane deposition easy and reproducible. Polymeric supports are common and

frequently integrated into the polymer membrane manufacturing process, but their lack of thermal

and mechanical robustness limits their application in some industrial environments. Ceramic

supports on the other hand, offer significant thermal resistance but their brittleness and difficulties

with sealing have seen a lack of industrial take-up.

Metallic supports are relatively unknown for both research and commercial use, although they have

been tested under industrial conditions [4]. Crucially metallic supports offer increased strength and

toughness as compared to ceramic or polymeric materials whilst still retaining a high degree of

thermal resistance, making them an ideal candidate for inorganic membrane supports. Furthermore,

sealing issues are not a concern as they can be easily integrated into metallic modules, for example

by welding, without the need for specialised seals. Porous metallic supports can be generated in the

same fashion as porous ceramic supports, that is, through forming a green or pre-sintered geometry

(such as a hollow fibre) which are typically held together using a polymeric binder. The geometry

can be formed either through static pressing or extrusion techniques, of which phase inversion has

Page 35: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

35

attracted the most interest for ceramic hollow fibres and will form the basis of the production

process for stainless steel hollow fibres.

2.3. POLYMERIC MEMBRANE SUPPORTS THROUGH PHASE

INVERSION

One of the most common techniques for generating polymeric membrane supports is the dry-wet

phase inversion process. This has been subsequently adapted to the production of porous ceramic

and metallic hollow fibres; as such this section will explore the underlying mechanisms of

formation, whilst section 2.4 and 2.5 will expand on how these underlying mechanisms affect

hollow fibre production for inorganic supports. The underlying mechanisms of the phase inversion

process are complex, encompassing the thermodynamics of mixing, the phase equilibria and the

kinetics of the interactions between the components. The process was first applied to obtain

cellulose acetate asymmetric membranes for desalination by Loeb and Sourirajan [5], which

sparked a flood of research into the production of polymeric membranes using this technique [5-9].

In general, all methods for preparing membranes via phase inversion involve the precipitation of a

polymer-rich phase from an initially homogeneous casting (polymer / solvent) solution. There are

therefore, three phases that participate in the process, the polymer-rich phase, the solvent phase, and

the non-solvent phase. Ternary phase diagrams are essential tools to discuss the thermodynamics of

the precipitation process. In the ternary phase diagram the corners of the triangle represent single

components, while any point inside the triangle represents the equilibrium of the three parts of the

mixture at their relative concentrations. Strathmann and Kock [10], explain that this three phase

system consists of two regions, the first a region where all components are miscible and the second,

a region where the system separates into a polymer-rich, generally solid phase, and a polymer-lean,

generally liquid phase. The phase boundary between these regions represents the so-called binodal

curve. The variation in the Gibbs free energy of mixing ΔG in the ternary system, according to

Karimi et. al. [11], can be expressed as Equation 2.1:

(2.1)

where, R is the gas constant; T is the absolute temperature, ni and øi are the number of moles and the

volume fraction of component i respectively, χ12 (u2) is a generalized non-solvent1/solvent2

interaction function depending on the volume fraction u2=ø2/(ø1+ø2) of a pseudo binary mixture,

Page 36: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

36

χ23(ø3) is a concentration dependent interaction function between the polymer and the solvent, in

contrast to χ13, between non-solvent and polymer which can be considered constant. The literature

shows that the non-solvent/polymer interaction parameter is especially difficult to determine

because χ13 has a considerable effect on the size and the location of the miscibility gap in the ternary

phase diagram [11-13].

Every composition inside the binodal curve will de-mix into two liquid phases which differs in

composition but that are in thermodynamic equilibrium with each other. A schematic representation

of the coagulation process can be seen in the Figure 2.3 [14].

Figure 2.3 – Composition paths of a cast film immediately after immersion demonstrating

instantaneous demixing (a); and delayed demixing (b), adapted from [14].

The process of phase inversion is however a dynamic one with several steps that are relevant to the

thesis. Initially the mixture consists only of polymer and solvent, often referred to as a spinning

dope. The polymer / solvent mixture then comes in contact with a non-solvent, in most cases water

is used, but other liquids like ethanol, glycerol etc. can be used as non-solvent. The critical

characteristic of the non-solvent is that it should be miscible (wholly or partly) with the solvent so

that contact between the spinning dope and the non-solvent causes the solvent to leave the spinning

dope, driven by the concentration gradient [15]. This generates polymer rich / solvent lean regions

which causes the polymer to coagulate. At the same time, the non-solvent mixes with the polymer /

solvent mixture, although to a much lower degree due to the inherent insolubility of the mixture.

This process is complex, both thermodynamically and kinetically; however, the formation of the

dual phase system happens because the single phase becomes thermodynamically unstable due to

Page 37: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

37

the diffusion of solvent and non-solvent. Depending on the relative diffusion rates of the solvent

into the non-solvent and vice versa, different micro-structures in the coagulated polymer can be

formed. Figure 2.3a shows that if the composition path, for t < 1s, crosses the binodal line, liquid-

liquid demixing and therefore polymer coagulation starts immediately after immersion. Figure 2.3b,

on the other hand shows that if all compositions directly beneath the top layer remain in the one-

phase region and are still miscible, no demixing occurs immediately after immersion. After a longer

time interval, compositions beneath the top layer will cross the binodal line and demixing will also

start in this case. Thus two distinctly different demixing processes can be distinguished –

instantaneous demixing and delayed demixing, which result in membrane morphologies that are

completely different [14, 16]. Instantaneous demixing is typically associated with large “finger-

like” macrovoids, whilst delayed demixing typically produces structures with a “sponge-like”

morphology of smaller interconnected pores. However, the kinetics of demixing is not the only

controlling factor as Strathmann et al. [10] also observed these two fundamentally different

structures depended on the rate of polymer precipitation by non-solvent induced phase separation,

and showed that slow precipitation rates produced membranes with a “sponge-like” morphology

(Figure 2.4a). Conversely, fast precipitation rates produced membranes with large “finger-like”

macrovoids in the substructure (Figure 2.4b).

Figure 2.4 – SEM image of a typical sponge-like structure (a) and a typical finger-like structure (b),

adapted from [17].

(a) (b)

Page 38: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

38

These rates of transfer are influenced by a number of factors, including the local concentrations of

solvent, polymer and non-solvent, the affinity of those materials to each other and their dynamic

characteristics. The viscosity of the solution also plays an important role because it modifies the

diffusion behaviour of the solvent and non-solvent [7, 10, 11, 18-20]. Radovanovic et. al. [21]

presents a method to calculate the critical initial composition in which the demixing changes from

instantaneous to delayed. Here, the diffusion equation for the mixture (eq. 2.2), and the diffusion

equation for the bath (eq. 2.3) must be solved simultaneously.

(2.2)

(2.3)

where, is the flux of component i in the polymer fixed frame of reference, νi is the component i

volume fraction in the bath, X is the position of the bath-film interface in the stationary coordinate

system, and y is a position coordinate that moves with the interface, defined by

(2.4)

If, in solving the diffusion equation for the mixture, the composition profile touches or crosses the

binodal line, instantaneous demixing will occur and finger-like macrovoids are expected to be

formed. If the composition profile does not touch the binodal, precipitation is slow and the

membrane takes relatively longer to form and a sponge-like structure is obtained. Guillen et al. [22]

presents a schematic step by step graph of the processes of demixing, which is reproduced in Figure

2.5.

Page 39: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

39

Figure 2.5 – Different membrane morphologies obtained by different demixing paths.

The large macrovoids are formed due to rapid separation of the solvent from the polymer producing

a solid phase that is polymer-rich [10, 23-25]. This fast separation triggers successive tensile

ruptures of the gelled polymer-rich phase creating large macrovoids [26]. The sponge-like structure,

on the other hand, is formed from the slower interactions between the solvent and the non-solvent

[10, 23-25, 27] that generate smaller tensions within the polymer and consequently the tensile

ruptures do not happen. Rather, Koenhen et al. [15] showed that the formation of sponge-like

structures happens when the solvent leaves the matrix slowly, allowing the polymer to coagulate in

clusters. These clusters go on to aggregate into a sponge-like layer which is never fully dense as the

non-solvent cannot re-dissolve and re-precipitate the polymer into a dense continuous layer.

The presence of macrovoids can, in some cases, be beneficial especially in ultrafiltration processes

[16]. But for the majority of applications these large finger-like pores are disadvantageous. The

presence of macrovoids decreases the mechanical resistance, debilitating the structure of the support

and can result in compaction and collapse of the membrane, therefore limiting their application

particularly for high pressures processes. In addition, the large size and the uneven surface of the

finger-like pores are not suitable for membrane layer deposition.

There are several parameters which can be altered to adjust the behaviour of a given system. For

example certain compounds, like viscosity modifiers, can be added to the polymer / solvent mixture

to alter its behaviour. Macrovoids are generally formed in systems where instantaneous demixing

takes place, except when the polymer additive concentration and the non-solvent concentration in

Page 40: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

40

the polymer solution exceed a minimum value [16, 25, 28, 29]. In such cases, the polymer solution

composition more closely matches the binodal composition and therefore favours the formation of

spongy structures.

Furthermore the effect of temperature in the cast solution must be considered, where an increase in

the temperature of the mixture will lead into a decrease in its viscosity. This change in viscosity will

affect the exchange rate of solvent and non-solvent during phase inversion. An increase in viscosity

will decrease the kinetics of the phase inversion, and if the change in viscosity is significant then,

for an otherwise identical system, the microstructure can change from finger-like macrovoids to a

sponge-like structure, and vice-versa for a decrease in viscosity. Thus, temperature must be

considered as an indirect influential parameter on membrane formation kinetics and ultimately the

surface and internal membrane morphology [30-33]. Zheng et. al. studied the influence of the

temperature on the viscosity of the polymer-solvent mixture. They showed that the increase in the

solution temperature, for the case of polysulfone mixed with N,N-Dimethylacetamide or N-Methyl-

2-pyrrolidone, decreased the viscosity and increased the solvent / non-solvent miscibility [34]. They

also showed that the increase in temperature lead to an increase in the kinetic parameter of

membrane formation. In addition, the temperature of the coagulation bath also has a tremendous

influence on the final size of the formed macrovoids [35]. Zheng et al. showed that the size of

finger-like macrovoids increased with the increase in the temperature of the coagulation bath [36].

Figure 2.6 is a schematic diagram which shows the variation of the mentioned parameters in the

kinetics of the phase inversion process. The increase in the temperature of the polymer / solvent

mixture will accelerate the process of diffusion of the solvent in direction to the non-solvent,

creating a solvent-lean region faster that at lower temperatures, this will result in the appearance of

finger-like macrovoids. In contrast, when the viscosity is increased, the rate of exchange between

solvent and non-solvent will be decreased, leading to a sponge-like structure. The change in

temperature of the coagulation bath also affects the microstructure, although to a lower degree than

the temperature and viscosity of the polymer / solvent mixture, but still an increase in the

temperature of the coagulation bath can lead to the formation of finger-like voids, whilst cooling the

coagulation bath can promote the formation of a sponge-like structure.

Page 41: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

41

Figure 2.6 – Effect of parameters in the kinetics of the phase inversion process.

2.4. POLYMERIC HOLLOW FIBRE PRODUCTION PROCESS

The process of coagulation and formation for hollow fibre membrane supports follows the same

process as presented in section 2.3, however due to the presence of non-solvent at both the inner

and outer side of the hollow fibre there are now two coagulation fronts. Thus, alternate asymmetric

structures can be created, particularly since the inner and outer surfaces are exposed to dry and wet

non-solvents during the phase inversion. The dry phase inversion takes place in the atmosphere by

evaporation of the volatile solvent and or by absorption of water as a non-solvent from the moisture

contained in the air. The wet phase inversion is carried out as above by contacting the polymer-

solvent mixture with the non-solvent, initiating an exchange between the solvent and the non-

solvent leading to a coagulated hollow fibres [37]. Wang et. al. reported that the dry inversion phase

helps to develop a dense layer at the outside of the hollow fibre which may or may not be desirable

depending on the final application [38].

The spinning of hollow fibres is conducted by the extrusion of the spinning mixture (polymer /

solvent) through a tube in orifice spinneret as depicted in Figure 2.7a. The complete setup of the

extrusion process can be seen in Figure 2.7b. The main variables present in the extrusion process

are air gap length, polymer extrusion rate and bore liquid (non-solvent). These variables are

important, because they control the amount of time that the outer side undergoes dry and then wet

phase inversion respectively. These have a significant impact on the final microstructure of the

hollow fibre, as demonstrated by Kingsbury and Li [17].

Page 42: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

42

Figure 2.7 - Cross-section of the tube-in-orifice spinneret (a); Scheme of the process of extrusion of

the Hollow fibre (b).

It was demonstrated by Wang [38-42] that the non-solvent plays a very important role in the final

microstructure of the hollow fibre, since its interaction with polymer-solvent system can accelerate

or hinder the diffusion of the solvent outside the mixture and as a consequence affect the entrance

of the non-solvent to the system. The fact that there are two mixing fronts (inside with the bore

liquid and outside with the coagulation bath) serves to complicate matters. To assist in

understanding the process Wang [39] presented the precipitation value, defined as the grams of non-

solvent needed to cause visual turbidity in a binary solution containing 100g of solvent and 2g of

polymer, which is a way to compare different systems. Different binary systems have different

precipitations values for the same non-solvent which means that the equilibrium can be reached

faster or slower, depending on the relative position of the precipitation value. Furthermore, in the

hollow fibre production process it is possible to have more than one non-solvent interacting with the

spinning mixture at the same time, which can lead into differences in the kinetics on both sides of

the hollow fibre.

Figure 2.8 shows the steps of the hollow fibre spinning process. At the time t = 0, the spinning

mixture leaves the spinneret and first contacts with the non-solvent present in the bore liquid, at this

time diffusion of the solvent in direction to the bore liquid (and vice versa) starts. As explained in

section 2.3, fast diffusion in the initial stage of spinning leads to the formation of finger-like voids,

which will be present at the inner surface of the hollow fibre. However, as a consequence of this

fast initial diffusion, the outer surface of the hollow fibre may become lean in solvent as the fibre

(a) (b)

Page 43: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

43

passes through the air gap, due to the concentration gradient between the polymer/solvent system

and the non-solvent bore liquid driving the remaining solvent inwards. When the fibre enters the

coagulation bath, the concentration of solvent inside the spinning mixture on the outer surface is

then low enough to promote a slow diffusion rate, which translates into the formation of a sponge-

like region on the outer surface of the hollow fibre. In the Figure 2.8 the red colour represents

solvent rich regions and the yellow colour indicates solvent lean regions.

Figure 2.8 - Coagulation phenomena in the hollow fibre production process.

However, if the air gap is not large enough then the solvent lean region on the outer surface of the

hollow fibre will not be created, thus allowing fast diffusion of solvent into the coagulation bath

resulting in the formation of finger-like macrovoids close to both inner and outer surfaces [20, 32,

34, 39-41]. Likewise if the solvent non-solvent exchange is especially rapid a thick polymer rich

region may form in the middle of the hollow fibre wall, effectively trapping the remaining solvent

in the outer half of the hollow fibre wall and ensuring the concentration gradient remains relatively

high. Then once the fibre hits the coagulation bath the concentration gradient for the solvent is still

large enough to initiate instantaneous demixing and the formation of finger-like macrovoids at the

hollow fibre’s outer surface. In either case, once the hollow fibre is inside the coagulation bath, the

diffusion process continues until the polymer phase is completely solidified and the equilibrium is

reached.

Page 44: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

44

2.5. EFFECT OF SOLID PARTICLES ON THE HOLLOW FIBRE

PRODUCTION PROCESS

Hollow fibre spinning technology has recently been adapted for creation the ceramic and metallic

hollow fibres by including the respective particles into the polymer / solvent mixture which are then

bound in place during the coagulation process. The nascent or coagulated hollow fibre, alternatively

called a green hollow fibre, is then subjected to high temperatures (>600 °C) to simultaneously

remove the polymer (also called a binder) and sinter the ceramic or metallic particles into a strong

porous solid [43-48]. The production of ceramic hollow fibres have been well studied in the

literature [45-47, 49-53] with a focus on application based testing. However, very few have

investigated the actual effect of the particles on the final morphology, performance and mechanical

robustness of the final hollow fibre. Kingsbury and Li [17] show in their work that by changing the

viscosity of the spinning dope it is possible to control the type of morphology that results after the

phase inversion process. Similar to polymeric membranes higher viscosities produce sponge-like

structures and lower viscosities produce more finger-like macrovoids. These conclusions, however,

do not correlate the effect that particle size and loading, amongst others, can have on the final

morphology of the hollow fibres produced.

The addition of particulate solids to a liquid such as a polymeric spinning dope, can have a marked

effect on the viscous properties of a fluid. Initially studied by Einstein [54] for diluted solutions

containing spherical particles, his theory showed a linear increase in viscosity with increasing

volume fraction of spheres in the solution up to 30 vol%. Since then several other authors have

advanced the theory [55-63] and there are relatively well developed methods to predict the viscosity

of such solutions. Current theory suggests that the particle size of the added solids does not alter the

viscosity for dilute solutions [58-60], but it is known that the shape of the particles affects the

viscosity of the spinning dope. Considering the previous work on polymeric membranes it is to be

expected that spinning dopes having same viscosity will produce hollow fibres with similar

morphology. However, in the work of Liu et. al. [46] it is possible to observe different

morphologies resulting from spinning dopes containing similar solids loadings. In this case the only

difference was that particles used were of different sizes, although crucially this finding was not

discussed by the authors. Likewise, the same effect was observable in the images published by

Kingsbury and Li [17] although again they remain undiscussed.

Metallic hollow fibres have only recently been produced [64-66], and very little has been discussed

on the effect of metallic particles during the spinning process. Luiten-Olieman et al. [65] observed

Page 45: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

45

that the level of stainless steel loading in the polymer/solvent changes the viscosity of the spinning

mixture. As expected higher steel loadings produced spinning dopes with higher viscosities which

in turn had a clear influence on the morphology of the hollow fibre. However, the effects of particle

size or shape were not explored. In subsequent publications the authors discussed the effect of low

loadings on the final hollow fibre morphology. Rather than focus on the green fibre morphology the

authors found that the final hollow fibre morphology was formed during the sintering step due to

significant shrinkage and densification [66].

2.6. SINTERING OF POROUS SOLIDS

Sintering is a processing technique to produce density controlled materials from ceramics and/or

metals by applying thermal energy. This process allows the formation of the shaped body using less

energy than a melting process, due to the lower temperatures used and the fact that the melting point

of the material is not reached. Porous materials have been produced conventionally by sintering

powder compacts, or green bodies (pre-sintered body with similar shape), to partial densification. In

the production of ceramic or metallic hollow fibres, the process of sintering is undertaken to

eliminate the binder and to encourage interconnection of the particles in order to create a solid body

with the desired characteristics. In this work only the sintering process for obtaining porous

materials will be discussed.

2.6.1. The Raw Powder

In order to make porous materials using ceramic or metallic powders as a starting material, it is

necessary to understand that the properties of the starting powders are important factors that

strongly influence the properties of the final porous product. The effect of particle size is explained

by Herring [67] who showed that when powders with similar shapes, but different sizes, are sintered

at the same conditions using the same sintering mechanism, the required time to obtain the same

degree of densification of the largest particle is directly proportional to the coefficient between the

radii of both particles, that is:

(2.5)

where β is the coefficient between the radii of the biggest and the smallest particle and ω is an index

dependent on the sintering mechanism. It is noteworthy that the law proposed by Herring is

generally not satisfied in sintering of powders because the mechanism of grain growth is different to

Page 46: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

46

the mechanism of densification. Nevertheless, this scale law can be used as a good approximation

and clearly demonstrates the effect of particle size on the sintering process. Table 2.1, adapted from

Ishizaki et al. 1998, shows the relationships between the properties of the raw powder and the

properties of the resulting porous materials.

Table 2.1- Influences that powder properties has on porous materials [64].

Properties of powder Properties of porous materials

Particle shape Pore shape, porosity, pore size distribution, tortuosity,

surface area.

Particle size Pore size, mechanical strength, surface area.

Particle size distribution Pore size, pore size distribution, porosity.

Porosity Porosity, pore size distribution, surface area.

Agglomeration Pore size, pore size distribution, porosity, pore shape.

Different processes are used to produce fine powders from different materials, but generally

powders made from ceramic and other brittle materials are fabricated by milling. This process is

cheap and easily can produce powders of particle size from millimetres to microns order. Ductile

materials, such as metals, are difficult to grind, therefore, most of the metallic powders are

produced by gas or water atomization methods [68-70]. The result of these processes is a spherical

powder with smooth surface and high sinterability.

2.6.2. Sintering Mechanism

Sintering processes can be divided in two main types: solid state sintering and liquid phase sintering

[71]. Solid state sintering occurs when the green body is densified wholly in a solid state at the

sintering temperature, while liquid phase sintering occurs when a liquid phase is present in the

powder compact during sintering. The schematic diagram (Figure 2.9) illustrates the two cases.

Page 47: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

47

Figure 2.9 - Scheme of various types of sintering, adapted from [67].

At temperature T1, solid state sintering occurs in an A-B powder compact with composition X1.

Liquid phase sintering happens at temperature T3, for the same green body. In addition other types

of sintering are also depicted. Viscous flow sintering occurs when the volume fraction of liquid is

high enough, so the full densification of the compact can be achieved by a viscous flow of grain-

liquid mixture without changing the shape of the grains in the process. The transient liquid phase

sintering is a combination of the effects of liquid phase sintering and solid state sintering, when the

compact is sintered above the eutectic temperature, but below the solid line, for example at

temperature T2. Since the sintering temperature is above the A-B eutectic temperature, a liquid

phase is formed through a reaction between A and B powders during the heating period, but during

sintering the liquid phase disappears and only a solid phase remains because the equilibrium phase

under the sintering conditions is the solid phase. In general, compared to solid state sintering, liquid

phase sintering allows better control of the microstructure and a reduction in production costs, but

can be detrimental to some mechanical properties.

The driving force for the sintering process is the reduction of the total interfacial energy, expressed

as γA, where γ is the specific surface energy and A is the total surface area of the compact. The

reduction in the total energy can be calculate through Equation (6).

(2.6)

Here the change in interfacial energy (Δγ) is due to densification whilst the change in interfacial

area is due to coarsening. For solid state sintering, Δγ is related to the replacement of solid-vapour

Page 48: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

48

interfaces by solid-solid interfaces. Figure 2.10 shows that the reduction in total interface energy

occurs via densification, the basic phenomena of sintering.

Figure 2.10 - Basic phenomena occurring during sintering.

The solid phase sintering mechanism can be divided into three stages as displayed in Figure 2.11.

The initial phase corresponds to the formation of necks at the contact point between particles.

During the intermediate stage the size of the necks increase and the pores alter their shape becoming

more spheroidal. In the final stage diffusion occurs between grain boundaries and pores disappear

or become isolated. For a porous material the sintering process should be interrupted during either

the initial or intermediate stage, with the end point becoming crucially important in determining the

properties of the porous sintered body.

Page 49: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

49

Figure 2.11 - Scheme of densification stages by sintering process [68-70].

The driving force for mass transport into the necks is provided by the difference in free energy or

chemical potential between the surface of the neck area (with negative curvature ρ) and the surface

of particle [75]. There are six mass transport paths in the sintering process [75], including

evaporation-condensation, diffusion and plastic flow [76] and by these transport processes necks

grow, and as result sintering proceeds. Figure 2.12 shows the schematic cross section of a sintering

model and together with Table 2.2 explains the routes for diffusion during sintering [75, 76]. This

process usually starts to occur when the ratio of the absolute temperature to the melting point, α, is

0.23.

Figure 2.12 - Sintering model of two particles: (a) shows a schematic representation of bonded

particles; x, y, a and ρ indicates neck radius, interpenetrated depth, particle radius, and radius of

Page 50: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

50

neck surface, respectively. (b) shows the different diffusion paths during the initial stage of

sintering.

Table 2.2 - Path routes for diffusion during sintering [77, 78].

No. Path Route Diffusion Source Diffusion Sink Densification

1 Surface diffusion Surface Neck No

2 Volume diffusion Surface Neck No

3 Evaporation-

condensation

Surface Neck No

4 Grain boundary diffusion Grain boundary Neck Yes

5 Volume diffusion Grain boundary Neck Yes

6 Volume diffusion Dislocation Neck No

The diffusion processes happen when, the equilibrium concentration of atomic or ionic vacancies at

the surface of the particle and necks varies with chemical potential at the respective location, which

is relatively higher than in the interior of the particles [77, 78]. Hence, diffusion occurs from inside

(volume diffusion) to the surface (surface diffusion) and to the crystalline grain boundary (grain

boundary diffusion). Surface diffusion usually takes place at a lower temperatures (α = 0.33 to

0.45), whereas at higher temperatures (α = 0.42 to 0.8), volumetric diffusion becomes active and the

growth of necks is increased due to an enhancement in mass transfer rate.

In the evaporation-condensation process, the material evaporates at convex particle points or on the

surface of necks with concave curvatures. As a result of this, round grains are produced. It is

noteworthy that surface diffusion and evaporation-condensation do not cause densification although

they do produce neck growth. Indeed, these processes actually prevent densification of the materials

because those mass transport mechanisms reduce the driving force for sintering by neck growth,

thus surface diffusion and evaporation-condensation are desirable to produce porous materials.

Sintering mechanisms for some amorphous materials, like glass and resin, are usually controlled by

plastic flow. This process plays a significant role in rapid shrinkage and densification at the initial

Page 51: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

51

stage of sintering. However, this process does not happen in sintering of metals and therefore, will

not be further investigated.

The kinetics of the multiple mass transport paths and by extension neck growth rate is complex and

dependent on the dominant sintering mechanism [77, 78]. Generally, the size of the neck radius,

denoted by x in the Figure 2.12, at time, t, can be determined by using equation 2.7:

(2.7)

The linear shrinkage, Δl/lo , as function of time is given by equation 2.8:

(2.8)

where a is the particle radius, m and n are numerical exponent dependent on the mechanism of

sintering and H is a function that contains geometrical and material parameters of the powder, also

dependent on the sintering mechanism. The values of the parameters dependent on the sintering

mechanism are displayed in Table 2.3 below.

Table 2.3 – Plausible values for the constants in Eq. 2.7 and 2. 8 for the initial stage of sintering.

Ds,Dl,Dgb are diffusion coefficients for surface, lattice, and grain boundary respectively. δs, δgb are

thickness for surface and grain boundary diffusion. γsv, specific surface energy, p0, vapour pressure,

m atomic mass, k Boltzmann constant, T absolute temperature, η viscosity and Ω is the atomic

volume [73, 74, 79-81].

Mechanism m n H

Surface diffusion 7 4

Lattice diffusion from surface 4 3

Vapour transport 3 2

Grain boundary diffusion 6 4

Page 52: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

52

Lattice diffusion from grain boundary 5 3

Viscous flow 2 1

For equations 2.7 and 2.8, it is assumed that only one kind of mass transport dominates the sintering

process, however it is known that this approximation cannot always be made, and sintering kinetics

with multiple paths of mass transport are available in the literature [82-84].

To produce porous materials with high open porosity, those sintering mechanisms without

densification are important. Importantly and perhaps counter-intuitively, materials sintered by those

mechanisms to have a highly open porosity can still have high mechanical strength, provide they

have well-developed necks [71]. In addition, large, well developed necks allow a higher fluid

permeability than poorly developed necks due to the difference in pore shape [68, 71]. Thus, the

sintering mechanism significantly influences pore geometry, porosity and by extension the

performance of the sintered body.

2.6.3. Effect of sintering parameters

In general, the densification rate increases with decreasing particle size and increasing sintering

temperature and time, as shown schematically in Figure 2.13.

Figure 2.13 - Effect of sintering parameters on densification.

Page 53: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

53

Since sintering is a thermally activated process, the variables affected by the change in temperature

(viscosity, diffusivity, etc.) are expressed as an exponential function of the temperature, but the

exact relation is different for each of the different sintering mechanisms. An increase in sintering

temperature leads into an enhancement of the densification rate relative to grain growth rate. To

obtain porous materials the sintering process commonly takes place at lower temperatures or for a

shorter time than those for dense materials. These conditions allow particle bonding without

significant densification; this is because at lower temperatures surface diffusion is more dominant

than volume diffusion. As temperatures increase volume diffusion becomes the main route for mass

transport, which suggests that sintering at low temperatures leads to produce porous materials with

well-developed necks and high open porosity, with the caveat that it takes longer periods of time to

sinter owing to the low diffusivity at low temperatures [85, 86].

The sintering atmosphere has a significant effect on the final sintered body, for example, an excess

in oxygen partial pressure causes oxidation of metals during sintering and so reducing atmospheres

are typically used. On the other hand, low oxygen partial pressure reduces oxide based ceramics,

which prefer oxidising atmospheres. Reducing atmospheres for sintering metals are often prepared

by introducing hydrogen gas into the sintering gas flow or by using a graphite furnace. Other

methods to reduce the presence of oxygen is by creating a nitrogen atmosphere during the sintering

process, but this can lead into the adsorption of the gas inside the matrix of the metal, especially for

stainless steels, changing the chemical and mechanical properties of the resulting body [87]. Li et.

al. [88] and Mariappan et. al. [89] showed that the sintering atmosphere not only influences the

oxidation or reduction of the powder, but also affects the microstructure, mechanical properties and

shrinkage of the final material. Finally, increasing the external pressure of the sintering atmosphere

has the effect of increasing the densification rate [90, 91].

The sintering atmosphere can alter both the kinetics of densification during the sintering, and the

composition of the material when exposed to high temperatures. For instance, German [92] reported

that nitrogen and argon produce a similar type of neutral sintering atmosphere; however, nitrogen

often results in the formation of chemical compounds that are detrimental to the densification

process. This effect is more accentuated in steels as the nitrogen can diffuse inside the crystal

lattice. Conversely, argon is completely inert and is unable to diffuse within the crystal structure of

the steel. Indeed, it has been shown in a number of published studies that sintering under argon

results in higher levels of densification [88, 89, 93, 94] for a number of other materials.

Page 54: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

54

Lastly, the choice of polymeric binder in forming the green body can also affect the sintering

process and be a determining factor in the final properties of the resulting sintered body. In

particular, the decomposition products formed during the debinding step (i.e. when the polymer is

decomposed) can alter the sintering atmosphere and contaminate the sintering process. Leo et al.

discussed the effect of the sulphur released from the PESf binder, on the chemical composition of

ceramic hollow fibres [44]. Likewise, the sulphur can break the passivated layer of the stainless

steel and promote its corrosion leading to a reduction in the mechanical strength of the hollow fibre

[95, 96].

2.7. CONCLUDING REMARKS

Membrane substrates made from porous metals are promising in that they offer several mechanical

and design advantages in comparison to polymeric or ceramic supports. However, to produce a

porous membrane support a polymeric binder must be used to hold the particles together in the

desired shape before and during sintering. This most commonly takes the form of hollow fibres

made via a dry-wet phase inversion spinning process. Control of the spinning conditions can lead to

the desired microstructure of the hollow fibre. Interactions in the ternary system, after the addition

of stainless steel particles is not well understood, however it is possible that the inclusion of the

metal powder will change the viscosity of the spinning mixture and as a consequence, change the

diffusion rate of solvent and non-solvent. Nevertheless, at this point is not possible to know if those

changes in the diffusion rate are going to be significant or negligible.

Secondly, the sintering step is very important to obtain the desired mechanical properties and a

competitive relation between mechanical properties and permeation of the support is expected. The

sintering atmosphere can significantly alter the resulting mechanical properties of the hollow fibre

when exposed to industrial conditions. Poor sintering can lead to weak hollow fibres incapable of

sustaining the working conditions. Conversely, an overly strong sintering step will result in a dense

hollow fibre which will be unable to permeate components. This project will focus on development

and testing of stainless steel hollow fibres for membrane supports, and how the variables of

production affect the performance of the support.

Page 55: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

55

2.8. REFERENCES

[1] P. Maarten Biesheuvel, H. Verweij, Design of ceramic membrane supports: permeability, tensile

strength and stress, Journal of Membrane Science, 156 (1999) 141-152.

[2] S. Smart, L. Ding, J.C. diniz da Costa, Inorganic membranes for syntesis gas processing, in:

Advanced membrane science and technology for sustainable energy and environmental

applications, Woodhead Publishing Limited, Cambridge, UK, 2011, pp. 214 - 254.

[3] R.W. Baker, Membranes and Modules, in: Membrane Technology and Applications, John

Wiley & Sons, Ltd, 2004, pp. 89-160.

[4] K. Brands, D. Uhlmann, S. Smart, M. Bram, J.C. Diniz da Costa, Long-term flue gas exposure

effects of silica membranes on porous steel substrate, Journal of Membrane Science, 359 (2010)

110-114.

[5] S. Loeb, S. Sourirajan, Sea Water Demineralization by Means of an Osmotic Membrane, in:

Saline Water Conversion?II, AMERICAN CHEMICAL SOCIETY, 1963, pp. 117-132.

[6] A.P. Borshchev, É.M. Aizenshtein, V.K. Gusev, Hollow-fibre membranes for ultrafiltration

(review), Fibre Chemistry, 26 (1994) 105-113.

[7] K.V. Peinemann, J.F. Maggioni, S.P. Nunes, Poly(ether imide) membranes obtained from

solution in cosolvent mixtures, Polymer, 39 (1998) 3411-3416.

[8] I. Cabasso, E. Klein, J.K. Smith, Polysulfone hollow fibers. I. Spinning and properties, Journal

of Applied Polymer Science, 20 (1976) 2377-2394.

[9] G.H. Koops, J.A.M. Nolten, M.H.V. Mulder, C.A. Smolders, Integrally skinned polysulfone

hollow fiber membranes for pervaporation, Journal of Applied Polymer Science, 54 (1994) 385-

404.

[10] H. Strathmann, K. Kock, The formation mechanism of phase inversion membranes,

Desalination, 21 (1977) 241-255.

[11] M. Karimi, W. Albrecht, M. Heuchel, M.H. Kish, J. Frahn, T. Weigel, D. Hofmann, H.

Modarress, A. Lendlein, Determination of water/polymer interaction parameter for membrane-

forming systems by sorption measurement and a fitting technique, Journal of Membrane Science,

265 (2005) 1-12.

Page 56: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

56

[12] G. Tkacik, L. Zeman, Component mobility analysis in the membrane-forming system water/n-

methyl-2-pyrrolidone/polyethersulfone, Journal of Membrane Science, 31 (1987) 273-288.

[13] Z. Li, C. Jiang, Determination of the nonsolvent–polymer interaction parameter χ13 in the

casting solutions, Journal of Membrane Science, 174 (2000) 87-96.

[14] M. Mulder, Basic principles of membrane technology, Kluwer Academic, Dordrecht,

Netherlands, 1991.

[15] D.M. Koenhen, M.H.V. Mulder, C.A. Smolders, Phase separation phenomena during the

formation of asymmetric membranes, Journal of Applied Polymer Science, 21 (1977) 199-215.

[16] C.A. Smolders, A.J. Reuvers, R.M. Boom, I.M. Wienk, Microstructures in phase-inversion

membranes. Part 1. Formation of macrovoids, Journal of Membrane Science, 73 (1992) 259-275.

[17] B.F.K. Kingsbury, K. Li, A morphological study of ceramic hollow fibre membranes, Journal

of Membrane Science, 328 (2009) 134-140.

[18] L.-W. Chen, T.-H. Young, Effect of nonsolvents on the mechanism of wet-casting membrane

formation from EVAL copolymers, Journal of Membrane Science, 59 (1991) 15-26.

[19] T.-H. Young, L.-W. Chen, Roles of bimolecular interaction and relative diffusion rate in

membrane structure control, Journal of Membrane Science, 83 (1993) 153-166.

[20] S.S. Tamhankar, M. Bagajewicz, G.R. Gavalas, P.K. Sharma, M. Flytzani-Stephanopoulos,

Mixed-oxide sorbents for high-temperature removal of hydrogen sulfide, Industrial & Engineering

Chemistry Process Design and Development, 25 (1986) 429-437.

[21] P. Radovanovic, S.W. Thiel, S.-T. Hwang, Formation of asymmetric polysulfone membranes

by immersion precipitation. Part I. Modelling mass transport during gelation, Journal of Membrane

Science, 65 (1992) 213-229.

[22] G.R. Guillen, Y. Pan, M. Li, E.M.V. Hoek, Preparation and Characterization of Membranes

Formed by Nonsolvent Induced Phase Separation: A Review, Industrial & Engineering Chemistry

Research, 50 (2011) 3798-3817.

[23] H. Strathmann, K. Kock, P. Amar, R.W. Baker, The formation mechanism of asymmetric

membranes, Desalination, 16 (1975) 179-203.

Page 57: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

57

[24] I. Cabasso, E. Klein, J.K. Smith, Polysulfone hollow fibers. II. Morphology, Journal of

Applied Polymer Science, 21 (1977) 165-180.

[25] K. Kimmerle, H. Strathmann, Analysis of the structure-determining process of phase inversion

membranes, Desalination, 79 (1990) 283-302.

[26] S.S. Prakash, L.F. Francis, L.E. Scriven, Microstructure evolution in dry–wet cast polysulfone

membranes by cryo-SEM: A hypothesis on macrovoid formation, Journal of Membrane Science,

313 (2008) 135-157.

[27] L. Broens, F.W. Altena, C.A. Smolders, D.M. Koenhen, Asymmetric membrane structures as a

result of phase separation phenomena, Desalination, 32 (1980) 33-45.

[28] R.M. Boom, T. van den Boomgaard, C.a. Smolders, Mass transfer and thermodynamics during

immersion precipitation for a two-polymer system: Evaluation with the system PES--PVP--NMP--

water, Journal of Membrane Science, 90 (1994) 231-249.

[29] J.-Y. Lai, F.-C. Lin, C.-C. Wang, D.-M. Wang, Effect of nonsolvent additives on the porosity

and morphology of asymmetric TPX membranes, Journal of Membrane Science, 118 (1996) 49-61.

[30] J.-H. Kim, K.-H. Lee, Effect of PEG additive on membrane formation by phase inversion,

Journal of Membrane Science, 138 (1998) 153-163.

[31] S. Yang, Z. Liu, Preparation and characterization of polyacrylonitrile ultrafiltration

membranes, Journal of Membrane Science, 222 (2003) 87-98.

[32] X. Wang, L. Zhang, D. Sun, Q. An, H. Chen, Effect of coagulation bath temperature on

formation mechanism of poly(vinylidene fluoride) membrane, Journal of Applied Polymer Science,

110 (2008) 1656-1663.

[33] E. Saljoughi, M. Amirilargani, T. Mohammadi, Effect of poly(vinyl pyrrolidone) concentration

and coagulation bath temperature on the morphology, permeability, and thermal stability of

asymmetric cellulose acetate membranes, Journal of Applied Polymer Science, 111 (2009) 2537-

2544.

[34] Q.-Z. Zheng, P. Wang, Y.-N. Yang, Rheological and thermodynamic variation in polysulfone

solution by PEG introduction and its effect on kinetics of membrane formation via phase-inversion

process, Journal of Membrane Science, 279 (2006) 230-237.

Page 58: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

58

[35] H.A. Tsai, L.D. Li, K.R. Lee, Y.C. Wang, C.L. Li, J. Huang, J.Y. Lai, Effect of surfactant

addition on the morphology and pervaporation performance of asymmetric polysulfone membranes,

Journal of Membrane Science, 176 (2000) 97-103.

[36] Q.-Z. Zheng, P. Wang, Y.-N. Yang, D.-J. Cui, The relationship between porosity and kinetics

parameter of membrane formation in PSF ultrafiltration membrane, Journal of Membrane Science,

286 (2006) 7-11.

[37] C. Barth, M.C. Gonçalves, A.T.N. Pires, J. Roeder, B.A. Wolf, Asymmetric polysulfone and

polyethersulfone membranes: effects of thermodynamic conditions during formation on their

performance, Journal of Membrane Science, 169 (2000) 287-299.

[38] D. Wang, K. Li, W.K. Teo, Polyethersulfone hollow fiber gas separation membranes prepared

from NMP/alcohol solvent systems, Journal of Membrane Science, 115 (1996) 85-108.

[39] D. Wang, K. Li, W.K. Teo, Preparation of Poly(ether sulfone) and Poly(ether imide) Hollow

Fiber Membranes for Gas Separation: Effect of Internal Coagulant, in: Membrane Formation and

Modification, American Chemical Society, 1999, pp. 96-109.

[40] D. Wang, K. Li, W.K. Teo, Preparation and characterization of polyetherimide asymmetric

hollow fiber membranes for gas separation, Journal of Membrane Science, 138 (1998) 193-201.

[41] L.K. Wang, J.P. Chen, Y.-T. Hung, N.K. Shammas, Membrane and Desalination Technologies,

Springer, Dordrecht, 2011.

[42] Y.H. Wang, J.G. Cheng, X.Q. Liu, G.Y. Meng, Y.W. Ding, Preparation and Sintering of

Macroporous Ceramic Membrane Support from Titania Sol-Coated Alumina Powder, Journal of the

American Ceramic Society, 91 (2008) 825-830.

[43] F.R. García-García, M.A. Rahman, B.F.K. Kingsbury, K. Li, Asymmetric ceramic hollow

fibres: New micro-supports for gas-phase catalytic reactions, Applied Catalysis A: General, 393

(2011) 71-77.

[44] A. Leo, S. Smart, S. Liu, J.C. Diniz da Costa, High performance perovskite hollow fibres for

oxygen separation, Journal of Membrane Science, 368 (2011) 64-68.

[45] S. Liu, K. Li, Preparation TiO2/Al2O3 composite hollow fibre membranes, Journal of

Membrane Science, 218 (2003) 269-277.

Page 59: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

59

[46] S. Liu, K. Li, R. Hughes, Preparation of porous aluminium oxide (Al2O3) hollow fibre

membranes by a combined phase-inversion and sintering method, Ceramics International, 29 (2003)

875-881.

[47] S. Liu, X. Tan, K. Li, R. Hughes, Preparation and characterisation of SrCe0.95Yb0.05O2.975

hollow fibre membranes, Journal of Membrane Science, 193 (2001) 249-260.

[48] B. Meng, X. Tan, X. Meng, S. Qiao, S. Liu, Porous and dense Ni hollow fibre membranes,

Journal of Alloys and Compounds, 470 (2009) 461-464.

[49] I.-H. Choi, I.-C. Kim, B.-R. Min, K.-H. Lee, Preparation and characterization of ultrathin

alumina hollow fiber microfiltration membrane, Desalination, 193 (2006) 256-259.

[50] D.F. Li, T.-S. Chung, R. Wang, Y. Liu, Fabrication of fluoropolyimide/polyethersulfone (PES)

dual-layer asymmetric hollow fiber membranes for gas separation, Journal of Membrane Science,

198 (2002) 211-223.

[51] K. Li, X. Tan, Y. Liu, Single-step fabrication of ceramic hollow fibers for oxygen permeation,

Journal of Membrane Science, 272 (2006) 1-5.

[52] S. Liu, G.R. Gavalas, Oxygen selective ceramic hollow fiber membranes, Journal of

Membrane Science, 246 (2005) 103-108.

[53] X. Tan, N. Liu, B. Meng, S. Liu, Morphology control of the perovskite hollow fibre

membranes for oxygen separation using different bore fluids, Journal of Membrane Science, In

Press, Accepted Manuscript.

[54] A. Einstein, Eine neue Bestimmung der Moleküldimensionen, Annalen der Physik, 324 (1906)

289-306.

[55] M. Mooney, The viscosity of a concentrated suspension of spherical particles, Journal of

Colloid Science, 6 (1951) 162-170.

[56] D.G. Thomas, Transport characteristics of suspension: VIII. A note on the viscosity of

Newtonian suspensions of uniform spherical particles, Journal of Colloid Science, 20 (1965) 267-

277.

Page 60: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

60

[57] N.A. Frankel, A. Acrivos, On the viscosity of a concentrated suspension of solid spheres,

Chemical Engineering Science, 22 (1967) 847-853.

[58] T. Kitano, T. Kataoka, T. Shirota, An empirical equation of the relative viscosity of polymer

melts filled with various inorganic fillers, Rheol Acta, 20 (1981) 207-209.

[59] A.B. Metzner, Rheology of Suspensions in Polymeric Liquids, Journal of Rheology, 29 (1985)

739-775.

[60] R.D. Sudduth, A generalized model to predict the viscosity of solutions with suspended

particles. III. Effects of particle interaction and particle size distribution, Journal of Applied

Polymer Science, 50 (1993) 123-147.

[61] P. Snabre, P. Mills, Rheology of concentrated suspensions of viscoelastic particles, Colloids

and Surfaces A: Physicochemical and Engineering Aspects, 152 (1999) 79-88.

[62] M. Schneider, J. Claverie, C. Graillat, T.F. McKenna, High solids content emulsions. I. A

study of the influence of the particle size distribution and polymer concentration on viscosity,

Journal of Applied Polymer Science, 84 (2002) 1878-1896.

[63] T. Hanemann, Influence of particle properties on the viscosity of polymer–alumina composites,

Ceramics International, 34 (2008) 2099-2105.

[64] S.-M. Lee, I.-H. Choi, S.-W. Myung, J.-y. Park, I.-C. Kim, W.-N. Kim, K.-H. Lee, Preparation

and characterization of nickel hollow fiber membrane, Desalination, 233 (2008) 32-39.

[65] M.W.J. Luiten-Olieman, L. Winnubst, A. Nijmeijer, M. Wessling, N.E. Benes, Porous stainless

steel hollow fiber membranes via dry–wet spinning, Journal of Membrane Science, 370 (2011) 124-

130.

[66] M.W.J. Luiten-Olieman, M.J.T. Raaijmakers, L. Winnubst, M. Wessling, A. Nijmeijer, N.E.

Benes, Porous stainless steel hollow fibers with shrinkage-controlled small radial dimensions,

(2011).

[67] C. Herring, Effect of Change of Scale on Sintering Phenomena, J. Appl. Phys., 21 (1950) 301.

[68] K. Ishizaki, S. Komarneni, M. Nanko, Porous materials: process technology and applications,

Kluwer Academic, Dordrecht, Netherlands, 1998.

Page 61: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

61

[69] W. Schatt, K.-P. Wieters, Powder metallurgy: processing and materials, European Powder

Metallurgy Association, Shrewbury, UK, 1997.

[70] R.M. German, Powder metallurgy science, Metal Powder Industries Federation, Princeton, N.J,

1994.

[71] S.-J.L. Kang, Sintering: densification, grain growth, and microstructure, Elsevier Butterworth-

Heinemann, Amsterdam, 2005.

[72] R.L. Coble, J. Sinter crystallines solids-Ⅰintermediate final state diffusion models, Applied

physics, 32 (1961) 787.

[73] G.C. Kuckzynski, Self-diffusion in Sintering of Metallic Particles American Institute of

Mining, Metallurgical and Petroleum Engineers (AIME) Transactions 185 (1949) 9.

[74] W.D. Kingery, M. Berg, Study of the Initial Stages of Sintering Solids by Viscous Flow,

Evaporation;Condensation, and SelfDiffusion, Journal of Applied Physics, 26 (1955) 1205-1212.

[75] W.D. Kingery, H.K. Bowen, D.R. Uhlmann, Introduction to ceramics, Wiley, New York, 1976.

[76] K. Okuyama, Sintering, in: Powder Technology Handbook, Third Edition, CRC Press, 2006.

[77] M.F. Ashby, A first report on sintering diagrams, Acta Metallurgica, 22 (1974) 275-289.

[78] F.B. Swinkels, M.F. Ashby, A second report on sintering diagrams, Acta Metallurgica, 29

(1981) 259-281.

[79] D.L. Johnson, I.B. Cutler, Diffusion Sintering: I, Initial Stage Sintering Models and Their

Application to Shrinkage of Powder Compacts, Journal of the American Ceramic Society, 46

(1963) 541-545.

[80] R.L. Coble, Initial Sintering of Hematite and Alumina, Journal of the American Ceramic

Society, 41 (1958) 55.

[81] Z.Z. Fang, Sintering of Advanced Materials - Fundamentals and Processes, in: M.N. Rahaman

(Ed.) Kinetics and mechanisms of densification, Woodhead Publishing, 2010, pp. 33-64.

[82] D.L. Johnson, T.M. Clarke, Grain boundary and volume diffusion in the sintering of silver,

Acta Metallurgica, 12 (1964) 1173-1179.

Page 62: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

62

[83] D. Johnson, New Method of Obtaining Volume, Grain Boundary, and Surface Diffusion

Coefficients from Sintering Data, J. Appl. Phys., 40 (1969) 192.

[84] J.G.R. Rockland, The determination of the mechanism of sintering, Acta Metallurgica, 15

(1967) 277-286.

[85] C. Falamaki, M.S. Afarani, A. Aghaie, Initial sintering stage pore growth mechanism applied

to the manufacture of ceramic membrane supports, Journal of the European Ceramic Society, 24

(2004) 2285-2292.

[86] E. Levänen, T. Mäntylä, Effect of sintering temperature on functional properties of alumina

membranes, Journal of the European Ceramic Society, 22 (2002) 613-623.

[87] H.G. Feller, R. Klinger, W. Benecke, Tribo-enhanced diffusion of nitrogen implanted into

steel, Materials Science and Engineering, 69 (1985) 173-180.

[88] S. Li, B. Huang, D. Li, Y. Li, S. Liang, H. Zhou, Influences of sintering atmospheres on

densification process of injection moulded gas atomised 316L stainless steel, Powder Metallurgy,

46 (2003) 241-245.

[89] R. Mariappan, S. Kumaran, T.S. Rao, Effect of sintering atmosphere on structure and

properties of austeno-ferritic stainless steels, Materials Science and Engineering: A, 517 (2009)

328-333.

[90] M. Qiu, J. Feng, Y. Fan, N. Xu, Pore evolution model of ceramic membrane during constrained

sintering, Journal of Materials Science, 44 (2009) 689-699.

[91] I. Stamenkovic, Aluminium titanate-titania ceramics synthesized by sintering and hot pressing,

Ceramics International, 15 (1989) 155-160.

[92] R.M. German, Chapter Fifteen - Sintering Practice, in: R.M. German (Ed.) Sintering: from

Empirical Observations to Scientific Principles, Butterworth-Heinemann, Boston, 2014, pp. 471-

512.

[93] H. Asgharzadeh, A. Simchi, Effect of sintering atmosphere and carbon content on the

densification and microstructure of laser-sintered M2 high-speed steel powder, Materials Science

and Engineering: A, 403 (2005) 290-298.

Page 63: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

63

[94] G.B. Schaffer, B.J. Hall, S.J. Bonner, S.H. Huo, T.B. Sercombe, The effect of the atmosphere

and the role of pore filling on the sintering of aluminium, Acta Materialia, 54 (2006) 131-138.

[95] J.F. Norton, D.J. Baxter, R. Santorelli, F. Bregani, The corrosion of AISI 310 stainless steel

exposed to sulphidizing/oxidizing/carburizing atmospheres at 600°C, Corrosion Science, 35 (1993)

1085-1090.

[96] P. Bolsaitis, K. Nagata, Kinetics of sulfidization of iron oxide with SO2-CO mixtures of high

sulfur potential, Metallurgical and Materials Transactions B, 11 (1980) 185-197.

Page 64: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

64

CHAPTER 3

EXPERIMENTAL

3.1. INTRODUCTION

This chapter is concerned with the details of the experimental aspects of this research. It begins with

the materials used to prepare the precursor of the stainless steel hollow fibre. This is followed by a

description of the techniques used to fabricate the nascent unsintered or ‘green’ hollow fibres and

the details of the sintering methods used to produce the final stainless steel hollow fibres. This

chapter also details the characterisation techniques used including thermogravimetric analysis

(TGA), mass spectroscopy coupled TGA, nitrogen adsorption, helium pycnometry, mercury

porosimetry, computerized tomography (CT) and scanning electron microscopy (SEM). The

mechanical properties of the hollow fibres were investigated using three point bending to obtain

mechanical strength. Finally, water and gas permeation tests (including pervaporation) were carried

out in order to determine the performance and stability of the membrane under a range of

conditions.

3.2. MATERIALS

During this work stainless steel AISI 316L of particle size 6, 10 and 16μm and AISI 410L of

particle size of 45μm, produced by Sandvik were used. The particle size referred to above is the

quoted size from the manufacturer, however this was independently verified used a Malvern

Mastersizer 2000 that measures particles sizes based on light scattering. Polyetherimide (PEI),

produced by SABIC Innovative Plastics and polyethersulfone (PESf) produced by Sigma Aldrich,

were used as polymeric binders in the formation of the ‘green’ hollow fibres. Synthesis grade N-

methyl-2-pyrrolidinone (NMP) from Sigma was used as the solvent, whilst reverse osmosis filtered

water (in house setup) was used as the non-solvent during the phase inversion step.

Page 65: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

65

3.3. VISCOSITY MEASUREMENTS

Viscosity measurements of the spinning dope were made using a TA ARES Rotational Rheometer.

Rotational viscometers use the principle that the torque required to turn an object in a fluid is a

function of the viscosity of that fluid. The torque required to rotate a disk in a fluid is measured and

the speed of rotation is known. The torque required to turn the spindle is quantified as the shear

stress acting across the surface area of the spindle. The fixed speed of the viscometer defines the

shear rate. Rotational velocity was varied between 0.22 to 47 rad.s-1.

3.4. PREPARATION OF MEMBRANE PRECURSORS

Membrane precursors, are prepared using the techniques adapted from the production of polymeric

and ceramic hollow fibres. In essence a polymer is used to bind the particles together to form the

hollow fibre and retain the shape during sintering. This is most commonly accomplished through a

phase inversion wet-dry spinning technique first reported for hollow fibre formation by Cabasso

and Smith in 1976 [1]. In this process the polymer is dissolved in the solvent over 2 days in a ratio

of 1:3. Once the polymer was completely dissolved, the stainless steel particles were added in the

ratio 8.5:3:1 (SS : solvent : polymer) and the mixture stirred for several hours until the mixture is

homogeneous and no lumps are observable. The mixture was degassed overnight, thus forming the

spinning dope.

The spinning dope was then extruded through an orifice spinneret into a non-solvent bath (water),

using a pressurised dry inert gas to push the slurry through. Upon reaching the bath coagulation of

the polymer / solids mixture occurred, consolidating the hollow fibre structure. The lumen of the

hollow fibre was formed by concurrently injecting the non-solvent through the bore of the spinneret

as the spinning dope was extruded into the non-solvent bath. The process is depicted in Figure 3.1

below.

Page 66: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

66

Nitrogen Bottle

Spinning dope

Coagulation Bath

Bore Liquid

Spinneret

Air

Ga

p

Figure 3.1 – Schematics of the hollow fibre production rig.

The air gap used to spin the hollow fibres was set to be 45mm. This air gap was found to be enough

to produce the desired morphology. After spinning the hollow fibres were left in the coagulation

bath for 24 hours to finish the coagulation process.

Table 3.1 below show the parameters used to prepare the green hollow fibres used in this work.

Table 3.1 – Fabrication parameters used in the production of Green Hollow Fibres

Particle Loading Particle Sizes Air Gap Binder Polymer / Solvent Ratio

70%

6 µm

50 mm PEI 1:3 10 µm

16 µm

45µm

70%

6 µm

50 mm PEI 1:4 10 µm

16 µm

45µm

70%

6 µm

50 mm PESf 1:3 10 µm

16 µm

45µm

50%

6 µm

50 mm PEI 1:3 10 µm

16 µm

45µm

10%

6 µm

50 mm PEI 1:3 10 µm

16 µm

45µm

Page 67: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

67

3.5. SINTERING

The sintering step was performed using a tubular furnace, MTI GSL-1600X, that allowed the flow

of gases during the sintering step, as shown in Figure 3.2, but the ends of the tube were blocked to

hinder the back flow of air and oxygen into the sintering region. In some cases the sintering of the

hollow fibres was divided in two processes, (i) the burning of the binder at 450°C for 60 minutes in

air, and (ii) the sintering of the particles at higher temperatures in nitrogen. For other samples a

nitrogen / argon mixture was used during the whole sintering process to preserve the carbon of the

polymer precursor, as explained by Salleh et. al. in 2011 [2]. In all cases the heating and cooling

rates were kept at 5°C min-1. The sintering temperatures used varied from 950°C to1100°C. The

sintering process is depicted below in Figure 3.2.

Figure 3.2 – Schematic of sintering furnace

Table 3.2 below shows the parameters used to sinter the hollow fibres in this work.

Page 68: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

68

Table 3.2 – Fabrication parameters used in the production of hollow fibres in this work

Particle Size Dwelling Time Sintering Temperature Sintering Atmosphere Binder

6 µm 1h

950°C

Nitrogen PEI 1000°C

1050°C

1100°C

10 µm 1h

950°C

Nitrogen PEI 1000°C

1050°C

1100°C

16 µm 1h

950°C

Nitrogen PEI 1000°C

1050°C

1100°C

45µm 1h

950°C

Nitrogen PEI 1000°C

1050°C

1100°C

6 µm 4h

950°C

Nitrogen PEI 1000°C

1050°C

1100°C

10 µm 4h

950°C

Nitrogen PEI 1000°C

1050°C

1100°C

16 µm 4h

950°C

Nitrogen PEI 1000°C

1050°C

1100°C

45µm 4h

950°C

Nitrogen PEI 1000°C

1050°C

1100°C

6µm 1h

950°C

Argon PEI 1000°C

1050°C

1100°C

10µm 1h

950°C

Argon PEI 1000°C

1050°C

1100°C

16µm 1h

950°C

Argon PEI 1000°C

1050°C

1100°C

45µm 1h

950°C

Argon PEI 1000°C

1050°C

1100°C

6 µm

1h 1050°C Nitrogen PESf 10 µm

16 µm

45µm

Page 69: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

69

3.6. THERMOGRAVIMETRIC ANALYSIS (TGA)

The polymers used and ‘green’ sample was analysed using a Shimadzu TGA-50 Thermogravimetric

analyser, in order to select the adequate debinding temperature, and simulate the sintering

conditions. By coupling the TGA with a Mass Spectrometer (Netzsch STA 449F1 TGA – DTA), it

was possible to observe the compounds released during these processes.

3.7. HELIUM PYCNOMETRY

Skeletal density and specific volumes of the membranes were measured using an AccuPyc 1340

Gas pycnometer at room temperature using helium (99.995% pure), as working fluid, due to its

molecular size. For this analysis, the machine initially degased the samples by applying vacuum,

and purged with helium before measuring the volume is measured ten times to obtain an average.

The density was calculated by measuring the sample mass and dividing this value by the volume

obtained from pycnometry.

3.8. NITROGEN ADSORPTION

Porosity and pore sizes can be inferred by using nitrogen adsorption techniques. Sing [3] describes

the process of using nitrogen adsorption for pore characterization.

Nitrogen adsorption experiments were carried out on a Micromeritics Tristar 3020 System

(Micromeritics Instrument Corporation) and the volumetric method was used to calculate the

amount of nitrogen adsorbed. Before this experiment, the samples were degassed at 250 °C for 24 h

under high vacuum. The degassed samples were placed in a glass cell station and the nitrogen

adsorption process was carried out at -196.15°C (77 K) using a liquid nitrogen bath to maintain

temperature constant. Pressure range between 0.001 kPa and 101 kPa was used to characterize pore

structure. BET method was used to calculate surface area [4].

A variant of the nitrogen adsorption experiments was carried out in the same machine using CO2 as

adsorbent. The temperatures used to do such test were room temperature (20°C), 0°C using a

Page 70: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

70

water/ice slurry to maintain the temperature content, and a slurry consistent in acetone and 50°C.

This process allowed to calculate the heat of adsorption by using the equations presented by

Thornton et al. [5].

3.9. MERCURY POROSIMETERY

Mercury porosimetry was used to measure pores in the meso- and macropore size regions. The

equipment used is a Micromeritics AutoPore IV9520. Briefly, the mercury is intruded under

pressure to the sample. This pressure values are directly converted into the corresponding pore size

using the Washburn equation. However, one should be aware that mercury porosimetry does not

actually measures the internal pore size, but it rather determines the largest connection (throat or

pore channel) from the sample surface towards that pore. Thus, mercury porosimetry results will

always show smaller pore sizes compared with other techniques [6].

For the analysis, the contact angle was set to be 130°, as recommended by Ellison et al.[7]. The key

assumption of cylindrical pores was used and the maximum pressure was 415MPa (60,000psia).

The lower limit of measurement is 3nm.

3.10. SCANNING ELECTRON MICROCOPY (SEM)

The surface morphological features of the produced hollow fibres membranes were examined using

a JEOL JSM-7001F SEM with a hot (Schottky) electron gun at accelerating voltage of 10 kV.

Samples were prepared by removing one centimetre of the uncoated ends of the tubular substrates

before breaking the membrane and selecting two segments for cross-section and surface

measurements, respectively. The segments were then mounted on SEM stubs and platinum coated

using a Baltec coating apparatus in high purity argon.

Energy dispersive X-ray spectroscopy (EDX) images were also obtained using a Zeiss Ultra55,

EDX INCAEnergy355 SEM – EDX machine. Surface, cross-section and samples embedded in

epoxy resin were analysed.

Page 71: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

71

3.11. COMPUTERIZED X-RAY MICROTOMOGRAPHY (µ-CT)

Computer microtomography was used as an additional technique to characterize the materials and

measure the pore size and porosity of the samples. This technique was also very helpful to visualise

the pore network and identify interconnected pores. For μCT the samples were imaged under the

same conditions, using the Skyscan 1172 high-resolution desktop X-ray microtomography system

(Skyscan, Belgium). For image processing and analysis, the skyscan software, CTAnalyser (CTAn)

was used.

3.12. MECHANICAL STRENGTH TESTS.

Three point bending tests performed on an Instron 5543 universal testing machine were used to

characterize the mechanical resistance of the hollow fibres produced in this work. The strain rate

was set at 1 mm min-1 and the sample span was 20 mm. The bending stress was calculated using

the 44

8

io

o

DD

FKD

Eq. 3.1 adapted

from [8]:

44

8

io

o

DD

FKD

Eq. 3.1

where σ is the bending stress (MPa), F is the load (N), K is the span (mm), Do and Di are the outer

and inner diameters (mm), respectively, as depicted in Figure 3.3.

Figure 3.3 – Three point bending test

Page 72: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

72

3.13. INDUCTIVELY-COUPLED PLASMA – OPTICAL EMISSION

SPECTROMETRY ANALYSIS

The chemical composition of the sintered hollow fibres was measured using Inductively-Coupled

Plasma – Optical Emission Spectrometry. This technique is used very commonly to determine the

chemical composition and traces of elements.

The ICP/OES is based upon the spontaneous emission of photons from atoms and ions that have

been excited in a RF discharge [9].

Only liquid and gas samples may be injected directly into the instrument, while solid samples

require extraction or acid digestion so that the elements will be present in a solution. The solution to

be tested is converted to aerosol and directed into the central channel of the plasma. At its core the

inductively coupled plasma (ICP) sustains a temperature of approximately 10,000 K, so the aerosol

is quickly vaporized. Elements present in the solution to be analysed are liberated as free atoms in

the gaseous state. Further collisional excitation within the plasma imparts additional energy to the

atoms, promoting them to excited states. Sufficient energy is often available to convert the atoms to

ions and subsequently promote the ions to excited states. Both the atomic and ionic excited state

species may then relax to the ground state via the emission of a photon. These photons have

characteristic energies that are determined by the quantized energy level structure for the atoms or

ions. Thus the wavelength of the photons can be used to identify the elements from which they

originated. The total number of photons is directly proportional to the concentration of the

originating element in the sample [9, 10].

The instrumentation associated with an ICP/OES system is relatively simple. A portion of the

photons emitted by the ICP is collected with a lens or a concave mirror. This focusing optic forms

an image of the ICP on the entrance aperture of a wavelength selection device such as a

monochromator. The particular wavelength exiting the monochromator is converted to an electrical

signal by a photo detector. The signal is amplified and processed by the detector electronics, then

displayed and stored by a computer.

The specifics of the method used in this work are:

Page 73: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

73

The method is suited for multi-element determination in solutions or after appropriate

sample preparation of solid samples brought into solution.

Detectable elements: approx. 70, determinable concentration range from a few µg/l up to 2

% in solution or 1 µg/g up to 100 % in solids.

Precision: 1 - 3 % for major elements, ± 10-30 % for traces.

Particularly suited for stoichiometry determinations, material controls, material

determinations

Test quantities at least 1-0 mg for major elements, 100-500 mg for traces.

Measurement times of generally a few minutes per sample; (for solid samples, a possibly

time-consuming sample preparation must be taken into account).

In this work the analysis was carried out in the Forschungszentrum Julich, in Germany, and the

machine available there has the following components:

TJA-IRIS-INTREPID: spectrometer with echelle optics and CID semiconductor detector,

axial and radial viewing, wavelength range 165 - 900 nm

TJA-IRIS-Advantage: spectrometer with echelle optics and CID semiconductor detector,

axial and radial viewing, wavelength range 170 - 900 nm

Thermo Scientific iCAP6500: spectrometer with echelle-optics and CID semiconductor

detector, axial and radial viewing, wavelength range 166-847 nm

The sample must be prepared by digestion methods such as microwave, high-pressure, fusion, and

acid digestion are employed for the sample preparation of solid sample material. Examples of

possible materials are steels, nickel-base alloys, nonferrous metals, light metals, ores, rock,

minerals, coals, ashes, glasses, ceramics, water and waste water samples, nutrient solutions, oils,

bulk analysis of thin films, high-temperature superconductors, mixed oxides, perovskites,

electroceramic materials, zirconium oxides and others.

3.14. PERMEATION TESTS

3.14.1. Single gas permeation tests

The gas permeation rig is depicted in Figure 3.4. A hollow fibre membrane was placed in a module

of small volume attached to a larger vessel of known volume. The larger vessel of known volume

Page 74: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

74

was pressurised with nitrogen gas. The experiment began by opening the valve connecting the

membrane module and the pressure vessel. The change in pressure in the permeate side was

subsequently monitored over a period of time. The low pressure transducer (MKS Baratron 122BA-

0100AB) measured the pressure in the permeate side, and the high pressure transducer (MKS

Baratron 122BA-10000BB) measured the pressure in the feed and the retentate side. The permeance

can be calculated then using the following expression:

Eq. 3.2

Where (P/L) is the permeance in mol s-1 m-2Pa-1, V is the volume of the system in m3, A is the

surface area of the membrane in m2, R is the universal constant of gases in J K−1 mol−1 ,T is the

temperature in K and t is time in s.

Known Volume

P

Low pressure transducer

PC

3 way valve / purge

Pressurised Gas

P

High pressure transducer

Membrane Module

Membrane

Figure 3.4 – Gas permeation rig setup

3.14.2. Binary gas permeation tests

Binary gas permeation tests were carried out using the setup shown in the schematic picture of

Figure 3.5. Briefly, gas was mixed by joining two gas lines with controlled volumetric flow each.

The mass flow of each gas was controlled using a volumetric flow control, and the fraction of each

gas was calculated using the universal law of gas. The feed composition was controlled after by

taking samples and injecting into a Shimadzu GC-2014 gas chromatograph (GC).

Page 75: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

75

Feed

Retentate

Permeate

F

Bubble flow meter

F

Bubble flow meter

Gas Chromatograph

Figure 3.5 – Binary gas permeation rig

A single hollow fibre was placed in a module. The module temperature was controlled using a

furnace with automatic temperature switch. The permeate and retentate mass flow were measured

using a bubble flow meter each, and the composition was measured using a GC. A mass balance

was used to calculate the separation factor (α) [11], as described in the equations below.

Eq. 3.3

Eq. 3.4

Eq. 3.5

Eq. 3.6

where, is the mass flow, x is the fraction of the component i in the stream, and is the trans-

membrane pressure difference.

3.15. UNCERTAINTY WITHIN EXPERIMENTS

The measurement of physical magnitudes is always subject to uncertainties and variations due to

several factors [12]. Hence, a brief discussion of the uncertainties is provided herein after. The first

Page 76: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

76

source of uncertainty is related to the raw materials used to produce the hollow fibres. Since the

production process of these materials were subject to process variables leading to variations in

chemical properties and particle size. For instance, SS powders from Sandvik are reported to have

typical particle size distribution as shown in Table 3.3. This particle size distribution would result in

variation in the uniformity in which densification occurs, since the particle size plays a role in the

mass transfer during densification [13] as will be discussed later in chapter 5.

Table 3.3– Particle size distribution as reported by the manufacturer

Typical Particle Size Distribution (µm)

D 90 (%) D 50 (%) D 10 (%)

90 % - 6µm < 6 < 3.5 < 2.1

90 % - 10µm < 10 < 6 < 3.1

90 % - 16µm < 16 < 9 < 3.5

90 % - 45µm < 45 < 30 < 9.4

Another source of uncertainty is in mixing the materials for the preparation of the hollow fibres.

This uncertainty comes related to measuring the amount of each component of the hollow fibre. As

bulk materials were mixed, even after using precision balances, measurement errors of 1% or higher

in the composition of the green hollow fibres are expected.

The effect of particle size and particle size distribution may play a role on the measurement of the

viscosity, as described by Eagland and Kay [14], but again these variation can be negligible

compared to other parameters used during fabrication of the hollow fibres. Uncertainties regarding

the fabrication process are mostly associated to parameters used to control the spinning conditions.

Nitrogen pressure is used to promote flow of the spinning dope into the spinneret and water flow

through the bore. This pressure is controlled by using manometers that have been proved to provide

small errors of between 0.02 and 0.04% [15].

Another source of error is in the measurement of the air gap. The air gap was measured using a

metallic ruler, and even though the precision of the ruler is 1 mm, errors of 2-5 mm are expected to

Page 77: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

77

measure accurately the position of the coagulation bath surface. This can result in changes in the

morphology due to change in the time in which the hollow fibre reach the coagulation bath.

Uncertainties related to the sintering process are mostly due to temperature control and atmosphere

control. The temperature of the furnace was controlled using a PID temperature controller and

fluctuations of ± 1°C are possible according to the manufacturer. The other source of variability is

in the control of the sintering atmosphere. Volumetric control was used to ensure that the flow of

gas was constant, and this control had relative large variations around 5 - 10%. To overcome this

situation large flux of inert gas was used in order to avoid back diffusion and minimize the effect of

leaks during sintering.

Helium pycnometry is essential to know the density of the samples, and its measurements were

subsequently used in the mercury porosimeter. It is known that typical errors in density for the

helium pycnometry are between 2 and 5% [6]. The mercury porosimetry technique calculated the

change of volume due to increase in pressure to pore size and pore volume. Thus, the accurate

values of pressure and volume are required. Pores as small as 3 nm can be measured and the

resolution of the machine, as reported by the manufacturer, for the intrusion volume is at least of

0.1µL and the accuracy is set to be ±1% of maximum penetrometer stem volume. However the

technique is highly sensitive to the amount of sample that is used during the test, as a very low

loading of sample may result in inaccurate readings due to the relative large size of the sample

chamber compared to the penetrometer stem, especially at higher pressures [6]. In addition surface

contamination can result in measurement errors.

One of the large uncertainties in this work are related to the measurements of the mechanical

strength of the hollow fibres and errors over 30% are quite common [16]. The load cell used to

measure the forces necessary to break has an uncertainty of 0.01%, and this is found to be too small

to produce significant errors on the values measured. The morphology and internal defects,

however, are to be considered as cause of variation between the values recorded for maximum

flexural stress and strain, since defects can help reduce the mechanical resistance of the hollow

fibre. The same can be said of hollow fibres that present high degrees of eccentricity, and when

possible testing those fibres was avoided. Eccentricity results in thinner and thicker sections of the

hollow fibre wall producing a weak part that will break with lower mechanical stresses [17].

Experimental calculation errors were allocated to the difficulties in measuring the dimensions of the

hollow fibres. In turn, these errors also affect the calculation for maximum stresses held by the

hollow fibre. The errors in measuring the outer diameter of the hollow fibre are related to the

Page 78: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

78

precision of the calliper, and this uncertainty is of 0.01mm. The other source of uncertainly in the

outer diameter of the hollow fibre is due to difference in circularity with variations as high as

0.15mm were recorded. Probably the most important source of uncertainty in measurement was

related to the inner diameter of the hollow fibre, since it is particularly difficult to measure due to its

small dimension and values for the inner diameter resulted in variations of around 0.1mm. The

combination of all these uncertainties resulted in variation of the maximum flexural stress of around

20 % for the same type of hollow fibre, however each measurement was shown with error bars

showing variation.

In the single gas permeation experiments there are uncertainties related to the volume, pressure

transducer and surface area of the hollow fibre. The measurement resolution for these pressure

transducers are of around 0.001% on digital output and the accuracy due to non-linearity is of

0.25% of the reading. The volume used was a calibrated double ended 400 cm3 vessel DOT – 4B

provided by Swagelok with a standard deviation of the nominal volume on ± 5%, and the volume

was measured upon testing using the universal gas law to confirm the volume size. All these errors

combined resulted in an error smaller than 0.5%. The large error in this measurement is related to

measurement of the permeation area, errors in the measurement of the diameters and the length

result in a surface area with an error of around 5%. The calculation of permeance uses all the

parameters mentioned, and the combined errors are calculated to be around 10%.

Binary gas permeation tests shared errors in measurement the permeate area with the single gas test,

since the same hollow fibre was tested, but different to single gas permeation tests the binary

permeation tests were carried out using bubble gas flow meter. This process adds further

uncertainty to the process of reading the mass flow, by combining parallaxes errors of reading the

bubble flow meter and errors in operation the chronometer, and those errors are believed to be

between 2 and 7% [18]. In addition the use and calibration of GC and the measurement of the

surface area below the peak of each gas has a large error associated and errors of 15% were found,

however these values are within expected and according to literature [19].

3.16. ERRORS IN CONCENTRICITY OF THE SURFACES IN HOLLOW

FIBRE PRODUCTION

The wall thickness of the hollow fibre can be controlled by changing the flow of water inside the

bore. An increase in the bore liquid flow rate will create more pressure in the walls, resulting in a

larger inner hole and subsequently in a thinner hollow fibre wall. The opposite will result in a

Page 79: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

79

smaller hole limited to the size of the needle used to inject water. As mentioned before, the hollow

fibre is produced by extruding the mixture of polymer in a concomitant manner with the injection of

water to promote initial coagulation, as shown in a schematic drawing displayed in Figure 3.6

below.

Figure 3.6 – Schematics of the spinning process used to produce hollow fibres.

This production process is susceptible to alignment problems that are caused mainly by two

possible situations. The first situation is depicted in Figure 3.7a and occurs when the surface of the

spinneret and the coagulation bath are not parallel. This causes the spinning dope to leave the

spinneret with an angle causing the hole in the centre to be slightly oval and out of concentricity, it

is noteworthy to mention that even a small misalignment can produce this.

Page 80: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

80

Figure 3.7 – Schematics showing the cause of errors of concentricity in the extrusion process of a

hollow fibre, (a) caused by tilting the spinneret and (b) caused misalignment between the non-

solvent injection needle and the spinneret hole.

The second situation occurs when the spinneret is not properly assembled or the needle is bent

causing the hole and the outer wall not to be concentric, as shown in Figure 3.7b. Small

misalignments between the non-solvent injection needle and the spinneret hole can be easily

produced during the assembly process and are very hard to identify before spinning. This problem

will result in a perfect circular hole but non-concentric to the hollow fibre outer surface.

3.17. REFERENCES

[1] Cabasso I, Klein E, Smith JK. Polysulfone hollow fibers. I. Spinning and properties. Journal of

Applied Polymer Science 1976; 20:2377-94.

[2] Salleh WNW, Ismail AF. Carbon hollow fiber membranes derived from PEI/PVP for gas

separation. Separation and Purification Technology 2011; 80:541-8.

[3] Sing K. The use of nitrogen adsorption for the characterisation of porous materials. Colloids and

Surfaces A: Physicochemical and Engineering Aspects 2001; 187–188:3-9.

[4] Brunauer S, Emmett PH, Teller E. Adsorption of Gases in Multimolecular Layers. Journal of the

American Chemical Society 1938; 60:309-19.

[5] Thornton AW, Hilder T, Hill AJ, Hill JM. Predicting gas diffusion regime within pores of

different size, shape and composition. Journal of Membrane Science 2009; 336:101-8.

a b

Page 81: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

81

[6] Giesche H. Mercury Porosimetry: A General (Practical) Overview. Particle & Particle Systems

Characterization 2006; 23:9-19.

[7] Ellison AH, Klemm RB, Schwartz AM, Grubb LS, Petrash DA. Contact angles of mercury on

various surfaces and the effect of temperature. Journal of Chemical & Engineering Data 1967;

12:607-9.

[8] Dewolf JT, Beer FP, Johnston ER. Mechanics of materials. New York: McGraw-Hill Higher

Education; 2009.

[9] Hou X, Jones BT. Inductively Coupled Plasma‐Optical Emission Spectrometry. Encyclopedia of

Analytical Chemistry 2000.

[10] Hou X. Inductively Coupled Plasma-Optical Emission Spectrometry. 2006.

[11] Brunetti A, Barbieri G, Drioli E, Lee KH, Sea B, Lee DW. WGS reaction in a membrane

reactor using a porous stainless steel supported silica membrane. Chemical Engineering and

Processing 2007; 46:119-26.

[12] Drosg M. Dealing with uncertainties: a guide to error analysis. Berlin: Springer Berlin

Heidelberg; 2007.

[13] Kingery WD, Berg M. Study of the Initial Stages of Sintering Solids by Viscous Flow,

Evaporation&#x2010;Condensation, and Self&#x2010;Diffusion. Journal of Applied Physics 1955;

26:1205-12.

[14] Eagland D, Kay M. The rheological properties of concentrated polymer dispersions: I. The

effects of concentration, particle size, and size distribution upon the shear dependence of viscosity.

Journal of Colloid and Interface Science 1970; 34:249-61.

[15] Zhokhovskii MK. Experimental determination of errors in piston manometers at high

pressures. Meas Tech 1959; 2:505-9.

[16] Baratta FI, Matthews WT, Quinn GD. Errors associated with flexure testing of brittle materials.

DTIC Document; 1987.

[17] Freiman SW, Mecholsky JJ. The fracture of brittle materials: testing and analysis. Hoboken,

N.J: John Wiley & Sons; 2012.

Page 82: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

82

[18] Dondi F, Betti A, Bighi C. Evaluation of errors in gas chromatography. Journal of

Chromatography A 1971; 60:1-13.

[19] Barwick VJ. Sources of uncertainty in gas chromatography and high-performance liquid

chromatography. Journal of Chromatography A 1999; 849:13-33.

Page 83: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

83

CHAPTER 4

FORMATION OF ASYMMETRIC HOLLOW FIBRE

PRODUCED VIA PHASE INVERSION OF A STAINLESS

STEEL / POLYMER / SOLVENT / NON-SOLVENT

SYSTEM.

4.1. ABSTRACT

Unsintered stainless steel hollow fibres were prepared via a dry-wet phase inversion process and the

impact that a variety of carefully selected parameters have on both the process and the subsequent

hollow fibre morphology were studied. In particular, this work investigated the effect of the choice

of binder, the stainless steel particle size, the stainless steel solid load, and the addition of a

commonly used viscosity enhancer Polyvinylpyrrolidone (PVP) on the hollow fibre precursor

morphology.

The choice of the polymeric binder must carefully considered based on the particular application

that the hollow fibre will have, and on the spinning conditions used. In this case two polymeric

binders were trialled, polyethersulfone (PESf) and polyetherimide (PEI) with a N-methyl

pyrrolidone (NMP) solvent. Whilst the addition of stainless steel particles altered the relative

viscosity of the spinning dope, the polymer solvent interactions proved to be the most significant in

forming the morphological features of the green hollow fibres.

Viscosity of the spinning dope is commonly used to infer the morphology that will result after

phase inversion. Low viscosity spinning dopes are associated with the formation of more finger-like

macrovoids. By contrast spinning dopes with higher viscosity promote the formation of larger

sponge-like regions. Here we show that using viscosity as a predictive measure can only be applied

to relatively similar systems. When the viscosity is modified (i.e. increased) by using PVP, the

result is contrary to expectations, as hollow fibres containing PVP showed more and larger finger-

Page 84: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

84

like macrovoids. This counter-intuitive finding was attributed to altered solvent / non-solvent

exchange dynamics as a result of the interaction between PVP and the non-solvent.

The effect of particle loading and particle size on the properties of the spinning dope and

subsequent hollow fibre morphology was studied in detail. In this chapter the influence of particle

size and loading on the solvent / non-solvent exchange kinetics was considered analogous to fluid

flow through a packed bed. Higher particles loadings act as a restriction to the phase inversion

process, resulting in a binary morphology with finger-like macrovoids near the lumen and a sponge-

like region near the outer surface. Lower solid loading resulted in a ternary structure composed of

finger-like macrovoids – sponge-like structure – finger-like macrovoids as you progress radially

from the hollow fibre lumen. The effect of particle size was also studied and the changes in

morphology were dramatic despite maintaining constant solids loading. Large particles did not

significantly affect solvent velocity and as a result, hollow fibres with large amount of finger-like

pores were observed. In comparison a particle size reduction from 45μm to 6μm yielded a 98%

reduction in solvent velocity and this was reflected in the production of a large sponge-like region.

Page 85: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

85

4.2. INTRODUCTION

Porous supports are commonly used for microfiltration, nanofiltration, and as substrates for

molecular sieving membranes for both liquid and gas separation processes. Conventionally, porous

substrates must have characteristics of high porosity, and tailored narrow pores sizes. Commonly

porous substrates are made from inorganic materials such as alumina, titania, zirconia or carbon,

particularly to meet the high temperatures and relatively good mechanical requirements otherwise

not available in polymeric supports [1]. However, many industrial processes require high pressure,

high temperatures accompanied by vibration and mechanical stresses [2]. As a result, porous

ceramic supports tend to fail generally at the seal and membrane interface [3]. Therefore, membrane

mechanical robustness is becoming an important industrial challenge.

Several different geometries including flat substrates [4, 5], tubular substrates [6] and hollow fibres

[7-10] have been produced through tape casting [11], pressing [12], extrusion [13] and dry-wet

phase inversion spinning techniques [14, 15]. Flat membranes are the most commonly investigated

geometry due to the relative simplicity of producing defect free, thin film membranes [16] and the

ease with which they can be incorporated into laboratory scale membrane modules. However, flat

membranes have the lowest surface area to volume ratio and the largest sealing perimeter of any of

the investigated geometries, which gives rise to larger module sizes and plant footprints, increasing

capital costs and reducing their industrial practicability. By contrast hollow fibres have the highest

surface area to volume ratio and smallest sealing perimeter, resulting in smaller unit operations and

lower capital costs. Therefore, robust hollow fibre substrates are of significant interest to the

membrane community and recently there have been a limited number of studies reporting the

fabrication of porous metallic hollow fibres based on nickel and stainless steel (SS).

The development of porous metallic hollow fibres was adapted from ceramic hollow fibre

processing techniques. In essence a polymer is used to bind the particles together to form the hollow

fibre and retain the shape during sintering. This is most commonly accomplished through a phase

inversion wet-dry spinning technique first reported for hollow fibre formation by Cabasso and

Smith in 1976 [17]. In this process the polymer is dissolved in a suitable solvent, followed by

suspension of the solids (in this case metal particles) to form the spinning dope. The spinning dope

is then extruded through an orifice spinneret into a non-solvent bath where coagulation of the

polymer / solids mixture occurs, solidifying the hollow fibre structure. The lumen of the hollow

Page 86: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

86

fibre is formed by concurrently injecting the non-solvent through the bore of the spinneret as the

spinning dope is extruded into the non-solvent bath. During this process the morphology of the

hollow fibre is formed. This morphology, it will be later responsible for the characteristics of pore

size, pore size distribution, and even in some extent mechanical robustness. The microstructure and

morphology of the green hollow fibre are principally controlled by the polymer / solvent / non

solvent interactions during the wet-dry phase inversion process. These interactions, which include

both dynamic diffusion rates and phase equilibrium, are complex and have been the subject of

numerous studies since the 1970s [18-20]. The final hollow fibre porosity can be adjusted via the

sintering process; however its morphology is primarily set during the phase inversion process [21-

23].

During the phase inversion process, there are two primary microstructures that can be formed

namely a sponge-like or larger finger-like structure [21, 24]. The sponge-like structure is generally

favoured for high strength membrane substrates with highly tailored pore sizes and narrow pore size

distributions. This morphology is formed in situations where the polymer / non solvent interactions

are relatively weak and the rate of diffusion of the solvent to the outside of the hollow fibre is

reduced [25]. The finger-like macrovoids, on the other hand, are formed when the polymer / non-

solvent interactions are relatively strong and the rate of diffusion of solvent out of the hollow fibre

is fast [18]. This creates large polymer-lean regions where the solvent / non solvent mixing can take

place.

In the literature several authors present works using the phase inversion technique in systems that

contain polymer, solvent and non-solvent, and particles in suspension. The presence of particles not

only affects the viscosity of the spinning dope, but it is also thought to affect the rate in which the

solvent – non-solvent interactions occur. The change in those diffusion rates will lead to different

hollow fibre morphology, and consequently different membrane characteristics. The objective of

this chapter is to investigate the effect that the addition of solid, spherical stainless steel particles

will have in the characteristics of the spinning dope and the resulting green hollow fibre

morphology by changing ratios in the particle, solvent, polymer and non-solvent system. In

particular, this chapter will investigate the effect of the choice of binder, the stainless steel particle

size, the stainless steel solid load, and the addition of a commonly used viscosity enhancer

Polyvinylpyrrolidone (PVP) on the hollow fibre precursor morphology.

Page 87: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

87

4.3. CHOICE OF BINDER

The selection of the binder is an important step in achieving both the desired macrostructure of the

green hollow fibre as well as the final morphology and mechanical strength of the sintered support.

Polyethersulfone (PESf), is the most common polymer used in the production of inorganic hollow

fibres [26-29], however it has been demonstrated that the sulphur contained in this polymer can

cause membrane poisoning under some circumstances [30]. Polyetherimide (PEI) on other hand has

been successfully used to produce hollow fibres without compromising the selective characteristics

and mechanical properties of the membrane [30]. As the two most common binders used in the

existing literature, they form the basis of this investigation.

In each case the spinning dope was formed according to the techniques described in Chapter 3.

Briefly, the polymer of choice was dissolved in the solvent (1-methyl-2-pyrrolidinone (NMP) from

Sigma) in the ratio of 1:3 in mass. Once the polymer was completely dissolved, the stainless steel

particles were added in the ratio 8.5:3:1 (SS: solvent: polymer) in mass and the mixture stirred until

homogeneous after which it was degassed to form the spinning dope.

The first step in the study, even before the green hollow fibre is produced, is to measure the

viscosity of both the virgin polymers and of the polymer and stainless steel particle mixtures. This

was accomplished using a TA ARES Rotational Rheometer. An understanding of the properties of

the spinning dope yields a more accurate prediction and/or design of the final macrostructure of the

green hollow fibres. In this case it conveys information about how each spinning dope (i.e. PESf

and PEI) will perform under identical spinning conditions. The viscosity results for each polymer

and polymer/SS mixture are displayed in Figure 4.1 as a function of increasing rotational force.

Page 88: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

88

[rad/s]

1 10 100

* [P

a.s

]

10

100

1000

PESf

PESf + 10µm SS

PEI

PEI + 10µm SS

Figure 4.1 – Viscosity results as a function of rotational speed for PESf (circle) and PEI (triangle)

dissolved in NMP in the ratio of 1:3 (closed symbols) and the same mixtures including stainless

steel particles (open symbols).

The results from Figure 4.1 show that the viscosity of the PEI solution is 3 times greater than for the

PESf solution. Interestingly, the viscosity of the PESf solution is unchanged with increasing

rotational speed, strongly suggesting that it behaves as a Newtonian fluid, whilst for PEI the

viscosity decreases marginally as rotational speed increases suggesting marginal shear-thinning

behaviour. As expected, addition of stainless steel particles to both polymer solutions increased the

mixture viscosity, by way of additional polymer/particle interactions [31-34]. In both cases the

viscosity was increased by approximate 2.4 times. The addition of the stainless steel particles did

not alter either the Newtonian behaviour of the PESf mixture or the shear-thinning behaviour of the

PEI mixture, although it did appear to enhance the later. Based on this characterisation it is

hypothesized that the larger viscosity of the PEI+SS mixture will see a more sponge-like structure

formed in the green hollow fibres as compared to the PESf+SS mixture when spun under the same

conditions. This is because higher viscosities slow down the rate of solvent diffusion out of the

hollow fibre during the demixing process [35-37]. However, it must be noted that the faster the

extrusion speed of the hollow fibre from the spinneret the closer the final macrostructure of the

PEI+SS and PESf+SS mixtures will be, owing to the decreased viscosity at higher shear rates for

the PEI+SS mixture.

Next, the PEI+SS and PESf+SS mixtures were spun into hollow fibres according to the methods

described previously in Chapter 3. Briefly, the spinning dope was extruded through an orifice

Page 89: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

89

spinneret (at a rate of 3 mm/sec) into a non-solvent bath (water), using a pressurised (5 bar) dry

inert gas to push the dope through. Coagulation of the polymer / solids mixture occurred when the

mixture reached the bath, consolidating the hollow fibre structure. The lumen of the hollow fibre

was formed by concurrently injecting the non-solvent (at a flow rate of 45 ml/min) through the bore

of the spinneret as the spinning dope was extruded into the non-solvent bath. Identical conditions

were used for both PEI+SS and PESf+SS mixtures, which should yield very different morphologies

based on the viscosity data.

The phase inversion process is the fundamental mechanism by which the morphology of the hollow

fibre is formed. Whilst sintering can change the morphology between the green and final fibre, it

takes high temperatures and long hold times to induce substantial alterations. Figure 4.2 shows the

morphology obtained by using PESf (a) and PEI (b) to produce stainless steel hollow fibres.

Figure 4.2 – Morphology resulting after sintering of the stainless steel hollow fibre prepared with

PESf (A) and PEI (B).

As predicted from the viscosity data, there is significant difference in the morphology of the

PESf+SS (Figure 4.2 (A)) and PEI+SS (Figure 4.2 (B)). The sample which utilised PESf as a binder

showed larger voids or finger-like pores and greater apparent porosity. In particular, large pores

(between 0.5 and 30μm in diameter) are located along both the outer and inner surfaces of hollow

fibre, whilst very large finger-like voids (from 50 to 130μm in diameter) are located in the centre of

the fibre. In comparison, samples which utilized PEI as a binder showed significantly fewer finger-

Lumen Lumen

Page 90: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

90

like voids and lower apparent porosity. In particular, no large pores where evident at the outer

surface of the hollow fibre, with a sponge-like structure extended past the centre point of the cross-

section. There were a number of finger-like voids towards the lumen side of the fibre, but these

were smaller in size (in the range of 15 to 50μm in diameter) when compared to the PESf+SS

sample. Notably, in the immediate area next to the inner surface, the sponge-like structure was

again prevalent for both the PESf+SS and PEI+SS hollow fibres, although this region was thicker in

the PEI+SS sample (approximately 200 μm thick as compared to 50 μm thick for PESf+SS).

It is noteworthy to observe that these results were obtained using identical spinning parameters and

indicate that the exchange rate of solvent and non-solvent in both polymers is significantly different

during the demixing / coagulation process. This remarkable difference is a consequence of a

complex series of interactions. In particular both the difference in apparent viscosities and the

difference in the precipitation value, that is, the amount of non-solvent required to initiate

coagulation of PESf or PEI [38] play a significant role. As discussed by Wang et al. [22, 23, 25],

PEI has a lower precipitation value than PESf, meaning that a smaller amount of non-solvent is

required to initiate coagulation (see Table 4.1 for Flory interaction parameters). Therefore, if the

other spinning parameters, such as air gap and extrusion rate, remain constant, PEI allows slower

diffusion of the solvent towards the non-solvent bore liquid, which creates a solvent lean region at

the outer hollow fibre wall. When the hollow fibre contacts the coagulation bath, the mass transfer

of solvent into the bath from the solvent-lean outer surface occurs at a slower rate due to the

reduced concentration gradient. On the contrary, PESf, has a higher interaction value and thus has a

higher diffusion rate of the solvent towards to the non-solvent in the bore liquid, which

consequently does not promote a solvent lean region at the outer hollow fibre wall. Therefore, when

the nascent PESf hollow fibre [7] reaches the coagulation bath, the outer wall is still a solvent rich

region. This leads to a high flux, due to the high concentration gradient, of solvent from the outer

surface of the hollow fibre into the coagulation bath forming finger-like macrovoids [7, 22, 23, 25,

39] as previously described for the inner surface.

Table 4.1– Flory interaction parameters for water (1) / solvent (2) / polymer (3) systems[40]

Polymer χ12 χ13 χ23

PES 0.431 2.5 0.434

PEI 0.431 3.43 0.27

Page 91: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

91

To better understand the morphology of the PESf+SS and PEI+SS samples, mercury porosimetry

was undertaken to fully characterise the pore-size distribution and porosity of the sintered hollow

fibres. Briefly, this was done using a Micromeritics AutoPore IV9520 where the contact angle was

set to be 130°, as recommended by Ellison et al. [41], and the maximum pressure was 400MPa. The

results of the porosimetry analysis are shown in Figure 4.3.

Pore Radius (µm)

0.0010.010.11101001000

Po

rosity

(%)

0

5

10

15

20

25

30

35

Samples with PEI

Samples with PESF

Figure 4.3 – Mercury porosimetry results for the fibres prepared using PESf and PEI polymers.

The difference in both porosity and pore size distribution between the PESf+SS and PEI+SS

samples, as determined by mercury porosimetry, is as striking as the low magnification SEM

images in Figure 4.2. The results clearly indicate that the PESf+SS sample had a porosity of ~23%,

while of the PEI+SS sample had a porosity of ~28.5%. Furthermore, the pore sizes within the

PESf+SS sample were predominately distributed around a single, large peak between 2 and 0.5 μm,

although a range of pores are measured between 30 and 2μm. By contrast the pore size distribution

of the PEI+SS was clearly trimodal with a small peak between 2 and 1μm, a larger peak between 1

and 0.1 μm and the final peak around 0.1 to 0.08 µm. The porosity contribution of each pore size

was 6.5%, 8% and 10% respectively. It is important to note here the polymer is still present during

mercury porosimetry testing and that the final porosity and pore size distribution of the sintered

fibre will be different from the results presented here. Indeed, the mid-range and smaller pores (i.e.

<1 μm) are attributed to either gaps between the polymer and SS particles (i.e. imperfect wetting of

the SS by the polymer) and/or pores within the polymer itself arising from the phase inversion

Page 92: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

92

process. The larger micron sized pores are related to the large, finger-like voids present in both

fibres and confirm the SEM observations that the PESf+SS sample contained many more than the

PEI+SS sample. Therefore, these results are congruent with the SEM analysis and provide more

evidence that the solvent / non solvent exchange rates are significantly different for the PESf and

PEI systems as previously discussed.

4.4. EFFECT OF THE VISCOSITY OF THE SPINNING DOPE

In the production of polymeric hollow fibres it has been reported that the change in viscosity of the

spinning dope will lead to a change in the morphology of the hollow fibre produced [36, 37]. Fast

interactions are associated with low viscosity spinning dopes and larger finger-like macrovoids;

slow interactions are associated with high viscosity spinning dopes and smaller sponge-like pores.

Changing the viscosity of the spinning dope can be achieved in two distinct ways. The first involves

changing the ratio of existing components that make up the spinning dope, whilst the second

involves the addition of a viscosity modifier. Both of these methods will have impacts beyond

altering the kinetics of the solvent/non-solvent interactions, such as changing ternary (or even

quaternary phase diagram) and in turn, the coagulation point of the system. Hence it is complex to

isolate only the impact of the viscosity on the nascent hollow fibre morphology. Nonetheless, both

methods were investigated in this section including adjusting the ratio of polymer to solvent in the

dope and the addition of Polyvinylpyrrolidone (PVP) as a viscosity modifier.

4.4.1. Polymer to Solvent Ratio

In this first instance, 4 spinning dopes were produced: 2 using 68 wt% of SS particles of 6µm

diameter and polymer and solvent ratios of 1:3 and 1:4; and 2 without SS particles and the same

polymer and solvent ratios. Figure 4.4 shows the impact of these changes on the viscosity.

Page 93: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

93

Steel loading [%

w]

0 20 40 60 80

Vis

cosity [

Pa.s

]

1

10

100

1:3 PEI:NMP - 6µm

1:4 PEI:NMP - 6µm

Figure 4.4 – Viscosity of the spinning dope with PEI / NMP ratios of 1:3 (solid circles) and 1:4

(hollow circles) as a function of stainless steel loading (6μm sized particles)

The results show that as the polymer to solvent ratio is increased from 1:4 to 1:3, the viscosity of

the resulting spinning dope increases from 3 Pa.s to 13 Pa.s, for the neat polymer and from 10 Pa.s

to 40 Pa.s for the dope containing SS particles. This represents a 3.3 fold increase in the neat

polymer and a 3 fold increase in the polymer + SS dope, both for a 25% increase in polymer

content. In the case of the neat polymer the increase in viscosity is similar to previous literature

studies [42, 43]. This is an interesting result as the relative increase in viscosity for the neat polymer

is ~10% greater than for the polymer + SS dope which implies the inclusion of SS particles may

have a mild dampening effect on potential changes to the overall viscosity of the spinning dope. In

either case the high viscosity of the samples prepared with a solvent to polymer ratio of 1:3, are

expected to have slower interactions, fewer finger-like pores and a greater degree of sponge-like

morphology.

Page 94: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

94

Figure 4.5 – SEM images of green fibres produced using spinning dopes containing stainless steel

and polymer to solvent ratios of 1:3 (a) and 1:4 (b)

Figure 4.5 shows SEM images of the cross sections of green hollow fibres formed from spinning

dopes containing stainless steel particles. The 1:3 sample in Figure 4.5 (a) shows a well-defined

region of finger-like macrovoids close to the lumen of the hollow fibre, while close to the outer

surface the morphology is sponge-like. In contrast, the 1:4 sample in Figure 4.5 (b) shows two

finger-like macrovoid regions at both the inner and outer surfaces of the hollow fibre. This implies

the lower viscosity lead to faster solvent / non-solvent interactions, in agreement with results

presented by [44-49].

Helium picnometry was used to determine that the specific volume of the samples containing 6μm

stainless steel particles, with a polymer to solvent ratio of 1:3 is 0.2 cm3g-1, while samples with a

ratio of 1:4 show average specific volume of 0.26 cm3.g-1. This represents a 30% increase in the

specific volume of the nascent hollow fibre which is consistent with a reduction in polymer content

of 20% and an increase in the solvent content of ~6.5%. In conjunction, mercury porosimetry was

performed to understand the porosity and pore size distribution of the nascent hollow fibres as a

function of the polymer to solvent ratio (Figure 4.6).

(a) (b)

Page 95: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

95

Pore Radius (µm)

0.00010.0010.010.11101001000

Po

rosity

(%)

0

10

20

30

40

50

60

1:3

1:4

a

Pore Radius (µm)

0.0010.010.11101001000

1:4

1:3 b

Mass Ratio of Polymer and Solvent

c

1:3 1:4

Po

rosity

(%v

ol)

28

30

32

34

36

38

40

42

44

46

48

Mass Ratio of Polymer and Solvent

c

Figure 4.6 – Mercury porosimetry (a), porosity distribution (b) and maximum porosity (c) of the

hollow fibres containing 6μm stainless steel particles prepared using polymer to solvent ratios of

1:3 and 1:4.

The results of the mercury porosimetry analysis show that samples produced using more solvent

have a larger total porosity (30% as compared with 47%, which represents a 57% increase) and

greater percentage of pores larger than 0.1μm. Indeed, for the 1:4 sample, almost half of the total

porosity is due to pores > 0.1μm, which is in direct contrast to the 1:3 sample where almost all

porosity is associated with pores < 0.1μm. Interestingly, the tri-modality seen in the porosity

distribution in Section 4.3 for the PEI+SS sample disappeared, replaced with bi-modal distribution

with a broad flat peak for dp>0.1μm and a very sharp one around 0.08μm. This is despite both

samples having identical polymer to solvent ratios; however, the loss of the moderate sized pores

(1<dp<0.1μm) is associated with the behaviour of the polymer itself and not the arrangement of

stainless steel particles within the nascent hollow fibre. To this end it is not considered a significant

result as the polymer will be burnt away during the sintering process and should not impact on the

final sintered morphology. It does however, serve to highlight the complexity of the hollow fibre

Page 96: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

96

spinning process. Importantly the total porosity remained within error at ~30%. These findings

agree with the predictions from theory, in so much as in a system with more solvent available for

exchange with non-solvent, the higher the kinetic potential of the system which promotes

instantaneous demixing and larger finger-like macrovoids. By comparison when relatively small

amounts of solvent are available for exchange, the system switches from instantaneous to delayed

demixing which favours the formation of sponge-like structures.

4.4.2. Viscosity Modifier

The second way of altering the viscosity of the spinning dope is through the addition of a viscosity

modifier. This is typically a secondary polymer which, when added in small amounts (typically

<2% of the total system) can significantly alter the rheological behaviour (in this case viscosity) of

the spinning dope. In this case, 4 samples with polymer to solvent ratios of 1:3 and 1:4 and with and

without stainless steel particles were used as controls. They were modified by adding PVP at

compositions of 0.5, 1 and 1.5wt% of the polymer and solvent mass, in line with previous studies

on ceramic hollow fibres [30, 50-53].

The viscosity profiles of all the samples remained constant over time and demonstrated Newtonian

behaviour above a sheer rate of ~20 rad/s. Below this sheer rate the behaviour was mildly sheer

thinning, but quickly plateaued as the sheer rate increased. These minor changes at such low sheer

rates were not considered significant in the scheme of spinning hollow fibres and so only the

Newtonian behaviour was analysed. As can be seen in Figure 4.7 the addition of PVP increases the

viscosity in all samples in an exponential fashion; however the greatest influence can be seen in the

samples containing stainless steel particles and a greater solvent to polymer ratio.

The first step in this investigation is to measure the effect that the addition of this additive will have

on the viscosity. Preliminary measurements were made on the polymeric mixture of PEI and NMP

in the ratios of 1:3 and 1:4. The viscosity was increased from around 13.8 Pa.s to 15.7 Pa.s in

polymeric mixtures containing 1:3 ratio, while the increase was from 2.85 Pa.s to 3.42 Pa.s in

mixtures containing 1:4 ratio, when 0.5%w of PVP was added. The viscosity increase was larger

when 1% PVP was added, jumping to 17.9 Pa.s for 1:3 and 4.09 for 1:4 ratios. Finally, the viscosity

measured for the polymeric mixtures when 1.5% PVP was added was around 18.04 Pa.s and 5.08

Pa.s for polymer to solvent ratios of 1:3 and 1:4 respectively.

Page 97: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

97

% of PVP

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

Vis

co

sity

(Pa

.s)

1

10

100

1:3 PEI:NMP

1:4 PEI:NMP

1:3 PEI:NMP + 68% SS

1:4 PEI:NMP +68%SS

Figure 4.7 – Spinning dope viscosity as a function of PVP content.

Figure 4.8 – SEM images of the morphology resulting of the phase inversion process of samples

produced using polymer to solvent ratio of 1:3 and contain 0% PVP (A), 0.5% PVP (B),, 1% PVP

(C), and 1.5% PVP (D).

Page 98: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

98

These results show the significant effect that the addition of PVP has on the viscosity of the

spinning dope, however, it is important to verify if this is reflected in the morphology of the hollow

fibres. SEM images of the cross sections of green stainless steel hollow fibres formed from spinning

dopes containing a polymer to solvent ratio of 1:3 and 1:4 are displayed in Figure 4.8 and Figure

4.9 respectively. For both sets of samples the apparent ratio of finger-like voids to sponge-like

macrostructures either remains constant or actually increases with increasing addition of PVP. For

instance, in the sample containing a polymer to solvent ratio of 1:3 there is a clear increase in the

number of finger-like voids at the outer surface when comparing the 1.5% PVP sample (Figure 4.8

(d)) against the control (Figure 4.8 (a)). The inner lumen in this case remains relatively unaffected

with a significant number of finger-like voids in all cases. Samples produced using the polymer to

solvent ratio of 1:4 (Figure 4.9) also demonstrated a similar trend of an increasing number of finger-

like macrovoids as the PVP content increased. However, in this instance there was less of a marked

change between the control sample (no PVP) and those containing PVP, as the control sample

already displayed finger-like macrovoids on the outer surface and inner lumen of the hollow fibre.

Figure 4.9 – SEM images of the morphology resulting of the phase inversion process of samples

produced using polymer to solvent ratio of 1:4 and contain 0% PVP (A), 0.5% PVP (B),, 1% PVP

(C), and 1.5% PVP (D).

Page 99: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

99

Helium picnometry measurements were also performed to help differentiate the resulting

morphologies. The results of these measurements returned that there is no significant difference

between the specific volumes of the samples with a polymer to solvent ratio of 1:3, with values

averaging ~0.2 cm3g-1, with a standard deviation of only 0.0018 cm3g-1. For the 1:4 samples the

value of the specific volume is larger at ~0.27 cm3g-1, with a standard deviation of 0.048 cm3g-1,

indicating more variability. However, on the whole it appears that PVP addition has no significant

impact on the specific volume of the nascent hollow fibres and rather the polymer to solvent ratio is

the primary causative agent.

a

% of PVP

0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6

Po

rosity

(%v

ol)

10

100

1:3 PEI:NMP + 68%SS

1:4 PEI:NMP + 68%SS

a

b

Pore Radius (µm)

0.00010.0010.010.11101001000

Po

rosity

(%)

0

10

20

30

40

50

60

1:3 + 0.5% PVP

1:4 + 0.5% PVP

b

Pore Radius (µm)

0.00010.0010.010.11101001000

Po

rosity

(%)

0

10

20

30

40

50

60

1:3 + 1% PVP

1:4 + 1% PVP

c

d

Pore Radius (µm)

0.00010.0010.010.11101001000

Po

rosi

ty (

%)

0

10

20

30

40

50

60

70

1:3 + 1.5% PVP

1:4 + 1.5% PVP

d

Figure 4.10 – Total porosity as determined by mercury porosimetry (a) porosity distribution of

nascent hollow fibres with 0.5% (b) 1% (c) and 1.5% (d) PVP added as a viscosity enhancer.

Page 100: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

100

The mercury porosimetry results displayed in Figure 4.10 are very interesting. Analysis of the total

porosity shows that PVP addition enhances the total porosity (Figure 4.10 (a)) of the unsintered

hollow fibres only in the samples where the polymer to solvent ratio is 1:4. Furthermore these

samples demonstrated almost double the porosity of the corresponding 1:3 samples, with the

exception of the 1% PVP sample where there was only a very small increase in porosity (~10%) as

the polymer to solvent ratio increased from 1:3 to 1:4. By contrast increasing addition of PVP had

no impact on the total porosity of the samples with a polymer to solvent ratio of 1:3, with the total

porosity remaining at ~30%.

The impact of PVP addition on the porosity distribution (Figure 4.10 (b-d)) was, however, more

dramatic. In particular, there was a noticeable increase in the percentage of larger pores (> 1μm)

and moderate pores (1<dp<0.1μm) as PVP addition increased. This was particularly true for samples

with a polymer to solvent ratio of 1:4. For the samples with a polymer to solvent ratio of 1:3 the

change in the percentage of larger and moderate sized pores was minimal, however; there was a

significant upward shift in the size of the small pores between the control and the sample containing

1.5% PVP. Taken together this data corroborates the observations made in the SEM analysis of the

nascent hollow fibre cross-sections, in that the addition of PVP increases the percentage of finger-

like macrovoids and decreases the formation of sponge-like structures.

These results show that despite the fact that PVP acts to increase the spinning dope viscosity, the

impact on the final hollow fibre morphology was both detrimental and counter-intuitive.

Conventional wisdom suggests that increasing the viscosity of the spinning dope should reduce the

speed of demixing and thereby promote the formation of sponge-like structures rather than finger-

like voids. However, it addition to its viscosity modification effects, PVP appears to act to alter the

phase diagram, shifting the coagulation point towards a high solvent content.

4.5. EFFECT OF THE ADDITION OF PARTICLES TO THE SPINNING

DOPE

The addition of solid particles to a solution and the subsequent impacts on solution viscosity has

been studied since early 1900s. Einstein [31] proposed a linear relation between the amount of

particles added to a suspension and its increase in viscosity. Later, other works proved that the

linear relation is only valid for low solids loadings, where particle to particle interaction is

Page 101: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

101

negligible, and other non–linear models considering these interactions have been proposed [32, 33,

54-57]. Brouwers [57], proposed that the maximum volume fraction of solid spheres that can be

added to a solution, and still retain the fluid properties is around 0.64. As the volume fraction is

increased beyond this point particle to particle interactions play an ever greater role, increasing the

viscosity in an exponential manner.

Traditional theory suggests that the viscosity is only affected by the volume fraction of the particles

and therefore solutions prepared with same volumetric fraction of solids should present similar

viscosity, independent of the particle size utilised. There are, however, two noted exceptions. For

particles where dp < 1 µm, colloid-chemical forces become important causing non-Newtonian flow

behaviour [58], which results in an increasing relative viscosity as particle size is decreased. This

region is further complicated by the fact that increasing the shear rate decreases the relative

viscosity (for a given particle size) but this tends to a limiting value [54]. For particles where dp >

10µm, de Bruijn [58] believes that inertial effects, due to the restoration of particle rotation after

collision, result in an additional energy dissipation and a consequent increase in relative viscosity

with increasing particle diameter.

Despite the multitude of work on impacts of particle addition on solution viscosity, there has been

limited work on the impact of particle size on the phase inversion process and resultant morphology

in the production of green ceramic or metallic hollow fibres. In this section of the thesis, a study

was conducted to see how the addition of particles to the spinning dope will affect the viscosity and

consequent morphology of the hollow fibres produced with different particle loading (10%w, 50%w

and 70%w) and different particle diameters (6, 16 and 45μm).

4.5.1. Particle Loading

Stainless steel particles of differing diameters (6, 16 and 45µm) were added to a spinning dope

mixture of PEI and NMP (1:3 solvent to polymer ratio) at various solids loadings (10, 50 and 70

wt%). The viscosity of the resulting solutions is presented in Figure 4.11, alongside the Einstein

[31] and Thomas [54] predictive relations.

Page 102: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

102

Particle loading [%w]

0 20 40 60 80

Vis

cosity [

Pa.s

]

10

100

6µm

10µm

16µm

45µm

Einstein

Thomas

Figure 4.11 – Viscosity as a function of particle loading for particles of different sizes

There is a clear increase in the viscosity of the spinning solution, as the solids loading is increased

from 10% to 70%. This increase shows a linear tendency for loads up to 50% in mass, and agrees

with the equation presented by Einstein [31]. Between 50% and 70% there is a change in the slope

of the trend, is consistent with prediction made by Thomas [54]. The volumetric fraction of stainless

steel particles in 70 wt% solution was calculated to be around 0.49. This value, whilst high, it is still

within the range in which the particle to particle interactions are considered negligible [32, 55]. The

difference in the observed results can be attributed to interactions between the particles and the

fluid, colloidal forces, and inertial effects [58].

As viscosity is so closely associated with the kinetics of the phase inversion process and by

extension the final green fibre morphology, the solutions with greater particle fractions should

exhibit fewer finger-like pores than those with lower solids loadings. Figure 4.12 displays SEM

cross sectional images of a green hollow fibre made from 6µm particles at solids loadings of 10, 50

and 70 wt%.

Page 103: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

103

Figure 4.12 – Morphology resulting from spinning hollow fibres using slurry with particle size of

6µm and loading of 10% (A), 50% (B) and 70% (C)

Figure 4.12A and B show a hollow fibre with finger-like macrovoids on lumen and outer sides. This

macrostructure is characteristic of fast happening phase inversion processes. At low solids loadings

and by extension reduced viscosities, the inter-diffusion of solvent and non-solvent happen rapidly,

creating areas rich in solvent and lean in polymer, producing macrovoids. Whereas, the hollow fibre

with the largest solids loading (70 wt%) exhibits finger-like pores only at the inner surface of the

fibre (Figure 4.12C). The outer surface meanwhile clearly displays a sponge-like morphology and

no macrovoid formation. In this case the inter-diffusion of solvent and non-solvent at the inner

surface is fast, leading to the formation of polymer rich and polymer lean regions that will result in

macrovoids. However, on the outer side of the hollow fibre, the interaction between solvent and

non-solvent is slower due to a reduced concentration gradient, and therefore a more homogenous

metastable region is formed which prevents macrovoid formation and promotes a sponge-like

morphology.

Specific volume measurements of the green hollow fibres were made using helium picnometry with

increasing solids loading resulting in a denser hollow fibre. Hollow fibres made using 10 wt%

solids present an average specific volume of 0.64 cm3g-1, which increased to 0.257 cm3g-1 as the

solids load is increased to 50 wt%. This represents a 60% decrease in specific volume. The fibres

become even denser when the solids load is increased to 70 wt% recording a specific volume of

0.196 cm3g-1, which represents a decrease of 24%. Mercury porosimetry analysis was conducted to

Page 104: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

104

whether the solids loading increased the density in and of itself or whether a particular pore size

range was specifically affected (Figure 4.13).

Pore Radius (µm)

0.00010.0010.010.11101001000

Po

rosity

(%)

0

10

20

30

40

50

60

Solid Load of 10%

Solid Load of 50%

Solid Load of 70%

(a) Pore radius (µm)

0.0010.010.11101001000

70% Particle Loading

50% Particle Loading

10% Particle Loading

(b)

Particle loading (%w

)

0 10 20 30 40 50 60 70 80

Po

rosity

(%v

ol)

10

100

(c) Particle loading (%

w)

0 10 20 30 40 50 60 70 80

Ave

rag

e p

ore

ra

diu

s (

µm

)

0.1

1

(d)

Figure 4.13 – Mercury porosimetry (a), porosity distribution (b), maximum porosity (c) and

average pore radius (d) of samples produced using different particle loading.

In this case it is clear that samples produced using fewer particles displayed larger porosity and an

average pore size, compared to those hollow fibres with greater solids loadings. The porosity

decreases from 53% to 42% with an increase in the particle loading from 10 to 50 wt%. Similarly

an increase in particle loading to 70 wt% (from 50 wt%) decreased the porosity a further 30%. The

porosity distribution correlates well with the SEM observations in that the hollow fibres with lower

solids loadings have a larger fraction of finger-like pores. This is particularly noticeable in the 10

wt% sample where 1/10th of the total porosity (5 percentage points) is associated with pores >

50µm. It is also remarkable that the size of the smaller pores is similar in all three cases, but this is

postulated to be because, in the case of the green fibre, the smaller pores (<0.1µm) are determined

by the polymer and not by the steel loading.

Page 105: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

105

4.5.2. Particle Size

The effect of particle size on the viscosity of a mixture of PEI and NMP (in the ration of 1:3) and

stainless steel particles of different particles sizes (at 68 wt%) was measured, and the results is

plotted in Figure 4.14.

Particle size (µm)

0 10 20 30 40 50

Vis

co

sity

(Pa

.s)

10

100

Figure 4.14 – Viscosity of solutions containing stainless steel particles of 6, 16 and 45 μm particle

diameter.

There is a clear trend of increasing viscosity with particle size, which agrees with the observations

of de Bruijn [58] who showed that above a particle diameter of 10µm relative viscosity increases

with increasing particle diameter, due to the inertial effects which serve to restore particle rotation

after collision. These act to dissipate energy through the fluid meaning that more energy is needed

to generate the same shear rate, meaning that viscosity of the fluid is increased. To determine

whether the increase in viscosity had the expected impact on hollow fibre morphology a series of

cross sectional SEM images were taken and are shown in the Figure 4.15.

Page 106: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

106

Figure 4.15 – Green hollow fibres produced with stainless steel particles of 6µm (A), 10µm (B),

16µm(C) and 45µm(D).

SEM analysis indicates that hollow fibres produced with smaller particle sizes show smaller and

fewer macrovoids that the samples with larger particle sizes. Similarly, the magnified images inset

into each quadrant of Figure 4.15 demonstrates that samples with larger particle sizes have larger

and more prevalent finger-like macrovoids, particularly at the inner surface of the hollow fibre.

Interestingly, all hollow fibre samples have a substantial sponge-like region close to the outer

surface, regardless of particle size, although it appears to be thicker for the smaller particle sizes.

These results are in contrast to the viscosity data presented in Figure 4.14 where increasing the

particle size increased the spinning dope viscosity and implies that the situation may be more

complex than traditional theory suggests. Indeed, as the SS particles should have no impact on the

ternary phase diagram these results suggest that particle addition influences the kinetics of the

solvent / non-solvent interactions. However, it was first necessary to confirm that larger particles

produced hollow fibres with a greater fraction of macrovoids, so both helium picnometry and

mercury porosimetry (Figure 4.16) were undertaken to quantitatively characterise the hollow fibre

Page 107: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

107

porosity distribution. The variation in density of the samples measured by helium picnometry was

marginal, varying from 5.05 through to 5.11 g.cm-3 for samples made with 45µm particles down to

6µm particles respectively. This equates to 69.4% to 70.2% respectively of the calculated

theoretical density of the components of the green hollow fibre.

a

Pore Radius (µm)

0.00010.0010.010.11101001000

Po

rosity

(%)

0

5

10

15

20

25

30

35

6 µm

10 µm

16 µm

45 µm

a

b

Pore Radius (µm)

0.0010.010.11101001000

6 µm

10 µm

16 µm

45 µm

b

c

Particle size (µm)

0 10 20 30 40 50

Ma

xim

um

Po

rosity

(%v

ol)

10

100

c

Figure 4.16 – Mercury porosimetry (a), porosity distribution (b) and maximum porosity (c) of green

hollow fibres made using particles of different sizes.

The mercury porosimetry data correlates well with the picnometer results with the porosity of all

hollow fibres in the range of 30% regardless of particle size Figure 4.16(c). The porosity

distribution Figure 4.16(b) obtained from the mercury porosimetry also correlates well with the

cross-sectional SEM images in Figure 4.15. Samples with larger particle sizes (16 and 45μm)

displayed a significant porosity peak at large pore sizes between 50-100μm, which are attributed to

finger-like macrovoids. These pores contribute around 1/6th (5 percentage points) of the total

porosity for these samples. The remaining porosity was associated with small pores below 0.1μm,

Page 108: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

108

which is attributed to the sponge-like region. By contrast, hollow fibres made with smaller particle

sizes (6 and 10μm) displayed porosity peaks at small pore sizes between 0.08 and 2μm which are

attributed to the sponge-like regions. These peaks accounted for all the porosity of the hollow fibres

made with smaller particle sizes confirming the observation from SEM analysis that the sponge-like

region appeared greater in these samples compared to those made with larger particle sizes.

Interestingly, samples in the middle range of particle size (10 and 16µm) show a trimodal pore size

distribution, presenting additional pores in the range between 1 and 0.1 microns taking about 12%

of the porosity.

4.6. GENERAL DISCUSSION

The production of green stainless steel hollow fibres through dry-wet phase inversion is a complex

process where each parameter can affect more than one property of the final hollow fibre. It is

crucial to understand the green fibre production process because the final morphology of the hollow

fibre after sintering bears a significant resemblance to the green fibre morphology (discussed later

in Chapter 5). Furthermore, the morphology of the final sintered fibre greatly affects the mechanical

and performance properties of the subsequent membrane or membrane support (Chapter 5-7).

Therefore, as it is frequently easier and cheaper to adjust the hollow fibre spinning process rather

than the sintering conditions, the production of a suitable green hollow fibre is the first and arguably

most important step in producing stainless steel membranes and membrane supports with desirable

characteristics.

It is noteworthy that in the production of polymeric hollow fibres, viscosity is commonly used to

compare polymer-solvent solutions, as a way to measure the dispersion and mobility of the

polymeric chains in the solvent, and thus predict the morphology that will result from the phase

inversion process. In fact, lower viscosities commonly result in a large amount of macrovoids, due

to the relative ease of the exchange between solvent and non-solvent favoured by the large

dispersion of the polymeric chains in the abundant solvent. When the viscosity is increased, for

example by increasing the relative amount of polymer, the result is a decrease in the time that it

takes for the system to shift to delayed demixing. This phenomenon is associated with slower

diffusion rates of the solvent and non-solvent, which is related to the close packing of the polymer

chains resulting in a more difficult movement of the solvent through these chains, hindering the

exchange. When looking specifically at the effect that the polymer to solvent ratio has on the final

morphology, the case of 1:4 polymer to solvent ratio results in more finger-like macrovoids as

compared to the 1:3 situation. In this case the chemical potential is not depleted in the exchange

Page 109: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

109

between solvent and non-solvent occurring in the lumen, and plenty of solvent is still concealed in

the spinning dope while the forming hollow fibre reaches the coagulation bath. This allows the

formation of macrovoids at both surfaces of the hollow fibre, and a clear sponge like region in the

middle of the hollow fibre cross section. On the contrary, when the polymer to solvent ratio is

changed to 1:3, the chemical potential rapidly decreases after the first contact with the non-solvent

in the bore, to the point that promotes delayed demixing. Hence when the hollow fibre reaches the

coagulation bath, a sponge like region close to the outer surface of the hollow fibre is formed.

The results of this chapter demonstrate that the viscosity of the spinning dope containing particles

can be only partially used to infer the morphology that will result from the phase inversion process.

However the use of viscosity as a predictive parameter is limited only to very similar systems, as in

most cases the interaction between solvent, polymer and particle loading is too complex to

accurately predict the morphology by simply comparing the viscosity of two very different spinning

dopes. In the case where only polymers are changed this rule of thumb can be confidently applied,

the viscosity of the sample will be related more to how well the polymer disperses in the solvent,

and how well the solvent will be able to move within the polymer. In the case of lower viscosity

polymers, the solvent / non-solvent interaction was able to take place so rapidly that a solvent

depleted region was formed in the mid-wall region, impeding the movement of the solvent at the

outer surface towards the non-solvent in the bore. A schematic is shown below in Figure 4.17

depicting the stages of phase exchange in low and high viscosities polymers.

Page 110: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

110

Figure 4.17 – Schematic of the phase inversion process of low(A, B) and high (C, D) viscosity

polymer.

Figure 4.17 shows the differences between samples produced with polymers of low viscosity (A

before the hollow fibre reaches the coagulation bath and B after) and polymers with higher viscosity

(C before the hollow fibre reaches the coagulation bath and D after). Figure 4.17 A shows that at the

beginning of the process the solvent is diffusing towards the lumen of the hollow fibre, and due to

the rapid diffusion of the solvent, a solvent lean region is formed. In this region the spinning dope is

solidifying (or in a quasi-solid state) impeding further diffusion of the solvent towards the non-

solvent in the bore. This creates a region with a high concentration of solvent between the middle of

the hollow fibre wall and the outer surface. In Figure 4.17B it is possible to see that when the

hollow fibre reaches the coagulation bath the diffusion towards the bore is minimal or non-existent,

and all the remaining solvent diffuses towards the outer surface in an instantaneous demixing

process. This is the case of hollow fibres made with PESf, that present large macrovoids towards

both surfaces of the hollow fibre.

Page 111: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

111

In the case of higher viscosity polymers shown in the Figure 4.17 C and D, the solid region that

divides the wall of the hollow fibre doesn’t form. This is due to the fact that the diffusion of the

solvent inside the spinning dope is not as fast and a solvent lean region doesn’t form. Therefore

when the hollow fibre reaches the coagulation bath there is still solvent flow towards both inner and

outer surfaces of the hollow fibre. This means that there is a solvent lean region close to the outer

surface of the hollow fibre, as compared to the PESf case. This in turn promotes delayed demixing

and sponge-like structure formation.

Therefore, the choice of the polymeric binder must carefully considered based on the particular

application that the hollow fibre will have, and on the spinning conditions used. In addition to

viscosity, different polymers have different solubilities in the same solvent, thus different

interactions can be expected during phase inversion, resulting in different morphologies for the

same spinning conditions. Table 4.1 shows Flory interaction parameters calculated for spinning

dopes made using PESf and PEI. The polymer solvent interaction parameter, χ23, is of particular

relevance with higher values indicating less affinity between solvent and polymer, which in turn

means that the resistance to movement of the solvent through the polymer is lower and hence less

energy is required to initiate phase separation. Therefore, in this case the greater the value of χ23, the

faster the solvent / non solvent exchange kinetics. This is in agreement with the images shown

Figure 4.2 where more finger-like macrovoids are formed for the sample made with PESf compared

with PEI. The higher χ23 value for PESf means that it requires more energy to be dissolved into the

NMP, compared to PEI. Consequently the NMP requires longer times and higher solvent

concentrations to redissolve the PESf once coagulation has occurred. This reinforces the mechanism

described in Figure 4.17 where the PESf-rich region is formed in the middle of the hollow fibre wall

inhibiting solvent diffusion. It is important to highlight that these kinetic characteristics are a

property of each ternary system and determine which phase inversion process is taking place at any

given time. This controls if the process switches from an instantaneous to delayed demixing.

Importantly, it is not possible to switch from a delayed to instantaneous demixing process, since

instantaneous demixing occurs for larger difference in Gibbs free energy [40].

PVP has been used in both the polymeric and ceramic hollow fibre fields to increase the viscosity of

the spinning dope and otherwise control the morphology of the hollow fibre. Typically, for the

production of polymeric membranes, as more PVP is added to the system the viscosity increases

and a delayed demixing process is favoured. However, the results obtained in this chapter are

contrary to this and the addition of PVP to the spinning dope apparently favours an instantaneous

demixing process. Whilst, the addition of PVP in fact resulted in an exponential increase in

Page 112: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

112

viscosity as shown in Figure 4.7, the SEM and mercury porosimetry analysis in Figure 4.8, Figure

4.9 and Figure 4.10 show an increase in finger-like macrovoids. This was particularly noticeable in

the samples containing 1:3 polymer to solvent ratio, despite similar increases in relative viscosity

being recorded for the 1:4 polymer to solvent ratio samples. In this case, the impact of PVP is

negligible compared to the effect that the higher potential created by the larger amount of solvent in

the spinning dope has on the morphology.

In essence PVP and PEI are miscible polymers, but they present different hydrophobicity and have

different interaction parameters with NMP and non-solvent [59], altering the way in which the

coagulation process happens. At the moment of contact between the non-solvent and the spinning

dope, the solvent / non-solvent exchange begins to coagulate the PEI and PVP; however these

polymers themselves phase separate with a PVP rich region forming closest to the non-solvent

boundary (i.e. the inner surface of the hollow fibre) due to the solubility of this polymer into the

non-solvent [59-61]. Immediately next to the PVP rich region is a PEI rich region which is lean in

both solvent and PVP which also partially coagulates. This dual layer structure acts as barrier

trapping the remaining NMP inside the hollow fibre, maintaining a high NMP concentration

gradient within the metastable spinning dope. This has two important effects; firstly it impedes the

initial solvent / non-solvent exchange kinetics as whilst NMP is a good solvent for PEI, readily

redissolving the PEI in the PEI rich region, NMP is not a good solvent for PVP so the PVP rich

region remains as a barrier preventing further exchange. Secondly, PVP is partially soluble in water,

so the PVP rich layer is gradually removed exposing the metastable spinning dope to the non-

solvent again. As the solvent concentration gradient is still high and most of the PVP has been

removed from the system, the solvent / non-solvent exchange proceeds via instantaneous demixing,

forming finger-like voids in the process. Importantly in this work, a relatively small air gap is

utilised which means that the outer PVP rich skin is formed before the inner PVP rich skin can

dissolve in the non-solvent. This ensures that the NMP concentration gradient remains high in both

the inner and outer halves of the hollow fibre wall so that finger-like voids are formed at both the

inner and outer surface.

The impact of particle size and loading on the morphology of the green hollow fibre is remarkable.

When particles are introducing to the ternary system, the viscosity is increased as a function of the

interaction between the particles and the solution. If the volume fraction of particles is large

enough, the voidage between those particles will be reduced. Einstein [31] and other subsequent

works [32-34, 62] proved that the particle size is not relevant to the effect that the particle loading

has on viscosity; however the volume fraction of the particle introduced yields a significant increase

Page 113: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

113

in the viscosity. The effect of the addition of particles was calculated [32], and an increase in

viscosity of between 2 and 2.5 times is expected through the addition of stainless steel particles.

This correlates well with the data in Figure 4.11. For lower particle loadings the exchange between

solvent and non-solvent during phase inversion was not disturbed, which resulted in two finger-like

regions, near both the inner and outer surfaces and a small sponge-like ring in the middle of the

hollow fibre cross section (Figure 4.12A). A similar morphology was obtained for particle loadings

of 50 vol%, but an increase in the thickness of the sponge-like ring was observed. When the particle

loading was further increased to 70 vol%, the morphology changed completely and finger-like pores

were only observed near the lumen of the hollow fibre, with a large sponge-like region formed next

to the outer surface. This accompanied by the steady decrease in porosity which is linked to the

growth of the sponge-like region associated with the increase in particle loading as shown in Figure

4.13. Hence, there is an overall decrease in the kinetics of the phase inversion process which can

only be attributed to particle addition and more specifically the resistance that the particles induce

on the mass transfer process.

Interestingly, Figure 4.15 shows a significant increase in finger-like macrovoids with increasing in

particle size, despite maintaining the same particle loading in the spinning dope. The implication

here is that the size of the particles must be considered as a factor in the phase inversion process. In

this case it is easiest to consider the influence of particle on the solvent / non-solvent exchange

kinetics analogous to fluid flow through a packed bed.

If we consider that the solvent velocity is within the laminar flow region, due to the relatively high

viscosity of the polymeric solution, the Carman – Kozeny equation (equation 4.1) can be used to

model effect that particles of different sizes will have on the kinetics of the phase inversion process:

(4.7)

where ΔP is the pressure drop (or resistance) across the bed of mono-dispersed spheres, H is the

thickness of the bed, µ is the fluid viscosity (in this case the solvent viscosity), U is the fluid

velocity, ε is the voidage, and x is the particle diameter. In this case ΔP becomes the osmotic

pressure of the solvent relative to the non-solvent. In the situation considered in Section 4.5.2, the

voidage remains constant as particle loading is unchanged between samples, whilst particle

diameter, viscosity and fluid velocity are the only parameters which change. Figure 4.18 was

produced to show variations in particle size affects the relative solvent velocity in the solvent / non-

solvent exchange process. For comparison purposes the fluid velocities were normalised against the

Page 114: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

114

fastest flow, which was associated with the largest particles (45μm). The impact of particle size on

solvent velocity is dramatic with a 64% reduction in particle size (from 45 - 16µm) yielding

velocities 88% lower. The effect is even greater for smaller particles with the 10 and 6 µm samples

yielding velocities of only 5 and 2% of the reference (45μm) velocity respectively.

Particle size (µm)

0 10 20 30 40 50

Velo

city r

atio (

VP

S /

V4

m)

0

20

40

60

80

100

120

Figure 4.18 – Relative effect of particle size in the solvent – non-solvent exchange velocity.

This velocity profile correlates well with the morphology results shown in Figure 4.15, where it is

possible to see that samples produced with 45µm particles contain more macrovoids (that are also

larger in size) compared to samples made with smaller particle sizes. Further the fraction and size of

macrovoids is reduced as SS particle size decreases. Clearly, the addition of SS particles cannot

influence the ternary phase diagram and so the change in morphology must be a direct result of the

altered solvent / non-solvent exchange kinetics.

4.7. CONCLUSION

The process of stainless steel hollow fibre production is non-trivial as the change in either

composition or spinning parameters will lead to a change in the final morphology of the hollow

fibre. It was shown that when comparing spinning dopes prepared with different polymers, viscosity

can be used as parameter to predict the morphology. In particular, systems with lower viscosity will

result in more macrovoids, compared to systems with higher viscosity that will have a larger sponge

like region. Samples produced with the same polymer but with different solvent to polymer ratios

can follow this viscosity based heuristic. However when viscosity enhancers, such as PVP, are used

Page 115: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

115

this rule is no longer applicable. Typically, for the production of polymeric membranes, as more

PVP is added to the system the viscosity increases and a delayed demixing process is favoured.

However, the results obtained in this chapter are contrary to this and the addition of PVP to the

spinning dope apparently favours an instantaneous demixing process as hollow fibres containing

PVP showed more and larger finger-like macrovoids. This counter-intuitive finding was attributed

to altered solvent / non-solvent exchange dynamics as a result of the interaction between PVP and

the non-solvent.

The effect of particle loading and particle size on the properties of the spinning dope and

subsequent hollow fibre morphology was studied in detail. Here the influence of particle size and

loading on the solvent / non-solvent exchange kinetics were considered analogous to fluid flow

through a packed bed. Higher particles loadings act as a restriction to the phase inversion process,

resulting in a binary morphology with finger-like macrovoids near the lumen and a sponge-like

region near the outer surface. Lower solid loading resulted in a ternary structure composed of

finger-like macrovoids – sponge-like structure – finger-like macrovoids as you progress radially

from the hollow fibre lumen. The effect of particle size was also studied and the changes in

morphology were dramatic despite maintaining constant solids loading. Large particles did not

significantly affect solvent velocity and as a result, hollow fibres with large amount of finger-like

pores were observed. In comparison a particle size reduction from 45μm to 6μm yielded a 98%

reduction in solvent velocity and this was reflected in the production of a large sponge-like region.

4.8. REFERENCES

[1] Takaba H, Matsuda E, Nair BN, Nakao SI. Molecular modeling of gas permeation through an

amorphous microporous silica membrane. J Chem Eng Jpn 2002;35:1312-21.

[2] Smart S, Lin CXC, Ding L, Thambimuthu K, Diniz da Costa JC. Ceramic membranes for gas

processing in coal gasification. Energy Environ Sci 2010;3:268-78.

[3] Duke M, Rudolph V, Lu GQ, Diniz da Costa JC. Scale-up of molecular sieve silica membranes

for reformate purification. AIChE Journal 2004;50:2630-4.

[4] Fujii T, Yano T, Nakamura K, Miyawaki O. The sol-gel preparation and characterization of

nanoporous silica membrane with controlled pore size. Journal of Membrane Science

2001;187:171-80.

Page 116: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

116

[5] Klein LC, Gallagher D. Pore structures of sol-gel silica membranes. Journal of Membrane

Science 1988;39:213-20.

[6] Gallagher D, Klein LC. Silica membranes by the sol-gel process. Journal of Colloid and

Interface Science 1986;109:40-5.

[7] Dong J, Lin YS, Kanezashi M, Tang Z. Microporous inorganic membranes for high temperature

hydrogen purification. Journal of Applied Physics 2008;104.

[8] Estella J, Echeverri̕a JC, Laguna M, Garrido JJ. Silica xerogels of tailored porosity as support

matrix for optical chemical sensors. Simultaneous effect of pH, ethanol:TEOS and water:TEOS

molar ratios, and synthesis temperature on gelation time, and textural and structural properties.

Journal of Non-Crystalline Solids 2007;353:286-94.

[9] Davis PJ, Deshpande R, Smith DM, Brinker CJ, Assink RA. Pore structure evolution in silica

gel during aging/drying. IV. Varying pore fluid pH. Journal of Non-Crystalline Solids

1994;167:295-306.

[10] Nair BN, Keizer K, Suematsu H, Suma Y, Kaneko N, Ono S, et al. Synthesis of gas and vapor

molecular sieving silica membranes and analysis of pore size and connectivity. Langmuir

2000;16:4558-62.

[11] Eriksson M, Klein LC, Lidén E, Lindqvist K. Preparation of nanoporous silica-zirconia

layers by in situ sol-gel method. Materials Science and Technology 2006;22:611-4.

[12] Nair BN, Elferink JW, Keizer K, Verweij H. Preparation and Structure of Microporous Silica

Membranes. Journal of Sol-Gel Science and Technology 1997;8:471-5.

[13] Klein LC, Bloxom T, Woodman R. Unsupported alkoxide-derived silica membranes. Colloids

and Surfaces 1992;63:151-61.

[14] Klein LC, Yu C, Woodman R, Pavlik R. Microporous oxides by the sol-gel process: synthesis

and applications. Catalysis Today 1992;14:165-73.

[15] Klein LC, Gallo TA. Densification of sol-gel silica: Constant rate heating, isothermal and step

heat treatments. Journal of Non-Crystalline Solids 1990;121:119-23.

Page 117: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

117

[16] Bocchetta P, Sunseri C, Bottino A, Capannelli G, Chiavarotti G, Piazza S, et al. Asymmetric

alumina membranes electrochemically formed in oxalic acid solution. Journal of Applied

Electrochemistry 2002;32:977-85.

[17] Gallo TA, Klein LC. Effect of dehydration on the viscosity of sol-gel processed silica. Journal

of Non-Crystalline Solids 1988;100:429-34.

[18] Okubo T, Takahashi T, Nair BN, Sadakata M, Nagamoto H. Formation mechanism of crack-

free porous YSZ membrane. Journal of Membrane Science 1997;125:311-7.

[19] Kordas G, Klein LC. Effects of water content of gels on sol-gel glass structures. Journal of

Non-Crystalline Solids 1986;84:325-8.

[20] Gallo TA, Klein LC. Apparent viscosity of sol-gel processed silica. Journal of Non-Crystalline

Solids 1986;82:198-204.

[21] Klein LC, Gallo TA, Garvey GJ. Densification of monolithic silica gels below 1000°C. Journal

of Non-Crystalline Solids 1984;63:23-33.

[22] Nair BN, Keizer K, Elferink WJ, Gilde MJ, Verweij H, Burggraaf AJ. Synthesis,

characterisation and gas permeation studies on microporous silica and alumina-silica membranes

for separation of propane and propylene. Journal of Membrane Science 1996;116:161-9.

[23] Elferink WJ, Nair BN, De Vos RM, Keizer K, Verweij H. Sol-gel synthesis and

characterization of microporous silica membranes: II. Tailor-making porosity. Journal of Colloid

and Interface Science 1996;180:127-34.

[24] Klein LC, Garvey GJ. Monolithic dried gels. Journal of Non-Crystalline Solids 1982;48:97-

104.

[25] Nair BN, Elferink WJ, Keizer K, Verweij H. Sol-gel synthesis and characterization of

microporous silica membranes I: SAXS study on the growth of polymeric structures. Journal of

Colloid and Interface Science 1996;178:565-70.

[26] Liu T, Zhang D, Xu S, Sourirajan S. Solution-Spun Hollow Fiber Polysulfone and

Polyethersulfone Ultrafiltration Membranes. Separation Science and Technology 1992;27:161-72.

Page 118: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

118

[27] Wang D, Li K, Sourirajan S, Teo WK. Phase separation phenomena of

polysulfone/solvent/organic nonsolvent and polyethersulfone/solvent/organic nonsolvent systems.

Journal of Applied Polymer Science 1993;50:1693-700.

[28] Wang D, Li K, Teo WK. Polyethersulfone hollow fiber gas separation membranes prepared

from NMP/alcohol solvent systems. Journal of Membrane Science 1996;115:85-108.

[29] Wang D, Li K, Teo WK. Preparation of Poly(ether sulfone) and Poly(ether imide) Hollow

Fiber Membranes for Gas Separation: Effect of Internal Coagulant. Membrane Formation and

Modification: American Chemical Society; 1999. p. 96-109.

[30] Leo A, Smart S, Liu S, Diniz da Costa JC. High performance perovskite hollow fibres for

oxygen separation. Journal of Membrane Science 2011;368:64-8.

[31] Einstein A. Eine neue Bestimmung der Moleküldimensionen. Annalen der Physik

1906;324:289-306.

[32] Mooney M. The viscosity of a concentrated suspension of spherical particles. Journal of

Colloid Science 1951;6:162-70.

[33] Frankel NA, Acrivos A. On the viscosity of a concentrated suspension of solid spheres.

Chemical Engineering Science 1967;22:847-53.

[34] Chong JS, Christiansen EB, Baer AD. Rheology of concentrated suspensions. Journal of

Applied Polymer Science 1971;15:2007-21.

[35] Strathmann H, Kock K. The formation mechanism of phase inversion membranes.

Desalination 1977;21:241-55.

[36] Tsay CS, McHugh AJ. Mass transfer modeling of asymmetric membrane formation by phase

inversion. Journal of Polymer Science Part B: Polymer Physics 1990;28:1327-65.

[37] Kimmerle K, Strathmann H. Analysis of the structure-determining process of phase inversion

membranes. Desalination 1990;79:283-302.

[38] Tamhankar SS, Bagajewicz M, Gavalas GR, Sharma PK, Flytzani-Stephanopoulos M. Mixed-

oxide sorbents for high-temperature removal of hydrogen sulfide. Industrial & Engineering

Chemistry Process Design and Development 1986;25:429-37.

Page 119: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

119

[39] da Costa JCD, Lu GQ, Rudolph V. Membrane-based gas separation: Potential energy recovery

and greenhouse abatement applications. Developments in Chemical Engineering and Mineral

Processing 1997;5:89-100.

[40] Karimi M, Albrecht W, Heuchel M, Kish MH, Frahn J, Weigel T, et al. Determination of

water/polymer interaction parameter for membrane-forming systems by sorption measurement and

a fitting technique. Journal of Membrane Science 2005;265:1-12.

[41] Ellison AH, Klemm RB, Schwartz AM, Grubb LS, Petrash DA. Contact angles of mercury on

various surfaces and the effect of temperature. Journal of Chemical & Engineering Data

1967;12:607-9.

[42] Wang D, Li K, Teo WK. Preparation and characterization of polyetherimide asymmetric

hollow fiber membranes for gas separation. Journal of Membrane Science 1998;138:193-201.

[43] Kim I-C, Lee K-H, Tak T-M. Preparation and characterization of integrally skinned uncharged

polyetherimide asymmetric nanofiltration membrane. Journal of Membrane Science 2001;183:235-

47.

[44] Frommer M, Lancet D. The Mechanism of Membrane Formation: Membrane Structures and

Their Relation to Preparation Conditions. In: Lonsdale HK, Podall HE, editors. Reverse Osmosis

Membrane Research: Springer US; 1972. p. 85-110.

[45] Frommer MA, Messalem RM. Mechanism of Membrane Formation. VI. Convective Flows and

Large Void Formation during Membrane Precipitation. Product R&D 1973;12:328-33.

[46] Strathmann H, Kock K, Amar P, Baker RW. The formation mechanism of asymmetric

membranes. Desalination 1975;16:179-203.

[47] Matz R. The structure of cellulose acetate membranes 1. The development of porous structures

in anisotropic membranes. Desalination 1972;10:1-15.

[48] Cabasso I, Klein E, Smith JK. Polysulfone hollow fibers. II. Morphology. Journal of Applied

Polymer Science 1977;21:165-80.

[49] Ray RJ, Krantz WB, Sani RL. Linear stability theory model for finger formation in asymmetric

membranes. Journal of Membrane Science 1985;23:155-82.

Page 120: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

120

[50] Liu S, Tan X, Li K, Hughes R. Preparation and characterisation of SrCe0.95Yb0.05O2.975

hollow fibre membranes. Journal of Membrane Science 2001;193:249-60.

[51] Liu S, Li K. Preparation TiO2/Al2O3 composite hollow fibre membranes. Journal of

Membrane Science 2003;218:269-77.

[52] Liu S, Li K, Hughes R. Preparation of porous aluminium oxide (Al2O3) hollow fibre

membranes by a combined phase-inversion and sintering method. Ceramics International

2003;29:875-81.

[53] Haworth P, Smart S, Glasscock J, Diniz da Costa JC. High performance yttrium-doped BSCF

hollow fibre membranes. Separation and Purification Technology 2012;94:16-22.

[54] Thomas DG. Transport characteristics of suspension: VIII. A note on the viscosity of

Newtonian suspensions of uniform spherical particles. Journal of Colloid Science 1965;20:267-77.

[55] Eagland D, Kay M. The rheological properties of concentrated polymer dispersions: I. The

effects of concentration, particle size, and size distribution upon the shear dependence of viscosity.

Journal of Colloid and Interface Science 1970;34:249-61.

[56] Farr RS, Groot RD. Close packing density of polydisperse hard spheres. The Journal of

Chemical Physics 2009;131:244104-7.

[57] Brouwers HJH. Viscosity of a concentrated suspension of rigid monosized particles. Physical

Review E 2010;81:051402.

[58] de Bruijn H. General discussion. Discussions of the Faraday Society 1951;11:86.

[59] Roesink HDW. Microfiltration : membrane development and module design. Enschede1989.

[60] Wu L, Sun J, He C. Effects of solvent sort, PES and PVP concentration on the properties and

morphology of PVDF/PES blend hollow fiber membranes. Journal of Applied Polymer Science

2010;116:1566-73.

[61] Han M-J, Nam S-T. Thermodynamic and rheological variation in polysulfone solution by PVP

and its effect in the preparation of phase inversion membrane. Journal of Membrane Science

2002;202:55-61.

Page 121: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

121

[62] Tam CM, Tremblay AY. Membrane pore characterization—comparison between single and

multicomponent solute probe techniques. Journal of Membrane Science 1991;57:271-87.

Page 122: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

122

CHAPTER 5

CHARACTERISATION OF THE STAINLESS STEEL

HOLLOW FIBRES PRODUCED VIA PHASE INVERSION

AND THE EFFECT OF SINTERING PARAMETERS

5.1. ABSTRACT

Stainless steel hollow fibres were prepared and sintered under several conditions. The effect of

sintering temperature, particle size, particle loading, sintering atmosphere, different binders, and the

use of viscosity modifiers were studied and the effects on final morphology and mechanical

properties of the hollow fibres is described. The morphology obtained in the phase inversion

process is essentially maintained during sintering for almost all cases, with the exception of low

particle loading, where the large finger-like macrovoids collapsed due to the lack of enough

particles to promote neck formation and densification. The calcination temperature has a significant

impact on the properties of the membranes. With the increment of the sintering temperature, the

pore size and total porosity decreased while the amount of small pores, the mechanical robustness

and the flexibility increased. Hollow fibres produced with smaller particle sizes showed more

densification, faster neck formation and higher mechanical strength, while larger particles produced

hollow fibres that were more porous, have smaller necks between particles and had lower

mechanical performance. The use of different atmospheres also altered densification; argon helps

improve densification while nitrogen promotes slower densification for the same sintering

conditions. The effect that different binders have on the morphology and mechanical behaviour is

also shown. PESf produced fibres with larger total porosity, especially macrovoids. PEI in the other

hand was able to produce comparatively more small pores. The effect of the ratio between polymer

and solvent in the spinning dope was also analysed, and it was found that after sintering, the larger

Page 123: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

123

and more copious macrovoids produced in samples that contain more solvent were preserved after

sintering. However, the amount of pores under 0.5µm of radius was significantly decreased. Similar

effects were observed with the introduction of PVP where a higher PVP content led to lower

porosity for smaller pores and high total porosity.

In general hollow fibres could stand significant stresses under a three point bending test. High

porosity of small pores, under 0.5µm of radius, corresponded with a tolerance to higher stresses.

The position of the sponge-like region also plays a role in the maximum stresses that the hollow

fibre can stand. A sponge like region near the outer end is related to higher resistance; conversely

macrovoids near the outer surface result in a weaker hollow fibre. A similar, but less pronounced

behaviour was also observed for strain. In addition, macrovoids were found to be responsible to

initiate and propagate cracks when positioned near the outer surface. In contrast, when a sponge-

like structure is located near the outer surface, the strain at break was improved.

Page 124: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

124

5.2. INTRODUCTION

Porous materials have been utilised as filter elements and membranes since the beginning of the

nuclear age, but more recently they have found application as supports for inorganic membranes for

both liquid and gas separation processes. Conventionally, porous substrates are made from

inorganic materials such as alumina, titania, zirconia or carbon, particularly to produce robust

structures capable of meeting the high temperature requirements that are otherwise unavailable in

polymeric supports [1]. However, many industrial processes which require high pressures and

temperatures are also accompanied by vibration and mechanical stresses [2]. As a result, porous

ceramic supports tend to fail generally at the seal and membrane interface [3]. Therefore, the

mechanical robustness of such membrane supports is becoming an important industrial challenge.

To address this issue, porous metallic supports have been proposed and investigated as robust

alternative substrates for inorganic thin film membranes. Indeed, a flat silica membrane coated on a

porous stainless steel (SS) support demonstrated proof of concept by operating for over 1000 hours

in a working coal-fired power station [4]. Several different geometries including flat substrates [5,

6], tubular substrates [7] and hollow fibres [8-11] have been produced through tape casting [12],

pressing [13], extrusion [14] and dry-wet phase inversion spinning techniques [15, 16]. Flat

membranes are the most commonly investigated geometry due to the relative simplicity of

producing defect free, thin film membranes [17] and the ease with which they can be incorporated

into laboratory scale membrane modules. However, flat membranes have the lowest surface area to

volume ratio and the largest sealing perimeter of any of the investigated geometries, which gives

rise to larger module sizes and plant footprints, increasing capital costs and reducing their industrial

practicability. By contrast hollow fibres have the highest surface area to volume ratio and smallest

sealing perimeter, resulting in smaller unit operations and lower capital costs.

There are several parameters of engineering consideration in the preparation of robust porous

hollow fibres using stainless steel. In particular sintering effects play a major role in the final

mechanical strength associated with particle contact and neck formation [18, 19], mass transfer and

densification [20] and precise sintering control is necessary to avoid full densification [21, 22]. Of

particular attention, the characteristics of the precursor materials such as particle size, particle size

distribution, particle loading, binder thermal stability, coupled with sintering conditions

(atmosphere) and the energy provided to the system (temperature and time) must be considered

Page 125: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

125

[23]. Sintering of metal particles has long been studied [19, 23-26], however the focus has always

been to produce dense structures, thus the effect of sintering temperature, sintering time, sintering

atmosphere on porous hollow fibre production are not well known.

Therefore, the aims of this chapter are to investigate the effects of the stated sintering conditions on

the final hollow fibre morphology and mechanical performance. This chapter focuses on the effect

of sintering temperature, dwelling time and atmosphere, as part of the sintering conditions. In

addition, the effect of different binders, particle size and particle loading are also investigated to

predict the mechanical strength and pore properties of the hollow fibres. All stainless steel hollow

fibres were initially pre-calcined at 450oC in air to burn off the polymeric binder of the green fibre,

followed by sintering in an inert atmosphere at higher temperatures up to 1100°C.

5.3. SINTERING TEMPERATURE

The hollow fibres studied in this thesis were sintered at 950, 1000, 1050 and 1100°C with a

dwelling time of one hour.

The SEM images in Figure 5.1 clearly display different patterns resultant from sintering hollow

fibres at different temperatures. Although SEM is a surface analysis technique, and not a bulk

technique to determine densification, it can aid in the analyses of the samples in question. The

difference in densification of the hollow fibre surface between the four samples is self-evident.

Samples produced at 950°C (Figure 5.1a) show minimum particle coalescence, and those particles

appear to be weakly linked to each other though small necks. Large pores of irregular shapes and

sizes are present and are attributed to the lack of coalescence. As the sintering temperature

increased to 1000°C (Figure 5.1b), the coalescence between particles is more obvious with larger

agglomerates, like continuous strips of several particles melded together into wiggly type structures,

clearly observable. The extra densification induced by an increase of 50°C suggests that the

porosity has significantly decreased and pores are starting changing shape and become smaller. All

these characteristics are evidence that a surface diffusion process is taking place [18].

Samples sintered at 1050°C (Figure 5.1c) show further melding between particles, but large

particles are still distinguishable. However, the wiggly type structures are less evident, suggesting

that particles continued to aggregate further leading to an increase in densification. As a result, the

number of surface pores in comparison to samples produced at lower temperatures greatly reduced.

Page 126: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

126

At this stage, it is noteworthy to observe that smaller pores are starting to coalesce, suggesting a

change in the sintering mechanism from surface diffusion to a densification process [23]. Finally,

Figure 5.1d shows a sample sintered at 1100°C, where individual particles or groups of particles are

not observable. Large spheroidal pores are clearly present, despite a distinct decrease in the number

of surface pores. As such, it is clear that the sintering mechanism at higher temperatures promoted

diffusion between grain boundaries and densification [23, 27].

Figure 5.1 - Surface morphology of the hollow fibres sintered at temperatures various

temperatures (a) 950°C, (b)1000°C, (c) 1050°C, and (d) 1100°C for 1 hour.

To shed further light on the bulk structural properties of the hollow fibres, mercury porosimetry was

carried out on samples (Figure 5.2a) where it was clear that porosity decreases as a function of

temperature. For instance, the maximum porosities decreased from 27% at 950°C to 16% at

1100°C. These results indicate that there has been a densification of the porous hollow fibre, thus

confirming the SEM observations of the hollow fibre surface. However, it is interesting to observe

that the loss of porosity tends to be linear (around 2%) for every 50°C temperature increase from

950 to 1050°C, but from 1050°C to 1100°C led to a porosity loss of 6%, 3 fold higher than previous

rates. This non-linearity suggests that a different densification mechanism has occurred and

combined with the SEM analysis confirms grain boundary and / or volume diffusion.

Page 127: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

127

The pore size distribution varied from 1 to 2,000 or 2,500 nm (Figure 5.2b). Further analyses of

these results shows that an increase in sintering temperature led to a shifting of the pore size

distribution, towards pores >500nm in diameter. This is particularly evident in the hollow fibre

calcined at 1100°C, again confirming the transition towards particle aggregation where a

densification mechanism becomes more prevalent.

Sintering Temperature (°C)

900 950 1000 1050 1100 1150

Ma

xim

um

Po

rosity

(%)

10

15

20

25

30

Pore size diameter [nm]

050010001500200025003000

Diffe

rentia

l Intr

usio

n [m

L/g

/nm

]

0.0

2.0e-5

4.0e-5

6.0e-5

8.0e-5

1.0e-4

1.2e-4

1.4e-4 950 o

C

1000 o

C

1050 o

C

1100 o

C

Figure 5.2 – (a) Porosity (5%< error<14%) and (b) pore size distribution.

Further bulk structural analysis was carried out by single gas testing as displayed in Figure 5.3. The

nitrogen permeation (i.e. flux normalised for transmembrane pressure) decreased as the hollow fibre

sintering temperature increased, thus confirming the matrix densification evidenced in Figure 5.2.

Noticeably, there is a large gap in permeability values between samples produced at 1000°C and

1050°C and runs contrary to the linear decrease in porosity observed in Figure 5.2a, more closely

matching the non-linear relationship between 1050 and 1100°C as observed in the same figure.

These differences could be attributed to differences in the percolation pathways for gas permeance.

For instance, if a pathway is blocked, then nitrogen will not diffuse through it as per the single gas

results, but the same dead-end pore is still accessible to mercury resulting in a misleadingly high

porosity. These results suggest that the temperature effect is significant in the densification of the

smaller pores after 1000oC, which are brought together by the increase in heating energy which

ultimately begins to close these small pores. By the same token, larger pores require higher energy

levels (or longer sintering times) to have enough mass transfer to initiate closure and so tend to

remain open at the same temperatures.

The permeance tests of the SS hollow fibres show high nitrogen permeance rates, which are

comparable with those for previously published SS hollow fibres [28] and ceramic substrates [29,

30]. Essentially, nitrogen flow rates followed a viscous flow transport mechanism, exemplified by

a

A

b

B

Page 128: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

128

the large pore sizes formed between the SS particles (Figure 5.1). The permeance was almost

constant as a function of pressure difference across the membrane, though it decreased as the

sintering temperature increased. For instance, permeance decreased three fold, from 1x10-4 to 3x10-

5 mol m-2 s-1 Pa-1, as the sintering temperature rose from 950 to 1100oC, respectively. This reduction

is attributed to the reduction in pore volumes, particularly those smaller pore sizes as observed in

Figure 5.2b.

Pressure drop [kPa]

20 30 40 50 60 70

Pe

rme

ance

[m

ol m

-2 s

-1 P

a-1

]

3.0e-5

4.0e-5

5.0e-5

6.0e-5

7.0e-5

8.0e-5

9.0e-5

1.0e-4

1.1e-4

950oC

1000oC

1050oC

1100oC

Figure 5.3 - Nitrogen Permeation results

To fully understand how the change in sintering temperature and the resulting change in

morphology affects the mechanical strength of the hollow fibre, three point bending tests were

carried out and the results are shown in Figure 5.4. In general terms, both the flexural strain and

stress parameters increased with increasing sintering temperature due to the enhanced neck growth

between particles and overall densification as observed in Figure 5.1. In other words, at higher

temperatures, there is more energy available to enhance the mass diffusion process to the neck

region between particles.

However, the increase is not linear if the average stress and strain for each temperature are

considered in Figure 5.4. For instance, at 1100°C an increase in flexural stress is not evident, giving

similar average results as for those hollow fibres sintered at 1050oC. Indeed, the maximum flexural

stress as shown in Figure 5.4 appears to level out above 1050°C. This can be explained by the fact

that at this point necks are already fully grown and they reached their maximum strength. These

results can be matched by the SEM micrographs, as the samples sintered at 1050°C (Figure 5.1c)

Page 129: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

129

and 1100°C (Figure 5.1d) show similar surface patterns distinguishable from those at lower

temperatures. Therefore, the changes in flexural stress are marginal once this stage of densification

is reached.

The flexibility, on the other hand, is more sensitive to the densification process of the sample, this

means that a small increase in density will show a sensible improvement in the flexibility. Figure

5.4 confirms this point, as the flexural strain increased steeply from 950 to 1000 °C and 1050 to

1100 °C, but no significant changes were observed between 1000 and 1050 °C. Consideration is

given initially to those samples sintered at 950 and 1000 °C, where there was around 2% loss in

porosity (Figure 5.2a), indicating marginal densification of the porous matrix. The morphology

changed from particles joined together by small necks (Figure 5.1a) to strips of particles coalesced

together (Figure 5.1b). Therefore, this process led to the formation of larger, more interlinked

structures more able to utilise the flexural properties of the original stainless steel which yielded

increased flexural strain. In other words, neck growth was more important in imparting flexural

ductility than densification because the original contact points between the particles were so small

and weak. However, increasing the sintering temperature from 1050 to 1100 °C yielded a major

decrease in porosity, which further corresponded with a significant increase in flexural strain. In

other words, densification of the matrix played a major role as opposed to neck growth which has

all but stopped at this stage. Interestingly, the lack of changes in flexural strain between the hollow

fibres sintered between 1000 and 1050 °C correspond with an increase in neck growth into further

wiggly coalesced particles but loss of porosity was marginal. Hence, once the necks reached a

minimal size, further growth did not provide sufficient bulk mechanical changes and flexural strain

remained similar between 1000 and 1050 °C. However, once densification became the major impact

of continued sintering, the flexural strain correspondingly increased.

Sintering Temperature

950 °C 1000 °C 1050 °C 1100°C

Ma

xim

al F

lexu

ral S

tre

ss [M

Pa

]

200

300

400

500

600

700

800

900

1000

Sintering Temperature

950 °C 1000 °C 1050 °C 1100°C

Ma

xim

al F

lexu

ral S

tra

in [m

m/m

m]

0.000

0.005

0.010

0.015

0.020

0.025

0.030

(a)

)

(b)

Page 130: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

130

Figure 5.4 - Effect of sintering temperature on the ultimate flexural stress (a) and flexural strain(b).

Finally, there is a significant variation from average points for all tested hollow fibres. These results

may suggest that the pore formation in the hollow fibre is not homogeneous along the axis nor

within the walls of the hollow fibres. This point is supported by the computerised tomography scans

of the SS hollow fibres found in Figure 5.12.

5.4. SINTERING ATMOSPHERE

Light microscope images were taken in order to show the effect that different atmospheres have on

the sintering process. Figure 5.5a shows that samples produced in nitrogen exhibit a larger number

of pores spread out throughout the surface. The pores are generally in the region of <5m, though a

few larger pores possibly >10m are clearly present. By comparison, the sample sintered in argon

in Figure 5.5b, had far fewer pores generally, although a few very large pores in excess of 20m

were observable. As these images were taken from the same green hollow fibre sintered in different

atmospheres there is some evidence the sintering atmosphere affects the morphology of the bulk

structure of the hollow fibre. In other words, the sintering atmosphere may interact either with the

matrix of the hollow fibre or interfere in the sintering process itself and thereby induce different

morphologies.

Figure 5.5 – Light Microscope pictures of samples produced with AISI316L 10µm and sintered at

1050°C for 1 hour in a) Nitrogen atmosphere b) Argon atmosphere.

In order to quantitatively verify the light microscope observations of the effect of sintering

atmospheres, mercury porosimetry was carried out with the results shown in Figure 5.6. It is

interesting to observe completely different patterns of porosity. First, the porosity of nitrogen

a

a

a

b

Page 131: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

131

sintered hollow fibres was higher than the analogous argon sintered sample for both 950 and

1050°C. Second, at lower sintering temperatures the porosity is focussed around a single peak

corresponding to a pore size of 0.9μm for argon and 0.3μm for nitrogen. This closely aligns with the

findings in Section 5.3 where pore coalescence does not occur until higher sintering temperatures.

Third, the majority of the porosity can be found in pores > 200 μm and is associated with finger like

macrovoids, although there is also some porosity in the pore size range of 3 - 5μm and 0.01 -

0.1m. Turning to the argon sintered hollow fibre, there is almost no porosity below 1m, however,

there are distinct pore sizes around 3, 10, 15, 50 and 70 μm. Hence, if we consider only the sponge-

like region, the argon sintering atmosphere tends to form regions of larger pores and correlates well

with the visual observations of the surface of the hollow fibres shown in Figure 5.5.

Pore radius [µm]

0.0010.010.1110100

Po

rosity

[%l]

0

5

10

15

20

25

30

Argon 950°C

Nitrogen 950°C

a Pore radius [µm]

0.0010.010.1110100

Po

re d

istr

ibutio

n [%

of p

ore

s]

0

5

10

15

20

25

30

Argon 950°C

Nitrogen 950°C

b

Pore radius [µm]

0.0010.010.11101001000

Po

rosity

[%]

0

5

10

15

20

25

30

Nitrogen

Argon

c Pore Radius [µm]

0.0010.010.11101001000

Nitrogen

Argon

d

Figure 5.6 – Cumulative and frequency porosimetry distributions for samples sintered at 950°C (a,

b) and 1050°C (c, d) in Argon and Nitrogen.

In order to verify if the atmosphere has a chemical interaction with the metal during the sintering of

the hollow fibre, chemical analysis using Inductively-Coupled Plasma – Optical Emission

Spectrometry was carried out. The results of this analysis are shown in Table 5.1 for 316L and 410L

SS used to prepare hollow fibres in this thesis. The chemical analysis show a big increase in the

Page 132: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

132

carbon content of the hollow fibre, when compared to the data provided by the manufacturer of the

powder. The carbon content of the hollow fibres was an order of magnitude greater (from 0.03% to

~0.55% (316L SS) and 0.29% (410L SS)) compared to the original stainless steel particles,

independent of the sintering atmosphere. The carbon increase is directly attributed to sintering

process of the hollow fibre which, particularly the removal of the polymeric binder. Although the

polymer was initially burnt off in the initial pre-calcination step in air, and not visually observed in

the porous structures of the samples sintered at higher temperatures (Figure 5.5), the results in Table

5.1 clearly indicate that residual carbon from the binder reacted with the SS particles.

Table 5.1 – Chemical Analysis of samples sintered under different atmospheres.

Sample Condition Carbon (% w) Sulphur (%w) Oxygen (%w) Nitrogen (%w)

316L Original

composition* 0.03 0.03 (min) N/D N/D

410L Original

composition* 0.03 N/D N/D N/D

316L

10µm

Sintered in N2 0.534 0.065 5.712 0.018

316L

10µm

Sintered in Ar 0.551 0.040 5.808 0.016

410L

15µm

Sintered in N2 0.294 0.008 5.881 0.002

410L

15µm

Sintered in Ar 0.287 0.003 5.956 0.008

*Manufacturers specification sheet

Further analysis was carried by SEM EDX of two distinct regions as shown Figure 5.7(a and d) to

evaluate the composition of the steel. The EDX analysis in Figure 5.7(b, c, e and f) show that both

regions contained the major components of SS such as Cr, Mo, Fe, Mn, Ni and other minor

components plus carbon. It is interesting to observe that the peak intensity of carbon to other SS

component is higher in spectrum 2 than spectrum 1, regardless of sintering atmosphere. Hence, the

relative peak intensity of carbon to other components seems to be varied on the matrix surface.

These results suggest that during the sintering process, there is a chemical reaction between the

Page 133: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

133

carbon derived from the carbonisation of the polymeric binder with the surface of the SS particle

precursor. Hence, at the interface of the polymer binder and SS particle, the concentration of carbon

is higher than inside the particles. Further, the peak corresponding to carbon in the spectrum 2 is

stronger and accompanied by a strong peak of chromium at 5.5keV possibly meaning that

chromium carbide was formed as reported elsewhere [31, 32]. As this sintering was carried under

inert gas atmosphere, then, in principle residual carbon is retained in the matrix of the hollow fibres

and high temperature sintering allows further carbon diffusion along the concentration gradient and

away from the particle surface (process known as carburization [33]). Correspondingly, in spectrum

1, where the intensity of carbon peak is much lower, as to is the chromium peak at 5.5 keV.

Figure 5.7 –SEM image of a 316L 10m sample sintered in (a) Argon and (d) Nitrogen for two

hours. Respective EDX spectra for Argon (b) spectrum 1 and (c) spectrum 2; Nitrogen (e) spectrum

1 and (f) spectrum 2.

Page 134: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

134

Closer inspection of the image identifies that the carbide rich regions are a darker grey than the

carbide lean regions, whilst the pores themselves are a deep black. Histogram analysis of the image

provides the relative areas associated with each colour which in turn allows us to estimate the area

belonging to each region and these results are shown in Table 5.2. Firstly, these results again

correlate with the light microscopy and porosimetry analysis in that the nitrogen sintered sample is

significantly more porous than the sample sintered in argon. Secondly, the argon sintered sample

contains ~14% more carbide-rich regions (by area) than the sample sintered in nitrogen (57% as

compared to 49%).

Table 5.2 – Percentage of surface area associated with each region of EDX analysis relative to

sintering atmosphere

Sintering

Atmosphere

Porous Region Carbide-Rich Region Carbide-Lean Region

Argon 8% 57% 36%

Nitrogen 17% 49% 34%

The effect of sintering with inert gases on the mechanical properties was also measured and the

results are plotted in Figure 5.8 The flexural stress (Figure 5.8a) increased with temperature and the

hollow fibres sintered in argon always gave higher flexural stresses than the analogous sample

sintered in nitrogen. These results correlate well with the porosity values in Figure 5.6, with the

least porous sample (argon sintered) possessing the greatest flexural strength. These results are

similar to findings for the hollow fibres sintered in an oxidising atmosphere (i.e. air) as discussed in

section 5.2 above and indicate that, for the steels in question, the structure of the hollow fibre is

more crucial in determining its strength than minor changes in its chemical make-up.

Similarly to the discussion in section 5.2, the flexural strain has a different trend from the flexural

stress as depicted in Figure 5.8b. In this case, increasing the temperature from 950 to 1050°C had

no impact on the flexural strain for the argon sintered sample, whilst for the nitrogen sintered

sample it was increased around two-fold. These results correlate well with the micrographs in

Figure 5.5 as the argon sintered samples had very large pores as compared to the nitrogen sintered

hollow fibres. Under static load and coupled with the fact that the hollow fibre wall is very thin

(~250 m), large pores are like structural defects which induce mechanical weakness and reduce the

Page 135: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

135

ability to oppose deformation effects. Lastly, the carbide phase identified in Figure 5.7 is more

brittle than the original stainless steel [34], so even though the argon sintered sample is denser than

the nitrogen sintered sample (see Figure 5.6), the strain at break is less.

Sintering Temperature [°C]

900 950 1000 1050 1100

Ulti

ma

te F

lexu

ral S

tre

ss [M

Pa

]

0

200

400

600

800

1000

1200

Nitrogen

Argon

Sintering Temperature [°C]

940 960 980 1000 1020 1040 1060M

axim

um

Str

ain

[m

m/m

m]

0.004

0.006

0.008

0.010

0.012

0.014

0.016

0.018

Nitrogen

Argon

Figure 5.8 - Flexural Strength (a) and strain (b) of AISI 316L hollow fibres product of different

sintering atmospheres.

To some extent, the results discussed in this section, which are mainly focused on porous SS

materials differ, from those published in the literature for fully dense materials [35-37]. In

particular, argon is not considered a good sintering atmosphere because once the pores close it

remains trapped promoting the formation of isolated pores within the stainless steel matrix [35]. On

the other hand nitrogen can diffuse through the stainless steel [38-41], not only react to form

nitrides, which indirectly allows the nitrogen to escape from the closed pores and therefore achieve

higher densification. However, the densification mechanisms may be analogous and should be

considered further. For instance, argon and nitrogen sintering atmospheres produced hollow fibres

with different final microstructures and mechanical properties. The results seem to suggest that

sintering in nitrogen inhibited the densification process and that this is associated with the formation

of chromium carbide. However, it is well known that nitrogen adsorbs onto the SS surface,

inhibiting mass diffusion and neck formation at the beginning of sintering [35, 36, 42]. This slows

down the sintering process and so the argon sintered samples appear relatively denser.

In this thesis the precursor materials differ from those dense SS materials previously studied. The

hollow fibre contains a higher concentration of polymeric binder as well as using SS particles which

have been previously processed and sintered. As both argon and nitrogen are inert gases, they

a b

Page 136: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

136

should in principle retain similar amounts of excess carbon in the final hollow fibre. Indeed, it

would also be expected that both the interfaces between SS particles and the interior of the pore

spaces would be relatively rich in carbon regardless of the sintering atmosphere used. The increased

carbide formation in the argon sintered sample may result from residual carbon (not removed in the

pre-sintering step) becoming trapped within the rapidly closing pores, later reacting with the

chromium as densification and sintering progresses. In contrast, the pores of the nitrogen sintered

sample remain more open and possibly allow the residual carbon to be removed in the gas phase

before it can react with the SS surface.

5.5. DWELLING TIME

The effect of dwelling time was addressed in the literature review (Chapter 2) which has a direct

impact in the densification of materials [18, 23, 43]. In this thesis, 316L samples sintered in

nitrogen at dwell times of one and four hours were analysed. SEM images in Figure 5.9 show the

effect of dwelling time on the morphology of the hollow fibres. Both hollow fibre cross sections

display representative pore formation, containing sponge like formation with small pores, and large

finger-like structures due to the solvent and non-solvent interactions in the fibre during coagulation,

a common feature of the phase inversion process. Visually, it is difficult to say with certainty that

there are any significant differences with the dwell time. Hence, mercury porosimetry was used as a

bulk analysis technique.

Figure 5.9 - SEM image of a hollow fibre sintered at 1050°C under nitrogen for (a) 1 hour and (b)

4 hours.

a b

Page 137: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

137

The results of the mercury porosimetry in Figure 5.10 indicate there is only a marginal decrease in

porosity as the dwell time increases from 1 to 4 hours, particularly for the 6 and 45μm samples.

Interestingly these results also show that the effect of dwell time upshifted the pore size distribution

for all particle sizes. All the samples sintered for 4 hours had a significant fraction of larger pores

most likely associated with the finger like pores observed in Figure 5.9. Indeed, for samples with

larger particles (i.e. 16 - 45μm) sintered for 4 hours the majority of pores were >10μm, with more

than a quarter of the porosity associated with pores >100μm. The lack of pores <1μm agrees with

the sintering theory in that longer dwell times allow densification to proceed further with the small

pores amongst the first to disappear. The exception here was the samples made from 6μm particles

where the majority of pores where ~1μm regardless of sintering time. As sintering time increased

the pores there was an increase in larger pores which corresponds well with the known pore

coalescence model.

Pore radius [µm]

0.0

00

1

0.0

01

0.0

1

0.1

110

10

0

10

00

Tota

l P

oro

sity [

%vo

l]

0

5

10

15

20

25

30

6µm - 1H

6µm - 4H

a Pore radius [µm]

0.0

00

1

0.0

01

0.0

1

0.1

110

10

0

10

00

Tota

l P

oro

sity [

%v

ol]

0

5

10

15

20

25

30

16µm - 1H

16µm - 4H

b

Pore radius [µm]

0.0

00

1

0.0

01

0.0

1

0.1

110

10

0

10

00

Tota

l P

oro

sity [

%v

ol]

0

5

10

15

20

25

30

45µm - 1H

45µm - 4H

c

Figure 5.10 – Porosimetry of samples made of particles of 6µm (a), 16µm (b) and 45µm(c) results

as a function of dwelling time for hollow fibres sintered at 1050oC.

Figure 5.11 shows the three-point bending test as a function of the dwelling time. The results show

that both flexural stress and strain were the same independently if the hollow fibres were sintered at

one or four hours. These results strongly suggest that the sintering process quickly attained the same

level of mechanical strength and then plateaued for longer dwell times. Again this corresponds well

Page 138: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

138

with previous analysis in that the porosity of the hollow fibres is a significant factor in the

mechanical properties, and since the porosity doesn’t change as a function of dwell time, neither do

the mechanical properties. Hence, the only significant changes observed with the dwell time are

related to the pore size distribution of the hollow fibres rather than their mechanical strength.

Particle size

6 µm 16 µm 45 µm

Ma

xim

um

fle

xura

l str

ess [M

Pa

]

0

200

400

600

800

1000

1 hour dwelling time

4 hours dwelling time

a

Particle size

6 µm 16 µm 45 µm

Ma

xim

um

fle

xura

l str

ain

[m

m/m

m]

0.000

0.002

0.004

0.006

0.008

0.010

1 hour dwelling time

4 hour dwelling time

b

Figure 5.11 - Effect of dwelling time on mechanical strength (a) and strain (b).

5.6. STAINLESS STEEL PARTICLE SIZE

Figure 5.12 shows the change in morphology that occurs after sintering hollow fibres with SS

particles ranging from 6 to 45m. The SEM micrographs show that the hollow fibres prepared with

the smaller particle sizes of 6, 10 and 16 m resulted in similar structures based on their cross

sections. Finger like structures are still observed although these cross sections are dominated by

sponge like structures. However, the hollow fibres prepared with the larger particle size of 45 m

resulted in a very inhomogeneous structure, containing many very large pores in the centre of the

hollow fibre wall with pore dimensions in excess of 100m. Closer examination shows that

individual particles were still visible, forming agglomerates with small necks only. This is not the

case for the hollow fibres prepared with the smaller particles where significant neck growth and

densification was observable. Hence, the particle size influences the membrane densification

process.

Page 139: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

139

Figure 5.12 – CT scans of hollow fibres produced with stainless steel of 6µm (a), 10µm (b), 16µm

(c) and 45µm (d) particle sizes sintered at 1050°C for one hour.

The dimensions of the sintered hollow fibres were measured and compared with their analogous

green fibres. Figure 5.13a shows a reduction of both outer (OD) and inner (ID) diameters after

sintering. The reduction was consistently between 65-70% irrespectively of the precursor particle

size. This clearly indicates that the process of sintering brings the SS precursor particles together as

the binder is degraded. As the particles come closer, the heating process allows the particles to

coalesce and the porous structure to densify. The SS hollow fibre shrinkage process is radial, i.e.,

towards the inner surface of the hollow fibre. However, the shrinkage is relatively low, particularly

for the large particle size. This is noticeable in Figure 5.13b, as the wall thickness of SS hollow

fibre (45m) as green or sintered materials were almost the same. This result reflects the CT scan

images in Figure 5.12, showing very large pores for the hollow fibre prepared (45m). In principle,

if the sintering process proceeds far enough (i.e. towards densification) it should reduce the size and

number of macrovoids, as it is observed for the hollow fibres prepared with 6 or 10 m SS particles

[44]. The lack of wall shrinkage for the 45m SS particles in Figure 5.13b suggests that the

Page 140: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

140

particles were too large and that the surface area to volume ratio of the particles was too large and

hence inter-particle mass transfer was too slow (or did not proceed far enough during the sintering

step) to enable much densification.

Particle diameter (µm)

0 10 20 30 40 50

Dia

me

ter

(µm

)

400

600

800

1000

1200

1400

1600

1800

2000

2200

2400

O/D Green

I/D Green

O/D Sintered

I/D Sintered

a

Particle diameter (µm)

0 10 20 30 40 50W

all

Thic

kne

ss (

µm

)

200

250

300

350

400

450

500

550

Green

Sintered

b

Figure 5.13 – Comparison of the dimensions of green and sintered (1050°C) hollow fibres, in terms

of diameters (a) and wall thickness (b)

Figure 5.14a shows the mercury porosimetry results for hollow fibres prepared with varying SS

particle size. There are some interesting variations observed with these results. First, there is a

major difference in total porosity between the small particles (6 and 10 µm) and the larger SS

particle sizes of 16 and 45 µm. The smaller particles gave more compacted structures with low

porosity, the majority of which is associated with pores <1μm. As the particle increases from 10 to

16 µm, the porosity is increased by at least twofold (and further for 10 to 45 μm). Hence, these

results strongly suggest that there are mass transfer limitations during sintering for SS particles with

sizes in excess of 10 µm. These results correlate well with the CT scan images in Figure 5.12. The

pore size distribution in Figure 5.14b elucidates for larger particle sizes of 16 and 45 m, the

porosity is contained in the large pores (50<dp<500m).

Page 141: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

141

Pore radius (µm)

0.00010.0010.010.11101001000

Po

rosity

(%)

0

10

20

30

40

50

60

6µm

10µm

16µm

45µm

a

Pore radius (µm)

0.0010.010.11101001000

6µm

10µm

16µm

45µm

b

Figure 5.14 - Mercury porosimetry results comparing hollow fibres produced in function of

porosity (a) and pore distribution (b).

Three point bending tests were also carried out in order to verify the influence of the morphology

on the mechanical strength of the produced hollow fibres, and the results are shown in Figure 5.15.

There is an order of magnitude reduction in ultimate flexural stress as the SS particle sizes increase

from 6 to 45 µm. Both flexural stress and strain parameters follow similar trends as a function of the

particle size. A short deviation in this trend occurs for the hollow fibres with SS particles 16 µm,

where the flexural strain remains in the same range as the SS 10 µm particle sizes, instead of

decreasing as it happed for the flexural stress.

Particle Size

6µm 10µm 16µm 45µm

Ma

xim

al F

lexu

ral S

tre

ss (

MP

a)

0

200

400

600

800

1000

1200

a

Particle Size

6µm 10µm 16µm 45µm

Ma

xim

al F

lexu

ral S

tra

in (

mm

/mm

)

0.003

0.004

0.005

0.006

0.007

0.008

0.009

b

Figure 5.15 – Effect of particle size on the ultimate flexural stress (a) and Flexural strain (b).

Page 142: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

142

There is a strong correlation between the morphology of the hollow fibres and their mechanical

properties as a function of the particle sizes. The hollow fibres prepared with smaller particle sizes

(6 and 10 µm) formed similar morphologies, and their mechanical properties are also similar. As the

particle size increased to 16 µm, the morphology changed significantly (see Figure 5.14) as the

porosity doubled, and in particular there was a major increase in the region of large pores. Hence,

these results strongly suggest that mass diffusion during sintering tends to be almost constant for

particles at or below 10 µm. Above this particle size, mass diffusion limitations increased

significantly, which combined with an inherent decrease in packing density of larger particles gives

an increase in porosity. These morphological changes are accompanied by the weakening of the

mechanical structure of the hollow fibre, as the flexural stress reduced by 50%. This trend carries

through to the largest particle size tested with very large pore sizes as clearly observed in Figure

5.12. The lack of structural densification coupled with large pore sizes (which act as structural

defects initiating breakage under stress) explain the significant loss of mechanical strength for the

hollow fibres prepared with SS particle sizes of 45 µm.

5.7. POLYETHERIMIDE TO SOLVENT RATIO

Chapter 4 discussed how the composition of the spinning dope affected of the final green hollow

fibre morphology. So, it is also an important question to address as to whether the binder

concentration (PEI) of the hollow fibres also affects the properties of the final sintered hollow fibre.

The SEM images in Figure 5.16 illustrate the morphological characteristics of the hollow fibre

obtained after sintering of the green bodies prepared differing polymer to solvent ratios of 1:3

(right) and 1:4 (left).

Page 143: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

143

Figure 5.16 - Hollow fibres sintered at 1050°C produced using polymer to solvent ratio of 1:4 (left)

and 1:3 (right) in the spinning dope

Helium pycnometry was also carried out and showed that they higher PEI loading (1:3) gave an

increased density of 6.06 g-1 cm3 as compared to 5.81 g-1 cm3 for the lower PEI loading (1:4)

sample. Further structural analysis was conducted by mercury porosimetry as shown in Figure 5.17.

The porosimetry results agree with the helium pycnometry, showing that samples containing the 1:4

polymer to solvent ratio present larger porosity compared to samples produced using 1:3 ratio. It is

interesting to observe the porosity distribution is relatively similar for the 1:3 and 1:4 samples. In

both cases it is bimodal with ~20% of the porosity (5-7 absolute percentage points) associated with

pores >1μm whilst the remaining 80% (between 15-20 absolute percentage points) was associated

with pores < 0.1 µm pore sizes. There was a slight broadening of the distance between the peaks as

the PEI content increased from 1:4 to 1:3. That is, the peak associated with small pores downshifted

from 0.3 to 0.2μm whilst the peak associated with larger pores upshifted from 60 to 200μm.

Pore radius (µm)

0.0010.010.11101001000

Po

rosity

(%)

0

5

10

15

20

25

30

1:3 - PEI : NMP

1:4 - PEI : NMP

a

Pore radius (µm)

0.0010.010.11101001000

1:3 - PEI : NMP

1:4 - PEI : NMP

b

Figure 5.17 - Mercury porosimetry of sintered samples produced using polymer to solvent ratio of

1:4 and 1:3.

Page 144: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

144

In principle, these morphological features are directly related to the viscosity of the samples, which

is altered by changing the solvent to PEI ratio for the production of green fibres. It is noteworthy

that since larger pores and porosity produced using 1:4 PEI ratio during the spinning process (see

Chapter 4), these large pores (i.e. finger-like macrovoids) keep the steel particles further apart thus

hindering particle to particle contact. Conversely, samples made with 1:3 PEI ratio show lower

porosity and smaller pore sizes associated with a greater sponge like region. Hence, closer contact

of the SS particles in the sponge-like structures enhanced the sintering process by creating smaller

pores and a more consistent sintering [21], although the variation based on the PEI to solvent ratio

is quite small. The flexural stress was kept almost constant independently of the PEI loading as

shown in Figure 5.18. However, the flexural strain variation was very small and within

experimental error. Hence, the mechanical properties of the hollow fibres were not significantly

altered, which is not unsurprising given the similar total porosity and porosity distributions.

Solvent to polymer ratio in the spinning dope

'1:4 '1:3

Ma

xim

al F

lexu

ral S

tre

ss (

MP

a)

700

800

900

1000

1100

1200

1300

a

Solvent to polymer ratio in the spinning dope

'1:4 '1:3

Ma

xim

al F

lexu

ral S

tra

in (

mm

/mm

)

0.0080

0.0082

0.0084

0.0086

0.0088

0.0090

b

Figure 5.18 - Result of 3 point bending test carried out on fibres produced using polymer ratio of

1:4 and 1:3 as function of maximum bending strength (a) and strain (b)

5.8. VISCOSITY MODIFICATION WITH POLYVINYLPYRROLIDONE

Polyvinylpyrrolidone (PVP) is a viscosity modifier used in the preparation of the spinning dope as

discussed in Chapter 4. To investigate the effect of PVP on the sintering process, SEM images are

displayed in Figure 5.19. The morphology obtained for the sintered SS hollow fibres was very

similar in all cases, regardless of the amount of PVP added. Further, the morphology mimics the

Page 145: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

145

morphology of green fibres (see Chapter 4) as large finger-like macrovoids were preserved at the

inner and outer surfaces of the hollow fibre separated by a thin sponge-like ring at the centre of the

hollow fibre wall. Helium pycnometry results showed that the variation in the density was minor, in

the region of ±10%, for all samples, although it did generally increase with the amount of PVP in

the doping mixture. The effect was marginally more pronounced for the SS hollow fibres prepared

with 1:4 PEI ratio where the density increased from 5.5 to 6.1 g cm-3 as PVP content went from 0.5

to 1.5%. In comparison, the density increased from 5.83 to 6.07g cm-3 as the PVP content increased

from 0.5% to 1.5% for samples containing a PEI to solvent ratio of 1:3.

Figure 5.19 - SEM images of the hollow fibres produced using PVP in the spinning dope.

The mercury porosimetry plots in Figure 5.20 show interesting results. Firstly, samples prepared

with a 1:4 PEI to solvent ratio typically recorded a higher porosity than for the analogous 1:3 ratio

samples. Secondly, the porosity of the samples typically decreased as the PVP content increased,

although this effect was more pronounced for the 1:4 ratio samples with the porosity approximately

equal (~24%) for the 1.5% PVP samples regardless of the PEI to solvent ratio. PVP addition also

impacted on the porosity distribution of the samples, although the most significant changes were

Page 146: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

146

observed when comparing between samples with differing PEI to solvent ratios at constant PVP

addition. For example, whilst the fraction and pore size of the larger pores was relatively unchanged

between samples, particularly at a PVP content of >0.5%, there was a greater percentage of pores

between 1-5μm for the 1:3 ratio samples as compared to the 1:4 samples where the majority of

small pores were between 0.1-1μm. Although, as PVP content increased the smaller pores in the 1:3

sample trended towards the 1:4 sample, that is, became smaller. These results suggest a strong

interaction between the solvent and the PVP, and show that the PVP increased the viscosity of the

spinning dope reducing the mobility of the particles during the coagulation process and allowing the

steel particles to remain together favouring sinterability.

Pore radius (µm)

0.0

01

0.0

1

0.1

110

10

0

10

00

Tota

l P

oro

sity (

%vo

l)

0

10

20

30

1:3 - 0.5% PVP

1:4 - 0.5% PVP

a Pore radius (µm)

0.0

01

0.0

1

0.1

110

10

0

10

00

Tota

l P

oro

sity (

%v

ol)

0

10

20

30

1:3 - 1% PVP

1:4 - 1% PVP

b

Pore radius (µm)

0.0

01

0.0

1

0.1

110

10

0

10

00

Tota

l P

oro

sity (

%vo

l)

0

10

20

30

1:3 - 1.5% PVP

1:4 - 1.5% PVP

c

Pore radius (µm)

0.0010.010.11101001000

Po

re d

istr

ibutio

n (

% o

f to

tal p

oro

sity)

0

2

4

6

8

10

12

14

1:3 - 0.5% PVP

1:4 - 0.5% PVP

d Pore radius (µm)

0.0010.010.11101001000

Po

re d

istr

ibutio

n (

% o

f to

tal p

oro

sity)

0

2

4

6

8

10

12

14

16

18

20

1:3 - 1% PVP

1:4 - 1% PVP

e

Pore radius (µm)

0.0010.010.11101001000

Po

re d

istr

ibutio

n (

% o

f to

tal p

oro

sity)

0

5

10

15

20

25

1:3 - 1.5% PVP

1:4 - 1.5% PVP

f

Figure 5.20 - Mercury porosimetry of samples produced using green fibres containing polymer to

solvent ratios of 1:3 and 1:4 and PVP content from 0.5% (a, d), 1% (b, e) and 1.5% (c, f).

Page 147: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

147

The flexural stress and strains in Figure 5.21 remained similar and within the experimental variation

for all samples independently of the PEI to solvent ratio or PVP loading. The mechanical testing

results reflect the morphological results. For instance, the helium pycnometry gave similar

densities. In addition, the morphological variations observed from the mercury porosimetry were

also small. Therefore, the mechanical properties were consistent with the morphological results.

a

% of PVP on spinning dope

0.0 0.5 1.0 1.5 2.0

Ma

xim

um

Fle

xura

l Str

ess (

MP

a)

300

400

500

600

700

800

1:3 PEI : NMP

1:4 PEI : NMP

a

b

% of PVP on spinning dope

0.0 0.5 1.0 1.5 2.0

Ma

xim

um

Str

ain

(m

m/m

m)

0.002

0.004

0.006

0.008

0.010

1:3 PEI : NMP

1:4 PEI : NMP

b

Figure 5.21 - Mechanical strength (a) and strain (b) of the hollow fibres produced using 1:3 and

1:4 PEI to NMP ratios and adding PVP as viscosity modifier.

These results are not coincident with results of previous works for polymeric hollow fibre, where

the addition of PVP shows a striking increase in the fraction of the sponge-like region [45-47]. This

difference could be explained by the fact that either the PVP content is not large enough to

satisfactorily alter the phase inversion kinetics or that the presence of particles alters the diffusion

path of the solvent within the polymer during the phase inversion, as discussed in Chapter 4.

5.9. POLYETHERIMIDE VS POLYETHERSULFONE

PESf has been generally used as the polymeric binder of choice in the preparation of ceramic

membranes as discussed in Chapter 2. The same applies to a limited number of SS hollow fibres

reported in the literature. The use of PEI was studied 5 years ago by Leo and co-workers [48] for

Page 148: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

148

ceramic membranes for the first time, showing benefits of using binders that didn’t contain sulphur.

Hence, the effect of both polymeric binders was investigated in this thesis. A TGA analysis was

carried out in an argon atmosphere for the pure binders as shown in Figure 5.22, in order to select a

suitable debinding temperature. The first initial mass loss occurs below 100 °C and represents the

removal of the residual solvent and water from the material. The second mass loss represents the

onset of the debinding step at ~450 °C for PEI and ~500 °C PESf. The presence of the sulphur in

the PESf binder makes it more resistant to thermal degradation and explains the higher burn off

temperature [49].

This implies that the particle to particle contact, which influences the mass transfer in the sintering

process, will begin at a lower temperature for the PEI+SS sample as compared to the PESf+SS

sample. This could mean that marginally lower sintering temperatures and shorter hold times are

required for the PEI+SS sample as compared to the PESf+SS sample. Conversely, if the same

conditions are used, then the PEI+SS sample should be marginally denser and stronger than the

PESf+SS sample. However, as the difference in burn-off occurs at low temperatures (relative to the

melting point of stainless steel), it is possible that the low temperature may not significantly affect

the overall morphological or mechanical properties of the subsequent hollow fibres. The final mass

loss stage occurs between 800-1000 °C and represents between 1.4 and 2% of total mass loss

respectively. The final mass following sintering is very close (94.5% for PESf and 93.9% for PEI).

The theoretical amount of stainless steel contained in the samples is of around 88%. This difference

can be caused by residual polymer (like carbon from pyrolysis due to the lack of oxygen in the test)

and experimental error (±5%). These results demonstrate that there is no significant difference in

using either PEI or PESf in terms of polymer decomposition.

Page 149: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

149

Temperature (°C)

0 200 400 600 800 1000 1200 1400

Ma

ss V

ari

atio

n (

%)

93

94

95

96

97

98

99

100

101

Stainless Steel + PEI

Stainless Steel + PESf

Figure 5.22 - TGA results for hollow fibres produced with PEI and PESf carried out under argon

atmosphere.

Nevertheless, the morphology obtained by using PESf (a) and PEI (b) to produce SS hollow fibres

varies as shown in Figure 4.2. The sample which utilised PESf as a binder showed larger

macrovoids and greater apparent porosity. In particular, large pores are located along both the outer

and inner surfaces of hollow fibre, whilst even larger voids are located in the centre of the fibre. In

comparison, samples which utilized PEI as a binder showed significantly fewer finger-like voids

and lower apparent porosity. In particular, no large pores where evident at the outer surface of the

hollow fibre, rather the sponge-like structure extended from the outer surface inwards past the

centre point of the cross-section. There were a number of finger-like voids towards the inner shell

of the fibre, but these were smaller in size when compared to the PESf+SS sample. Notably, in the

immediate area next to the inner surface, the sponge-like structure was again prevalent for both the

PESf+SS and PEI+SS hollow fibres, although this region was thicker in the PEI+SS sample.

The morphological variations in Figure 4.2 clearly indicate that the different binders produced

different green hollow fibres, as the sintered fibre tends to mimic the morphology of the green fibre.

Hence, this difference is consequence of a complex series of interactions, but is due to the both the

difference in apparent viscosities during the green fibre spinning and coagulating as discussed in

Chapter 4 and difference in the precipitation value. In other words, the amount of non-solvent

Page 150: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

150

required to initiate coagulation of PESf or PEI [50] affects the morphological formation and it is a

function of the polymer, solvent and non-solvent interaction.

To better understand the morphology of the PESf+SS and PEI+SS samples, mercury porosimetry

was undertaken to fully characterise the pore size distribution and porosity of the sintered hollow

fibres. The results clearly indicate that the PESf+SS sample had a porosity of ~52%, almost double

that of the PEI+SS sample which had a porosity of 27%. Furthermore, the pore radius within the

PESf+SS sample were predominately distributed around a single, large peak between 5 and 0.5 μm,

although smaller peaks were evident at 7μm, and between 200 – 100nm. A very broad increase in

porosity is observable over the 200 to 7 μm range which is attributed to the very large macrovoids

evident in Figure 5.23(b). The porosity contribution of the pores was 5%, 7%, 33% and 5% for the

macrovoids, 7μm, 5μm and 200nm pores respectively. By contrast the pore size distribution of the

PEI+SS was clearly trimodal with a small peak between 7 and 6μm, a larger peak between 5 and 2

μm and the final peak around 500 to 200 nm. The porosity contribution of each pore size was 3%,

8% and 15% respectively. The micron sized pores are related to the large, finger-like voids present

in both fibres and confirm the SEM observations that the PESf+SS sample contained many more

than the PEI+SS sample. The pores <500nm in size are related to the sponge-like structures evident

from the SEM analysis and again confirm that the majority of the PEI+SS sample was related to this

feature. Therefore, these results are congruent with the SEM analysis in Figure 4.2 and provide

more evidence that the solvent / non solvent exchange rates are significantly different for the PESf

and PEI systems as previously discussed.

a

Pore radius (µm)

0.0010.010.11101001000

Po

rosity

(%)

0

10

20

30

40

50

60

PEI

PESf

a

b

Pore radius (µm)

0.0010.010.11101001000

PEI

PESf

b

Figure 5.23 - Mercury porosimetry results for the fibres prepared using PESf and PEI polymers.

Page 151: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

151

To this end three point bending tests were performed on the PESf+SS and PEI+SS sintered hollow

fibres and the results are displayed in Figure 5.24. There is a significant difference in the flexural

stress. The PEI+SS samples show a much higher mechanical strength (1000 ± 62 MPa) when

compared with the PESf samples (500 ± 100 MPa), a difference of 100%. By contrast the PESf+SS

sample demonstrated slightly higher flexural strain (as measured indirectly by the strain at break)

with values of 0.01125 ± 0.0009 mm/mm observed, whereas for the PEI+SS samples the strain was

approximately 9.5 ± 0.00015 mm/mm. The flexural strain variations of both samples are within

experimental error, and are not significantly different.

Polymer on Green fibre

10µm PEI 10µm PESf

Ma

xim

al F

lexu

ral S

tre

ss (

MP

a)

300

400

500

600

700

800

900

1000

1100

a

Polymer on the Green fibre

10µm PEI 10µm PESf

Ma

xim

al F

lexu

ral S

tra

in (

mm

/mm

)

0.000

0.005

0.010

0.015

0.020

b

Figure 5.24 - Results of 3 point bending test carried out in the hollow fibres. (a) Maximum Flexural

Stress, (b) Maximum bending strain.

Therefore it was necessary to analyse the PESf+SS and PEI+SS hollow fibres for sulphur

contamination using Inductively-Coupled Plasma – Optical Emission Spectrometry and the results

are listed in Table 5.3. The results of the chemical analysis clearly indicate that the PESf+SS hollow

fibres have an increased sulphur content, both with respect to the original 316L stainless steel

particles and the PEI+SS post-sintered hollow fibres. Both SS hollow fibres resulted in sulphur

content around 0.07wt%. Although this values is higher than that of the original PESF fibre of

0.03wt%, there is a discrepancy for the PEI+SS hollow fibres as PEI does not contain sulphur.

Hence, sulphur contamination occurred during the sintering process as the same furnace has been

used. These results suggest that a minor concentration of sulphur reacts with the SS particles during

sintering.

Page 152: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

152

Table 5.3 – Chemical composition of the original 316L SS particles and the post-sintered PEI+SS

and PESf+SS hollow fibres.

Sample C %w Σ S %w σ O %w σ N %w σ

316L* 0.03 Max. 0.03 Max. N/D -- 0.10 Max.

316L – PEI 0.534 0.232 0.065 0.073 5.784 0.140 0.015 0.004

316L – PESf 0.603 0.047 0.070 0.004 3.107 0.131 1.023 0.045

* Manufacturers specification sheet

5.10. PARTICLE LOADING

The effect of SS particle size loading in the final hollow fibre was also investigated. Samples

produced using 50% and 70% particle loadings (by mass) were sintered. The SEM images in Figure

5.25 show the morphological difference between the two samples. The 50% particle loading

resulted in a misshapen hollow fibre post sintering due to excessive shrinkage and generally a lack

of sufficient particles to hold the shape. Contrary to this, the sample with 70% particle loading

clearly displays a round shape for both the inner and outer shells. Lower particle loadings of 10%

were also attempted, though it was not possible to sinter the hollow fibres. This is in line with the

minimum amount of particles needed for sintering hollow fibres as described by Luiten-Olieman

[51]. These results clearly indicate that the excess binder reduced the proximity of the SS particles

and hindered sintering by preventing neck formation and coalescence. In this case the SS particles

were affected by gravity with the hollow fibre collapsing or slumping during the debinding process.

This was particularly true for the 50% loading samples which developed a flat bottom during

sintering. In this process particles are not homogeneously sintered leading to the formation of large

cavities. Contrary to this, the high 70% particle loading allows for closer proximity of the SS

particles which sinter and coalesce radially towards the centre of the hollow fibre inner shell as the

hollow fibre shrinks during sintering. Hence, the shape of the hollow fibre is maintained during the

sintering process.

Page 153: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

153

Figure 5.25 - SEM images of the hollow fibres produced using solid loads of 50% (a) and 70% (b).

Helium pycnometry results for these two samples resulted in densities of 5.11g.cm-3 (50% particle

loading) and 6.05g.cm-3 (70% particle loading). These results show that by decreasing the particle

loading 20% the density decreases 15.5%. These results match the SEM images in Figure 5.25, thus

indicating that cavities in the 50% sample reduced the density of the hollow fibre. Further, helium

pycnometry results also match the mercury porosimetry displayed in Figure 5.26. For instance, the

50% particle loading sample had the highest maximum porosity of over 36%. In addition, 26% (out

of 36%) of the porosity corresponded to very large pores in the region between 65 and 5µm, while

the remaining 10% was associated with pores between 5µm and 0.5µm. The 70% particle loading

sample had a broader porosity distribution with larger pores between 400 and 20µm associated with

~5% (of ~20%) of the porosity, whilst the remaining 15% (of ~20%) of the porosity was associated

with pores between 0.1 and 0.08µm.

a b

Page 154: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

154

Pore radius (µm)

0.0010.010.11101001000

Po

rosity

(%)

0

10

20

30

40

70%

50%

a

Pore radius (µm)

0.0010.010.11101001000

70%

50%

b

Figure 5.26 - Mercury porosimetry of samples producing with 50% and 70% particle loading. (a)

Porosity function of pore radius and (b) pore size distribution.

Mechanical three point bending test were also carried out as shown in Figure 5.27. The results from

the mechanical test correlate well with the porosimetry results. For instance, the samples that show

lower porosity and smaller pores gave higher mechanical strength and could withstand higher

strains. In addition, a higher solids loading allows more particle to particle contact, increasing the

rate of densification during sintering process. By comparison the hollow fibres with lower SS

particle loading resulted in both a lower flexural strain and stress due to the high porosity and large

pores, including the large cavities observed in the SEM images in Figure 5.25.

a

Particle Loading (%)

45 50 55 60 65 70 75

Ma

xim

um

Fle

xura

l Str

ess (

MP

a)

0

200

400

600

800

1000

1200

a

b

Particle Loading (%)

45 50 55 60 65 70 75

Ma

xim

um

Fle

xura

l Str

ain

(m

m/m

m)

0.001

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

b

Figure 5.27 - Three point bending test result for sintered hollow fibres containing 50% and 70%

solid loads. (a) Maximum Flexural strength (b) Maximum flexural strain.

Page 155: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

155

5.11. DISCUSSION OF MORPHOLOGY EFFECTS

A major observation from the results described above is that the morphology of the green fibre

influences the morphology of the sintered hollow fibre. In other words, the finger-like pores and

macrovoids in the green fibres are also found in the sintered fibres. Similarly, sponge-like regions in

the green fibres remain after the sintering process. In order to provide a measure of pore size, the

mercury porosimetry (Figure 5.2, Figure 5.6, Figure 5.10, Figure 5.14, Figure 5.17, Figure 5.20,

Figure 5.23 and Figure 5.26) clearly shows that there are three distinct pore size regions. The first

region is for pore radius below 0.5µm which are associated with the sponge-like region and in the

case of pores <0.01μm, the surface roughness of the stainless steel particles. The next region is for

pore radius between 0.5 and 10 µm, which is still associated with the sponge like region. The third

region is associated with the larger macropores and finger-like structures as also observed in Figure

5.9 (i.e. inner shell). Therefore, in terms of the discussion in this section, the terms small (r 0.5

µm), medium (10 r 0.5 µm), and large (r > 10 µm) pores will be used to elucidate the

morphological arguments.

Regardless of the morphology of the green fibre, sintering temperature has a major effect in the

final morphology of the sintered fibre. This effect can be clearly seen in the micrographs of Figure

5.1 and porosity in Figure 5.2. These results suggest that increasing the temperature, reduces the

porosity and therefore forms denser hollow fibre structures. In addition, by increasing the dwell

time for the same temperature, also reduces the porosity (see Figure 5.10) but the impact is

significantly smaller than for temperature. Hence, total thermal energy input plays a major role in

the final morphology. To explain these effects, a morphological sintering mechanism is

schematically displayed in Figure 5.28. Initially in the green fibre, the SS particles are in contact

with the binder or with each other. In principle, the green fibre formation shows that porosity is not

homogeneous, and that SS particles are not evenly dispersed in the fibre structure (Figure 5.28a).

During the sintering process, the binder decomposes and leaves the matrix as evidenced by the mass

loss in the TGA results (Figure 5.22). The fibres shrink accordingly (Figure 5.13) although the

overall shape is preserved. As there is no mass loss associated with the SS particles, both TGA and

geometrical shrinkage results strongly suggest that the SS particles are getting closer to each other

as depicted in Figure 5.28b. This will lead to (i) particle to particle contact, or (ii) particles

separated by pores or any residual carbon.

As this process continues, SS particles form necks. However, necks can be formed only when

particles are contacting each other. This is clearly seen in Figure 5.1a at 950°C. As the temperature

Page 156: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

156

increases to 1000°C, there is densification (see Figure 5.1b). However, this densification does not

follow the pattern for the densification if pure SS particles without binders were used. For instance,

pure SS particles would form necks to all adjacent particles [52, 53]. In the case of the hollow

fibres, there is a formation of small strips of 3-4 SS particles (see Figure 5.1a and Figure 5.1b). This

point is schematically shown in Figure 5.28c, and attributed to the segregation effect of the binder

in isolating several particles together. Hence, only those SS particles that are in close contact with

each other form necks. As sintering temperatures increase to 1050 and 1100°C (i.e. with the

addition of additional thermal energy) these agglomerated particle strips start coalescing in a

random fashion as observed in Figure 5.1c and Figure 5.1d. By this stage the residual carbon should

have been fully decomposed; however, the TGA results (Figure 5.22) show that not all carbon was

removed at these high temperatures. Further chemical analysis in Table 5.1 clearly indicates that the

SS particles underwent a reaction process with the amount of carbon on the particles (and within the

steel matrix) increased.

Figure 5.28 - Evolution of the stainless steel hollow fibre during sintering process.

The morphology of the sintered fibres suggests that the mass transfer influences the final particle to

particle contact, coalescence, or pore formation. The driving force behind the sintering process is

the minimisation of surface energy. In other words, the reduction of the surface area and

consequently the reduction of the surface free energy of the system [54]. The stage of densification

can be directly evaluated by measuring the size of the necks between particles. However, it is

important to remember that there are several mass diffusion paths from the particle to the neck

Page 157: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

157

formation, and not all of them lead to densification [43]. For instance, Figure 5.1 shows the

different stages of densification for all samples. Sintering can start at temperatures as low as ¼ of

the melting temperature (according to the manufacturer for 316L SS the melting temperature is in

the range between 1390 and 1430°C), though at lower temperatures the mass transfer process is

exceedingly slow. Commonly sintering processes at temperatures below the melting point occur in a

solid state, however as temperature rise to values near the melting point a transient liquid phase

sintering can occur [43].

Figure 5.1a shows the early stages of sintering, as particles are linked together by small necks. At

this stage the particles are clearly defined. This is a characteristic of an initial stage of sintering

defined by Ashby [54, 55]. The result of the process is a highly porous solid in line with the results

in Figure 5.2a. As the temperature is raised the densification is increased as seen in Figure 5.1b,

where the increase in sintering temperature led to larger necks. The same effect is observed for

subsequent increases in temperature to 1050 and 1100°C, in Figure 5.1c and d. However, Figure

5.1c start showing some characteristics of liquid phase sintering, attributed to a transient liquid

phase sintering [23]. A schematic displayed in Figure 5.29 clearly shows the sintering temperature

effect. As the temperature increases, the particles come together leading to the loss in porosity.

Hence, the energy supplied tends to accelerate neck formation, growth and particle coalescence as

the temperature increases.

Figure 5.29 – Schematic effect of sintering temperature

Page 158: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

158

There are several equations that are used to explain the densification process, however one of the

most common approaches used to describe the densification process [56] follows the general

equation:

(5.1)

where x is the neck radius, a is the particle radius, t is time, m and n are numerical exponent

dependent on the mechanism of sintering and H is a function that contains geometrical and material

parameters of the powder, also dependent on the sintering mechanism. The values of the parameters

dependent on the sintering mechanism are displayed in the Chapter 2, Table 2.3, which considers

that sintering is a heat activated process, thus affected by temperature. In turn, diffusivity and

viscosity are also affected by temperature. Let us consider the mass transfer calculations based on

the parameters of Table 2.3 for m, n and H to be used in Eq. 2 where results are displayed in Figure

5.30, together with the neck radius as ascertained from the SEM images in Figure 5.1. These results

clearly indicate that the mass transfer mechanisms leading to neck formation and sintering are

predominantly controlled by (i) lattice diffusion, (ii) grain boundary diffusion, and (iii) surface

diffusion.

At 950°C, Figure 5.1 shows the formation of small necks while the shape of the original particles

was effectively maintained. These results correlate well with a lattice diffusion and grain boundary

diffusion as observed in Figure 5.30. So the diffusion of atoms towards the neck occurs mainly at

the inter-particle contact point or surface. As the temperature increases to 1000°C, the necks are

almost the same size as the precursor SS particles, and form small strips of 3-4 particles joined

together. However, there are some cases of strips coalescing and forming bigger agglomerates.

Hence, Figure 5.30 shows that the variation of particle radius for 1000°C is broadening and surface

diffusion starts occurring. As the temperature increases further to 1050°C, the strips cluster together

in various degrees of coalescence leading to a distribution of neck radii. Indeed, in some parts

neither individual particles nor strips are distinguishable. At this intermediate stage, the viscosity of

the SS decreases, becoming slightly more fluidic and surface diffusion becomes more prominent.

Further increases in temperature led to a significant densification where particles disappear and

necks are no longer discernible. This final stage is controlled by surface diffusion.

Page 159: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

159

Sintering Temperature (°C)

900 950 1000 1050 1100 1150

Ne

ck r

ad

ius (

µm

)

0

1

2

3

4

5

6

Surface dif fusion

Lattice dif fusion from surface

Grain boundary dif fusion

Lattice dif fusion from Grain bounda

Neck radius Measured from SEM

Figure 5.30 – Comparison of neck radius measured with prediction of Eq. 5.2 as a function of

sintering temperature (particle size 6µm).

In principle, Eq. 5.2 is applicable for all particles in close contact to each other. However, the SEM

images show that for the hollow fibres this is not the case. In fact, the SEM images shows that the

structure of the sintered fibre follows to a large degree the morphology of the green fibre, with large

and small pore sizes, though overall the sintering process leads to the densification of the hollow

fibre. The results in Figure 5.30 are well below the predicted neck size associated with the lattice

diffusion. These results are not suggesting that lattice diffusion is dominant, but this discrepancy is

related to the inter-particle space as generally observed in all SEM images. Further, as temperature

increases to 1050 and 1100°C, the diffusion coefficient goes over the upper boundary for surface

diffusion. This is the result of increased fluidity, faster surface diffusion and coalescence of the SS

strips.

The diffusion coefficients of the average neck sizes in Figure 5.30 were plotted against the porosity

of the samples as shown in Figure 5.31, to highlight the effect of diffusion. First, there is a

correlation as the diffusion coefficient increases the hollow fibre densifies particularly at the small

pore radius (0.5m). This is clearly observed for the average diffusion coefficients that porosity

decreases from 45 % to 8 % as the diffusion shifts from 10-16 to 10-10 m2 s-1, accompanied by an

increase in temperature from 950 to 1100°C, respectively. As there is a major reduction in the small

pore sizes, these results suggest that the sponge-like structure is densifying at a faster rate, which is

mainly dominated by surface diffusion. However, surface diffusion becomes dominant only around

Page 160: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

160

1050°C. Similar trends can also be observed for the total porosity as displayed in Figure 5.31b.

These values are the total pore volume divided by the volume of the sample. However, there is a

difference in the spread of the data at 1000 and 1050°C as compared with Figure 5.31a. This

difference is attributed to the large pore sizes (i.e. finger-like pores). As these pores are too large,

the inter-particle distances are too wide thus requiring major pore filling for the particles to come

close together. However, the mass transfer mechanism is controlled by surface diffusion and

particles must be linked to this to occur. Hence, surface diffusion is the limiting transport

phenomena towards the densification of large pores.

ba

950°C

1050°C

1000°C

1100°C

950°C

1000°C

1050°C

1100°C

ba

Diffusion coefficient (m2 s

-1)

1e

-17

1e

-16

1e

-15

1e

-14

1e

-13

1e

-12

1e

-11

1e

-10

1e

-9

Pore

s u

nder

0.5

µm

(%

vol)

0

10

20

30

40

50

950°C

1050°C

1000°C

1100°C

Diffusion Coefficient (m2 s

-1)

1e

-17

1e

-16

1e

-15

1e

-14

1e

-13

1e

-12

1e

-11

1e

-10

1e

-9

Tota

l poro

sity (

%vo

l)

16

18

20

22

24

26

28

950°C

1000°C

1050°C

1100°C

a ba b

Figure 5.31 – Average diffusion coefficient for neck radius showed in Figure 5.30 in function of

small pores (a) and total porosity (b)

The effect of solid loading on the morphology of the stainless steel hollow fibre is displayed in

Figure 5.25. Initially it is observable that the hollow fibre produced with lower amount of stainless

steel particles (Figure 5.25a) show different aspects than the one produced with higher loading

(Figure 5.25b). A major noticeable difference is that the outer diameter and the inner diameter of

the hollow fibre are not concentric. The outer shell has small eccentricities and inner shell no longer

presents a circular cross section. Contrary to this, the hollow fibre with a higher SS particle load

presents perfect concentric surfaces, and the cross section is circular for both outer and inner shells.

The departure from the symmetric geometrical shape observed for the green hollow fibre with lower

solids loading suggests that the large voids are consistent with morphological features generated

during sintering.

Page 161: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

161

Two factors are important here which may cause the observed deformation. First, by reducing the

SS particle loading, the loading of polymer increases. In principle, this means that the amount of

polymer between particles increases and likewise the amount of polymer around macrovoids (i.e.

finger-like pores). Second, during the sintering process, the polymer degradation may cause an in-

situ mechanical collapse in the region around the macrovoids, were SS particles released from the

decomposing binder can move freely under gravity. This can be clearly observed in Figure 5.25b as

the bottom of the hollow fibre becomes flat instead of round. This movement will gradually cease

when particles start to contact one another and start forming necks, thus once again limiting their

mobility. This deformation does not occur for hollow fibres with a high SS particle loading. When

the particle loading is large enough the viscoelastic effect of the polymer is considerably hindered

by the inter-particle contacts, thus the viscous deformation is obstructed, allowing the hollow fibre

to retain the morphology produced during phase inversion [51, 57].

The effect that particle sizes have on the sintering process is also evident in the Eq.2. The

densification rate (x/a) is inversely proportional to the particle size. This means that samples

produced using smaller particles will densify faster because the total surface area available to

initiate the sintering is larger for smaller particles. This difference in surface area can be seen also

as difference in surface curvature. Herring [25] proposed scaling laws relating the effect that

particle size has over densification. Those scaling laws express that when powders with similar

shapes but different sizes are sintered under the same conditions, and under the same sintering

mechanism, the time required to achieve the same densification is proportional to the ratio of the

particle difference elevated to a certain exponent that varies from 1 to 4 depending on the sintering

mechanism as shown in Eqs. 5.3 and 5.4 [25]:

(5.3)

(5.4)

Here α is a coefficient that depends on the sintering mechanism, being 3 for lattice diffusion, 4 for

grain boundary diffusion, 4 for surface diffusion and 1 for viscous flow [43].

Figure 5.14a shows that the samples produced with particles of 6µm are much denser than those

produced with larger particles of 45µm. In fact, shrinkage is one way to measure densification.

Figure 5.32 shows how density changes in a pack of particles as function of the sintering time and

particle size, considering that lattice diffusion from the surface is the dominant mechanism. In this

case the reduction of wall thickness of the hollow fibres was used to quantify the increase in

Page 162: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

162

density. Wall thickness was measured for samples sintered at 1050°C, with a dwelling time of one

hour (total sintering time of 6h) and using Eq 5.3 and 5.4, it was possible to plot wall thickness over

time as function of particle size. It is interesting to notice that this data can be interpreted in two

ways. Firstly, if the wall thickness is fixed then it is possible to see that there is a three order of

magnitude increment in the time needed for the largest particles to reach the same wall thickness of

smaller particles, for the same sintering temperature. Secondly, if the sintering time is fixed there is

a twofold reduction in the hollow fibre wall thickness if the particles used change from 45µm to 6

µm (7.5-fold reduction). Figure 5.32 suggests that the influence of particle size is much stronger

than the influence of dwelling time in determining density, especially in the initial and intermediate

stages of sintering.

Sintering time (h)

0.01 0.1 1 10 100 1000

Ho

llow

fib

re w

all

thic

kne

ss (

µm

)

200

250

300

350

400

450

500

6µm

10µm

16µm

45µm

Figure 5.32 – Effect of particle size and sintering time in the wall thickness of the hollow fibres as

predicted by Eq. 5.4.

5.12. DISCUSSION OF MECHANICAL EFFECTS

The results shown in sections 5.3 to 5.10 strongly suggest that the morphology affects the

mechanical properties of the sintered hollow fibres. Therefore, this section focuses on the analysis

of the mechanical properties as a function of the morphology of the sintered fibres. In this analysis,

the important parameters studied in this thesis are globalised, where mid-points for both mechanical

properties and morphologies are plotted in a single figure. The first global plot is shown in Figure

5.33, which plots the flexural stress versus the total porosity (a) and porosity under 0.5m (b).

Page 163: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

163

There are four quadrants based on averaged values of flexural stress of 604 MPa and porosity for

small pores of 32%. In principle, the optimised SS hollow fibres should have high mechanical

strength and high porosity which is the top-right quadrant. These two properties allow the hollow

fibre to have higher fluxes and withstand high pressure operating conditions. On the other side of

the spectrum, the bottom-left quadrant, the hollow fibres have low mechanical strength, low

porosity and would make poor membranes and membrane supports.

a

Total porosity (%vol)

0 10 20 30 40 50

Maxim

um

fle

xura

l str

ess (

MP

a)

0

200

400

600

800

1000

1200

Particle Size

Sintering Temperature

Dw elling Time

Particle loading

Atmosphere

1:3 Polymer /Solvent

1:4 Polymer / Solvent

Polymer

b

Porosity under 0.5µm radius (%)

0 20 40 60 80

Maxim

um

fle

xura

l str

ess (

MP

a)

0

200

400

600

800

1000

1200

Particle Size

Sintering Temperature

Dw elling Time

Particle loading

Atmosphere

1:3 Polymer /Solvent

1:4 Polymer / Solvent

Polymer

Figure 5.33 – Global maximal flexural stress of the hollow fibres produced versus total porosity (a)

and small porosity (under 0.5µm)(b) for all parameters studied.

The global graph in Figure 5.33 was further broken down for every condition as shown in Figure

5.34.

Page 164: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

164

Total porosity (%vol)

0 10 20 30 40 50

0

200

400

600

800

1000

1200

1:3 Polymer /Solvent

1:4 Polymer / Solvent

Polymer

0% PVP

0% PVPPEI

0.5% PVP

1.5% PVP

1% PVP PESf

0.5% PVP

1.5% PVP

1% PVP

c

Total porosity (%vol)

0 10 20 30 40 50

Maxim

um

fle

xura

l str

ess (

MP

a)

0

200

400

600

800

1000

1200

Particle Size

Sintering Temperature

6µm10µm

1100°C

1050°C

1000°C

16µm

950°C

45µm

a

Total porosity (%vol)

0 10 20 30 40 50

0

200

400

600

800

1000

1200

Dwelling time

Particle loading

Atmosphere70%

Argon

Nitrogen

6µm / 4H

6µm / 1H

16µm / 1H

16µm / 4H

50%45µm / 4H

45µm / 1H

b

Figure 5.34 – Maximum flexural stress as function of total porosity and porosity of pores under

0.5µm for: particle size and sintering temperature (a, d); dwelling time, particle loading and

sintering atmosphere (b, e); polymers(c, f)

Figure 5.34 shows a strong relationship between the amount of pores and the resistance to breakage.

It is noteworthy that Figure 5.34 d, e and f show that samples that have an increased quantity of

small pores, also have a mechanical resistance above the average. For instance Figure 5.34a shows

large volume of pores for samples made of 16µm and 45µm, but Figure 5.34d show that very little

of this porosity is associated with small pores < 0.5µm. In most cases a large porosity fraction

associated with small pores (< 0.5µm) is correlated with high densification of the sponge-like

region. In this region, due to its proximity, particles are able to form multiple connections (necks)

that allow an easy and fast densification process. This neck formation and densification process

ultimately allows hollow fibre to withstand higher mechanical loads. In the case of Figure 5.34d,

samples produced with small particles of 6µm and 10µm present a combination of large porosity

and high resistance to stress due to a well-developed network of necks linking the particles within

the sponge-like region. Samples produced with larger particles present less mechanical resistance

and fewer small pores. This combination is consistent with lower densification that translates into

smaller inter-particular necks, which can withstand lower mechanical loads.

A similar trend can be observed when analysing the sintering temperature at 950°C and 1000°C,

where low densification is responsible for lack of mechanical strength. Hollow fibres sintered at

1050°C show an adequate combination of porosity and mechanical strength following sufficient

Page 165: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

165

densification. Samples sintered at 1100°C show high mechanical resistance, however the decrease

in porosity for small pores denotes an advanced stage of sintering. At this stage small pores

coalesce and form larger pores, which is undesirable for membrane applications.

Figure 5.34e also show samples in the top right quadrant, samples produced using a particle loading

of 70% allows sufficient densification to withstand higher mechanical loads compared to samples

produced using only 50% particle loading. The amount of steel particles used to produce the hollow

fibres is of vital importance for the final morphology and dimension control of the hollow fibre. If

too few particles are used, the hollow fibre will not retain its shape during sintering [51]. Samples

produced with 50% of steel loading show large porosity, large pore sizes but low mechanical

strength. The relatively small amount of particles mean that they are not in close contact initially

and sintering only takes place after the binder has been removed. During debinding the hollow fibre

shrinks allowing the steel particles to contact each other initiating the sintering process, when the

particle loading is increased the time that it takes to initiate sintering is faster, which allows for a

longer effective sintering time. In turn this gives a tougher hollow fibre, with smaller pores and

higher degree of densification.

Figure 5.34e also shows the effect of sintering atmosphere and it is noteworthy that none of the

models for densification take in account the influence that the sintering atmosphere has on

densification. The solubility of the gas was reported to have an effect on the density especially in

the final stages of sintering [58]. Argon is neutral and is considered one of the best atmospheres for

porous SS 316L constructs [59]. Nitrogen on other hand, has been proven to delay densification

[35] which is important in reference to sintering of porous bodies. This agrees with the results

presented in Figure 5.6, where samples prepared with nitrogen show larger porosity, which is

particularly associated with larger pores. Li et al. [35] explains that this is due to the fact that

chromium oxide and chromium nitride are liable to form, when sintering under nitrogen, which will

hinder the sintering densification process. Figure 5.8 agrees, showing a higher mechanical strength

of the hollow fibres sintered in argon, compared to those sintered in nitrogen at 950°C. This can be

explained by the fact that samples sintered under argon require less energy to densify because less

energy will be expended on breaking the chromium oxide and chromium nitride surface layers.

However, when the temperature increases to 1050°C the mechanical strength is not significantly

different for either sintering atmosphere. Now the samples made under argon start to reduce in

porosity as can be seen in Figure 5.6b where the small pores have begun to coalesce, actually

increasing the pore size. Whereas samples sintered under nitrogen continue to experience neck

growth which correspondingly increases the flexural strength. Above this temperature the larger

Page 166: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

166

coalesced pores begin to close and the sintering gas can remain trapped inside. For argon this will

hinder the densification, because the pressure in the pore will make the further shrinkage of closed

pores difficult. However this does not happens in samples sintered under nitrogen, because nitrogen

atoms can dissolve into the matrix of the SS and the pores can continue to close. Therefore the final

stage of densification is faster under nitrogen than argon [42]. Whilst this phenomenon is important

for the production of porous and dense bodies, it is unlikely that this stage of densification was

reached in this thesis.

In the Figure 5.34f, the influence of the morphology promoted by use of polymers and viscosity

modifiers is shown. Samples produced using PEI, and little to no PVP appear in the top right

square. These samples were able to form sponge-like regions which covered a greater cross-

sectional area when compared to the other polymers in the study. In turn, the sponge-like structure

was able to densify faster producing a stronger hollow fibre. On the other hand, samples produced

with PESf and / or with large amounts of PVP promoted more finger-like macrovoids during phase

inversion. The presence of high quantities of macrovoids not only weaken the hollow fibre making

it more prone to rupture, but also shift the position of the sponge like region from the outer layer of

the hollow fibre towards the centre of the hollow fibre wall. This will change the effective diameter

of the layer that withstands the loads reducing its inertial momentum, thus increasing the stresses

within the tubular membrane. This can be seen in Figure 5.35 where samples made from 6µm

particles and sintered for one hour show high mechanical resistance and high porosity (with a large

fraction of small pores), whereas samples made using 1.5% of PVP show low mechanical resistance

but high total porosity (associated with large pores). In both cases the sponge-like region is

highlighted, and attention should be given to the position of this sponge-like region.

The fact that the hollow fibres in this Chapter show vastly different mechanical properties yet are

made from what is essentially the same stainless steel (316L) warrants further investigation. The

mechanical strength of the material, here measured as bending strength, is related to the porosity of

the hollow fibre and to the size of the necks that form between SS particles during sintering [60]. It

is unsurprising therefore to observe that less porous samples had a higher bending strength than

more porous hollow fibres. Indeed, the porosity and strength data appear to exist in an inverse

relationship, for instance the porosity of the PESf+SS sample was almost double that of the PEI+SS

sample whilst it’s bending strength was almost half. However, the actual relationship is more

complex. Previous studies have shown that the strength of a porous metal sample can be predicted

according to its relative density using the relationship in Equation 5.5 [61]:

Page 167: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

167

(5.5)

Where σpl is the yield stress of the porous solid, σys is the yield stress of the original material, ρ is

the density of the porous material, ρs is the density of the original solid material and c is a constant

that depends on the material being investigated. In the case of porous metals, c has been found to be

0.3 [62].

a

b

Figure 5.35 – Position of the sponge-like region for samples made from 6µm SS particles and

sintered for 1 hour (a) and made from 6µm SS particles and 1.5% PVP and sintered for 1 hour (b)

There are relatively few results available in the literature to compare these values of mechanical

stress, however Table 5.4 show values obtained by other authors.

Table 5.4 – Comparison of mechanical strength with literature values

Page 168: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

168

Author Material Stress Range

[MPa]

Sintering

temperature range

[°C]

Luiten-Olieman et al. [28] SS + PESf 450 – 650 1050 – 1100

Lee et al. [63] Nickel + PolySulphone 75 – 160 1100 – 1300

Liu et al. [30] Al2O3 + PESf 25 – 250 1300 – 1600

This work SS + PEI 600 – 1120 1000 – 1100

Table 5.4 shows that the hollow fibres contained in the upper right quadrant of Figure 5.33 and

Figure 5.33 were able to stand higher values of stress than those produced in other available works

in literature. The fact that SS has an inherently higher mechanical resistance than either nickel or

alumina hinders any attempt to further correlate different morphologies with different mechanical

strengths. Importantly, when hollow fibres made of SS are compared the values obtained in this

work are higher. This can be explained by the fact that Luiten-Olieman et al. [28] prepared hollow

fibres using PESf, resulting in a morphology where the sponge-like region is situated in the middle

region of the wall thickness, compared to the one produced in this work using PEI that allowed a

sponge like region near the outer surface of the hollow fibre for the same spinning and sintering

conditions.

The second global analysis was conducted on the flexural strain at break values and was plotted in

Figure 5.36, which also shows the how the total porosity (a) and the porosity for pores under 0.5µm

(b) affect the strain. Similar to the previous global plot of Figure 5.33, the Figure 5.36a and b are

divided into 4 quadrants. For Figure 5.36a the average values used as boundaries are 27% for the

total porosity and 0.008mm mm-1 for the strain at break. For Figure 5.36b the boundaries are 32%

of pores with radius <5μm and the strain boundary is again 0.008 mm mm-1.

Page 169: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

169

Total porosity (%v ol)

0 10 20 30 40 50

Maxim

um

fle

xura

l str

ain

(m

m/m

m)

0.000

0.005

0.010

0.015

0.020

0.025

0.030

Particle Size

Sintering Temperature

Dw elling Time

Particle loading

Atmosphere

1:3 Polymer /Solvent

1:4 Polymer / Solvent

Polymer

a

Pore radius under 0.5µm (%pore volume

)

0 20 40 60 80

Ma

xim

um

fle

xura

l str

ain

(m

m/m

m)

0.000

0.005

0.010

0.015

0.020

0.025

0.030

Particle Size

Sintering Temperature

Dwelling Time

Particle loading

Atmosphere

1:3 Polymer /Solvent

1:4 Polymer / Solvent

Polymer

b

Figure 5.36 - Global maximal flexural strain of the hollow fibres produced versus total porosity (a)

and small porosity (under 0.5µm) (b) for all parameters studied.

To better analyse the results in Figure 5.36, the figures were divided into 3 overarching parameter

spaces as shown in Figure 5.37. Figure 5.37 a and d show the effect that sintering temperature and

particles size have on the strain at break of the hollow fibre. Figure 5.37 b and e compare the effect

of dwelling time, particle loading and sintering atmosphere, while Figure 5.37 c and f shows the

effect of the polymer to solvent ratio, PVP addition and two different binders.

In these graphs it is possible to note that, contrary to the stress analysis, the number of points in the

desired quadrant (top right) is far fewer. For Figure 5.37a and d there again is a dependence of

fraction of small pores on the ductility of the hollow fibre, these results agree with literature that

states that strain at break is increased when porosity is decreased [61, 64]. Samples produced with

smaller particles and sintered at 1050°C show the best results here. As shown before the flexural

strength decreases with an increase in particle size, and samples produced with particles of 6µm

resist maximal flexural stresses an order of magnitude larger than those that samples produced with

45µm. The same trend is present for the ductility, however the difference is not as striking, around

two times larger. Again higher sintering temperatures produced hollow fibres that had larger strain

at break values (i.e. were more ductile), however, this comes at a cost of losing porosity associated

with smaller pores.

Page 170: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

170

a b c

Pore radius under 0.5µm [%pore volume]

0 20 40 60 80

Ma

xim

um

fle

xura

l s

tra

in [m

m/m

m]

0

5e-3

1e-2

2e-2

2e-2

3e-2

3e-2

Particle Size

Sintering Temperature

Pores under 0.5µm [%pore volume]

0 20 40 60 80

0

5e-3

1e-2

2e-2

2e-2

3e-2

3e-2

Dw elling Time

Particle loading

Atmosphere

6µm / 4H

6µm / 1H

70%

Nitrogen

16µm / 1H

Argon

16µm / 4H

45µm / 4H

45µm / 1H50%

Pores under 0.5µm [%pore volume]

0 20 40 60 80

0

5e-3

1e-2

2e-2

2e-2

3e-2

3e-2

1:3 Polymer /Solvent

1:4 Polymer / Solvent

Polymer

PEI

0% PVP

0% PVP

0.5 % PVP

0.5% PVP

1% PVP1.5% PVP

PESf

1.5% PVP

1% PVP

a b c

Total porosity [%vol]

0 10 20 30 40 50

0

5e-3

1e-2

2e-2

2e-2

3e-2

3e-2

1:3 Polymer /Solvent

1:4 Polymer / Solvent

Polymer

PEI

PESf

0% PVP

0.5% PVP

1% PVP1.5% PVP

0% PVP

0.5% PVP

1.5% PVP

1% PVP

Total porosity [%vol]

0 10 20 30 40 50

0

5e-3

1e-2

2e-2

2e-2

3e-2

3e-2

Dw elling Time

Particle loading

Atmosphere

Nitrogen

Argon

70%

50%

6µm / 1H6µm / 4H

16µm / 1H

16µm / 4H

45µm / 1H

45µm / 4H

Total porosity (%vol)

0 10 20 30 40 50

Maxim

um

fle

xura

l str

ain

(m

m/m

m)

0

5e-3

1e-2

2e-2

2e-2

3e-2

3e-2

Particle Size

Sintering Temperature

1100°C

1000°C

1050°C

950°C45µm

16µm6µm

10µm

a b c

d ef

Figure 5.37 – Maximum flexural strain as function of porosity of pores under 0.5µm for: particle

size and sintering temperature (a); dwelling time, particle loading and sintering atmosphere (b);

polymers(c)

From Figure 5.37b and e it is clear that none of the analysed samples fall in the upper right

quadrant. Indeed, there is no clear effect of dwelling time on the strain at break values, adding to the

conclusion that the densification has not significantly increased by adding another 3 hours to the

sintering process. Rather the extra energy input is used to close the small pores [21, 24, 43]. On

other hand, particle loading and sintering atmosphere are important to tailor fibre ductility. Samples

sintered under nitrogen demonstrated strain at break values almost 3 times greater than samples

produced with argon. In addition, total porosity and porosity of small pores are higher when

nitrogen is used. Figure 5.37f show several samples within the desirable boundary, with a clear

effect of PVP addition on strain. Indeed, adding PVP to the spinning dope reduced the ductility of

the fibre in all cases. The choice of polymer however does not show significant difference in strain

at break values.

Page 171: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

171

The results shown in Figure 5.37 suggest that there is no overarching effect of densification on the

fibre ductility. Fibres that present higher degrees of densification (with a low fraction of small

pores) appear less ductile and commonly in this case the cause of failure is tearing of the necks [60].

For all the other samples apparently there is a significant effect from the amount of large

macrovoids on the strain at break value. Macrovoids will act as defect within the hollow fibre

helping first to concentrate tensions and later to propagate the cracks [37, 65]. Hadrboletz and

Weiss [66] showed that typically cracks initiate at pores located at or near the surface of the sample

because of the stress intensity is higher than at a pore in the interior.

The schematic in Figure 5.38 shows the effect of applying a bending force in a radial direction on a

hollow fibre. In particular it shows how the hollow fibre will respond, both in terms of ductility and

crack initiation, when the macrovoids are located either at the outer surface Figure 5.38 (left) and /

or the inner surface Figure 5.38 (right). When a force is applied radially towards the centre of the

hollow fibre a complex series of tension and compression forces are translated through the

structure. For example tension occurs at the surface of the hollow fibre farthest from the force and

compression at surface where the force is applied. In addition the opposite side of the hollow fibre

wall feels the opposite force, so if the surface closest to the force undergoes compression, the other

side of the hollow fibre (in the case of Figure 5.38(right) this would be the top inner surface)

undergoes compression and vice versa for the other side of the hollow fibre. In the case where

tension also corresponds with macrovoids then crack initiation and migration is promoted. Whereas

compressive forces tend to close the macrovoid and cracks are not initiated. If the macrovoids are

located close to the outer surface of the hollow fibre the tension is concentrated and cracks are more

easily initiated than if the tension and macrovoid correspond to the inner hollow fibre wall.

Figure 5.38 – Schematic on the effect of macrovoids in strain and crack initiation.

Page 172: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

172

5.13. CONCLUSIONS

The sintering process is inherently complex and so the production of a porous body with all the

desired characteristics of a SS membrane or membrane support presents challenging task. In most

instances the morphological structure obtained after sintering mimics the morphology obtained after

the phase inversion process. There are two main morphologies present including the characteristic

finger-like macrovoids and the denser sponge-like region. However, this chapter demonstrates that

by adjusting a variety of parameters it is possible to achieve a morphology and performance close to

the desired one. However, one must be very careful because there are inherent trade-offs and

altering a parameter to increase of some characteristics will frequently lead to a reduction in other

desired parameters.

Temperature affects the mass transport rate from the particles to the necks. Hollow fibres sintered

under higher temperatures (1050 and 1100°C) show larger inter-particle necks. In some cases the

densification was so advanced that no neck or indeed no individual particles were distinguishable.

This morphology was associated with a sharp decrease in porosity (compared to hollow fibres

sintered at a lower temperature) and increase in mechanical strength. Conversely samples sintered at

lower temperature (950°C) show small necks, large bulk porosity, and decreased mechanical

strength.

By using particles of different sizes the densification may be accelerated or retarded for the same

sintering conditions. Particle size has an inverse impact in densification, with smaller particles

promoting faster densification than larger particles. This results from the fact that smaller particles

have more total surface area available for diffusion. However, this can be further impacted by

particle loading in that using less SS in the spinning dope allows the large pores, and ultimately the

entire morphology, to collapse due to a lack of particles to create necks with. This was true even

with small particles as to form necks, particles must be in direct contact; however, at lower particle

loadings this was not achieved.

The effect of using different polymers affected the green morphology as described in Chapter 4.

PESf produced more macrovoids, both at the outer and inner surfaces of the hollow fibre with a thin

sponge-like ring in the middle of the hollow fibre wall. Conversely in PEI the sponge-like region

extended from the centre of the hollow fibre all the way to the outer surface. Furthermore, by

increasing the amount of polymer to solvent ratio in the spinning dope from 1:3 to 1:4, the number

of macrovoids also increased, particularly at the outer surface which, in the case of PEI, shifted the

sponge-like region towards the middle of the hollow fibre wall. Likewise when PVP was used to

Page 173: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

173

modify the viscosity, the sintered hollow fibres preserved the morphology created during the phase

inversion process. Briefly, as the PVP content was increased the number of macrovoids also

increased, leading to an increase in the overall porosity of the hollow fibre.

Mechanical resistance of the hollow fibres was closely related to morphology and densification.

Samples with large porosity showed low mechanical strength, especially if the porosity was related

to large pores. As the fraction of the porosity associated with smaller pores increased so too did the

mechanical strength. Another important morphological feature for mechanical strength is the

location of the sponge-like region within the hollow fibre wall. If this region is near the outer

surface, the hollow fibre was more resilient to larger stresses. If the sponge like region was located

towards the lumen however, for the same densification, the mechanical strength decreased. Higher

densification led to stronger hollow fibres, due to the presence of larger necks that could stand

higher loads.

In the same way the maximum strain that the hollow fibre can achieve before breaking is also

related to morphology and densification. The presence of large macrovoids near the outer surface of

the hollow fibre led to early breakage. The large macrovoids act to concentrate the tension

experienced by the fibre under flexural stress which leads to crack initiation and propagation. When

the macrovoids are near the lumen the effect is not so critical, producing stronger and more flexible

hollow fibres. Densification also has an impact on the strain, however the impact was reduced in

comparison to the impact on strength. Fibres that present higher a higher fraction of porosity for

pores under 0.5µm could withstand higher rates of strain, due to the fact that smaller pores were less

critical for crack propagation and larger necks could strain more before failing.

5.14. REFERENCES

[1] Takaba H, Matsuda E, Nair BN, Nakao SI. Molecular modeling of gas permeation through an

amorphous microporous silica membrane. J Chem Eng Jpn 2002; 35:1312-21.

[2] Smart S, Lin CXC, Ding L, Thambimuthu K, Diniz da Costa JC. Ceramic membranes for gas

processing in coal gasification. Energy Environ Sci 2010; 3:268-78.

[3] Duke M, Rudolph V, Lu GQ, Diniz da Costa JC. Scale-up of molecular sieve silica membranes

for reformate purification. AIChE Journal 2004; 50:2630-4.

Page 174: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

174

[4] Duke MC, Campbell R, Cheng X, Leo A, Diniz da Costa JC. Characterization and pervaporation

study on ethanol separation membranes. Drying Technology 2009; 27:538-41.

[5] Fujii T, Yano T, Nakamura K, Miyawaki O. The sol-gel preparation and characterization of

nanoporous silica membrane with controlled pore size. Journal of Membrane Science 2001;

187:171-80.

[6] Klein LC, Gallagher D. Pore structures of sol-gel silica membranes. Journal of Membrane

Science 1988; 39:213-20.

[7] Gallagher D, Klein LC. Silica membranes by the sol-gel process. Journal of Colloid and

Interface Science 1986; 109:40-5.

[8] Dong J, Lin YS, Kanezashi M, Tang Z. Microporous inorganic membranes for high temperature

hydrogen purification. Journal of Applied Physics 2008; 104.

[9] Estella J, Echeverria JC, Laguna M, Garrido JJ. Silica xerogels of tailored porosity as support

matrix for optical chemical sensors. Simultaneous effect of pH, ethanol:TEOS and water:TEOS

molar ratios, and synthesis temperature on gelation time, and textural and structural properties.

Journal of Non-Crystalline Solids 2007; 353:286-94.

[10] Davis PJ, Deshpande R, Smith DM, Brinker CJ, Assink RA. Pore structure evolution in silica

gel during aging/drying. IV. Varying pore fluid pH. Journal of Non-Crystalline Solids 1994;

167:295-306.

[11] Nair BN, Keizer K, Suematsu H, Suma Y, Kaneko N, Ono S, et al. Synthesis of gas and vapor

molecular sieving silica membranes and analysis of pore size and connectivity. Langmuir 2000;

16:4558-62.

[12] Eriksson M, Klein LC, Lidn E, Lindqvist K. Preparation of nanoporous silica-zirconia layers

by in situ sol-gel method. Materials Science and Technology 2006; 22:611-4.

[13] Nair BN, Elferink JW, Keizer K, Verweij H. Preparation and Structure of Microporous Silica

Membranes. Journal of Sol-Gel Science and Technology 1997; 8:471-5.

[14] Klein LC, Bloxom T, Woodman R. Unsupported alkoxide-derived silica membranes. Colloids

and Surfaces 1992; 63:151-61.

Page 175: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

175

[15] Klein LC, Yu C, Woodman R, Pavlik R. Microporous oxides by the sol-gel process: synthesis

and applications. Catalysis Today 1992; 14:165-73.

[16] Klein LC, Gallo TA. Densification of sol-gel silica: Constant rate heating, isothermal and step

heat treatments. Journal of Non-Crystalline Solids 1990; 121:119-23.

[17] Bocchetta P, Sunseri C, Bottino A, Capannelli G, Chiavarotti G, Piazza S, et al. Asymmetric

alumina membranes electrochemically formed in oxalic acid solution. Journal of Applied

Electrochemistry 2002; 32:977-85.

[18] Kingery WD, Berg M. Study of the Initial Stages of Sintering Solids by Viscous Flow,

Evaporation; Condensation, and SelfDiffusion. Journal of Applied Physics 1955; 26:1205-12.

[19] Johnson DL, Cutler IB. Diffusion Sintering: I, Initial Stage Sintering Models and Their

Application to Shrinkage of Powder Compacts. Journal of the American Ceramic Society 1963;

46:541-5.

[20] Cornwall RG, Messing GL, German RM. Sintering technology. New York: Marcel Dekker;

1996.

[21] Okuyama K. Sintering. Powder Technology Handbook, Third Edition: CRC Press; 2006.

[22] Ishizaki K, Komarneni S, Nanko M. Porous materials: process technology and applications.

Dordrecht, Netherlands: Kluwer Academic; 1998.

[23] Rockland JGR. The determination of the mechanism of sintering. Acta Metallurgica 1967;

15:277-86.

[24] Kuckzynski GC. SELF-DIFFUSION IN SINTERING OF METALLIC PARTICLES AIME

TRANS 1949; 185:9.

[25] Herring C. Effect of Change of Scale on Sintering Phenomena. J Appl Phys 1950; 21:301.

[26] Coble R. Sintering Crystalline Solids. I. Intermediate and Final State Diffusion Models. J Appl

Phys 1961; 32:787.

[27] Johnson DL, Clarke TM. Grain boundary and volume diffusion in the sintering of silver. Acta

Metallurgica 1964; 12:1173-9.

Page 176: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

176

[28] Luiten-Olieman MWJ, Winnubst L, Nijmeijer A, Wessling M, Benes NE. Porous stainless

steel hollow fiber membranes via dry-wet spinning. Journal of Membrane Science 2011; In Press,

Accepted Manuscript.

[29] Liu S, Li K. Preparation TiO2/Al2O3 composite hollow fibre membranes. Journal of

Membrane Science 2003; 218:269-77.

[30] Liu S, Li K, Hughes R. Preparation of porous aluminium oxide (Al2O3) hollow fibre

membranes by a combined phase-inversion and sintering method. Ceramics International 2003;

29:875-81.

[31] Jiang DE, Carter EA. Carbon dissolution and diffusion in ferrite and austenite from first

principles. Physical Review B 2003; 67:214103.

[32] Norton JF, Baxter DJ, Santorelli R, Bregani F. The corrosion of AISI 310 stainless steel

exposed to sulphidizing/oxidizing/carburizing atmospheres at 600°C. Corrosion Science 1993;

35:1085-90.

[33] Rouillard F, Moine G, Tabarant M, Ruiz J. Corrosion of 9Cr Steel in CO2; at Intermediate

Temperature II: Mechanism of Carburization. Oxidation of Metals: 1-14.

[34] Weiss B, Stickler R. Phase instabilities during high temperature exposure of 316 austenitic

stainless steel. Metallurgical and Materials Transactions B 1972; 3:851-66.

[35] Li S, Huang B, Li D, Li Y, Liang S, Zhou H. Influences of sintering atmospheres on

densification process of injection moulded gas atomised 316L stainless steel. Powder Metallurgy

2003; 46:241-5.

[36] Mariappan R, Kumaran S, Rao TS. Effect of sintering atmosphere on structure and properties

of austeno-ferritic stainless steels. Materials Science and Engineering: A 2009; 517:328-33.

[37] Chawla N, Deng X. Microstructure and mechanical behaviour of porous sintered steels.

Materials Science and Engineering A 2005; 390:98-112.

[38] Feller HG, Klinger R, Benecke W. Tribo-enhanced diffusion of nitrogen implanted into steel.

Materials Science and Engineering 1985; 69:173-80.

Page 177: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

177

[39] Santos CAd, et al. Surface modifications and the mechanical properties of carbon steels

implanted with nitrogen. Journal of Physics D: Applied Physics 1984; 17:551.

[40] Manova D, Eichentopf IM, Heinrich S, Mändl S, Richter E, Neumann H, et al. Interplay of

cold working and nitrogen diffusion in austenitic stainless steel. Nuclear Instruments and Methods

in Physics Research Section B: Beam Interactions with Materials and Atoms 2007; 257:442-6.

[41] Manova D, Mändl S, Neumann H, Rauschenbach B. Influence of grain size on nitrogen

diffusivity in austenitic stainless steel. Surface and Coatings Technology 2007; 201:6686-9.

[42] Ji CH, Loh NH, Khor KA, Tor SB. Sintering study of 316L stainless steel metal injection

molding parts using Taguchi method: final density. Materials Science and Engineering A 2001;

311:74-82.

[43] Kang S-JL. Sintering: densification, grain growth, and microstructure. Amsterdam: Elsevier

Butterworth-Heinemann; 2005.

[44] Kingsbury BFK, Li K. A morphological study of ceramic hollow fibre membranes. Journal of

Membrane Science 2009; 328:134-40.

[45] Chuang WY, Young TH, Chiu WY, Lin CY. The effect of polymeric additives on the structure

and permeability of poly(vinyl alcohol) asymmetric membranes. Polymer 2000; 41:5633-41.

[46] Yoo SH, Kim JH, Jho JY, Won J, Kang YS. Influence of the addition of PVP on the

morphology of asymmetric polyimide phase inversion membranes: effect of PVP molecular weight.

Journal of Membrane Science 2004; 236:203-7.

[47] Zhao S, Wang Z, Wei X, Tian X, Wang J, Yang S, et al. Comparison study of the effect of

PVP and PANI nanofibers additives on membrane formation mechanism, structure and

performance. Journal of Membrane Science 2011; 385–386:110-22.

[48] Leo A, Smart S, Liu S, Diniz da Costa JC. High performance perovskite hollow fibres for

oxygen separation. Journal of Membrane Science 2011; 368:64-8.

[49] Li X-G, Shao H-T, Bai H, Huang M-R, Zhang W. High-resolution thermogravimetry of

polyethersulfone chips in four atmospheres. Journal of Applied Polymer Science 2003; 90:3631-7.

Page 178: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

178

[50] Tamhankar SS, Bagajewicz M, Gavalas GR, Sharma PK, Flytzani-Stephanopoulos M. Mixed-

oxide sorbents for high-temperature removal of hydrogen sulfide. Industrial & Engineering

Chemistry Process Design and Development 1986; 25:429-37.

[51] Luiten-Olieman MWJ, Raaijmakers MJT, Winnubst L, Wessling M, Nijmeijer A, Benes NE.

Porous stainless steel hollow fibers with shrinkage-controlled small radial dimensions. 2011.

[52] Ryan G, Pandit A, Apatsidis DP. Fabrication methods of porous metals for use in orthopaedic

applications. Biomaterials 2006; 27:2651-70.

[53] W. Schafbauer FS-K, S. Baumann, W.A. Meulenberg, N.H. Menzler, H.P. Buchkremer and D.

Stöver. Tape Casting as a Multi Purpose Shaping Technology for Different Applications in Energy

Issues. Materials Science Forum 2012; 706-709:1035-40.

[54] Ashby MF. A first report on sintering diagrams. Acta Metallurgica 1974; 22:275-89.

[55] Swinkels FB, Ashby MF. A second report on sintering diagrams. Acta Metallurgica 1981;

29:259-81.

[56] Fang ZZ. Sintering of Advanced Materials - Fundamentals and Processes. In: Rahaman MN,

editor. Kinetics and mechanisms of densification: Woodhead Publishing; 2010. p. 33-64.

[57] Luiten-Olieman MWJ, Raaijmakers MJT, Winnubst L, Bor TC, Wessling M, Nijmeijer A, et

al. Towards a generic method for inorganic porous hollow fibers preparation with shrinkage-

controlled small radial dimensions, applied to Al2O3, Ni, SiC, stainless steel, and YSZ. Journal of

Membrane Science 2012; 407–408:155-63.

[58] Coble RL. Sintering Alumina: Effect of Atmospheres. Journal of the American Ceramic

Society 1962; 45:123-7.

[59] Upadhyaya GS. Sintered metallic and ceramic materials: preparation, properties, and

applications. New York: Wiley; 2000.

[60] Ducheyne P, Aernoudt E, De Meester P. The mechanical behaviour of porous austenitic

stainless steel fibre structures. Journal of Materials Science 1978; 13:2650-8.

Page 179: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

179

[61] Dewidar MM, Khalil KA, Lim JK. Processing and mechanical properties of porous 316L

stainless steel for biomedical applications. Transactions of Nonferrous Metals Society of China

2007; 17:468-73.

[62] Gibson LJ, Ashby MF. Cellular solids : structure and properties. 2nd ed., 1st pbk. ed. with corr.

ed. Cambridge :: Cambridge University Press; 1999.

[63] Lee S-M, Choi I-H, Myung S-W, Park J-y, Kim I-C, Kim W-N, et al. Preparation and

characterization of nickel hollow fiber membrane. Desalination 2008; 233:32-9.

[64] Lindstedt U, Karlsson B. Microstructure and Mechanical Behaviour of Single Pressed and

Vacuum Sintered Gas and Water Atomised 316L Stainless Steel Powders. Powder Metallurgy

1998; 41:261-8.

[65] Polasik SJ, Williams JJ, Chawla N. Fatigue crack initiation and propagation of binder-treated

powder metallurgy steels. Metallurgical and Materials Transactions A 2002; 33:73-81.

[66] Hadrboletz A, Weiss B. Fatigue behaviour of iron based sintered material: a review.

International Materials Reviews 1997; 42:1-44.

Page 180: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

180

CHAPTER 6

PRODUCTION AND CHARACTERISATION OF

COMPOSITE CARBON STAINLESS STEEL, COMPOSITE

CARBON ALUMINA HOLLOW FIBRES AND CARBON

ALUMINA HOLLOW FIBRES.

6.1. ABSTRACT

This chapter studies a novel method to produce composite hollow fibres consisting of carbon and

stainless steel (CSS); carbon and alumina and stainless steel (CASS); and carbon and alumina (CA).

The novelty in this process is the use of a combination of sintering and pyrolysis in the same

thermal treatment. Specifically nitrogen is used as the debinding and sintering atmosphere in order

to induce pyrolysis of the polymeric binder to maintain the carbon phase, resulting in a composite

hollow fibre. The study involves the production of CSS hollow fibres using particles of 6, 10, 16

and 45µm in diameter. CASS hollow fibres were also prepared using SS particles (6µm) and

smaller alumina particles (0.5µm). The alumina to SS particle ratio was varied between 25, 50 and

75vol%, and included a CA (i.e. no stainless steel) control. Systematic analysis using

thermogravimetric analysis (TGA), mercury porosimetry, nitrogen adsorption and scanning electron

microscopy (SEM) was carried out to understand the morphological evolution during the

sintering/pyrolysis process. Additionally the mechanical properties (flexural stress and strain) were

also analysed using a 3-point bending test. A strong correlation was found between particle size,

morphology and mechanical properties. It is postulated in this chapter that superior morphology is

conferred to SS hollow fibres by retaining carbon in the hollow fibre structure. A further test of this

postulation was the addition of small alumina particles. From the morphological point of view, the

results show that there is a strong correlation between small particles and the formation of sponge-

like regions. Smaller SS particles (6m) and alumina particles played a major role the formation of

bi-modal pore size distribution, generally conferring an increase in porosity in the sponge-like

Page 181: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

181

region. As such superior mechanical properties were achieved when the sponge like region

dominated at the expense of the macrovoid region.

Page 182: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

182

6.2. INTRODUCTION

Inorganic membranes are a promising separation technology due to their thermal, chemical and

mechanical stability. Conventionally, inorganic membranes are conventionally composed of a

coarse porous substrate, a series of smoother interlayers and a thin active layer, all of which are

produced in separated stages. The porous substrate is produced first from either ceramic or metallic

particles and sintered into its final form, in a similar way as hollow fibres previously reported in

Chapter 5. Once the porous substrate is finished, an interlayer (or series of interlayers) is deposited

on the surface of the substrate to allow for a smooth deposition of the active layer [1]. These

manufacturing stages always require sintering at high temperatures which significantly increases

production costs and the time taken to produce the final membrane. So it is highly desirable to

reduce the number of stages necessary to produce an inorganic membrane, which should reflect as

important cost savings for the overall membrane fabrication process [2].

The production of hollow fibre membranes is generally straight forward as hollow fibres per se can

be produce very cheaply as green fibres, followed by sintering. Depending on the application, the

hollow fibre can act as both the substrate and active layer, dispensing with the need to add extra

coating layers and additional manufacturing stages. Examples of hollow fibres in this category

include dense ceramic perovskites [3-9] and porous hollow fibres made from silica [10], alumina

[11-17], titania [15] and a combination of both [18]. Metals were reported to produce dense or

porous hollow fibres [19-25]. In Chapter 5, it was shown that porous SS hollow fibres could not

separate gases, as the pore sizes were too large and the pore size distribution too broad. Hence, this

chapter tests the postulation that composite hollow fibres made of carbon and SS, where the

polymeric binder is carbonised instead of combusted, can produce structures capable of gas

separation (tested in Chapter 7). In this way, the key idea is to retain the carbonised binder which

fills in the large pore spaces in the SS hollow fibre matrix.

This postulation stems from the fact that carbon hollow fibre membranes derived from PEI and

PESf precursors have been produced using a similar two-step process of combined phase inversion

and pyrolysis, and have proven to have interesting properties for gas separation [26-28]. In these

works the authors mentioned that the carbon hollow fibres showed very low mechanical strength,

though no values are provided. Hence, this chapter addresses the carbon SS hollow fibres by

combining the SS particles for structure with the gas permeation properties of carbon derived from

Page 183: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

183

the carbonisation of pyrolysis of the polymeric binder as schematically shown in Figure 6.1. This

postulation is also extended to ternary materials mixture of carbon and SS particles plus alumina

particles, which were included for modification of the structure and mechanical properties of the

hollow fibres.

To test this postulation, the hollow fibres are prepared in the same manner as the SS hollow fibres

in Chapter 4 and sintered as in Chapter 5. Briefly, a mixing dope solution is spun through a tube-in-

hole spinneret, followed by coagulation in a water bath. Subsequently, the hollow fibres are sintered

in an inert atmosphere, which is used to carbonise the binder, similar to carbon hollow fibres from

PEI precursors [27]. In this chapter, the morphological features and mechanical properties of the

composite hollow fibres are studied as a function of the concentration of materials, whilst the effect

of different SS particle sizes is also tested.

Figure 6.1 – Example of the process of production of the carbon composite hollow fibre

6.3. CARBON STAINLESS STEEL (CSS) HOLLOW FIBRES

Figure 6.2 shows the sintered CSS hollow fibres produced by mixing SS particles of 6, 10, 16 and

45µm into the spinning dope which contained a 1:3 ratio of PEI to solvent. All sintered CSS hollow

fibres were black in colour, contrary to the metallic colour of the SS hollow fibres in Chapter 5.

This is visual proof that the majority of the polymeric binder was carbonised and remained in the

hollow fibre structure.

Page 184: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

184

Figure 6.2 – Picture of the carbon stainless steel produced of 6µm (a), 10µm (b), 16µm (c) and

45µm (d).

The SEM images in Figure 6.3 show that the geometrical shape of the CSS hollow fibre was

generally maintained, although minor variations can be observed. The morphology was mainly

sponge-like at both the inner and outer surfaces, whilst the middle of the CSS hollow fibre

presented in large round macrovoid pores. Again, these morphological features, for the most part,

mimic the green fibre. The inset images show further morphological details. The large macrovoids

range from 25 to 60 m whilst the small pores are lower than 1m.

Figure 6.3 – Carbon Stainless steel hollow fibre made with particles of 6µm.

Further magnification of the SEM images in Figure 6.4 shows different morphological features

between the CSS hollow fibres prepared with small (6 m) and large (45 m) particles. In the case

a b c

d

Page 185: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

185

of the 6m CSS hollow fibres, there is a honey comb like structure where the SS particles are not

clearly distinguishable. This SEM image suggests that carbon derived from the carbonisation of the

polymeric binder remained both in the SS inter-particle regions, and at the same time covered the

SS particles. This is in line with the postulation proposed in this chapter, and with the schematic

depicted in Figure 6.1. It is also interesting to note that the finger-like macrovoids common in SS

hollow fibres are not observed in CSS hollow fibres, although the large round macrovoids in the

middle of hollow fibre wall were retained (see Figure 5.9 in Chapter 5). In the case of the 45m

CSS hollow fibres, the large SS particles are clearly distinguishable as are the carbonaceous regions

between the particles. However, the formation of necks is not clearly observed, contrary to the

findings in Chapter 5. This is attributed to the inter-particle carbon which provides a barrier for SS

particles to contact each other and form necks.

Figure 6.4 – Magnification of SEM images for hollow fibres made of 6µm (a) and 45µm (b)

particles showing the presence of carbon within the matrix.

The TGA analysis emulating the sintering/pyrolysis process of the CSS hollow fibres is displayed

in Figure 6.5 with interesting results. The starting composition of the green fibre is 89.5 wt% SS

particles and 10.5 wt% polymer. The CSS hollow fibres prepared with the smaller SS particles (6,

10 and 16 m) follow the same trend of mass loss up to ~750°C, mass gain until ~900°C and mass

loss above this temperature. In comparison the 45m CSS hollow fibre had no mass gain between

750 and 900°C. To understand these results more clearly, Table 6.1 lists the mass change calculated

from the TGA data within the appropriate temperature range. Initially, the mass loss is consistent

with the decomposition of PEI which started at ~440°C and finished at ~640°C. The average mass

losses are ~4% for all samples. From there, there is a major divergence as the samples prepared

with the smaller SS particles gain mass, up to a peak value of 2.9 wt%, contrary to the larger 45m

CSS hollow fibre which continued to lose mass across this range.

Page 186: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

186

Temperature (°C)

400 500 600 700 800 900 1000 1100

Ma

ss v

ari

atio

n (

%)

94

95

96

97

98

99

1006µm

10µm

16µm

45µm

Figure 6.5 – TGA analysis of the carbon stainless steel hollow fibres

However, it is important to note that these hollow fibres were prepared with different steels which

may explain the divergence of results. For, instance the smaller SS particles (6, 10 and 16 m)

consisted of the austenitic steel 316L whilst the 45m particles were ferritic steel (410L). Nitrogen

is known to cause nitridation in stainless steels [29] and values of 1% wt of nitrogen are normally

able to be adsorb in austenitic stainless steels, and values up to 3%wt have been reported in some

cases. In this process nitrogen is adsorbed at high temperature on the surface of the SS particles and

has been associated with enhanced corrosion resistance and mechanical properties [30]. Chromium

nitrides are known to form in the surface of the steel, and the formation rate is dependent on the

surface area [31]. Rawers et al. [32] reported that the total nitrogen concentration in the austenitic

steel were higher than those displayed for ferritic steel, an observation that is also supported by the

work of Hunter and Eagar [33] on ferritic stainless steel. Hence, the TGA results are in line with

results in the literature suggesting the uptake or otherwise of nitrogen, depending on the type of SS

used.

Table 6.1 – TGA mass loss of CSS hollow fibres.

SS Particle

Size

(m)

Mass Change at

640°C (wt%)

Peak Mass

Change at 800-

900°C (wt%)

Mass Change from

Peak Value to

1050°C (wt%)

Total Mass

Loss

(wt%)

Mass Remaining

at 1050°C (wt%)

6 -4.3

+0.8 -2.9

-6.3 93.7

10 -3.8

+0.8 -2.0

-5.2 94.8

Page 187: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

187

16 -4

+0.5 -1.9

-5.6 94.4

45 -4.2 -5.8 94.2

As the pyrolysis used nitrogen as the inert gas, these results suggest that the increase in mass that

begins at ~700oC and peaks at ~900oC may be associated with the reaction of nitrogen with the

316L SS particles. Nevertheless, from 900oC to the final TGA temperature of 1050oC, the CSS

hollow fibres prepared with the smaller SS particles followed a much faster mass loss rate than the

45m CSS hollow fibre. These results suggest that the mass gain from the nitrogen reaction is

reversed as nitrogen decoupled from the SS particles, in agreement with the findings of Rawers et

al. [32] who showed that CrN is desorbed within this temperature range. The final total mass loss

for all samples was ~5.8wt%, indicating that the sintered CSS hollow fibres contained in the region

of 95.6wt% SS particles and 4.4wt% carbon, equivalent to 42% retention of the carbon in the

polymeric binder. The experimental variation in mass loss of 5.2 to 6.3wt% are all within 10%

experimental variation which was considered acceptable given the experimental variations inherent

in mixing large masses (in kg) of polymeric binder, solvent and SS particles, and the formation of

non-homogeneous porosity during the phase inversion process.

Figure 6.6 shows the pore radius as function of the porosity (pore volume / volume of the sample)

and the pore size distribution derived from mercury porosimetry for all samples. The 6m CSS

hollow fibres (Figure 6.6a) display a bimodal pore size distribution dominated by larger pores

around 90µm which are related to macrovoids and a secondary porosity peak at 0.04µm which is

related to small pores both within the carbon and the CSS matrix itself and surface roughness of the

SS particles. As the particles size increases to 10m (Figure 6.6b) and 16m (Figure 6.6c), the

bimodal distribution disappears, replaced by a multi-modal distribution. Interestingly, the smaller

pores close and several new peaks appear at 0.1m and higher for the 10m SS particles, and even

higher at 4 and 20m for the 16m SS particles. This suggests a degree of pore coalescence, but

also the larger particles naturally form larger pores due to a decreased packing density. For the

larger 45m particles (Figure 6.6d), the fraction of smaller pores almost disappears. It is replaced

with a minor peak at 1m, although the porosity is dominated by a peak between 20 and 100m.

The mechanical characteristics of the hollow fibre were also studied using a 3 point bending test as

previously utilised in Chapter 5. The maximum flexural stress in Figure 6.7a shows a clear trend

relating to the effect of particle size. For instance, samples made with 45µm particles could

withstand flexural stresses of around 35MPa, whilst the 6µm samples could withstand almost four

Page 188: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

188

times this value (139 MPa). In a similar fashion, the flexural strain at break in Figure 6.7b shows

that the ductility of the CSS hollow fibre prepared with 6m SS particles was greater than the 45m

CSS hollow fibre, with average values of 0.005 and 0.0035 mm.mm-1, respectively. The mechanical

trends are quite clear in that increasing the SS particle size decreased mechanical performance.

Pore radius (µm)

0.0010.010.11101001000

Po

rosity

(%vo

l)

0

10

20

30

40

Po

re s

ize

dis

trib

utio

n (

% o

f p

ore

s)

0

5

10

15

20

25

30

Total porosity [%sample v olume]

Pore distribution [%pores]

a

Pore radius (µm)

0.0010.010.11101001000

Po

rosity

(%v

ol)

0

10

20

30

40

Po

re s

ize

dis

trib

utio

n (

%o

f p

ore

s)

0

5

10

15

20

25

30

Total porosity [%sample v olume]

Pore distribution [%pores]

b

Pore radius (µm)

0.0010.010.11101001000

Po

rosity

(%v

ol)

0

10

20

30

40

Po

re s

ize

dis

trib

utio

n (

%o

f p

ore

s)

0

5

10

15

20

25

30

Total porosity [%sample volume]

Pore distribution [%pores]

c

Pore radius (µm)

0.0010.010.11101001000

Po

rosity

(%v

ol)

0

10

20

30

40

Po

re s

ize

dis

trib

utio

n (

%o

f p

ore

s)

0

5

10

15

20

25

30

Total porosity [%sample v olume]

Pore distribution [%pores]

d

Figure 6.6 – Porosity and pore size distribution of carbon stainless steel hollow fibres produced

with particles of 6µm (a), 10µm (b), 16µm (c) and 45µm (d).

Particle Diameter (µm)

0 10 20 30 40 50

Ma

xim

um

fle

xura

l str

ess (

MP

a)

0

20

40

60

80

100

120

140

160

a

Particle diameter (µm)

0 10 20 30 40 50

Ma

xim

um

fle

xura

l str

ain

(m

m/m

m)

0.0030

0.0035

0.0040

0.0045

0.0050

0.0055

b

Figure 6.7 – Maximum flexural stress (a) and strain (b) that the carbon stainless steel hollow fibres

can stand before breaking as function of particle size.

Page 189: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

189

6.4. CARBON ALUMINA STAINLESS STEEL (CASS) HOLLOW

FIBRES

This section focuses on the preparation and characterisation of ternary composite hollow fibres

containing carbon derived from the pyrolysis of the polymeric binder, alumina and SS particles

hereafter referred to as CASS hollow fibres). The hollow fibres were prepared as per procedures for

the binary material CSS hollow fibres described in section 6.3, using 0.5µm alumina particles and

6µm SS particles. The choice of the 6µm SS particles was based on the superior morphological

features and mechanical properties as shown in section 6.3 above.

The CASS hollow fibre is a further test of the postulation of that filling in the inter-particle space

observed in the primary SS hollow fibres in Chapter 5 is able to produce active membranes in a

single step, rather than just membrane supports. As the alumina particle size is significantly smaller

than that of the SS particles, it is postulated that the superior pore size control is possible. To test

this postulation, CASS hollow fibres were produced by controlling the ratio of alumina to SS

particles (25%, 50% and 75 vol%) added to the spinning dope. The total volume of solids loading

was maintained for all samples in order to, in theory, obtain similar viscosities in the spinning dope

and by extension the same cross sectional morphology in the green fibres. The vol% was calculated

based on the density of SS and alumina particles, and the ratios of vol% and wt% for SS and

alumina particles, and polymeric binder are listed in Table 6.2. The CASS hollow fibres were

sintered following the same process described for CSS hollow fibres in section 6.3. The sintered

CASS hollow fibres are shown in Figure 6.8. Similar to the CSS hollow fibres, the pyrolysis

process allowed for the carbonisation of the polymeric binder, which was reflected as in the black

colour of the final sintered hollow fibres.

Page 190: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

190

Figure 6.8 – Picture of the hollow fibres obtained with 25 vol% (a), 50 vol% (b), 75 vol% (c) and

100 vol% (d) alumina loading sintered at 1050oC.

Table 6.2 – Composition of the green fibre obtained after spinning in vol% (left) and wt% (right)

Polymer

[vol%]

SS

[vol%]

Al2O3

[vol%]

Polymer

[wt%]

SS

[wt%]

Al2O3

[wt%]

CSS 42.5 57.5 0.0 10.5 89.5 0.0

CASS 25% 44 42 14 12.5 75 12.5

CASS 50% 44 28 28 14.3 57.1 28.6

CASS 75% 44 14 42 16.7 33.3 50.0

CA 60 0.0 40 32.7 0.0 67.3

The SEM images of the CASS hollow fibres in Figure 6.9 show that the increase in the content of

alumina and consequent decrease in the amount of SS resulted in the formation different

morphology. When large amounts of alumina 50 to 75 vol% are used, the morphological features of

the CASS membrane tend to approach to that of the CA alumina (i.e. no SS particles). These results

strongly suggest that the morphological features tend to be controlled by the smaller alumina

particles.

a b

c d

Page 191: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

191

Figure 6.9 – Morphology of hollow fibres produced using alumina to SS particle ratios of 25 vol%

(a), 50 vol% (b), 75 vol% (c) and 100 vol% (d) and sintered at 1050°C for 1 hour

The TGA results of the CASS hollow fibres in Figure 6.10 also show similar results to CSS hollow

fibres in Figure 6.5. The TGA displays mass losses of 7.1, 7.4, 8.8 and 18.7% for the samples

prepared with 25, 50, 75 and 100 vol% alumina, respectively. The increase in mass loss with the

alumina content is attributed to the increase of the amount of the polymeric binder (see Table 6.2), a

requirement to keep the viscosity of the spinning dope constant. The total amount of carbon retained

in the hollow fibre was 5.4, 6.9, 7.9 and 14.0 wt% for the hollow fibres prepared with 25, 50, 75

and 100 vol% alumina, respectively. These results show that the amount of carbon retained in the

hollow fibres was equivalent to 45.4 wt% (±1.9%) of the PEI binder which correlates well with the

42wt% carbon retention observed for the CSS membrane.

Temperature (°C)

400 600 800 1000 1200

Ma

ss V

ari

atio

n (%

)

80

85

90

95

100

105

25% Al2O3

50%Al2O3

75%Al2O3

100% Al2O3

Figure 6.10 – TGA results for hollow fibres produced using alumina loadings between 25% and

100% loading in volume.

a

a

b

a

c d

a

Page 192: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

192

Porosity distributions measured using mercury porosimetry are displayed for all samples in Figure

6.11. There is a clear underlying trend of a bi-modal pore size distribution, consisting of large

macrovoids or finger-like pores around 100m and small pores around 0.8m. A second interesting

trend correlates to the significant increase in the intensity of the peak at 0.8m (i.e. fraction of

porosity associated with small pores) as a function of the alumina particle ratio, at the expense of

the large pore peak which reduced in intensity. A secondary trend is the narrowing of the small pore

peak as the alumina particle ratio increased. These results clearly demonstrate that mixing smaller

alumina particles (0.5m) with larger SS particles (6m) confers a degree of control over pore size.

Pore radius (µm)

0.0010.010.11101001000

Po

re d

istr

ibutio

n (

%p

ore

vo

lum

e)

0

10

20

30

40

50

Po

rosity

(%v

olu

me)

0

10

20

30

40

50

60

25% Al2O3 - Pore size distribution

25% Al2O3 - Porosity

a

Pore radius (µm)

0.0010.010.11101001000

Po

re d

istr

ibutio

n (

%p

ore

vo

lum

e)

0

10

20

30

40

50

Po

rosity

(%v

olu

me)

0

10

20

30

40

50

60

50%Al2O3 - Pore size distribution

50% Al2O3 - Porosity

b

Pore radius (µm)

0.0010.010.11101001000

Po

re d

istr

ibutio

n (

%p

ore

vo

lum

e)

0

10

20

30

40

50

Po

rosity

(%v

olu

me)

0

10

20

30

40

50

60

75% Al2O 3 - Pore size distribution

75%Al2O 3 - Porosity

c

Pore radius (µm)

0.0010.010.11101001000

Po

re d

istr

ibutio

n (

%p

ore

vo

lum

e)

0

10

20

30

40

50

Po

rosity

(%v

olu

me)

0

10

20

30

40

50

60

100%Al2O3 - Pore size distribution

100% - Porosity

d

Figure 6.11 – Pore size distribution for samples made using 25 vol% (a), 50 vol% (b), 75 vol% (c)

and 100 vol% (d) alumina.

Page 193: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

193

Relative pressure (p/po)

0.0 0.2 0.4 0.6 0.8 1.0 1.2

Qua

ntity

ad

so

rbe

d (

cm

3g

- 1 @

ST

P)

0

10

20

30

40

50

60

700% Al2O3

25%Al2O3

50%Al2O3

75% Al2O3

100% Al2O3

Figure 6.12 - Nitrogen adsorption results for hollow fibres containing alumina particles varying

from 0 to 100 vol%

In order to further verify the morphological features observed in Figure 6.11, nitrogen adsorption

measurements were carried out on all samples. As the surface area associated with large macrovoids

is small, nitrogen adsorption was not a suitable technique to characterise the SS hollow fibres in

Chapter 5. However, the CSS (0 vol% alumina), CASS and CA hollow fibres form micro (<0.2nm)

and meso (0.2dp1.0nm) pores which can be easily measured by nitrogen adsorption as displayed

in Figure 6.12. There is a clear trend of increasing the volume of nitrogen adsorbed as a function of

increasing alumina content in the hollow fibre from 0, 25, 50, 75 to 100 vol%. Another interesting

observation is that all CASS samples gave type II isotherm with a H4 hysteresis loop, which

indicates a combination of meso and macroporosity. A large increase in nitrogen adsorption was

recorded for the CA hollow fibres containing no SS particles and the isotherm changed shape to a

type I, with a minor H4 hysteresis loop, indicating a combination micro and mesoporosity [34].

The BET surface areas were calculated for all samples based on nitrogen adsorption measurements.

There is a significant increase in surface area as the alumina to SS ratio is increased from 0 to 100

vol% as displayed in Figure 6.13. Without the smaller alumina particles, the CSS hollow fibre (0

vol%), has a very small surface area which is mainly attributed to the large SS particles. Somewhat

surprisingly the carbon structure seems to have almost no impact on the pore structure of CSS

hollow fibre. However, by adding small alumina particles, the increase in surface area is quite

evident peaking at 160 m2 g-1 for the CA hollow fibre.

Page 194: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

194

Alumina loading (%vol

)

0 20 40 60 80 100 120

BE

T s

urf

ace

are

a (

m2g

-1)

0

20

40

60

80

100

120

140

160

180

Figure 6.13 – Surface area as function of the content of alumina.

Figure 6.14 shows the three point bending results to investigate the impact of alumina addition on

the mechanical properties of the hollow fibres. The results for the CSS hollow fibre (0 vol%

alumina) from section 6.3 are also displayed for comparison purposes. The maximum flexural stress

in Figure 6.14a shows no variation between the CSS and CASS 25% hollow fibres bending strength

of ~138 MPa retained. From thereon there is a sevenfold decrease as the samples prepared with 50

and 75 vol% of alumina recorded strengths of 25MPa and 19.6MPa, respectively. By comparison,

CA hollow fibres prepared with 100 vol% alumina showed increased bending strength of 72MPa,

though this value is only half that of the CSS and CASS (20 vol% alumina) hollow fibres. Flexural

strain measurements (Figure 6.14b) showed a similar trend, with a decrease in flexibility as the

amount of alumina increased to 50% which was the most brittle sample by far. Above this ratio,

flexibility was partially restored with further increases in alumina content. Average flexural strain

values of 0.005, 0.0037, 0.0014, 0.0025, and 0.0029 mm.mm-1 were measured for hollow fibres

made with 0, 25, 50, 75 and 100 vol% alumina respectively.

Page 195: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

195

Alumina loading (%vol

)

0 20 40 60 80 100 120

Ma

xim

um

fle

xura

l str

ess (

MP

a)

0

20

40

60

80

100

120

140

160

a

Alumina loading (%vol

)

0 20 40 60 80 100 120

Ma

xim

um

fle

xura

l str

ain

(m

m/m

m)

0.001

0.002

0.003

0.004

0.005

0.006

b

Figure 6.14 – Mechanical properties of hollow fibres (a) flexural stress and (b) flexural strain.

6.5. DISCUSSION

6.5.1. 6.4.1 Morphological effects of pore filling

An initial observation from the SEM image of the CSS hollow fibres (Figure 6.3) is that the

pyrolysed/sintered morphology followed to some extent the morphology of the green fibre produced

by phase inversion as discussed in Chapter 4. However, there are a few observations in the CSS

morphology that differs from the SS morphology in Chapter 5. One of the major differences is that

the finger-like structures in the CSS fibres were generally not present, particularly close to the inner

shell where they were commonly observed for the SS fibres, whilst large round macrovoids

prevailed in the centre of the hollow fibre wall. These results are very interesting and a schematic of

the CSS pyrolysis/sintering is displayed in Figure 6.15 to explain this process.

Page 196: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

196

Figure 6.15 – Schematic of the carbon stainless steel hollow fibre sintering / pyrolysis process.

Contrary to the SS hollow fibres process where the polymeric binder is burnt out during pre-

calcination in air (see Chapter 5), in the case of CSS hollow fibres the entire sintering process is

carried out in an inert atmosphere (nitrogen) where the polymeric binder is pyrolysed and the

resultant carbon is essentially kept within the CSS hollow fibre structure. The TGA results (Figure

6.5 and Table 6.1) clearly show mass loss from ~450 to ~600oC associated with the degradation of

volatile compounds in the PEI binder. Similar carbonisation results were reported for the production

of pure carbon hollow fibres derived from PEI [27, 28], with that temperature range corresponding

to the release of volatiles during pyrolysis, following by shrinkage of the hollow fibre. However,

4.4wt% carbon was kept after the pyrolysis/sintering process at 1050oC, out of 10.5wt% polymeric

binder in the green fibre. These results, coupled with the black colour of the hollow fibres in Figure

6.2, clearly demonstrate that carbon remained within the structure of the CSS hollow fibre.

During the pyrolysis/sintering process up to 1050°C, the CSS hollow fibre external and internal

diameter reduced by an average of 10 and 17% respectively as compared to the initial green fibre.

This led to further densification of the CSS hollow fibre, which was evidenced by a reduction in

wall thickness of 10%. Hence, the carbon remained in the SS inter-particle spaces, particularly for

the sponge-like regions. Importantly though, the large round macrovoids persisted in the centre of

the fibre (see Figure 6.3). The disappearance of the long finger-like structures in the CSS hollow

fibres is elucidated by the schematic in Figure 6.15 – Schematic of the carbon stainless steel hollow

fibre sintering / pyrolysis process.. In the case of the SS hollow fibres, the sintering process caused

the particles to form necks which were able to resist densification and maintain the finger-like

structures. In the CSS fibres however, carbon remains between the SS particles preventing sintering

Page 197: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

197

and favouring collapse of these pores during shrinkage before suitable necks could begin to form.

This is particularly the case because the finger-like pores were predominantly close to the inner

shell, where densification was greatest. In the case of the large round macrovoids in the centre of

the CSS hollow fibre wall, these pores are too large to begin with. So as the hollow fibre shrinks

and necks begin to form, further collapse of pore structures is increasingly constrained by the

comparatively rigid network of sintered particles. Meaning the formation of the sponge-like regions

at the inner and outer surfaces of the hollow fibre actually prevents sufficient collapse of the middle

region and thus the large circular macrovoids are preserved. Hence, the CSS hollow fibres formed a

combination of large round macrovoids and sponge like regions.

Nevertheless, it is interesting to observe that the addition of small alumina particles in the CASS

hollow fibres led to a relative decrease in the presence of finger-like pores. This is also clearly

observed in that the morphology of the CASS hollow fibres approached the morphology of the CA

hollow fibre (i.e. no SS particles) as the alumina to SS ratio increased. Hence, these results strongly

suggest that the smaller alumina particles (~0.5m) were more influential in determining the final

hollow fibre structure than the larger SS particles (~6m). These changes may also be correlated to

the total number of particles per unit volume of dope mixture which increased significantly as a

function of the alumina particles. For instance, by assuming a homogeneous doping mixture, SS

hollow fibres (100 vol% SS) may contain around 8.8x1015 particles per cm3 of doped mixture, and

this number increases to 7.6x1018 particles for CASS hollow fibres (50 vol% alumina) and to 1.53

x1019 particles for the CA hollow fibre (100 vol% alumina). Therefore, the addition of alumina with

smaller particle size provided the conditions for reduction on the overall kinetics of the phase

inversion process, despite the initial fast de-mixing when the spinning dope first contacted the non-

solvent (i.e. water) in the bore liquid during the spinning process, the solvent / non-solvent

exchange rapidly loses energy, due partially by the depletion of the solvent contained in the

mixture, but more importantly due to the large amount of small particles that acts as resistance to

the flow of solvent towards the non-solvent. As a result, a much slower de-mixing in the outer shell

is achieved once the hollow fibre reached the coagulation bath. Thus, the inner shell presented

finger-like structures whilst the outer shell showed only sponge-like structures. This is in line with

the results found for using small SS particles in Chapter 4. It is important to remember here that the

original CSS fibre possessed finger-like pores in the green hollow fibre but these collapsed during

the sintering process as explained above, whereas when alumina is present the smaller particles

already fill the space between the SS particles and prevent collapse of the finger-like regions during

sintering. In this way it appears that finger-like voids are present in the CASS fibres when they are

Page 198: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

198

no in the CSS fibre, but this is a function of the pyrolysis/sintering process and not the hollow fibre

formation during phase inversion.

In order to understand the effect of carbonisation of the pore structure of the hollow fibres using

different particle sizes, Figure 6.16 shows the overall morphological results of the CSS hollow fibre

based on the mercury porosimetry results in Figure 6.6. There is significant variation in terms of

porosity for the CSS hollow fibres prepared with the smaller particles of 6, 10 and 16m. Similar

morphological variations were also observed for the sintered SS particles in chapter 5, and

attributed to the morphological variations as ascertained by CT scanning (see Figure 5.12). In other

words, experimental variation for total porosity ranges from 28% (6m) to 33% (10m) and 36%

(16m). However, a common trend with the SS hollow fibres in chapter 5 is that the CSS hollow

fibres prepared with 45m SS particles resulted in large total porosity which was dominated by

large macrovoids. As the total mass of carbon left in the hollow matrix after pyrolysis and sintering

is only 4.4wt% as per TGA results, this amount is not enough to fill in the SS inter-particle space,

particularly for the larger 45m SS particles which generate much larger inter-particle spaces than

the smaller SS particles (6, 10 or 16m) as a consequence of having a lower packing density. For

this reason, macrovoids tend to be more prevalent in the CSS hollow fibres prepared with larger

particles than with the smaller particles.

Particle size (µm)

6 10 16 45

To

tal p

oro

sity

(%v

ol)

0

10

20

30

40

Total Porosity [%v ol]

Finger-like macrovoids [%v ol]

Sponge-like pores [%v ol]

Figure 6.16 – Overall morphological features of the CSS hollow fibres.

Page 199: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

199

It is noteworthy that the addition of the smaller alumina particles to produce the CASS hollow

fibres resulted in a bi-modal pore size distribution, independently of the ratio of alumina to SS

particles used. In addition, by increasing the amount of alumina particles, the porosity distribution

becomes narrower for the small pores sizes below 1m. Hence, these results show proof of the

postulation put forward on the effect of using small particles to control the pore size of the hollow

fibres. To further understand the effects of alumina in the ternary carbon/alumina/SS structures,

Figure 6.17 shows the overall morphological results of the CASS hollow fibre based on the mercury

porosimetry results in Figure 6.11, including the results for the CSS hollow fibre for comparison

purposes. The macrovoids are allocated as pores over 10µm in radius, and sponge-like structure for

pores under 1µm in radius. There are almost no pores in the 1-10μm region and so it was treated as

a transition region and excluded from the analysis.

Alumina load (%vol

)

0 20 40 60 80 100

To

tal P

oro

sity

(%v

ol)

0

10

20

30

40

50

60

Max Porosity [%v ol]

Finger-like Macrovoids [%v ol]

Sponge-like pores [%v ol]

Figure 6.17 – Morphological features of CASS

Interestingly the total porosity is always higher with alumina addition than without, which is

directly related to the increase of porosity in the sponge-like region. Hence, the utilisation of small

alumina particles combined with controlled carbonisation of the polymeric binder clearly favours

the sponge-like morphology in the final hollow fibre structure. This makes sense as the average

Page 200: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

200

diameter of alumina particles is 12 times smaller than that of SS particles. As a result, more alumina

particles can be packed in the SS inter-particle space, which is coupled with the carbonisation

process to form smaller pores and a sponge-like structure. However, the total porosity and fraction

of the sponge-like structure peaked at an alumina to SS particle ratio of 50 vol% but showed a

lower total porosity and fraction of sponge-like structure for the CSS (0% alumina) and CA (100

vol% alumina) hollow fibres in Figure 6.17. Hence, the postulated pore control is achieved by

balancing by the quantity of the ternary components in the mixture, the effect of particle size and

the carbonisation process.

The morphological evolution of the CASS hollow fibres is schematically represented in Figure

6.17. In the case of CSS hollow fibres (0 vol% alumina), the pore sizes are still relatively large and

are dominated by a significant distribution of large pores around 70 m (Figure 6.6a). These pores

are too large to be measured by nitrogen adsorption where the BET surface areas (Figure 6.13) for

the CSS hollow fibres resulted in very low values of ~3 m2 g-1. These results strongly suggest that

the contribution of carbon towards micro or mesoporosity is very low, and it is likely that carbon

simply coated the large SS particles (6 m) or the interconnecting space between the SS particles as

a dense layer rather than porous carbon. As alumina is added to form CASS hollow fibres, the BET

surface areas significantly increased from ~10 (25 vol% alumina) to ~40 (75 vol% alumina) m2 g-1.

This was attributed to the formation of mesopores evidenced by the lack of adsorption for relative

pressures up to 0.2 in the nitrogen isotherms (Figure 6.12) and the hysteresis observed which

increased as a function of alumina content. Finally, the CA hollow fibres (no SS particles) resulted

in the formation of significant microporosity reflected by an abrupt increase in nitrogen adsorption

for relative pressures below 0.2, and high surface areas 159 m2g-1. This suggests that in the CA

hollow fibres the carbon is microporous and no longer dense.

Page 201: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

201

Figure 6.18 – Effect of particle loading in the nitrogen adsorption.

The structural rearrangement of the pores as schematically displayed in Figure 6.18 represents the

variation of using binary and ternary mixtures on the final micro and mesoporosity of the hollow

fibres. Coupled with the presence of carbon, large SS particles tend to form larger pores and small

surface areas contrary to the small alumina particles which yielded smaller pores and larger surface

areas. It is important to note that there is a distinction between the total porosity measured by

mercury porosimetry and the micro and mesoporosity measured by nitrogen adsorption. Whilst

mercury can measure to total porosity (% of volume) of the hollow fibres for a very wide range of

pore radius from 1nm to over 100m, the porosity obtained by nitrogen adsorption (cm3 g-1) is

associated with pore sizes from 0.2nm to 10nm. In the case of the CA membrane (100 vol%

alumina) the high surface area measured by nitrogen adsorption (Figure 6.13) was mainly attributed

to the higher content of carbon retained in the matrix (see TGA Figure 6.10) which was associated

with the higher initial amount of polymeric binder to maintain a constant viscosity of the spinning

dope.

6.5.2. Mechanical properties of pore filling

Figure 6.19 compares the mechanical properties of the CSS hollow fibres against the SS hollow

fibres from Chapter 5 as a function of the precursor SS particle. A common feature of these results

is the large reduction of both flexural stress and strain for all CSS hollow fibres, independent of the

size of the SS particle. For instance, the flexural stress for the CSS hollow fibres reduced by 770,

Page 202: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

202

799, 418 and 40 MPa as compared with the analogous SS hollow fibres prepared from 6, 10, 16 and

45µm, respectively.

Particle diameter (µm)

0 10 20 30 40 50

Ma

xim

um

fle

xura

l str

ess (

MP

a)

0

200

400

600

800

1000

1200

CSSHF

SSHF

Particle diameter (µm)

0 10 20 30 40 50

Ma

xim

um

fle

xura

l str

ain

(m

m/m

m)

0.002

0.003

0.004

0.005

0.006

0.007

0.008

0.009

0.010

CSSHF

SSHF

Figure 6.19 – Comparison between (a) flexural stress and (b) flexural strain for the CSS and SS

hollow fibres.

These striking differences in mechanical strength are attributed to the characteristic of the

composite material. Firstly, carbon hollow fibres are categorised by their very weak mechanical

strength [26-28]. Secondly, the carbon contained in the space between the SS particles hinders the

densification process [35], evidenced by the porosity of all SS hollow fibres. Third and most

importantly, these results suggest that the carbon coated the SS particles and therefore inhibited

neck formation between the SS particles. As discussed in Chapter 5, neck formation was

fundamental to confer improved mechanical resistance, which is lacking in the CSS hollow fibres.

Nevertheless, there the reduction in mechanical strength with increasing SS particle size is far less

significant for the CSS hollow fibres when compared with the SS hollow fibres. For instance, the

mechanical strength of the CSS hollow fibres reduced by 29% as the particle size increased from 6

to 45, whilst the SS hollow fibres experience a major reduction of 82%. Similar trends are also

observed for flexural strain. These small variations in the mechanical properties of the CSS hollow

fibres strongly suggest that the mechanical strength and ductility are controlled by the composite

matrix and by the morphology. For instance, in Figure 6.16 the sponge like region reduces as the SS

particle size increases, and likewise correlates well with the reduction in the mechanical properties.

Similar findings were discussed for the SS hollow fibres in Chapter 5.

Page 203: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

203

Figure 6.14 displays the mechanical properties of the CASS hollow fibres, including the CSS (0

vol% alumina) and CA (100 vol% alumina) for comparison purposes and three distinct regions can

be identified. The first region contains hollow fibres produced using 0% and 25% alumina and

importantly the maximum flexural stress is almost equal despite the introduction of alumina

particles. In this region the strength is likely associated with the network formed between the steel

particles as schematically represented in Figure 6.20 (left). At these low alumina loading conditions,

it is possible that SS particles formed necks during sintering as the amount of carbon remaining was

relatively low. Under these conditions, carbon and / or alumina particles were not present in

sufficiently large amounts to separate all SS particles. Thus whilst necks were not easily observable

in the SEM images a fraction of the SS particles must have been in close enough contact to form

necks, which later lead to coalescence and some strengthening of the hollow fibre. Similar results

were reported for alumina metal structures where the small amount of alumina was not high enough

to impede the formation of a steel network responsible for the mechanical strength [36].

The second region identified in Figure 6.14 corresponds to alumina / SS ratios ≥ 1. In this region the

mechanical properties are greatly reduced as a network of interconnected stainless steel particles

could not form due to a lack of SS particle contact (schematically depicted in Figure 6.20 (centre)).

Larger volumes of alumina reduced the probability of particle to particle contact between the SS

particles, which in turn hindered neck formation. In this case the alumina formed a loosely network

of particles, which offered some mechanical resistance, but ultimately minimal alumina sintering

occurred due to the relatively low temperatures compared with the melting point of alumina. Most

importantly coalescence between alumina and SS particles was not observed. In fact, sintering

different powders requires more energy to promote neck formation and densification than

homogeneous powders [37, 38]. Prielipp et al. [39] discussed the effect of interconnecting networks

where ceramics provide the probable failure site at the largest flaw, and the metal provides the

fracture toughness. However, it is also interesting to observe a trend where the mechanical strength

in Figure 6.14 is inversely proportional to total porosity (see Figure 6.17). For instance, the CASS

hollow fibres (50 vol% alumina) conferred the maximum total porosity and the minimum

mechanical strength and ductility. Hence, the correlations between morphology and mechanical

properties as discuss in chapter 5 for SS hollow fibres are also applicable for CASS membranes.

Page 204: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

204

Figure 6.20 – Interaction between alumina, SS and carbon within the matrix of the hollow fibres

The third region of strength (Figure 6.20 right) occurs when there is an exclusion of SS particles,

essentially forming a carbon alumina hollow fibre. In this case the formation of a homogeneous

composite structure is observed due to the close inter-particle contact between the smaller alumina

particles (Figure 6.20 right). These conditions favour the formation of sponge-like regions with

lower total porosity and a reduced number of defect sites for crack initiation and migration, thus

improving both mechanical strength and ductility.

Finally, Table 6.3 shows a list of the mechanical strength for several ceramic and metallic hollow

fibres as reported elsewhere for comparison purposes. It is worth noting that even though several

studies have been conducted on the preparation of carbon hollow fibres from polymer pyrolysis, the

mechanical strength is reported only as qualitative property and not in actual values. Higher

mechanical strength for both SS and alumina hollow fibres were reported as compared to this thesis,

although they were sintered at higher temperatures. On the other hand, the optimal sintering

temperature of 1050oC in this work, coupled with the preparation methods developed in this thesis,

resulted in hollow fibres with higher mechanical resistances. For instance, CA hollow fibre reached

75 MPa, over threefold higher than the alumina hollow fibres sintered at 1300oC with a flexural

strength of 21 MPa [18].

Page 205: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

205

Table 6.3 – Comparison of mechanical strength values available in literature.

Author Material Sintering

Temperature Min Strength Max Strength

Liu et al. [18] Al2O3 1300 – 1600 °C 21 MPa 304 MPa

Liu et al. [11] Al2O3 1500 – 1600 °C 45 MPa 340 MPa

Lee et al. [21] Ni 1100 – 1300 °C 75 MPa 160 MPa

Yang et al. [20] Ni / YSZ 1200 – 1400 °C 40 MPa 160 MPa

Yang et al. [20] NiO / YSZ 1200 – 1400 °C 40 MPa 354 MPa

Luiten–Olieman et al.

[25] SS 1050 – 1200 °C 450 MPa 3500 MPa

Michelsen et al. [22] SS 1000 – 1300 °C 47 MPa 333 MPa

SS hollow fibres SS 950 – 1100 °C 303 MPa 820 MPa

CSS hollow fibres SS / C 1050 °C 35 MPa 139 MPa

CASS hollow fibres SS / Al2O3 / C 1050 °C 20 MPa 137 MPa

6.6. CONCLUSION

A combined pyrolysis/sintering process allowed composite hollow fibres to be made from carbon,

alumina and stainless steel. In most cases the ~4.4wt% of carbon out of the 10.5wt% of the original

polymeric binder was retained in the hollow fibre matrix. The morphological features of the CSS

hollow fibres followed those of the green fibres after the sintering/pyrolysis process, although the

presence of finger-like macrovoids in the inner shell disappeared which was attributed pore

collapse, carbon retention and densification. However, the round macrovoids in the middle wall of

the CSS hollow fibres remained as they were too large to be closed by these mechanisms. Using the

small 6m particles resulted in a bi-modal pore size distribution dominated by a sponge-like region

at both inner and outer surfaces of the hollow fibre. Increasing the particle size changed the pore

size distribution until it was no longer bi-modal.

Superior pore size tailoring was achieved by incorporating alumina particles into the spinning dope.

A major morphological feature was the formation of a bi-modal pore size distribution, independent

of the alumina to SS ratio. As the alumina content increased from 0 to 25vol%, there was a

Page 206: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

206

significant formation of finger-like structures in the CASS hollow fibres. These morphological

features became more distinct as the alumina content increased still further until a CA hollow fibre

was formed. Hence, the morphology of the sintered fibre was dominated by the smaller alumina

particles (0.5m) rather than the larger SS particles (6m). In addition, the increase in the amount

of alumina allowed the formation of a closely packed alumina and carbon region in the SS inter-

particle space, which was mainly sponge-like and increased the resultant surface area up to 150 m2

g-1. This was significantly greater than values for CSS hollow fibres which were typically below 3

m2 g-1. However, the increase in surface area was mainly attributed to the high content of polymeric

binder in the spinning dope, which was necessary to keep the viscosity of the spinning dope

constant. Hence, the pyrolysis/sintering process allows for the retained carbon to form micropores

within the alumina inter-particle space or at the interfaces between alumina and carbon. This was in

contrast to the CSS fibres where the retained carbon appeared to be essentially dense.

The mechanical properties of the CSS and CASS hollow fibres were greatly affected by their

composition. Carbon and mixtures of carbon and alumina formed anti-sintering barriers which

hindered the formation of necks between SS particles unless they were already in close contact. As

the probability for the latter was very small, the general trend was for a major reduction in both

flexural stress and strain of the CSS and CASS hollow fibres as compared to the values obtained for

SS hollow fibres in Chapter 5. Nevertheless, the mechanical properties could still be correlated to

the morphology. For instance, the CSS hollow fibres prepared with larger SS particles (45µm)

presented a relatively small sponge-like region, resulting in a poor mechanical strength and

ductility, contrary to the hollow fibres prepared with small SS particles (6µm and 10µm). Hence,

hollow fibres containing a greater sponge-like region delivered superior mechanical properties.

However, it was found that there was a major trade-off between porosity (or densification) and

mechanical properties, particularly for CASS hollow fibres which consistently resulted in bi-modal

pore size distribution. For low alumina loading (25vol%) the mechanical resistance was similar to

the CSS hollow fibre, where the close proximity of SS particles had allowed necks to form. In this

case, the presence of alumina had little effect on both mechanical strength and strain. For

intermediate alumina loadings (50 vol% and 75vol%) both mechanical strength and strain reached

minimum value, contrary to the porosity trend which reached a maximum value. This is attributed

to the fact that bonding between ceramics and metals particles requires higher energy than the

bonding between the same material, which in turn reduced densification evidenced by the high

porosity. For the CA hollow fibre (100 vol% alumina), mechanical strength and ductility was

partially restored towards the CSS values, indicating that sintering had started for these samples but

Page 207: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

207

required higher temperature or longer dwell times to attain values similar to pure alumina hollow

fibres.

6.7. REFERENCES

[1] Burggraaf AJ, Cot L. Fundamentals of Inorganic Membrane Science and Technology.

Burlington: Elsevier Science; 1996.

[2] De Falco M, Marrelli L, Iaquaniello G. Membrane reactors for hydrogen production processes.

New York: Springer; 2011.

[3] Haworth P, Smart S, Glasscock J, Diniz da Costa JC. High performance yttrium-doped BSCF

hollow fibre membranes. Separation and Purification Technology 2012; 94:16-22.

[4] Leo A, Liu S, Diniz da Costa JC. Production of pure oxygen from BSCF hollow fiber

membranes using steam sweep. Separation and Purification Technology 2011; 78:220-7.

[5] Leo A, Smart S, Liu S, Diniz da Costa JC. High performance perovskite hollow fibres for

oxygen separation. Journal of Membrane Science 2011; 368:64-8.

[6] Tan X, Liu N, Meng B, Liu S. Morphology control of the perovskite hollow fibre membranes

for oxygen separation using different bore fluids. Journal of Membrane Science 2011; 378:308-18.

[7] Zhang Y, Li Q, Li H, Cheng Y, Zhang J, Cao X. Sintering-resistant hollow fibers of

LaMgAl11O19 prepared by electrospinning. Journal of Crystal Growth 2008; 310:3884-9.

[8] Liu S, Gavalas GR. Oxygen selective ceramic hollow fiber membranes. Journal of Membrane

Science 2005; 246:103-8.

[9] Wang H, Werth S, Schiestel T, Caro J. Perovskite Hollow-Fiber Membranes for the Production

of Oxygen-Enriched Air. Angewandte Chemie International Edition 2005; 44:6906-9.

[10] Way JD, Roberts DL. Hollow Fiber Inorganic Membranes for Gas Separations. Separation

Science and Technology 1992; 27:29-41.

Page 208: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

208

[11] Liu S, Li K, Hughes R. Preparation of porous aluminium oxide (Al2O3) hollow fibre

membranes by a combined phase-inversion and sintering method. Ceramics International 2003;

29:875-81.

[12] Tan X, Liu S, Li K. Preparation and characterization of inorganic hollow fiber membranes.

Journal of Membrane Science 2001; 188:87-95.

[13] García-García FR, Rahman MA, Kingsbury BFK, Li K. Asymmetric Ceramic Hollow fibres:

New micro-supports for gas-phase catalytic reactions. Applied Catalysis A: General 2010; In Press,

Accepted Manuscript.

[14] Kingsbury BFK, Li K. A morphological study of ceramic hollow fibre membranes. Journal of

Membrane Science 2009; 328:134-40.

[15] Zhang X, Wang DK, Lopez DRS, Diniz da Costa JC. Fabrication of nanostructured TiO2

hollow fiber photo catalytic membrane and application for wastewater treatment. Chemical

Engineering Journal 2014; 236:314-22.

[16] Deng Z, Nicolas CH, Daramola MO, Sublet J, Schiestel T, Burger AJ, et al. Nanocomposite

MFI-alumina hollow fibre membranes prepared via pore-plugging synthesis: Influence of the

porous structure of hollow fibres on the gas/vapour separation performance. Journal of Membrane

Science 2010; 364:1-8.

[17] Nair BKR, Harold MP. Experiments and modeling of transport in composite Pd and Pd/Ag

coated alumina hollow fibers. Journal of Membrane Science 2008; 311:53-67.

[18] Liu S, Li K. Preparation TiO2/Al2O3 composite hollow fibre membranes. Journal of

Membrane Science 2003; 218:269-77.

[19] Meng B, Tan X, Meng X, Qiao S, Liu S. Porous and dense Ni hollow fibre membranes.

Journal of Alloys and Compounds 2009; 470:461-4.

[20] Yang N, Tan X, Ma Z. A phase inversion/sintering process to fabricate nickel/yttria-stabilized

zirconia hollow fibers as the anode support for micro-tubular solid oxide fuel cells. Journal of

Power Sources 2008; 183:14-9.

[21] Lee S-M, Choi I-H, Myung S-W, Park J-y, Kim I-C, Kim W-N, et al. Preparation and

characterization of nickel hollow fiber membrane. Desalination 2008; 233:32-9.

Page 209: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

209

[22] Michielsen B, Chen H, Jacobs M, Middelkoop V, Mullens S, Thijs I, et al. Preparation of

porous stainless steel hollow fibers by robotic fiber deposition. Journal of Membrane Science 2013;

437:17-24.

[23] Luiten-Olieman MWJ, Raaijmakers MJT, Winnubst L, Bor TC, Wessling M, Nijmeijer A, et

al. Towards a generic method for inorganic porous hollow fibers preparation with shrinkage-

controlled small radial dimensions, applied to Al2O3, Ni, SiC, stainless steel, and YSZ. Journal of

Membrane Science 2012; 407–408:155-63.

[24] Luiten-Olieman MWJ, Raaijmakers MJT, Winnubst L, Wessling M, Nijmeijer A, Benes NE.

Porous stainless steel hollow fibers with shrinkage-controlled small radial dimensions. 2011.

[25] Luiten-Olieman MWJ, Winnubst L, Nijmeijer A, Wessling M, Benes NE. Porous stainless

steel hollow fiber membranes via dry–wet spinning. Journal of Membrane Science 2011; 370:124-

30.

[26] Wan Salleh WN, Ismail AF. Effect of stabilization temperature on gas permeation properties of

carbon hollow fiber membrane. Journal of Applied Polymer Science 2013; 127:2840-6.

[27] Salleh WNW, Ismail AF. Carbon hollow fiber membranes derived from PEI/PVP for gas

separation. Separation and Purification Technology 2011; 80:541-8.

[28] Barbosa-Coutinho E, Salim VMM, Piacsek Borges C. Preparation of carbon hollow fiber

membranes by pyrolysis of polyetherimide. Carbon 2003; 41:1707-14.

[29] Nikonorova AI, Florensova FR. Nitriding of austenitic steels. Met Sci Heat Treat 1965; 7:636-

8.

[30] Reis R, Maliska A, Borges P. Nitrogen surface enrichment of austenitic stainless steel ISO

5832-1. Journal of Materials Science 2011; 46:846-54.

[31] Mariappan R, Kumaran S, Rao TS. Effect of sintering atmosphere on structure and properties

of austeno-ferritic stainless steels. Materials Science and Engineering: A 2009; 517:328-33.

[32] Rawers J, Dunning J, Asai G, Reed R. Characterization of stainless steels melted under high

nitrogen pressure. Metallurgical and Materials Transactions A 1992; 23:2061-8.

Page 210: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

210

[33] Hunter GB, Eagar TW. Ductility of stabilized ferritic stainless steel welds. MTA 1980; 11:213-

8.

[34] Sing KSW, Everett DH, Haul RAW, Moscou L, Pierotti RA, Rouquerol J, et al. Reporting

Physisorption Data for Gas/Solid Systems. Handbook of Heterogeneous Catalysis: Wiley-VCH

Verlag GmbH & Co. KGaA; 2008.

[35] Wu Y, German RM, Blaine D, Marx B, Schlaefer C. Effects of residual carbon content on

sintering shrinkage, microstructure and mechanical properties of injection molded 17-4 PH stainless

steel. Journal of Materials Science 2002; 37:3573-83.

[36] Konopka K, Olszówka-Myalska A, Szafran M. Ceramic–metal composites with an

interpenetrating network. Materials Chemistry and Physics 2003; 81:329-32.

[37] Cheng T, Raj R. Flaw Generation During Constrained Sintering of Metal-Ceramic and Metal–

Glass Multilayer Films. Journal of the American Ceramic Society 1989; 72:1649-55.

[38] Upadhyaya GS. Sintered metallic and ceramic materials: preparation, properties, and

applications. New York: Wiley; 2000.

[39] Prielipp H, Knechtel M, Claussen N, Streiffer SK, Müllejans H, Rühle M, et al. Strength and

fracture toughness of aluminum/alumina composites with interpenetrating networks. Materials

Science and Engineering: A 1995; 197:19-30.

Page 211: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

211

CHAPTER 7

GAS TESTING OF THE COMPOSITE HOLLOW FIBRES

7.1. ABSTRACT

In this chapter CSS, CASS and CA hollow fibres were tested for gas permeation of helium, carbon

dioxide and nitrogen. Gas permeation tests were carried out in dead-end mode for single gas

permeation from 20 to 100oC, and in a continuous open-ended mode for binary gas mixtures from

75 to 150oC. Adsorption of nitrogen and carbon dioxide was investigated to understand the

synergistic effect of the composite material surface on the transport of gases. This work shows that

nitrogen adsorption is not measurable for the CSS hollow fibres, whilst the isosteric heat of

adsorption for carbon dioxide is much higher for the hollow fibres containing only SS as opposed to

alumina particles. Finally, this chapter shows that the CSS hollow fibre separated nitrogen from

carbon dioxide at low feed concentrations of less than 20/80 carbon dioxide to nitrogen. This was

attributed to the strong carbon dioxide heat of adsorption, which controlled the surface diffusion

transport trough the CSS hollow fibre, whilst the non-adsorbable nitrogen permeation could flow

unimpeded through the porous structures. This gas separation behaviour is anomalous for

mesoporous or macroporous materials pore sizes, and only possible due to the synergistic effect of

the SS and carbon dioxide materials.

Page 212: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

212

7.2. INTRODUCTION

This chapter focuses on permeation tests carried out for the carbon stainless steel (CSS) hollow

fibre, carbon alumina stainless steel (CASS) hollow fibre and carbon alumina (CA) hollow fibre.

CSS and CASS hollow fibres were prepared with 6µm particles, and the latter contained a 50vol%

alumina to SS particle ratio. The gas permeation tests were carried out using helium, carbon dioxide

and nitrogen at temperatures of 25, 50 and 100°C, with transmembrane pressures of 50, 100, 150

and 200 kPa.

This chapter is an extension of this thesis, which was initially focused on the fundamental studies of

forming inorganic hollow fibres containing SS particles. This is evidenced in Chapter 2 where the

literature review canvassed published articles in the area of inorganic materials (particularly metals)

and hollow fibre processing techniques followed by the experimental chapters 3 to 6 on the

production and characterisation of the morphological and mechanical properties of the resultant

hollow fibres. As gas permeation involves transport through porous media, this chapter starts with a

short literature review on transport phenomena followed by the experimental work, analysis and

discussion of results for both single gas permeation and binary gas (including simulated power

station flue gas (15% N2 and 85% CO2)) mixture separations.

7.3. TRANSPORT PHENOMENA IN POROUS MEDIA

The major gas transport mechanisms through porous membranes include bulk flow, Knudsen flow,

surface diffusion and activated diffusion [1-5]. When considering macroporous (dp > 50 nm)

membranes the transport of gases is dominated by bulk or viscous flows, as the gas to gas

interactions are much more frequent than gas to pore wall interactions, and the permeance is

determined by:

(7.1)

where FP,0 is the Poiseuille flow (mol·m-2·s-1·Pa-1), εp is the porosity, τ is the tortuosity, η is the gas

viscosity (N s m-2), Pm is the pressure (Pa), L is the thickness (m) of the porous layer, and R is the

Page 213: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

213

gas constant (J mol-1 K-1), T is the absolute temperature (K), r is the pore radius (m) [3]. In this

transport regime the pore sizes are too large to adequately separate gas molecules. Under these

conditions, the selectivity, defined by the ratio of gas flow or permeance generally, is equal to one.

As the pore size decreases the frequency of gas to pore wall collisions increases relative to the

frequency of gas to gas collisions. If the mean free path λ for a molecule is substantially larger than

the pore diameter, then there are far more gas to pore wall collisions than gas to gas collisions and

the flow is characterised by the Knudsen model [6]. The pore size in question here is typically in the

microporous region (dp 2 nm) or the lower end of the mesoporous region (2 dp 50 nm). In this

case the permeance is determined by:

(7.2)

where P is permeance (mol m-2 s-1 Pa-1),ε is porosity, dp is pore diameter (m), δm is membrane

thickness (m), τ is tortuosity, R is the gas constant (J mol-1 K-1), T is temperature (K), and MW is

molecular weight (kg mol-1) [7]. It is interesting to note that the permeance is inversely proportional

to the square root of the temperature. Hence, permeance and or gas flow rates decrease with

temperature. Another property of Knudsen diffusion is that the selectivity, generally called ideal

Knudsen selectivity, is a function of the square root of the molecular weight of gases as show in Eq

7.3. In this case, gas separation is low generally between 2 and 5.

(7.3)

It is important to note that there is a transition region between the viscous and Knudsen regimes.

Pore sizes here are typically at the higher end of the mesoporous (2 dp 50 nm) size range,

although it is highly dependent on the gas in question. Selectivity here is difficult to adequately

characterise as it is typically greater than unity but less than the ideal Knudsen value. Unfortunately,

membranes with defects (i.e. larger holes in otherwise micro or mesoporous membranes) can

present similar transport properties and it is complex to correctly identify the combined transport

mode in each case. Such identification is outside the scope of the thesis.

Another important transport mechanism is surface diffusion which is associated with a gas species

preferentially adsorbing on surface of the porous material. In this case, it may be energetically

Page 214: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

214

preferential for the gas to permeate through the membrane surface via adsorption sites [8]. In other

words, provided that there is a driving force, gas molecules hop from one adsorption site to another.

The flux through a membrane by surface diffusion can be calculated by:

(7.5)

where JS,0 is the surface diffusion flux component (mol s-1 m-2), ρapp is the apparent density (kg m-3),

Ds is the surface diffusion coefficient (m2 s-1), and dq/dl is the surface concentration gradient (mol

kg-1 m-1). In this equation, the surface concentration gradient is a function of temperature. As

adsorption decreases with temperature, surface adsorption is more prevalent at lower temperatures

than at high temperatures. High separation factors can be achieved in this case in gas mixture

separation, as the strongly adsorbed gas may block the pore entrance and otherwise hinder the

diffusion of the weakly adsorbed gas, even if the weakly adsorbed gas is smaller.

Finally, activated transport was reported as a gas transport mechanism by Barrer [9] in 1934.

Activated diffusion occurs when the pore size of the membrane is in the order of Angstroms and the

permeance for smaller gases increases with temperature. Gases with kinetic diameters larger than

the apparent pore size of the membrane usually demonstrate decreased permeance with increasing

temperature. Therefore, the selectivity under activated diffusion increases with temperature [10].

Activation transport is also called molecular sieving, as gas separation occurs via the molecular

size. In activated diffusion, the permeance of gas is driven by the chemical potential of the

concentration gradient across the membrane described by Fick’s law as shown in Eq 7.6 or by the

modified Fick’s law in Eq 7.7

(7.6)

where D is the diffusion coefficient, c is the local concentration and x is the coordinate in the

direction of flux

(7.7)

where Δc is the concentration difference between both membrane surfaces and L is the membrane

thickness [11]. As Henry’s law applies for many micro porous materials, the flux can be rewritten

as Eq 7.8:

Page 215: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

215

(7.8)

considering and the overall apparent activation energy , where D0 is a

temperature independent proportionality constant K0 Henry’s constant at infinite temperature, Em is

the positive mobility energy (i.e. diffusion activation energy), and Qst is the isosteric heat of

adsorption [11]. In the case of activated transport, gas selectivities can reach values in excess of

1000 [12].

The isosteric heat of adsorption (Qst) of each component of a gas mixture is an important

thermodynamic variable to understand the conditions applicable in gas separation [13]. This

property is related to the derivative of the adsorbate enthalpy with respect to the amount adsorbed

[14]. Hence, the amount of heat released due to adsorption results in a diminishing of the adsorption

kinetics [14]. In real terms, the isosteric heat of a pure gas is the difference between the molar

enthalpy in the gas phase and the differential enthalpy in the adsorbed phase [15, 16], which is

given by the van ‘t Hoff equation:

(7.9)

where K is an isosteric gas constant, T is the temperature ΔH is the variation of enthalpy and R is

the universal gas constant. The van ‘t Hoff equation can be re-written as follows to determine Qst,

which is the isosteric heat of adsorption:

(7.10)

The value of K is a function of the isothermal behaviour of adsorption with the change in pressure.

In this sense the most widely used model is the one developed by Langmuir [17] that can be used

for all the isotherm types as follows:

(7.11)

where X is the amount adsorbed in the surface, KL is a constant and p is the pressure. In the case that

there is a linear relation between the amount adsorbed and pressure, this behaviour is consistent

with Henry’s law (equation 7.12):

Page 216: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

216

(7.12)

where, X is the quantity adsorbed, p is pressure, and KH is the Henry’s adsorption constant. For this

special case, when Henry’s law is evident, the value of KL is approximately equal to KH. This results

in equation 7.10 being simplified to be:

(7.13)

This means that the isosteric heat of adsoption can be determined by the slope of the plot between

and ln (KH).

Adsorption is a relevant phenomenon in transport processes across the membrane, and in this sense

knowing the amount of gas adsorbed into each sample may help understand the transport

mechanism governing the flux across the membrane. Adsorption can increase the selectivity of the

membrane by reducing the kinetics of the most adsorptive gas [18, 19].

7.3.1. Adsorption

Figure 7.1 shows that there is a linear relationship between the quantity of CO2 adsorbed and the

increase in pressure for the CSS, CASS and CA hollow fibres, which is consistent with Henry’s

law. This means that equation 7.13 can be used to calculate the isosteric heat of adsorption for CO2.

Whilst, the surface coverage is generally low per mass for all hollow fibres due to the relatively

small surface area examined per sample, CO2 adsorption per surface area (using results from

Chapter 6) follows the order of CSS>CASS>CA.

Page 217: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

217

CSS

Absolute Pressure (Pa)

0 2e+4 4e+4 6e+4 8e+4 1e+5

Quantity

adsorb

ed (

mol g

-1)

0

1e-6

2e-6

3e-6

4e-6CASS

Absolute pressure (Pa)

0 2e+4 4e+4 6e+4 8e+4 1e+5

0

1e-5

2e-5

3e-5

4e-5CA

Absolute pressure (Pa)

0 2e+4 4e+4 6e+4 8e+4 1e+5

1e-5

2e-5

3e-5

4e-5

5e-5

6e-5

0°C 20°C 50°C

a b c

CSS

Absolute pressure (Pa)

0 2e+4 4e+4 6e+4 8e+4 1e+5

Qu

an

tity

ad

so

rbe

d (

mo

l m

-2)

0

2e-7

4e-7

6e-7

8e-7

1e-6

CASS

Absolute pressure (Pa)

0 2e+4 4e+4 6e+4 8e+4 1e+5

0

2e-7

4e-7

6e-7

8e-7

1e-6

CA

Absolute pressure (Pa)

0 2e+4 4e+4 6e+4 8e+4 1e+5

0

2e-7

4e-7

6e-7

8e-7

1e-6

d e f

Figure 7.1 – Adsorption isotherms for CSS (a, d), CASS (b, e) and CA(c, f) for CO2

The Arrhenius plot of the Henry’s constant (lnK) versus T-1 is shown in Figure 7.2 where Qst was

calculated as 34.8, 17.7 and 10.8 kJ mol-1 for the CSS, CASS and CA hollow fibre materials,

respectively.

Page 218: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

218

1000 (RT)-1

0.36 0.38 0.40 0.42 0.44 0.46

ln (

KH)

-24

-22

-20

-18

-16

-14

-12

-10

CSS

CASS

CA

Figure 7.2 – Arrhenius plot of the Henry’s law constant for CSS, CASS and CA hollow fibres for

CO2.

CA

Absolute pressure (Pa)

0 2x104 4x104 6x104 8x104 105

Quantity

adsorb

ed (

mol g

-1)

0

10-5

2x10-5

3x10-5

4x10-5

5x10-5

6x10-5

7x10-5

0°C

16°C

50°C

RT-1

0.00036 0.00038 0.00040 0.00042 0.00044 0.00046

ln (

KH)

-22.8

-22.6

-22.4

-22.2

-22.0

-21.8

-21.6

-21.4

-21.2

-21.0

-20.8

Figure 7.3 – CA hollow fibres (a) nitrogen sorption and (b) Arrhenius plot.

Nitrogen adsorption for the CSS hollow fibre was very low and within the experimental error of the

machine. This is not unexpected; since very low amounts of nitrogen were adsorbed at 77 K (see

Chapter 6, Figure 6.12). For the CASS hollow fibres, small quantities of N2 adsorbed were

measureable at 0°C however at higher temperatures the adsorption decreased to a level that was

within the experimental error of the machine. Hence, nitrogen adsorption for the temperatures of

interest for the CSS and CASS hollow fibres is considered to be negligible. However, the CA

hollow fibres showed measureable quantities of nitrogen adsorbed which are plotted in Figure 7.3a,

though again at low coverage. These results show an almost linear relation and therefore Henry’s

law is suitable. Figure 7.3 b show the Arrhenius plot resulting in an isosteric heat of adsorption of

22.7 kJ mol-1.

Page 219: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

219

The Qst values for CO2 for the three different hollow fibre samples are plotted in Figure 7.4 against

the (a) amount of carbon retained in each hollow fibre, (b) the content of alumina, and (c) the BET

surface area. It is noteworthy that Qst was the highest for the fibres containing SS particles, and the

lowest for the fibres containing alumina, even though the latter had the largest amount of carbon

and surface area. This result is unusual as carbon dioxide generally as a good affinity with carbon

materials, and adsorption tends to increase as a function of the surface area as a general rule.

However, Chapter 6 shows that the CSS hollow fibres had the lowest surface area of 3 m2g-1, which

is 10 times and 53 times lower than the CASS and CA hollow fibres, respectively. This unusual

result seems to suggest that there is a synergy between the SS particles and carbon, which is not

noticeable in the case of alumina particles and carbon. Conversely, the Qst for nitrogen adsorption

for the CSS hollow fibre was not noticeable, though a high value of 22.7 kJ mol-1 was calculated for

the CA hollow fibres. Hence, the SS and alumina particles are playing different roles in the hollow

fibre matrix in terms how they combine with carbon to influence the CO2 and nitrogen adsorption.

Carbon content (%)

0 2 4 6 8 10 12 14

Iso

ste

ric h

ea

t o

f a

dso

rptio

n o

f C

O2 (

kJ m

ol-1

)

0

10

20

30

40

a Alumina content (w%)

0 20 40 60 80 100

Isoste

ric h

eat

of adsorp

tion for

CO

2 (

kJ

mol-1

)

0

10

20

30

40

b BET surface area (m

2g

-1)

0 20 40 60 80 100 120 140 160 180

Isoste

ric h

eat

of adsorp

tion for

CO

2 (

kJ

mol-1

)

0

10

20

30

40

c

Figure 7.4 – Effect of carbon content (a), alumina content (b) and BET surface area (c) on isosteric

heat of adsorption of CO2

7.4. SINGLE GAS PERMEATION FOR THE CSS, CASS AND CA

HOLLOW FIBRES

7.4.1. Modelling

Single gas permeation tests were used as a preliminary step to estimate the performance of hollow

fibres. In this thesis, tests were carried out using a dead-end system as described in Chapter 3,

Page 220: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

220

where permeate flow was calculated from the increase in pressure of a known control volume. In

order to understand the single gas permeation, the flux through a membrane of thickness l is

schematically shown in Figure 7.5. The key assumptions for this system are:

1 The system is at steady state

2 There are no mass losses/accumulation

3 There are no reactions occurring

4 The system is isothermal.

Figure 7.5 – Schematic representation of the flux (J) across a membrane of thickness l.

The conservation of mass across the membrane is given by:

IN – OUT + REACTION – ACCUMULATION = 0 (7.10)

As there are no reactions or mass loss/accumulation, the conservation of mass is re-written as

follows:

IN – OUT = 0 (7.11)

By applying an infinitesimal flux variation from x to Δx, the flux IN is Jx and OUT is Jx+Δx. Hence,

equation 7.11 can be written as:

(7.12)

For an infinitesimal surface when Δx →0, then:

(7.13)

which is the differential form of the flux equation:

Page 221: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

221

(7.14)

which is solved as follows:

(7.15)

As Fick’s law is applicable for gas transport through membranes:

(7.16)

and in the majority of cases, adsorption in inorganic membranes complies with Henry’s law (C = k

p), equation 7.16 can be re-written as:

(7.17)

It is important to note that the product D k is known as permeability P, so re-writing Equation 7.17

and combining with equation 7.15:

(7.18)

The boundary conditions for the system described in Figure 7.5 are:

For x = 0, p(x) = pH

For x=l, p(x) = pL

where pH is the high pressure on the feed side and pL is the low pressure in the permeate side. By

applying the boundary conditions to Equation 7.18

(7.19)

(7.20)

re-arranging Equation 7.20:

(7.21)

Page 222: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

222

As the testing system contains a volume control, then it is important to consider the gas entering the

volume control at x=l. Hence, the flux is given by the number of moles (n) that enters the control

volume as follows:

(7.22)

By considering the ideal gas law (pV=nRT), the flux of gas entering the volume control can be re-

written as:

(7.23)

where A is the surface area of the membrane (m2), t is time (s), n is the number of moles, V is the

known volume control, R is the universal gas constant (m3Pa K−1mol−1) and T is the absolute

temperature (K).

Combining Equation 7.21 and Equation7.23 gives

(7.24)

As this is a transient system, the boundary conditions are:

t = 0 ; p = po

t = t ; p = p(t)

By applying the boundary conditions to Equation 7.24

(7.25)

which gives the following solution:

(7.26)

Finally, rearranging Equation 7.26 gives the permeance parameter (P/l) which can be determined

from the slope of the natural log function of the pressures versus time, as the other parameters (A,

R, T and V) are constant and known.

Page 223: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

223

(7.28)

In order to compare the permeance of different gases in the same membrane an ideal selectivity can

be calculated as follows:

(7.29)

7.4.2. Results

The effective diameter of the hollow fibres was calculated using a logarithmic mean value between

the outer diameter and the inner diameter as described elsewhere for ceramic hollow fibre gas

permeation [20, 21]. Table 7.1 list the surface dimensions for the hollow fibres.

Table 7.1 – Hollow fibre dimensions.

Hollow Fibre Surface Area (m2) Length (m) Outer diameter (m) Inner diameter (m)

CSS 3.7 x 10-4 97 x 10-3 1.61 x 10-3 0.8 x 10-3

CASS 1.59 x 10-4 41 x 10-3 1.93 x 10-3 0.74 x 10-3

CA 1.87 x 10-4 61 x10-3 1.64 x 10-3 0.52 x10-3

Figure 7.6 shows the single gas permeance results for the CSS hollow fibres calculated based on Eq.

7.28 using nitrogen, carbon dioxide and helium at different temperatures and different pressures. A

general trend is that increasing the transmembrane pressure also increased the permeance. Contrary

to this trend, increasing the temperature resulted in the permeance decreasing. However, there are a

few unexpected trends if compared to conventional gas permeance results in porous inorganic

membranes. First, the permeance of nitrogen was similar at 25 and 50°C, despite the marginal

increase in temperature. Second, permeance values for helium and nitrogen gases at pressure

differences of 100 kPa showed scattered results though the values are all within the experimental

error of ±10%.

Page 224: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

224

Helium Nitrogen Carbon Dioxide cba

25 °C 50 °C 100 °C

P (kPa)

80 100 120 140 160 180 200 220 240 260

5.0e-9

1.0e-8

1.5e-8

2.0e-8

2.5e-8

3.0e-8

P (kPa)

80 100 120 140 160 180 200 220 240 260

Perm

eance (

mol m

-2s

-1 P

a-1

)

5.0e-9

1.0e-8

1.5e-8

2.0e-8

2.5e-8

3.0e-8

P (kPa)

80 100 120 140 160 180 200 220 240 260

5.0e-9

1.0e-8

1.5e-8

2.0e-8

2.5e-8

3.0e-8Helium Nitrogen Carbon Dioxide cba

25 °C 50 °C 100 °C

Figure 7.6 – Single gas permeation (± 10%) results for CSS hollow fibre for He (a), N2 (b) and CO2

(c) at transmembrane pressures of 1, 1.5, 2 and 2.5 bar.

The single gas permeance values for the CASS (Figure 7.7) and CA (Figure 7.8) hollow fibres also

follow similar trends for the CSS hollow fibre in Figure 7.6. However, the permeance for the CASS

and CA hollow fibres are three and two orders magnitude higher than that of the CSS hollow fibres,

respectively. Furthermore, the effects of temperature were far less pronounced for both the CASS

and CA hollow fibres when compared to the CSS hollow fibres. These results strongly suggest that

the CSS hollow fibres resulted in a denser matrix than the counterpart CASS and CA hollow fibres.

Page 225: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

225

25 °C 50 °C 100 °C

Helium Nitrogen Carbon Dioxidea b c

P (kPa)

40 60 80 100 120 140 160 180 200 220

Pe

rme

an

ce

(m

ol m

-2 s

-1P

a-1

)

1e-5

2e-5

3e-5

4e-5

5e-5

6e-5

7e-5

P (kPa)

40 60 80 100 120 140 160 180 200 220

1e-5

2e-5

3e-5

4e-5

5e-5

6e-5

7e-5

P (kPa)

40 60 80 100 120 140 160 180 200 220

1e-5

2e-5

3e-5

4e-5

5e-5

6e-5

7e-5

25 °C 50 °C 100 °C

Helium Nitrogen Carbon Dioxidea b c

Figure 7.7 – Single gas permeation (± 10%) results for CASS hollow fibre for He (a), N2 (b) and

CO2 (c) at transmembrane pressures of 1, 1.5, 2 and 2.5 bar.

Helium Nitrogen Carbon Dioxidea b c

P (kPa)

40 60 80 100 120 140 160 180 200 220

Perm

eance (

mol m

-2s

-1P

a-1

)

1.0e-6

1.5e-6

2.0e-6

2.5e-6

3.0e-6

P (kPa)

40 60 80 100 120 140 160 180 200 220

1.0e-6

1.5e-6

2.0e-6

2.5e-6

3.0e-6

P (kPa)

40 60 80 100 120 140 160 180 200 220

1.0e-6

1.5e-6

2.0e-6

2.5e-6

3.0e-6

25 °C 50 °C 100 °C

Helium Nitrogen Carbon Dioxidea b c

Figure 7.8 – Single gas permeation (± 10%) results for CA hollow fibre for He (a), N2 (b) and CO2

(c) at transmembrane pressures of 1, 1.5, 2 and 2.5 bar.

Page 226: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

226

100 kPa 150 kPa 200 kPa 250 kPa Knudsen

a b c

Temperature (°C)

20 40 60 80 100

Sele

ctivi

ty (

(P/l) N

2/(

P/l) C

O2)

1.0

1.5

2.0

2.5

3.0

3.5

4.0

Temperature (°C)

20 40 60 80 100

Sele

ctivi

ty (

(P/l) H

e /

(P

/l) C

O2)

1.5

2.0

2.5

3.0

3.5

4.0

Temperature (°C)

20 40 60 80 100

Sele

ctivi

ty (

(P/l) H

e/(

P/l) N

2)

1.0

1.5

2.0

2.5

3.0

3.5

4.0

100 kPa 150 kPa 200 kPa 250 kPa Knudsen

a b c

Figure 7.9 – Single gas selectivities (± 22%) for the CSS hollow fibres for (a) N2/CO2, (b) He/CO2

and (c) He/N2.

a b c

50 kPa 100 kPa 150 kPa 200 kPa Knudsen selectivity

a b ca b c

Temperature (°C)

20 40 60 80 100

Sele

ctivi

ty (

(P/l) N

2 / (

P/l) C

O2)

0.9

1.0

1.1

1.2

1.3

50 kPa 100 kPa 150 kPa 200 kPa Knudsen selectivity

Temperature (°C)

20 40 60 80 100

Sele

ctivi

ty (

(P/l) H

e/(

P/l) C

O2)

1.0

1.5

2.0

2.5

3.0

3.5

Temperature (°C)

20 40 60 80 100

Sele

ctivi

ty (

(P/l) H

e/(

P/l) N

2)

1.0

1.5

2.0

2.5

3.0

3.5a b c

Figure 7.10 – Single gas selectivities (± 22%) for CASS for N2/CO2 (a), He/CO2 (b), He/N2 (c) for

different pressures.

The selectivities based the ratio of the permeance of gases as a function of temperature and pressure

are shown for the CSS (Figure 7.9), CASS (Figure 7.10) and CA (Figure 7.11) hollow fibres. The

Page 227: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

227

ideal Knudsen selectivity calculated using equation 7.3 is also shown in these figures. The general

trend is that selectivities are generally well below the ideal Knudsen selectivity, particularly for the

CASS and CA hollow fibres, thus suggesting that parallel transport is occurring, either because the

pore size distribution is such that the fibres are in the transition regime, or more likely because

larger defects are present. In either case it is clear that viscous flow is dominant. The only exception

is the CSS hollow fibre which show values above the ideal Knudsen diffusion N2/CO2 of ~1.25.

Also a single point above the ideal Knudsen diffusion for H2/CO2 of ~ 3.32 is observed at 50°C and

100 kPa pressure difference.

a b c

50 kPa 100 kPa 150 kPa 200 kPa Knudsen selectivity

a b c

Temperature (°C)20 40 60 80 100

Sele

ctivi

ty (

(P/l) N

2 / (

P/l) C

O2)

0.8

0.9

1.0

1.1

1.2

1.3

1.4

Temperature (°C)20 40 60 80 100

Sele

ctivi

ty (

(P/l) H

e/(

P/l) N

2)

1.0

1.2

1.4

1.6

1.8

2.0

2.2

2.4

2.6

2.8

Temperature (°C)20 40 60 80 100

Sele

ctivi

ty (

(P/l) H

e/(

P/l) C

O2)

1.0

1.5

2.0

2.5

3.0

3.5

50 kPa 100 kPa 150 kPa 200 kPa Knudsen selectivity

Figure 7.11 – Single gas selectivities (± 22%) for CA for N2/CO2 (a), He/CO2 (b), He/N2 (c) for

different pressures.

7.5. BINARY GAS TEST FOR THE CSS HOLLOW FIBRES

In view of the better performance of the CSS hollow fibres in terms of selectivity of single gas

permeation as compared to the CASS and CA hollow fibres, the CSS hollow fibres were selected

for binary gas testing. By introducing two (or more) gases into the feed side of the membrane, a

more realistic situation of gas separation in an industrial scenario can be understood. The flow rates

were determined by a bubble flow meter and calculated as follows:

Page 228: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

228

(7.30)

Where F is the total flow rate, V is a fix volume and t is time. As the concentration of the gases

were determined by gas chromatography using a Shimadzu GC, the flow rate of a gas species “i”

was calculated by multiplying the total flow rate (F) and gas concentration (c) as follows:

(7.31)

The separation factor (α) for binary gas mixture separation was calculated based on the

concentration of the reciprocal gases in the permeate and retentate streams as follows:

(7.32)

where αx/y is the separation of gas x over y, and x and y are concentrations of the two gases in the

retentate and permeate streams [11, 22].

Figure 7.12 shows permeance results as a function of gas composition and temperature. It is

interesting to note that the permeance values are within the same order of magnitude as those

measured in the single gas experiment (Figure 7.6). In the case of the gas mixture, the permeance of

nitrogen increased when compared to single gas results, contrary to the permeance of CO2 which

was reduced. The results are scattered for N2 permeance, although the variation is small suggesting

that all results are within the experimental variation of ±15% for gas mixture permeation. However,

there is a significant variation of CO2 permeance as a function of the concentration of CO2 in the

feed stream, close to one order of magnitude difference between the permeances for a CO2 feed

concentration of 20 and 10% for the temperature sets of 50 and 75oC.

Page 229: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

229

CO2 Fraction (%)

0 10 20 30 40 50 60

Pe

rme

ance

(m

ol m

-2s

-1P

a-1

)

5.0e-9

1.0e-8

1.5e-8

2.0e-8

2.5e-8

N2 Fraction (%)

405060708090100

5.0e-9

1.0e-8

1.5e-8

2.0e-8

2.5e-8

75°C

100°C

125°C

150°C

Figure 7.12 – Binary gas permeance (±15%) for CO2 (bottom) and N2 (top) as a function of the flue

gas composition for a transmembrane pressure difference of 150 kPa.

CO2 Fraction [%]

0 10 20 30 40 50 60

(P/l

) N2 /

(P

/l) C

O2

0

10

20

30

40

50

N2 Fraction (%)

405060708090100

75°C

100°C

125°C

150°C

CO2 Fraction(%)

0 10 20 30 40 50 60

Se

pa

rati

on

fa

cto

r)

0

2

4

6

8

10

12

14

N2 Fraction (%)

405060708090100

75°C

100°C

125°C

150°C

Figure 7.13 – (a) Separation factor for CO2 and N2 and (b) Ratio between permeance of N2 and

CO2 as function of feed composition, and feed stream temperature

Figure 7.13a shows the binary N2/CO2 separation values ranging between 13.5 and 1.5. It is

noteworthy to observe that the CSS hollow fibre is preferentially permeating nitrogen over carbon

dioxide at a low CO2 feed fraction of 10%, and but this is greatly reduced at 20% (and beyond) for

all temperature points.

Page 230: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

230

Similar binary gas permeation was carried out for He/N2 is and He/CO2. The binary gas separation

factors are displayed in Figure 7.14. The general trend for these binary gas mixtures is that

separation factors were too low independent of temperature and concentration of gases in the feed

fraction. These results suggest that helium permeance is slightly faster than either nitrogen or

carbon dioxide. Although the results are scattered, these are generally within experimental variation

of ±20%.

CO2 Fraction [%]

0 10 20 30 40 50 60

(Separa

tion facto

r)

0.50

0.55

0.60

0.65

0.70

0.75

0.80

N2 Fraction [%]

405060708090100

75°C

100°C

125°C

150°C

He Fraction [%]

0 20 40 60 80 100

(Separa

tion facto

r)

0.55

0.60

0.65

0.70

0.75

0.80

0.85

0.90

CO2 Fraction [%]

020406080100

75°C

100°C

125°C

150°C

Figure 7.14 – Binary gas separation factor (±20%) for (a) He and N2 and (b) He and CO2

Page 231: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

231

7.5.1. Long term testing

In order to determine the stability of the hollow fibres, the CSS hollow fibres were tested in excess

of 1500 hours. Figure 7.15 shows results for helium permeation which were carried out while

testing the membrane for binary separation. This means that for each test point, the membranes

have been previously tested for a binary gas test at four temperature set points of 75, 100, 125 and

150°C. Permeance values around 3x10-8 mol m-2 s-1 Pa-1 were generally found, whilst increasing

temperature lead to a reduction in helium permeance. These results are in line with the helium

permeance displayed in Figure 7.6, and within expected experimental variation. These long term

results strongly suggest that the porous structure of the CSS hollow fibre is mechanically robust and

stable to thermal cycling. In addition, the porous structure was not affected by the temperature

cycling as helium permeance was stable for over 1500 hours.

Testing time (hours)

0 200 400 600 800 1000 1200 1400 1600

He

lium

pe

rme

ance

(m

ol m

-2s

-1P

a-1

)

2e-8

4e-8

6e-8

8e-8

1e-7

25°C

50°C

75°C

100°C

125°C

150°C

Figure 7.15 – Long-time testing results for Helium permeance (±20%).

7.6. DISCUSSION

Figure 7.6 to Figure 7.8 show that the permeance of a single gas through all hollow fibres decreases

with temperature. According to the transport principles discussed in section 7.3, this type of

transport cannot be molecular sieving (i.e. activated transport), and is most likely associated with

viscous flow (Equation 7.1) and / or Knudsen diffusion (Equation 7.2). If the gas selectivities are in

the region of the ideal Knudsen diffusion (Equation 7.3), then the transport of gases could comply

Page 232: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

232

with Knudsen diffusion. However, the gas selectivities for the CASS and CA hollow fibres were

below the ideal Knudsen selectivities (Figure 7.10 and Figure 7.11), this suggest that viscous flow

is dominant and the effective pore sizes are distributed in the meso and macropore regions.

Although it is important to remember that the selectivities are not unitary, which implies parallel

transport must be occurring, but only to a lesser degree. The gas selectivity results for CSS hollow

fibre in Figure 7.9 display a different picture. The N2/CO2 selectivity was consistently above the

ideal Knudsen selectivity, whilst He/CO2 was close with one point above the ideal Knudsen

diffusion. However, the He/N2 selectivity was below the ideal Knudsen selectivity. These results

suggest that percolation pathways for the CSS hollow fibres are controlled by large micropores

and/or small mesopores and that in addition CO2 transport is being affected by adsorption.

Let us first consider the porous structure of the hollow fibres. The permeance of gases via the CASS

and CA hollow are three orders of magnitude higher than the CSS hollow fibres. In principle these

results give a clear indication that the CASS and CA hollow fibres are very porous, whilst the CSS

hollow fibre has a denser structure with low porosity. In fact, these results correlate very well with

the mercury porosimetry results in Chapter 6, where the total porosity decreased in the order of

CASS (~51%)>CA (~34%)>CSS(~24%).

Nevertheless, Chapter 6 also showed different morphological regions of large voids and finger-like

pores, together with sponge-like regions, which warrants further discussion. Figure 7.16a shows the

effect of total porosity on the permeation of single gases. It is observed that there is no linearity, as

the fitted line is likely to intercept the x-axis around 20% total porosity for permeances around 10-

10 mol m-2 s-1 Pa-1, which is close to the minimum detectable permeation level of the experimental

set up. Hence a direct relationship between total porosity and permeance seems unlikely. If we

consider the structure of the hollow fibre, then it is clear why. The first consideration here is that all

gas molecules have also to diffuse via the sponge-like region of the hollow fibres, regardless of the

total porosity. This is a reasonable assumption based on the SEM and TEM images shown in

Chapter 6, as all hollow fibres contained a sponge-like region. Based on this assumption, Figure

7.16b shows the permeance as a function of the porosity associated only with sponge-like region

from Chapter 6. In this case a much better linear fit is observed. Hence, these results strongly

suggest that gas fluxes are controlled via the region with the smaller pore sizes, the sponge-like

region.

Page 233: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

233

Total porosity (%v ol)

20 30 40 50

Perm

eance (

mol m

-2s

-1P

a-1

)

1e-9

1e-8

1e-7

1e-6

1e-5

1e-4

Porosity of the sponge-like region (%vol)

5 10 15 20 25 30 35 40

N2CO2 He

Figure 7.16 – Effect of the (a) total porosity and (b) sponge-like porosity for the all hollow fibre for

gas permeation at 50°C and 150 kPa.

The second consideration is related to the effect of adsorption on the transport of gases through the

hollow fibres. It is known that single gas and multiple gas permeation delivers different results due

to the competitive adsorption [23]. The effect of adsorption can be clearly seen in Figure 7.12 and

Figure 7.13 for gas mixture. This effect was particularly noticeable for binary mixtures containing

carbon dioxide and nitrogen for the CSS hollow fibres. However, this effect was not evident for the

other gas separation mixtures containing helium and carbon dioxide, or helium and nitrogen as

shown in Figure 7.14.

It is interesting to observe in Figure 7.12 that the carbon dioxide permeance greatly reduced as the

feed fraction of CO2 reduced from 20 to 10 mol%. This reduction was almost of one order of

magnitude for low temperatures of 50 and 75oC, though this effect was of lesser significance as

temperatures increased to 100 and 125°C. As the permeation of nitrogen remained almost constant

(within experimental variation), the selectivities ranged from values ~10 to 40 as displayed in

Figure 7.13b. As selectivity is the ratio of gas permeance, and the N2 data is scattered, it is

interesting to show the results as gas purity in the permeate stream which is a preferable industrial

specification [24]. In addition, these results suggest that the CSS hollow fibres have the potential to

separate nitrogen from carbon dioxide in post-combustion carbon capture trains in coal power

Page 234: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

234

stations, where the carbon dioxide fraction is around 15vol% (PCO2=100 to ~150 kPa), in flue gas

temperatures of between 80 to 150°C [25, 26].

CO2 Fraction [%]

0 10 20 30 40 50 60

Purity

of N

2 in p

erm

eate

[m

ol.m

ol-1

]

0.6

0.7

0.8

0.9

1.0

1.1

N2 Fraction [%]

405060708090100

75°C

100°C

125°C

150°C

Figure 7.17 – Purity of N2 in the permeate stream (±20%)

The purity of N2 in the permeate stream as a function of the composition of binary gas mixture in

the feed stream is shown in Figure 7.17. These results clearly show that the CSS hollow fibre is

preferentially permeating nitrogen over carbon dioxide, concentrating CO2 in the retentate and

returning a more pure nitrogen permeate. Indeed, high purity nitrogen (~99.5 mol%) was obtained

in the permeate for a feed stream containing 88/12 nitrogen to carbon dioxide ratio at temperatures

from 75 to 125oC, though the nitrogen purity slightly reduced to 95.5 mol% at 150oC. As the feed

composition changed to 80/20, the nitrogen permeate purity reached 90-95 mol%. Further dilution

to 46/54 resulted in nitrogen purity of 64-80 mol%. The results are scattered within the

experimental variation of ±10%, though the trend in nitrogen purity is clearly observed as a function

of the feed concentration in the binary mixture.

It is apparent that carbon dioxide has an effect in gas permeation through the CSS hollow fibres.

This is further observed when the N2/CO2 selectivities for single gas (Figure 7.9a) and gas mixtures

(Figure 7.13b) tend to have similar trends. Further evidence on this point can be seen for the

permeation of helium, nitrogen and carbon dioxide as a function of time in Figure 7.18. Whilst

helium and nitrogen reached steady state permeation since the start of the experimental work,

steady state permeation for carbon dioxide was achieved after 30 minutes only. It is interesting to

observe that even at 100°C temperature testing, the reduction of carbon dioxide reached fourfold.

Page 235: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

235

These results strongly suggest that surface diffusion of carbon dioxide (Equation 7.5) is an

important transport mechanism for the CSS membranes.

Time (min)

0 20 40 60 80 100 120 140 160

Pe

rme

ance

(m

ol m

-2s

-1P

a-1

)

1e-9

1e-8

1e-7 CO2

N2

He

Figure 7.18 – Permeance (±10%) of He, CO2 and N2 as a function of time at 100oC.

It is noteworthy to observe that carbon dioxide adsorption per surface area was the largest for the

CSS hollow fibres, (Figure 7.1d), whilst nitrogen adsorption was not measurable suggesting that

nitrogen is a non-adsorbable gas for the CSS hollow fibres for temperatures above 0oC.

Nevertheless, the strong Qst of carbon dioxide of 38 kJ mol-1 strongly suggests that carbon dioxide

molecules once adsorbed require higher energy to diffuse along the porous surface and desorb into

the permeate. As nitrogen did not show measurable adsorption, it is assumed that nitrogen could

diffuse through the membrane pores more freely. To further explain the interesting results of gas

separation for N2/CO2 mixtures below 80/20, a schematic of the transport phenomena is displayed

in Figure 7.19. The strong carbon dioxide adsorption and surface diffusion at low CO2 feed

concentration allowed for the non-adsorbed N2 gas to diffuse freely through the porous structures of

the CSS hollow fibres. Therefore, carbon dioxide permeation was controlled by the low surface

diffusion mechanism whilst nitrogen was controlled by the faster viscous flow and/or Knudsen

diffusion mechanisms.

Page 236: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

236

Figure 7.19 – Schematics showing the effect that adsorption has on selectivity.

A further interesting point is the synergic effect of the composite materials. The results strongly

suggest that there is strong relationship between SS particles and carbon; however, by adding

alumina particles to form the composite CA hollow fibre, nitrogen adsorption became measurable,

delivering a higher Qst value for nitrogen than carbon dioxide. It could be argued that the special

effect observed with gas mixture separation for the CSS hollow fibre could be allocated to pore size

change or closure. However, the long term test for over 1500 hours clearly indicate that the

permeation of helium was almost constant thus suggesting that pore size variation was negligible.

This synergistic effect can also be related to the structure of the CSS hollow fibres, which resulted

in a denser structure than the CASS and CA hollow fibres. The SEM images clearly show in

Chapter 6 that the CSS hollow fibre is mainly sponge-like at both inner and outer shells, whilst have

voids in the central part of the hollow fibre wall. Further, the CSS hollow fibres did not have finger-

like structures as seen with small alumina particle addition. Therefore, the dense structure with a

high fraction of sponge-like region, coupled with the adsorption effect of SS and carbon,

contributed to the overall performance of the CSS membrane for N2/CO2 gas mixture separation.

Finally, it is important to compare the permeance results with values available in the literature. The

permeance for nitrogen in CSS hollow fibres (~10-8 mol m-2s-1Pa-1) was smaller than those reported

by Luiten-Olieman [27] for SS hollow fibres (~10-4 mol m-2s-1Pa-1 ). However, the SS hollow fibre

results (~10-5 mol m-2s-1Pa-1) presented in chapter 5 are in the same range as those reported by

Luiten-Olieman for samples sintered at high temperatures of 1200°C. Again, this is attributed to the

reduction of pore sizes. Liu et al. [28] reported nitrogen permeances for alumina hollow fibres being

Page 237: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

237

in the range of 10-5 mol m-2 s-1 Pa-1 whilst the carbon hollow fibres prepared by Ismail and David

[29] delivered nitrogen permeances ~10-10 mol m-2 s-1 Pa-1. The hollow fibres prepared in this thesis

deliver gas permeances in the range of those reported in the literature.

7.7. CONCLUSION

This chapter focuses on the testing of the CSS, CASS and CA composite hollow fibres for single

gas and binary gas mixture separation. Carbon dioxide and nitrogen adsorption tests showed low

coverage for all hollow fibres. However, nitrogen adsorption was not measurable for the hollow

fibres containing SS particles, but was observable for the hollow fibres prepared with alumina. By

comparison carbon dioxide demonstrated measurable adsorption for all hollow fibre types. The

isosteric heat of adsorption for CO2 followed the order of CSS>CASS>CA hollow fibres for values

at 34.8, 17.7 and 10.8 KJ mol-1, respectively. The isosteric heat of adsorption for nitrogen was only

measurable and calculated for the CA hollow fibre at 22.7 kJ mol-1.

Single gas permeation tests show essentially no gas separation for all hollow fibres with fluxes

generally complying with a viscous flow mechanism, as permeance decreased with temperature and

selectivities were close to unity. An exception was the N2/CO2 results which were marginally above

the ideal Knudsen diffusion for the CSS hollow fibre, thus suggesting that the flux for these gases

followed the Knudsen diffusion mechanism. Single gas permeation tests followed the order of

CA>CASS>CSS hollow fibres, as the permeation of the latter was two and three orders of

magnitude lower than the CASS and CA hollow fibres, respectively. Therefore, the CSS hollow

fibres were denser than the other hollow fibres, in line with the porosity results from Chapter 6.

Nevertheless, a major finding of this chapter is that the CSS hollow fibre could separate nitrogen

from carbon dioxide at low feed concentrations of less than 20/80 carbon dioxide to nitrogen. This

is relevant as this gas composition is in the range of flue gas compositions from coal power stations.

This was attributed to the strong carbon dioxide heat of adsorption, which controlled the surface

diffusion transport trough the CSS hollow fibre, whilst the non-adsorbable nitrogen permeation

could flow unimpeded through the porous structures. Therefore, the synergistic effect of the CSS

hollow fibre structure, coupled with carbon dioxide adsorption, resulted in the unusual gas

separation under the tested conditions, even though the pore sizes were sufficiently large to deliver

any significant separation values for other gas pairs.

Page 238: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

238

7.8. REFERENCES

[1] Ockwig NW, Nenoff TM. Membranes for Hydrogen Separation. Chemical Reviews 2007;

107:4078-110.

[2] Burggraaf AJ, Cot L. Fundamentals of Inorganic Membrane Science and Technology.

Burlington: Elsevier Science; 1996.

[3] de Lange RSA, Hekkink JHA, Keizer K, Burggraaf AJ. Permeation and separation studies on

microporous sol-gel modified ceramic membranes. Microporous Materials 1995; 4:169-86.

[4] de Lange RSA, Keizer K, Burggraaf AJ. Analysis and theory of gas transport in microporous

sol-gel derived ceramic membranes. Journal of Membrane Science 1995; 104:81-100.

[5] Diniz da Costa JC, Rudolph V, Lu GQ. Molecular Sieve Silica Membranes. Encyclopedia of

Nanoscience and Nanotechnology 2004; 5:723-41.

[6] Bhatia SK, Bonilla MR, Nicholson D. Molecular transport in nanopores: a theoretical

perspective. Physical chemistry chemical physics: PCCP 2011; 13:15350-83.

[7] Knudsen M. Die Gesetze der Molekularströmung und der inneren Reibungsströmung der Gase

durch Röhren. Annalen der Physik 1909; 333:75-130.

[8] Burggraaf AJ. Single gas permeation of thin zeolite (MFI) membranes: theory and analysis of

experimental observations. Journal of Membrane Science 1999; 155:45-65.

[9] Barrer RM. 89. The mechanism of activated diffusion through silica glass. Journal of the

Chemical Society (Resumed) 1934:378-86.

[10] Thornton AW, Hilder T, Hill AJ, Hill JM. Predicting gas diffusion regime within pores of

different size, shape and composition. Journal of Membrane Science 2009; 336:101-8.

[11] de Vos RM, Verweij H. Improved performance of silica membranes for gas separation. Journal

of Membrane Science 1998; 143:37-51.

Page 239: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

239

[12] Uhlmann D, Liu S, Ladewig BP, Diniz da Costa JC. Cobalt-doped silica membranes for gas

separation. Journal of Membrane Science 2009; 326:316-21.

[13] Mohr R, Rao MB. Isosteric Heat of Adsorption:  Theory and Experiment. The Journal of

Physical Chemistry B 1999; 103:6539-46.

[14] Duong DD. Adsorption analysis: equilibria and kinetics. London: Imperial College Press;

1998.

[15] Sircar S. Excess properties and thermodynamics of multicomponent gas adsorption. Journal of

the Chemical Society, Faraday Transactions 1: Physical Chemistry in Condensed Phases 1985;

81:1527-40.

[16] Shen D, Bülow M, Siperstein F, Engelhard M, Myers A. Comparison of Experimental

Techniques for Measuring Isosteric Heat of Adsorption. Adsorption 2000; 6:275-86.

[17] Langmuir I. THE ADSORPTION OF GASES ON PLANE SURFACES OF GLASS, MICA

AND PLATINUM. Journal of the American Chemical Society 1918; 40:1361-403.

[18] Brunauer S, Emmett PH, Teller E. Adsorption of Gases in Multimolecular Layers. Journal of

the American Chemical Society 1938; 60:309-19.

[19] Rao MB, Sircar S. Nanoporous carbon membrane for gas separation. Gas Separation &

Purification 1993; 7:279-84.

[20] Liu S, Gavalas GR. Oxygen selective ceramic hollow fiber membranes. Journal of Membrane

Science 2005; 246:103-8.

[21] Leo A, Liu S, Diniz da Costa JC. The enhancement of oxygen flux on Ba0.5Sr0.5Co0.8Fe0.2O3-δ

(BSCF) hollow fibers using silver surface modification. Journal of Membrane Science 2009;

340:148-53.

[22] Brunetti A, Barbieri G, Drioli E, Lee KH, Sea B, Lee DW. WGS reaction in a membrane

reactor using a porous stainless steel supported silica membrane. Chemical Engineering and

Processing 2007; 46:119-26.

Page 240: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

240

[23] Battersby S, Smart S, Ladewig B, Liu S, Duke MC, Rudolph V, et al. Hydrothermal stability of

cobalt silica membranes in a water gas shift membrane reactor. Separation and Purification

Technology 2009; 66:299-305.

[24] Yacou C, Smart S, Diniz da Costa JC. Long term performance cobalt oxide silica membrane

module for high temperature H2 separation. Energy & Environmental Science 2012; 5:5820-32.

[25] Aaron D, Tsouris C. Separation of CO2 from Flue Gas: A Review. Separation Science and

Technology 2005; 40:321-48.

[26] Brands K, Uhlmann D, Smart S, Bram M, da Costa JCD. Long-term flue gas exposure effects

of silica membranes on porous steel substrate. Journal of Membrane Science 2010; 359:110-4.

[27] Luiten-Olieman MWJ, Winnubst L, Nijmeijer A, Wessling M, Benes NE. Porous stainless

steel hollow fiber membranes via dry–wet spinning. Journal of Membrane Science 2011; 370:124-

30.

[28] Liu S, Li K, Hughes R. Preparation of porous aluminium oxide (Al2O3) hollow fibre

membranes by a combined phase-inversion and sintering method. Ceramics International 2003;

29:875-81.

[29] Ismail AF, David LIB. A review on the latest development of carbon membranes for gas

separation. Journal of Membrane Science 2001; 193:1-18.

Page 241: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

241

CHAPTER 8

CONCLUSIONS AND RECOMMENDATIONS

8.1. CONCLUSIONS

In this work SS hollow fibres and composite hollow fibres were produced using a combined phase

inversion technique and a sintering / pyrolysis process. Initially this work focused on the effect that

the addition of particles has on the morphology obtained. It was found that the addition of particles

to the spinning dope decreased the kinetics of the phase inversion process, whilst the effect on

kinetics was influenced to a large degree by the particle size and to a lesser degree by the particle

loading. Higher solid loading resulted in a decrease in the voids between the particles thus

diminishing the area available for solvent / non-solvent exchange. This resulted in a reduction on

the kinetics, and producing a larger region with sponge-like pores. A similar effect was noticed

when smaller particles were used. The resistance introduced due to the presence of particles

increased with the decrease in particle size, leading to the reduction of the velocity of exchange

between solvent and non-solvent of 98% when the particle size was reduced from 45 to 6 µm.

In the second part of this work it was shown that the sintering process of hollow fibres did mimic

the morphology produced during phase inversion. This sintering morphological effect was

attributed to inter-particle surface diffusion limitations associated with the production of porous

bodies. The mass transfer diffusion coefficients for the SS particles shifted from ~10-16 to ~10-10 m2

s-1 as the temperature increased from 950 to 1100 °C, respectively, leading to the faster

densification of the porous SS hollow fibre matrix. It was found that the particle size influenced the

mechanical properties and morphology of the sintered hollow fibre significantly. Smaller SS

particles favoured the densification as the mass transfer kinetics was enhanced, since the

densification rate is inversely proportional to the particle size, resulting in a stronger hollow fibre

with more tailored pore sizes in detriment to loss of porosity.

Page 242: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

242

The third part of this work dealt with a novel process to produce carbon stainless steel (CSS),

carbon alumina stainless steel (CASS), and carbon alumina (CA) hollow fibres. This was achieved

by exposing the green hollow fibre to an inert atmosphere during sintering, resulting in the

pyrolysis of the polymeric binder. The carbon retained within the pores between the SS particles

provided porous pore filling to the composite hollow fibre, resulting in a binary pore size

distribution with the reduction in average pore sizes at the lower range. The addition of smaller

alumina particles to the matrix further helped to reduce the pore sizes at the lower range by

allowing the formation of CA composite in the inter-particle spaces created by the SS particles.

In the last part of this work the CSS, CASS and CA hollow fibres were tested for single gas

permeation showing that the transport mechanism through the membrane was mostly in the

transition region between viscous flow and Knudsen diffusion, except for the CSS hollow fibre

which delivered flow in the Knudsen regime. It was noteworthy to observe that the CSS hollow

fibres could adsorb carbon dioxide but not nitrogen. Further, by adding alumina, the hollow fibres

then adsorbed nitrogen. In a binary gas separation test, the CSS hollow fibre was the only produced

hollow fibre capable of separating nitrogen from carbon dioxide for gas mixtures containing CO2

20 mol%. Although the CSS hollow fibre morphology constituted of meso and macropores, in

addition to large voids, this first-time observed phenomenon was attributed to the strong carbon

dioxide isosteric heat of adsorption during gas permeation. Therefore, carbon dioxide transport was

dominated by slow surface diffusion, whilst the non-adsorbing nitrogen could permeate unimpeded

through the CSS hollow fibre matrix, a promising technology for post carbon capture from flue

gases from coal-fired power plants.

8.2. RECOMMENDATIONS

Although this work was clearly focused on the morphology and mechanical properties of SS and

composite hollow fibres, the adsorption and permeation tests also showed interesting outcomes

which merits further research. Therefore, I would like to suggest the following recommendations

for future work:

Long term testing of the CSS hollow fibres for processing real industrial flue gas,

particularly on mixtures containing fly ash particles and water vapour. This will allow for

performance investigations of the CSS hollow fibres under industrial operating conditions.

Page 243: Morphological, mechanical and gas transport properties of ..."Fabrication of Nanostructured TiO2 Hollow Fiber Photocatalytic Membrane and Application for Wastewater Treatment ." Chemical

243

There is an array of solvent – polymer – non-solvent system which may warrant further

research, particularly to obtain different morphologies, in addition to varying the carbon

content and sintering pyrolysis process.

Further investigation should be carried out to understand the synergistic effect of carbon and

SS on gas adsorption, which differs from carbon and alumina.